Nomenclature
- NN
-
neural network
- LOS
-
line of sight
- API
-
adding a power integrator
- FTO
-
finite-time observer
- ERF
-
earth reference frame
- BRF
-
body-fixed reference frame
- CV
-
centre of volume
- AUV
-
autonomous underwater vehicle
- EL
-
Euler–Lagrange
- CFB
-
command-filter backstepping
- IAE
-
integrated absolute error
- ITAE
-
integrated time absolute error
- MIAC
-
mean integrated absolute control
- $ \mathbb{R}_{>0} $
-
set of positive real numbers
- $ I_{n} $
-
identity matrix of size $n$
- $ \mathbb{R}^{n} $
-
$n$ -dimensional Euclidean space
- $ o_{g}x_{g}y_{g}z_{g} $
-
earth reference frame (ERF)
- $ ox_{b}y_{b}z_{b} $
-
body-fixed reference frame (BRF)
- $ {[x,\,y]} $
-
positions of the CV in ERF
- $ \psi $
-
yaw attitude in BRF
- $ {[u,\,v,\,r]} $
-
velocities in BRF
- $ m_{i} $
-
mass, $ i=u,v,r $
- $ d_{i} $
-
damping term, $ i=u,v,r $
- $ \tau_{i} $
-
control input, $ i=u,r $
- $ \delta_{i} $
-
system uncertainties and external disturbances, $ i=u,v,r $
- $ \delta^{'}_{i} $
-
cross-coupling term, $ i=u,v,r $
- $ m^{'}_{i} $
-
uncertain part of $ m_{i}, i=u,v,r $
- $ d^{'}_{i} $
-
uncertain part of $ d_{i}, i=u,v,r $
- $ \delta_{dis_{i}} $
-
wind resistance, $ i=u,v,r $
- $ {[x_{d},\,y_{d}]}$
-
reference trajectory
1.0 Introduction
Recently, the stratosphere has received considerable attention from the modern aviation and aerospace industry that endeavors to exploit its stable atmospheric conditions [Reference Dancila and Botez1–Reference Yang, Yang and Deng3]. This dramatically accelerates the development of the long-dwell stratospheric airship [Reference Manikandan and Pant4], which is a typical lighter-than-air aircraft that plays various roles from telecommunication to space-like observation outpost, similar to satellites [Reference Agrawal, Gobiha and Sinha5–Reference Yang, Xu, Zhang, Zheng and Wang8]. To accomplish diverse mission objectives, driving the airship to reach and follow a time parameterised reference route, also termed as trajectory tracking control, is the most fundamental flight control task [Reference Han, Wang, Chen and Duan9–Reference Yang11]. However, high nonlinearities, strong couplings, modeling inaccuracies, and unpredictable disturbances render the trajectory tracking control design quite intractable.
To date, many powerful control methodologies have been applied to solve this problem, such as adaptive control [Reference Meng, Zhao and Yu12], backstepping method [Reference Han, Wang, Chen and Duan9, Reference Sabiha, Kamel, Said and Hussein13], and sliding mode control [Reference Xiao, Han, Wang, Zhou and Duan14]. Taking several kinds of uncertainty into consideration, Xiao et al. [Reference Xiao, Han, Wang, Zhou and Duan14] proposed an adaptive integral sliding mode controller for an airship. However, no effective modification technique was provided to eliminate control chattering, thus yielding its implementation impossible. In other works [Reference Zheng, Guan, Ma and Zhu6, Reference Wu, Wang, Duan, Xie and Wei10, Reference Yang15–Reference Zheng and Xie17], the unmodeled dynamics and external disturbances were identified and compensated by neural networks (NNs) or fuzzy logic systems (FLSs). However, the employment of NNs and FLSs will inevitably make the controller computationally expensive due to their inherent attributes. Furthermore, it should be emphasised that these controllers only apply to fully actuated airships, which, in general, cannot guarantee the tracking behaviour of underactuated ones.
In reality, most of the potential application scenarios of stratospheric airships are always associated with horizontal motion, and the stratospheric airship can automatically maintain the cruise altitude via an independent lift adjustment system alone [Reference Yang, Yang and Deng3, Reference Yang15, Reference Zheng and Xie17, Reference Yan, Weidong, Yugeng and Zili18]. Given this fact, this work focuses on the horizontal trajectory tracking design for one kind of airship. At present, several challenging issues concerning this subject are still open, three of which are discussed in this paper. The first one is the underactuated problem. The conventional teardrop-shaped airship, operating at a proper flight altitude, is a typically underactuated system [Reference Zheng, Guan, Ma and Zhu6, Reference Zheng and Xie17], primarily due to the non-existence of an independent actuator producing the lateral force to command the sway dynamics. This poses new challenges as the lateral underactuation imposes a non-integrable restriction on the acceleration of the airship, and therefore it has become an important topic of research [Reference Yang, Yang and Deng3, Reference Yamada, Adachi and Funahashi19–Reference Yang, Wu and Zheng22]. Toward underactuated airships and other types of underactuated vehicles, several controllers have been proposed, both of which have great reference values for us, including the waypoint navigation method [Reference Azinheira, Moutinho and De Paiva23], the transverse function control [Reference Dai, He and Lin20], and the line of sight (LOS) approach [Reference Zheng, Guan, Ma and Zhu6, Reference Wu, Wang, Duan, Xie and Wei10, Reference Zheng and Xie17, Reference Jia, Hu and Zhang21]. The controller grounded on the waypoint navigation method [Reference Azinheira, Moutinho and De Paiva23] demands pre-planning an optimal course that aligns with the direction of the wind, which may be suitable for hovering control rather than trajectory tracking control. The transverse function control also needs dynamic extension [Reference Dai, He and Lin20] to accommodate the lateral underactuation, which admittedly complicates the plant model. Although the LOS approach is deemed an efficient guidance law for underactuated vehicles, LOS-based trajectory tracking controllers require confining the tracking error of yaw angle $\psi_{e}$ to the interval $\left( -\frac{\pi}{2},\frac{\pi}{2}\right) $ for $\forall t\ge0$ . Generally, the error-constrained problem is unavoidably linked to rather complex nonlinear mappings. Some prime examples can be found in the works of Jia et al. [Reference Jia, Hu and Zhang21] and Wu et al. [Reference Wu, Wang, Duan, Xie and Wei10], both of which introduced barrier functions and error-dependent transformations to meet such a restrictive condition.
The second issue is about the performance specification. Note that some previous designs [Reference Zheng, Guan, Ma and Zhu6, Reference Han, Wang, Chen and Duan9, Reference Xiao, Han, Wang, Zhou and Duan14, Reference Yang15, Reference Yan, Weidong, Yugeng and Zili18, Reference Azinheira, Moutinho and De Paiva23] can only ensure the equilibrium point of error dynamics is of asymptotic stability, namely, that the settling time, a critical indicator in control design for airships, is infinity. To this end, the finite-time stability and stabilisation theory was established [Reference Meng, Zhao and Yu12, Reference Anjum and Guo24–Reference Rocha, Castaños and Moreno26], permitting a bounded convergence time. Until now, the finite-time control has been an active research area because of its good control qualities, such as the faster decay rate, higher tracking accuracy, and better disturbance rejection property, and has been rapidly applied in various fields, including strict-feedback or nonstrict-feedback nonlinear systems [Reference Sun, Liu, Qiu and Feng16, Reference Zhang, Tong and Li27, Reference Huang, Wang, Wang and He28], uncertain manipulators [Reference Shen, Tang, Hu, Guo, Li and Zhang29], and rigid spacecrafts [Reference Du, Zhang, Wu, Zhu, Li and Chu30, Reference Sun, Hou, Zong and Yu31]. Nevertheless, own to the coupled nonlinearities in the kinematic equation of airships, extending these finite-time controllers to the trajectory tracking control design for airships is nontrivial, especially for underactuated ones. For fully actuated airships, several finite-time trajectory tracking or path following control algorithms were constructed [Reference Wu, Wang, Duan, Xie and Wei10, Reference Zheng and Xie17], where the hard computation of time derivatives of virtual control laws was obviated through filter tools. Though the filter can avoid repeated differentiation, it still structurally increases the complexity of control systems. Attributed to the adding a power integrator (API) method, a range of finite-time control strategies were formulated for airships and other mechanical systems without filters [Reference Du, Zhang, Wu, Zhu, Li and Chu30–Reference Fu, Wang and Wang35]. It should be noted that the power terms used in these API-based methods are strictly constrained to be an even integer or a ratio of two odd integers. Furthermore, in the above finite-time controllers, the prior designation of the size of the residual set is infeasible in that the steady-state accuracy counts on some unknowable model parameters and uncertainty bounds.
The third issue is related to the output feedback. Careful reviews of the above results reveal that most entail full state measurement. However, such a demand is hardly guaranteed in some practical applications. For example, the velocity information of airships cannot always be available at each instant coming from considerations of sensor faults. Inspired by this observation, some notable works used high-gain observers [Reference Jia, Hu and Zhang21], sliding model observers [Reference Wu, Wang, Duan, Xie and Wei10], and fuzzy observers [Reference Zhang, Tong and Li27] to achieve output-feedback control. Although the maturity of the current output-feedback control, the finite-time trajectory tracking control design for airships subject to underactuated dynamics, unmeasured velocities, modeling imprecisions, and external disturbances is still a challenging control problem that needs more in-depth research.
Motivated by the above discussion, this paper proposes a novel approach by a combined application of the backstepping method, the API technique, and the idea of coordinate transformation. Our contributions are as follows:
-
1. Compared to the asymptotic control algorithms [Reference Zheng, Guan, Ma and Zhu6, Reference Han, Wang, Chen and Duan9, Reference Xiao, Han, Wang, Zhou and Duan14, Reference Yang15, Reference Yan, Weidong, Yugeng and Zili18, Reference Azinheira, Moutinho and De Paiva23], our approach allows tracking behaviour to be preassigned by the operator, i.e. it drives the position errors of the airship into a preset range near zero within a finite settling time.
-
2. By employing some useful lemmas, we relax the strong constraints placed on the power terms [Reference Du, Zhang, Wu, Zhu, Li and Chu30–Reference Fu, Wang and Wang35], broadening the set of possible design parameters. Moreover, unlike the sliding mode controllers [Reference Yang11, Reference Xiao, Han, Wang, Zhou and Duan14, Reference Yang15], our control signal is continuous and chattering-free.
-
3. A coordinate transformation is performed herein. Consequently, this work forsakes the extra dynamics needed in the transverse function control [Reference Dai, He and Lin20], and lifts the restriction of LOS-based controllers [Reference Wu, Wang, Duan, Xie and Wei10, Reference Jia, Hu and Zhang21]. Furthermore, the presented controller can be extended easily to a wide range of second-order mechanical systems as its design counts on a transformed equivalent model.
-
4. Our method is structurally less demanding; no tools for filtering [Reference Zheng, Guan, Ma and Zhu6, Reference Han, Wang, Chen and Duan9, Reference Xiao, Han, Wang, Zhou and Duan14, Reference Yang15, Reference Yan, Weidong, Yugeng and Zili18, Reference Azinheira, Moutinho and De Paiva23] are involved, and no arduous computation of analytic differentiation required in the backstepping technique [Reference Dai, He and Lin20] is performed. Furthermore, this work realises velocity-free control and is robust enough in that it establishes a finite-time observer (FTO) to reconstruct unmeasured velocities and unpredictable uncertainties in a integral manner. In contrast to NN or FLS approximation [Reference Zheng, Guan, Ma and Zhu6, Reference Wu, Wang, Duan, Xie and Wei10, Reference Yang15–Reference Zheng and Xie17, Reference Zhang, Tong and Li27], the FTO can sharply lighten the calculational burden, making it particularly appealing for control applications.
Section 2 presents the preliminaries and control objective. Section 3 elucidates the coordinate conversion to cope with lateral underactuation. The major control design procedure and Lyapunov analysis are given in Section 4. Section 5 delineates the simulation results. Section 6 concludes this brief.
2.0 Preliminaries and problem formulation
2.1 Preliminaries
The notation $\mathbb R_{>0}$ is referred to the set of positive real numbers, and the notation $I_{n}\in \mathbb R^{n\times n}$ represents the unit matrix. $|{\cdot}|$ is the absolute value of a scalar, while $\left\|\cdot\right\|$ is the Euclidean 2-norm of a vector or the induced 2-norm of a matrix. Given $\imath>0$ and $\ell=\left[ \ell_{1},\ell_{2},\ldots,\ell_{n}\right]^{T}\in\mathbb R^{n}$ , $|\ell|$ , $|\ell|^{\imath}$ , and $\lfloor \ell \rceil$ refer to $|\ell|=\left[ |\ell_{1}|,|\ell_{2}|,\ldots,|\ell_{n}|\right]^{T}$ , $|\ell|^{\imath}=\left[ |\ell_{1}|^{\imath},|\ell_{2}|^{\imath},\ldots,|\ell_{n}|^{\imath}\right]^{T}$ , and $\lfloor \ell \rceil=diag\left\lbrace \ell_{1},\ell_{2},\ldots,\ell_{n}\right\rbrace $ , respectively. Denote $sig^{\imath}\!\left(\ell\right)=\left[ sig^{\imath}\!\left(\ell_{1}\right),sig^{\imath}\!\left(\ell_{2}\right),\ldots,\right.$ $\left.sig^{\imath}\!\left(\ell_{n}\right)\right]^{T}$ , where $sig^{\imath}\!\left(\ell_{i}\right)=|\ell_{i}|^{\imath}sgn\!\left( \ell_{i}\right)\left(i=1,\ldots,n\right)$ , and $sgn\!\left( \cdot\right)$ is the standard signum function given by
At this stage, we provide some useful definitions and lemmas used later.
Consider the dynamical system
where x is a state vector, the time variable t varies from 0 to $\infty$ , $\mathbb U_{0}$ is a finite open set containing the origin $x=0$ , and $f({\cdot})\,:\,\mathbb R^{n}\to\mathbb R^{n}$ , well-defined on $\mathbb U_{0}$ , is a continuous differentiable nonlinear vector function.
Definition 1 (see the work of Sun et al. [Reference Sun, Liu, Qiu and Feng16]). If the equilibrium point $x=0$ of system (2) is referred to as a (locally) asymptotic stable node and for any initial state $x_{0}\in \mathbb U_{0} $ , there exist $\varepsilon\in\mathbb R_{>0}$ and a settling time function $T\!\left(\varepsilon,x_{0}\right)< \infty$ such that $\|x(t)\|\le\varepsilon, \forall t>T\!\left(\varepsilon,x_{0}\right)$ , then it is true that system (2) has a (locally) finite-time stable equilibrium point at $x=0$ . Furthermore, if $\mathbb U_{0}=\mathbb R^{n}$ , then $x=0$ is globally finite-time stable.
Lemma 1 (see the work of Sun et al. [Reference Sun, Liu, Qiu and Feng16]). Suppose there exists a Lyapunov function $V\!\left(x\right) $ defined in domain $\mathbb U_{0}$ , and the time derivative of $V\!\left(x\right) $ along the trajectory of system (2) satisfies
where $\left\lbrace q, p\right\rbrace\in\mathbb R_{>0}$ , and $0<g<1$ , then the system (2) is finite-time stable.
Lemma 2 (see the work of Sun et al. [Reference Sun, Liu, Qiu and Feng16]). For any $\left\lbrace x,y \right\rbrace\in\mathbb R $ , the following inequality holds:
where $\left\lbrace m,n,s \right\rbrace\in \mathbb R_{>0}$ ,
Lemma 3 (see the work of Zheng et al. [Reference Zheng, Feroskhan and Sun34]). Let $\xi_{i}\in \mathbb R$ , $i=1,2,\ldots,n$ . Then
Lemma 4 (see the work of Du et al. [Reference Du, Zhang, Wu, Zhu, Li and Chu30]). If $\vartheta_{1}>0$ and $0<\vartheta_{2}\le1$ , then
Lemma 5 (see the work of Du et al. [Reference Du, Zhang, Wu, Zhu, Li and Chu30]). For any $\varsigma\in \mathbb R_{+}$ and $z\in \mathbb R$ , we have
2.2 Airship model
Figure 1 displays the stratospheric airship with a typical streamline ballonet. The helium-filled ballonet generates an upward lift for the airship. The cargo bay fixed below the ballonet aims to house the on-board systems. The propulsive units mounted on both sides of the gondola furnish thrust for flight. The control surfaces (elevators and rudders) installed on the tail offer yawing and pitching moments.
To investigate the motion control of the airship, it is reasonable to establish the earth and body-fixed coordinate systems; see Fig. 1. The earth reference frame (ERF) has its origin $o_{g}$ at a fixed point on the earth, the $o_{g}x_{g}$ -axis points north, the $o_{g}y_{g}$ -axis points east, and the $o_{g}z_{g}$ -axis points to the earth’s centre perpendicular to the plane $o_{g}x_{g}y_{g}$ . The body-fixed reference frame (BRF) moving with the airship sets its origin o at the centre of volume (CV), the $ox_{b}$ -axis points to the nose of the airship, the $oy_{b}$ -axis points to the starboard side of the airship, and the $oz_{b}$ -axis lying on the longitudinal axisymmetric plane of the airship normal to the plane $ox_{b}y_{b}$ .
Neglect the aeroelastic influences and regard the airship as a rigid body. Taken from the airship modeling technique [Reference Khoury2, Reference Han, Wang, Chen and Duan9, Reference Wu, Wang, Duan, Xie and Wei10, Reference Xiao, Han, Wang, Zhou and Duan14, Reference Yang15, Reference Zheng and Xie17, Reference Yang, Wu and Zheng22, Reference Azinheira, Moutinho and De Paiva23, Reference Zheng, Feroskhan and Sun34], the airship model built around the horizontal motion can be directly given here, which is formulated by [Reference Zheng, Guan, Ma and Zhu6, Reference Yan, Weidong, Yugeng and Zili18, Reference Yang, Wu and Zheng36–Reference Yang, Wu and Zheng38]
In the above equations, x, y are the CV’s positions in the ERF, and $\psi$ is the yaw attitude in the BRF; u, v, r are the surge velocity, lateral velocity and yaw angular velocity with respect to the BRF, respectively; $m_{i}$ and $d_{i}$ ( $i=u,v,r$ ) denote the mass and damping term [Reference Khoury2, Reference Yamada, Adachi and Funahashi19, Reference Funk, Lutz and Wagner39], respectively; $m_{uv}=m_{u}-m_{v}$ ; $\tau=[\tau_{u},\tau_{r}]^{T}$ is the actuating signal. $\delta=[\delta_{u},\delta_{v},\delta_{r}]^{T}$ characterises all the modeling imperfections and disturbances. To facilitate subsequent discussions, some of most fundamental assumptions are delineated below.
Assumption 1 (see other works [Reference Yang, Yang and Deng3, Reference Wu, Wang, Duan, Xie and Wei10, Reference Yang, Wu and Zheng22]). The airship is flying at a proper cruising altitude with stable meteorological conditions, while maintaining buoyancy-weight balance. The pitch and roll angle are very small such that the associated dynamics can be neglected.
Assumption 2 (see other works [Reference Zheng, Guan, Ma and Zhu6, Reference Yamada, Adachi and Funahashi19, Reference Funk, Lutz and Wagner39]). In this work, we suppose that $\delta_{i}$ (i = u,v,r) takes the form
where $\delta^{\prime}_{i}$ is the cross-coupling term, $m^{\prime}_{i}$ and $d^{\prime}_{i}$ are the uncertain part of $m_{i}$ and $d_{i}$ , respectively, and $\delta_{dis_{i}}$ characterises the slow time-varying wind resistance.
Remark 1. From Equation (8), the airship, operating in three degrees of freedom, only has two independent actuating signals ( $\tau_{u},\tau_{r}$ ) in surge and yaw, which poses a prominent obstacle to steering the airship alone through a scheduled trajectory with stringent time and performance requirements. Indeed, it is physically apparent that most surface vehicles (SVs) and autonomous underwater vehicles (AUVs) propelling themselves on a horizontal plane are underactuated, and the motion equations of SVs and AUVs are pretty similar to that of the underactuated airship [Reference Dai, He and Lin20, Reference Jia, Hu and Zhang21]. When designing the motion control algorithms for SVs and AUVs, the performance specifications, such as the convergent time and the residual set, always exist. Therefore, this work is constructive for motion control design for SVs and AUVs to some extent.
Remark 2. Some effective altitude control techniques [Reference Yang, Yang and Deng3, Reference Wu, Wang, Duan, Xie and Wei10, Reference Yang, Wu and Zheng22], such as inflating and deflating valves, make Assumption 1 mild and realistic. Assumption 2 is frequently made in the works on horizontal motion control of airships [Reference Zheng, Guan, Ma and Zhu6, Reference Yamada, Adachi and Funahashi19, Reference Funk, Lutz and Wagner39]. It is also important to point out that the term $\delta^{\prime}_{i}$ ( $i=u,v,r$ ) in Equation (9) is used to depict the coupling effects of pitch and roll.
2.3 Control objective
In this article, our control objective is to generate a control law for $\tau$ to be employed by the stratospheric airship in underactuated mode, such that, despite the adverse influences of unmeasured velocities and uncertainties, the position of the airship $\eta=[x,y]^{T}$ tracks with the scheduled route $\eta_{d}=[x_{d},y_{d}]^{T}$ with a priori designate performance, i.e. such that the tracking error $e=\eta-\eta_{d}=[x_{e},y_{e}]^{T}$ fulfills
where $\epsilon$ represents the preassigned tracking accuracy, and $0<T_{f}<\infty$ denotes the finite settling time. Meanwhile, all the closed-loop signals maintain bounded $\forall t\ge0$ .
Assumption 3. The reference trajectory $\eta_{d}$ and its derivatives up to $\ddot{\eta}_{d}$ are bounded, continuous, and available for $\forall t\ge0$ .
3.0 Coordinate transformation
To fix the underactuated issue, a coordinate transformation is first conducted, specified as
where $\eta_{t}=\left[x_{t},y_{t}\right]^{T}$ denotes a new position, and except for having the definition in Equation (10), $\epsilon$ also is the distance between $\eta$ and $\eta_{d}$ (see Fig. 2). Calculating the second-order time-derivative of $\eta_{t}$ and taking Equation (8) into consideration yield
Then, substituting Equation (8) for Equation (12), we obtain
where
Denoting $\varkappa_{1}=\eta_{t}$ , $\varkappa_{2}=\dot{\eta}_{t}$ , $\digamma=\left[\digamma_{x},\digamma_{y}\right]^{T}$ , $\delta_{tf}=\left[\delta_{tf_{x}},\delta_{tf_{y}}\right]^{T}$ , and $\tau=\left[\tau_{u},\tau_{r} \right]^{T} $ , Equation (13) then can be rewritten in the synthetic form
where
Obviously, $R\!\left(\psi\right)R^{T}\!\left(\psi\right)=I_{2}$ , and $M_{\epsilon}$ is a positive-definite diagonal matrix if, and only if, $\epsilon\in\mathbb R_{>0}$ . Note that in this paper, we suppose that the velocity cannot be measured, and therefore, the term $\digamma=\left[\digamma_{x},\digamma_{y}\right]^{T}$ is actually unavailable for control design. To this end, we consider it as a part of uncertainties, and define the lumped disturbances $\delta_{lu}$ as $\delta_{lu}=\digamma+\delta_{tf}$ , which will be estimated by an observation mechanism designed later. For the convenience of observer design, the following assumption is provided.
Assumption 4 (see other works [Reference Zhang, Yu and Yan40–Reference Zhang, Wei, Wu and Cui42]). A bounded positive constant $B_{\delta_{lu}}$ exists such that the lumped disturbances $\delta_{lu}$ satisfy $\|\dot{\delta}_{lu}\|\le B_{\delta_{lu}}$ .
Remark 3. From Equation (18), the original motion model for the underactuated airship has been transformed into a fully actuated uncertain Euler–Lagrange (EL) model, formally defined as $M\!\left(q\right)\ddot{q}+C\!\left(q,\dot{q} \right)\dot{q}+G\!\left(q\right)=\tau$ , where $q\in \mathbb R^{n}$ is the generalised coordinate, $M\!\left(q\right)\in \mathbb R^{n\times n}$ is the known matrix, $C\!\left(q,\dot{q}\right)\in \mathbb R^{n}$ is the known nonlinear dynamic, $G\!\left(q\right)\in \mathbb R^{n}$ accounts for uncertain dynamics and disturbances, and $\tau\in \mathbb R^{n}$ is the control variable; essentially, the EL model can describe various physical systems [Reference Yang, Hua, Li and Guan33], such as robot manipulators [Reference Meng, Zhao and Yu12, Reference Anjum and Guo24, Reference Shen, Tang, Hu, Guo, Li and Zhang29] and spacecrafts [Reference Du, Zhang, Wu, Zhu, Li and Chu30, Reference Sun, Hou, Zong and Yu31]. Thus, the presented method can be readily extended to a series of mechanical systems in second-order form.
Remark 4. It is emphasised that the coordinate transformation does not weaken the control quality; conversely, it is a potent approach to resolve the underactuated problem and realise the preassigned accuracy simultaneously. This is especially clear if we now construct a control algorithm that succeeds in forcing the signal $e_{t_{1}}=\varkappa_{1}-\eta_{d}$ to shrink to zero within a finite time $T_{f}$ and maintain it there for $t\ge T_{f}$ , i.e., $\|e_{t_{1}}\|=0, \forall t\ge T_{f}$ , then the actual position error e will satisfy
Evidently, decreasing $\epsilon$ yields a higher tracking precision.
Remark 5. It is necessary to remark that this paper views the airship as a rigid body and therefore the linear velocities u and v and the angular velocity r cannot change suddenly. In addition, the rate of change of the uncertainties $\delta$ is restrained as the stratospheric climate is stable and has limited energy. As a result, Assumption 4 is reasonable in reality and is widely employed in current literature [Reference Zhang, Yu and Yan40–Reference Zhang, Wei, Wu and Cui42] to facilitate the output feedback control realisation.
4.0 Main result
In this section, we first build an FTO to offer the exact information of the unmeasured velocity $\varkappa_{2}$ and the lumped disturbances $\delta_{lu}$ in an integrated manner. Afterward, in conjunction with the transformed model, the designed FTO, the backstepping technique, and the API method, a novel trajectory tracking control algorithm for stratospheric airships is proposed. Finally, a Lyapunov analysis is carried out to prove the closed-loop system stability.
4.1 FTO
In this paper, the FTO is formulated by
where $\hat{\varkappa}_{1}$ , $\hat{\varkappa}_{2}$ , and $\hat{\delta}_{lu}$ are the estimates of $\varkappa_{1}$ , $\varkappa_{2}$ , and $\delta_{lu}$ , respectively. The parameters in Equation (21) satisfy $0<r_{i}<1$ , $z_{i}>1$ , $r_{i}=ir_{0}-\left( i-1\right) $ , $z_{i}=iz_{0}-\left( i-1\right) $ , $i=1,2,3$ , $0<r_{0}<1-\sigma_{1}$ , $0<z_{0}<1+\sigma_{2}$ , $\sigma_{1}\in\mathbb R_{>0}$ and $\sigma_{2}\in\mathbb R_{>0}$ are sufficiently small constants, and $\Upsilon\ge B_{\delta_{lu}}$ . The observer gains are selected to ensure the matrices
are Hurwitz. Based on the above contents, we obtained the main results of the FTO.
Theorem 1. Using the FTO (21) under Assumption 4, the velocity $\varkappa_{2}$ and the lumped disturbances $\delta_{lu}$ can be estimated accurately; more specifically, the estimation errors $e_{o_{1}}$ , $e_{o_{2}}$ , and $e_{o_{3}}$ can be driven to zero with a finite reaching time $T_{o}$ .
Proof of Theorem 1. Define $e_{o_{1}}=\varkappa_{1}-\hat{\varkappa}_{1}$ , $e_{o_{2}}=\varkappa_{2}-\hat{\varkappa}_{2}$ , and $e_{o_{3}}=\delta_{lu}-\hat{\delta}_{lu}$ as observer errors; therefore, together with Equation (18), the observer error dynamics can be computed as
The reminder of this proof is quite similar to that of Theorem 1 given by Basin et al. [Reference Basin, Yu and Shtessel43], and therefore, it is omitted here for space.
Remark 6. The FTO (21) essentially is a uniform robust exact differentiator. Historically, the concept of uniform exact convergence was proposed by Cruz-Zavala et al. [Reference Cruz-Zavala, Moreno and Fridman44] for the first time. Note that chattering, measurement noise, sampling step and small delay are out of the scope of this paper.
4.2 Control algorithm design
The entire design procedure is elaborated as follows.
Step 1. Design a stabilising function for $e_{t_{1}}$ . To begin with, let
Consider the simple quadratic Lyapunov function candidate $V_{1}=\frac{1}{2}e^{T}_{t_{1}}e_{t_{1}}$ . Evaluating the time derivative of $V_{1}$ by using Equations (18) and (24) results in
Adopting the virtual control law
for $e_{t_{1}}$ produces
where $\kappa_{1}\in \mathbb R^{2}$ and $\alpha\in\!\left(0,1\right) $ . Define the intermediate variable $\varpi$ as
and together with Lemma 4, we get
In view of Lemma 2, Equation (29) can be rewritten as
Substituting Equation (30) for Equation (27), the derivative $\dot{V}_{1}$ becomes
Step 2. Design a fixed-time control law for $\tau$ . To this end, select the complete Lyapunov function candidate as
where $V_{2_{i}}$ , $i=1,2$ , takes the form
In the sequel, we demonstrate that $V_{2_{i}}$ is a scalar positive function through analysing the following cases.
-
Case 1: $e^{\dagger}_{t_{2_{i}}}> e^{*}_{t_{2_{i}}}\ge0$ . From $sig^{\imath}({\cdot})=|{\cdot}|^{\imath}sign({\cdot}) $ and $s\in\left[ e^{\dagger}_{t_{2_{i}}},e^{*}_{t_{2_{i}}} \right] $ we thus get
(34) \begin{equation} s\ge e^{*}_{t_{2_{i}}}\ge 0,\ \ sig^{\frac{1}{\alpha}}\!\left( s\right)=s^{\frac{1}{\alpha}},\ \ \text{and}\ \ sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) = e^{*\frac{1}{\alpha}}_{t_{2_{i}}}. \end{equation}This implies that(35) \begin{equation} V_{2_{i}}=\int_{e^{*}_{t_{2_{i}}}}^{e^{\dagger}_{t_{2_{i}}}}\Big|s^{\frac{1}{\alpha}}-e^{*\frac{1}{\alpha}}_{t_{2_{i}}}\Big|^{2-\alpha}sign\!\left(s^{\frac{1}{\alpha}}-e^{*\frac{1}{\alpha}}_{t_{2_{i}}} \right)ds= \int_{e^{*}_{t_{2_{i}}}}^{e^{\dagger}_{t_{2_{i}}}}\!\left( s^{\frac{1}{\alpha}}-e^{*\frac{1}{\alpha}}_{t_{2_{i}}}\right) ^{2-\alpha}ds, \end{equation}where we used the fact that the power function $f\!\left(x \right)=x^{\frac{1}{\alpha}} $ is strictly increasing when $x>0$ . This, together with the well-known mean value theorem, gives(36) \begin{equation} V_{2_{i}}=\left( s^{*\frac{1}{\alpha}}_{1}-e^{*\frac{1}{\alpha}}_{t_{2_{i}}}\right) ^{2-\alpha}\!\left(e^{\dagger}_{t_{2_{i}}}-e^{*}_{t_{2_{i}}} \right)>0 \end{equation}with $s^{*}_{1}\in\left( e^{*}_{t_{2_{i}}},e^{\dagger}_{t_{2_{i}}}\right) $ . -
Case 2: $e^{\dagger}_{t_{2_{i}}}\ge0> e^{*}_{t_{2_{i}}}$ . Rewrite Equation (33) as
(37) \begin{equation} V_{2_{i}}=\int_{0}^{e^{\dagger}_{t_{2_{i}}}}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds+\int_{e^{*}_{t_{2_{i}}}}^{0}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds. \end{equation}Regarding the first term on the right side of Equation (37), if $e^{\dagger}_{t_{2_{i}}}=0$ , then $\int_{0}^{e^{\dagger}_{t_{2_{i}}}}sig^{2-\alpha} \!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-\right.$ $\left.sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds=0$ . Otherwise, we can easily verify that there exists a strictly positive constant $s^{*}_{2}\in\left( 0,e^{\dagger}_{t_{2_{i}}}\right) $ such that(38) \begin{equation} \begin{aligned} &\int_{0}^{e^{\dagger}_{t_{2_{i}}}}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds=\int_{0}^{e^{\dagger}_{t_{2_{i}}}}\big|s^{\frac{1}{\alpha}}+\big| e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}\big|^{2-\alpha}sign\!\left(s^{\frac{1}{\alpha}}+\big| e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}} \right)ds=\\ &\int_{0}^{e^{\dagger}_{t_{2_{i}}}}\!\left(s^{\frac{1}{\alpha}}+\big| e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}} \right)^{2-\alpha}ds=\left(s^{*\frac{1}{\alpha}}_{2}+\big| e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}} \right)^{2-\alpha}e^{\dagger}_{t_{2_{i}}}>0, \end{aligned} \end{equation}where $e^{*}_{t_{2_{i}}}<0\Rightarrow sign\!\left(e^{*}_{t_{2_{i}}}\right)=-1$ , $\forall s\in\left[ 0, e^{\dagger}_{t_{2_{i}}}\right] \Rightarrow sign(s)=1 $ , and the mean value theorem have been used.As for the second term, let us now consider the integration by substitution technique. By making the substitution $s=-g$ , we have
(39) \begin{equation} \begin{aligned} &\int_{e^{*}_{t_{2_{i}}}}^{0}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds=\int_{0}^{\big|e^{*}_{t_{2_{i}}}\big|}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( -g\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)dg=\\ &\int_{0}^{\big|e^{*}_{t_{2_{i}}}\big|}sig^{2-\alpha}\!\left(|g|^{\frac{1}{\alpha}}sign\!\left({-}g \right)-\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}sign\!\left(e^{*}_{t_{2_{i}}} \right) \right)dg. \end{aligned} \end{equation}Further consider that $sign\!\left(e^{*}_{t_{2_{i}}}\right)=-1$ and $\forall g\in\left[ 0, \big|e^{*}_{t_{2_{i}}}\big|\right] $ means $ sign({-}g)=-1$ and $\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}>|g|^{\frac{1}{\alpha}}$ . Consequently, Equation (39) simply becomes(40) \begin{equation} \begin{aligned} &\int_{0}^{\big|e^{*}_{t_{2_{i}}}\big|}sig^{2-\alpha}\!\left(|g|^{\frac{1}{\alpha}}sign\!\left({-}g \right)-\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}sign\!\left(e^{*}_{t_{2_{i}}} \right) \right)dg=\int_{0}^{\big|e^{*}_{t_{2_{i}}}\big|}sig^{2-\alpha}\!\left(\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}-|g|^{\frac{1}{\alpha}}\right) dg=\\[4pt] &\int_{0}^{\big|e^{*}_{t_{2_{i}}}\big|}\!\left(\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}-|g|^{\frac{1}{\alpha}}\right)^{2-\alpha}sign\!\left(\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}-|g|^{\frac{1}{\alpha}} \right)dg= \int_{0}^{\big|e^{*}_{t_{2_{i}}}\big|}\!\left(\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}-|g|^{\frac{1}{\alpha}}\right)^{2-\alpha}dg. \end{aligned} \end{equation}Proceeding similarly to get Equation (36) leads to(41) \begin{equation} \int_{e^{*}_{t_{2_{i}}}}^{0}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds=\left( \big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}-g_{1}^{*\frac{1}{\alpha}}\right)^{2-\alpha}\big|e^{*}_{t_{2_{i}}}\big|>0 \end{equation}with $g^{*}_{1}\in\left(0,\big|e^{*}_{t_{2_{i}}}\big|\right) $ . Summarising the results in Equations (38) and (41) gives $V_{2_{i}}>0$ in the case of $e^{\dagger}_{t_{2_{i}}}\ge0> e^{*}_{t_{2_{i}}}$ . -
Case 3: $0\ge e^{\dagger}_{t_{2_{i}}}> e^{*}_{t_{2_{i}}}$ . Clearly, in this case, $0\le\big|e^{\dagger}_{t_{2_{i}}}\big|<\big|e^{*}_{t_{2_{i}}}\big|$ holds. Let us again use the integration by substitution technique. Setting $s=-g$ results in
(42) \begin{equation} V_{2_{i}}=\int_{e^{*}_{t_{2_{i}}}}^{e^{\dagger}_{t_{2_{i}}}}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left(s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds=\int_{\big|e^{\dagger}_{t_{2_{i}}}\big|}^{\big|e^{*}_{t_{2_{i}}}\big|}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( -g\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)dg. \end{equation}Following the same lines to obtain Equations (40) and (41), it can also be shown that(43) \begin{equation} V_{2_{i}}=\int_{\big|e^{\dagger}_{t_{2_{i}}}\big|}^{\big|e^{*}_{t_{2_{i}}}\big|}\!\left(\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}-|g|^{\frac{1}{\alpha}}\right)^{2-\alpha}dg=\left(\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}}-g_{2}^{*\frac{1}{\alpha}}\right)^{2-\alpha}\big(\big|e^{*}_{t_{2_{i}}}\big|-\big|e^{\dagger}_{t_{2_{i}}}\big|\big)>0 \end{equation}with $g^{*}_{2}\in\left(\big|e^{\dagger}_{t_{2_{i}}}\big|,\big|e^{*}_{t_{2_{i}}}\big|\right) $ . -
Case 4: $e^{*}_{t_{2_{i}}}\ge0>e^{\dagger}_{t_{2_{i}}}$ . We now switch the upper and lower bounds of integral (33) and thus obtain
(44) \begin{equation} \begin{aligned} V_{2_{i}}&=-\int_{e^{\dagger}_{t_{2_{i}}}}^{e^{*}_{t_{2_{i}}}}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds\\ &=-\int_{0}^{e^{*}_{t_{2_{i}}}}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds-\int_{e^{\dagger}_{t_{2_{i}}}}^{0}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds. \end{aligned} \end{equation}The first term on the right side of Equation (44) is identically equal to zero if $e^{*}_{t_{2_{i}}}=0$ ; if not, proceeding as before, it can be simplified as(45) \begin{equation} \begin{aligned} &-\int_{0}^{e^{*}_{t_{2_{i}}}}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds=-\int_{0}^{e^{*}_{t_{2_{i}}}}\Big|s^{\frac{1}{\alpha}}-e^{*\frac{1}{\alpha}}_{t_{2_{i}}}\Big|^{2-\alpha}sign\!\left( s^{\frac{1}{\alpha}}-e^{*\frac{1}{\alpha}}_{t_{2_{i}}}\right)ds=\\ &\left( e^{*\frac{1}{\alpha}}_{t_{2_{i}}}-s_{3}^{*\frac{1}{\alpha}}\right) ^{2-\alpha}e^{*}_{t_{2_{i}}}>0, \end{aligned} \end{equation}where $s^{*}_{3}\in\left(0,e^{\dagger}_{t_{2_{i}}}\right) $ and we have used the fact $\forall s\in\left[ 0,e^{*}_{t_{2_{i}}}\right] \Rightarrow s^{\frac{1}{\alpha}}\le e^{*\frac{1}{\alpha}}_{t_{2_{i}}}$ .And the second term, given the substitution $s=-g$ , satisfies
(46) \begin{align} &-\int_{e^{\dagger}_{t_{2_{i}}}}^{0}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( s\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)ds=\int_{\big|e^{\dagger}_{t_{2_{i}}}\big|}^{0}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( -g\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)dg=\nonumber\\ &-\int_{0}^{\big|e^{\dagger}_{t_{2_{i}}}\big|}sig^{2-\alpha}\!\left(sig^{\frac{1}{\alpha}}\!\left( -g\right)-sig^{\frac{1}{\alpha}}\!\left( e^{*}_{t_{2_{i}}}\right) \right)dg=\int_{0}^{\big|e^{\dagger}_{t_{2_{i}}}\big|}\left(g^{\frac{1}{\alpha}}+\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}} \right)^{2-\alpha}dg=\nonumber\\ & \left(g^{*\frac{1}{\alpha}}_{3}+\big|e^{*}_{t_{2_{i}}}\big|^{\frac{1}{\alpha}} \right)^{2-\alpha}\big|e^{\dagger}_{t_{2_{i}}}\big|>0, \end{align}where $g^{*}_{3}\in\left(0,\big|e^{*}_{t_{2_{i}}}\big|\right) $ and we have used the facts $sign\!\left({-}g \right)=-1 $ and $sign\!\left(e^{*}_{t_{2_{i}}} \right)=-1 $ . Taking Equations (44) and (45) into account, we know that $V_{2_{i}}>0$ when $e^{*}_{t_{2_{i}}}\ge0>e^{\dagger}_{t_{2_{i}}}$ .
Evidently, the above discussion guarantees that $V_{2_{i}}$ is positively defined. Differentiating $V_{2_{i}}$ with respect to time and applying Lemma 5 lead to
Noting that
and
we have
The second term in the right side of Equation (50) satisfies
with
where Equations (24), and (26), (28), and Lemma 4 have been used. Coupling Equations (51) and (52), we have
where Lemma 2 has been used. Then, the differentiation of $V_{2_{i}}$ and the utilisation of Equation (53) yield
Together with Equations (18), (23), (24), (27), and (54), the definition of V, and the dynamics of $e^{\dagger}_{t_{2}}$
we get
with $\gamma_{1_{i}}=\frac{2^{1-\alpha}(1+\left(2-\alpha\right)\alpha)+1}{1+\alpha}$ and $\gamma_{2_{i}}=\frac{2^{1-\alpha}\alpha}{1+\alpha}+\kappa_{1_{i}}^{\frac{1}{\alpha}}\!\left(2-\alpha\right)2^{2-2\alpha}+\frac{\left(2-\alpha\right)2^{1-\alpha}}{1+\alpha}\kappa_{1_{i}} ^{1+\alpha}$ , $i=1, 2$ . According to the structure of $\dot{V}$ , the trajectory tracking control law for $\tau$ is designed as
where $\rho\!\left(t\right)=\left\lbrace \begin{array}{cl}((T_{\rho}-t)/T_{\rho})^\frac{1}{\alpha_{\rho}}(\rho_{0}-\rho_{\infty})+\rho_{\infty}&,0\le t\le T_{\rho}\\ \rho_{\infty}&,t>T_{\rho}\end{array}\right. $ with $\left\lbrace\rho_{0},\rho_{\infty},\alpha_{\rho},T_{\rho} \right\rbrace\in\mathbb R_{>0} $ , and $ \kappa_{2} \in \mathbb R^{2}_{>0}$ . The main results of this brief are included in the following theorem.
Theorem 2. Consider the stratospheric airship (8) subject to underactuated dynamics, unmeasured velocities, modeling imperfections, and exogenous disturbances, with the coordinate transformation (11), the FTO (21), the virtual control law (26), and the control action (57). Suppose that Assumptions 1–4 hold and the control parameters are selected such that
where $ \kappa^{\prime}_{1_{i}} \in R_{>0}$ , $i=1,2$ . Then, the position error $e=\eta-\eta_{d}$ converges to a preassigned small vicinity of the origin within a finite time $T_{f}$ , while all the closed-loop states maintain bounded for $\forall t\ge0$ .
Proof of Theorem 2. Let us prove first that these signals do not go to infinity when $t\in\left[0,T_{o}\right] $ . Substituting the control action (57) with design parameters satisfying the condition (58), we have
where the fact
has been used. Noticing that the signals $e_{o_{1}}$ and $e_{o_{2}}$ decay to zero after a fixed-time time $T_{0}$ , hence $\left\lbrace e_{o_{1_{i}}},e_{o_{2_{i}}}\right\rbrace\in\mathbb{L}^{\infty}$ . Accordingly, there exists a constant $B_{e_{o}}\in\mathbb{R}_{>0}$ such that $\sup_{t\in\left[0,\infty \right) }\big\lbrace |e_{o_{1_{i}}}|,\left|e_{o_{2_{i}}}\right|,$ $\left.|e_{o_{1_{i}}}|^{r_{2}},\left|e_{o_{2_{i}}}\right|^{z_{2}}\right\rbrace \le B_{e_{o}}$ , $i=1,2$ . This, combined with Young’s inequality, gives
In view of Equations (61) and (62), Equation (59) can be restated as
Apparently, a constant $B_{s}^{[0,T_{o}]}\in\mathbb{R}_{>0}$ exists, which is an upper bound on the sum of the last three terms in Equation (63) when $t\in[0,T_{o}]$ . Besides, Recalling the definition of function $V_{2i}$ , we can easily see
and therefore the complete Lyapunov function candidate V satisfies $V\le2e^{T}_{t_{1}}e_{t_{1}}+2\varpi^{T}\varpi$ . Since $\alpha\in\left( 0,1\right)\Rightarrow \frac{1+\alpha}{2}\in\left(\frac{1}{2},1\right) $ , this further leads to
When $V>1$ , summarising the results in Equations (59), (63), (64) and (65) gives
where Lemma 3 has been used, $\lambda_{0}=\min\!\left(\frac{1}{2}\right)^{\frac{1+\alpha}{2}}\times\left\lbrace \kappa^{\prime}_{1_{i}},\kappa_{2i}\right\rbrace $ , $i=1,2$ . Multiplying Equation (66) by $e^{ \lambda_{0}t}$ and integrating it over $\left[0,t\right] $ , we have
Besides, in the case of $V\le1$ , the system states $e_{t_{i}}$ , $i=1,2$ , and $\varpi$ are obviously bounded. As a result, we can conclude that these signals will not drift to infinity as $t\in\left[0,T_{o}\right] $ .
Now, we proceed to prove Theorem 2. Theorem 1 reveals that $\hat{\varkappa}_{2}=\varkappa_{2}$ and $\hat{\delta}_{lu}=\delta_{lu}$ for $\forall t\ge T_{0}$ . Along with the fact that $\varrho\!\left(t\right)\in\mathbb{R}_{>0} $ , Equation (59) becomes
Consequently, all of the system states remain bounded for $t\ge T_{o}$ . Besides, it is noteworthy that:
-
1. the finite-time tracking task is fulfilled. Given Lemma 1, we can conclude that both $e_{t_{1}}$ and $\varpi$ decay to zero within a finite time $T_{f}$ estimated by
(69) \begin{equation} T_{f}\le\frac{2V^{\frac{1-\alpha}{2} }\!\left(0\right)}{\lambda_{0}(1-\alpha)}+T_{o}; \end{equation} -
2. a priori assignment of tracking accuracy is assured. Recalling Equation (20), we get
(70) \begin{equation} \|e_{t_{1}}\|=0, \forall t\ge T_{f}\Rightarrow\max\left\lbrace|x_{e}|,|y_{e}| \right\rbrace\le\epsilon, \forall t\ge T_{f}, \end{equation}indicating that we can specify the accuracy bound with the parameter $\epsilon$ .
Remark 7. In the related works on API method [Reference Du, Zhang, Wu, Zhu, Li and Chu30–Reference Fu, Wang and Wang35], the power is circumscribed to be an even integer or a ratio of two odd integers. However, with the benefit of Lemma 4, such curtailments are removed in our work.
Remark 8. In contrast to the sliding mode control designs [Reference Yang11, Reference Xiao, Han, Wang, Zhou and Duan14, Reference Yang15, Reference Zhang, Wei, Wu and Cui42], the control action (57) is chattering-free. In this note, although the signum operator $sign({\cdot})$ is employed, the highly undesirable control activity is avoided as the fractional power term $sig^{\varsigma}\!\left(z\right)=|z|^\varsigma sign\!\left( z\right) $ is a non-smooth but continuous function of z, where $\varsigma>0$ . This is vital for the long-time flight of the airship as control chattering may shorten the lifespan of the devices. Moreover, from the definition of $\rho\!\left( t\right)$ , we can find that $\rho\!\left( t\right)$ is a strictly decreasing positive function and it takes a very small value $\rho_{\infty}$ when $t>T_{\rho}$ . In this sense, we can avoid unnecessary control effort to some extent by incorporating it into the control action (57).
5.0 Simulation
5.1 Simulation condition
In this section, numerical simulations are performed to intuitively evaluate the effectiveness of the preceding theoretical results with the running time $T_{final}=100$ s. The physical parameters of the stratospheric airship are borrowed from Zheng et al. [Reference Zheng, Guan, Ma and Zhu6] and Zhang et al. [Reference Yan, Weidong, Yugeng and Zili18], and all of them are listed in Table 1.
In this simulations, the airship starts from $\left[x\!\left(0\right),y\!\left(0\right)\right]^{T}= \left[340{\rm m},390{\rm m}\right]^{T}$ with initial heading $\psi\!\left(0\right)=0.1$ rad, and initial velocity $\nu\!\left(0\right)= \left[u\!\left(0\right),v\!\left(0\right),r\!\left(0\right)\right]^{T}$ $=\left[0{\rm m/s},0{\rm m/s},0{\rm rad/s}\right]^{T}$ , and is required to track the reference trajectory
with the preassigned accuracy $\epsilon=0.2$ . To test the robustness, the disturbance $\delta_{dis}$ induced by environmental forces is mathematically assumed as [Reference Zheng, Guan, Ma and Zhu6, Reference Wu, Wang, Duan, Xie and Wei10]:
To fulfill this mission, the control parameters are selected as $p_{1}=q_{1}=24$ , $p_{2}=q_{2}=216$ , $p_{3}=q_{3}=864$ , $r_{1}=0.8$ , $r_{2}=0.6$ , $r_{3}=0.4$ , $z_{1}=1.2$ , $z_{2}=1.4$ , $z_{3}=1.6$ , $\Upsilon=5$ , $\epsilon=0.2$ , $\alpha=0.7$ , $\kappa_{1}=\left[2,2\right]^{T}$ , $\kappa_{2}=\left[0.1,0.1\right]^{T}$ , $\rho_{0}=1$ , $\rho_{\infty}=0.01$ , $\alpha_{\rho}=2$ , and $T_{\rho}=1$ . The initial conditions of the FTO (21) are set to zero.
5.2 Simulation result
Applying the FTO (21) and the control action (57) to the airship model (8), we reach some simulation outcomes, illuminated by Figs 3-13. From Figs 3-6, we see that the airship can move to the desired trajectory swiftly and smoothly, and the position errors $x_{e}=x-x_{d}$ and $y_{e}=y-y_{d}$ decay toward a close vicinity of zero within a finite time. Then, the airship flies along the reference trajectory precisely, irrespective of underactuated dynamics, modeling imprecisions, and exceptional disturbances. The simulation results for velocities are shown in Fig. 7, with the three curves corresponding to velocities in the surge, sway and yaw, respectively, all of which are bounded for $\forall t\ge0$ but do not enter a steady state. The reason is twofold. First, a circular path is allocated to track, yielding the desired velocities time-varying. Second, the persistent perturbations (9) consisting in the airship model (8) affect the system dynamics directly. Figures 8-9 plot the unmeasured velocity $\dot{\eta}_{t}$ (i.e., $\varkappa_{2}$ ), the lumped disturbances $\delta_{lu}$ , and their observed values, which show that the FTO (21) can supply the exact observations of $\dot{\eta}_{t}$ and $\delta_{lu}$ in a finite time. Figure 10 presents the necessary control action, from which we see that the surge force and yaw torque are continuous, and no control chattering exists. It should be noted that, compared with the actuating signal $\tau_{r}$ in the steady-state phase, it appears more aggressive in the initial stage. The reason is that the mass $m_{r}$ is huge (12,176kg $\cdot$ m2), and the velocity is initialised to zero. As a result, a large control torque is required in the initial stage to accelerate the airship to the reference route. Nevertheless, it decreases quickly, as shown in Fig. 10. The trajectory tracking responses for different initial positions are depicted in Figs 11-13. It can be observed that the finite-time convergence is assured, and the requirement for tracking accuracy, featured in Equation (70), is also met. The above conclusions successfully affirm that a good tracking performance is achieved under our method.
Aimed at comparison, a standard command-filter backstepping (CFB) controller
formulated by Han et al. [Reference Han, Wang, Chen and Duan9] is introduced in this paper, where the control parameters $k_{1}=k_{2}=eye(3)$ are identical to those used by Han et al. [Reference Han, Wang, Chen and Duan9]. The comparison results are illustrated in Figs 14-17, where FI control shorts for our method. Moreover, to further display the comparative simulations, we summarise the quantisation indexes in Table 2, where the integrated absolute error, IAE (defined as IAE $=\int_{0}^{T_{final}}|j_{e}\!\left(t\right) |dt,j=x,y$ ), the integrated time absolute error, ITAE (defined as ITAE $=\int_{0}^{T_{final}}t|j_{e}\!\left(t\right) |dt,j=x,y$ ), and the mean integrated absolute control, MIAC (defined as MIAC $=\frac{1}{T_{final}}\int_{0}^{T_{final}}|\tau_{j}\!\left(t\right) |dt,j=u,i$ ), devote to assess steady-state performance, transient performance, and control effort, respectively. As seen in Figs 14-17 and Table 2, it is clear that the tracking performance is not satisfactory under the CFB controller and large position errors emerge in the steady state, although the CFB controller ensures the boundedness of position errors. Particularly, the robustness of the CFB controller cannot be warranted in default of a compensation mechanism versus modeling imprecisions and disturbances, as demonstrated in Figs 15-16. In addition, compared to our method, the CFB controller has the biggest IAE, ITAE and MIAC values. This indicates that our method can offer a better tracking quality with less control energy consumption.
5.3 Discussion
This paper studies the horizontal trajectory tracking control problem of airships. The system model is built upon Assumption 1, which is widely made in the related literature [Reference Yang, Yang and Deng3, Reference Zheng, Guan, Ma and Zhu6, Reference Wu, Wang, Duan, Xie and Wei10, Reference Zheng and Xie17, Reference Yan, Weidong, Yugeng and Zili18, Reference Yang, Wu and Zheng22, Reference Azinheira, Moutinho and De Paiva23]. Assumption 2 is introduced to give some characterisation of unmodeled dynamics and external disturbances, as the exact model of airships is not always attainable in reality, and the persistent wind field directly affects the motion of airships and thus cannot be neglected in the design of a tracker. From Equation (8), we easily find that the airship, flying in the horizontal plane, is a typical underactuated system. To this end, we start our research with a coordinate transformation (11). After a series of reasoning shown in Equations (12)–(17), a fully actuated EL model (18) is obtained. From Figs 3-6, we see that the coordinate transformation (11) does solve the underactuated problem. We then present an FTO (21) that requires Assumption 4 to realise the velocity-free control and the dynamical compensation. The main property of the FTO (21) is given in Theorem 1, and the reasonability of Assumption 4 is given in Remark 5. Figures 8-9 demonstrate the effectiveness of the FTO (21). Note that in the current literature on API technique [Reference Du, Zhang, Wu, Zhu, Li and Chu30–Reference Fu, Wang and Wang35], certain control parameters are restricted to ratios of positive odd integers. Based on Lemma 4, we relax this restriction. The selection for $\alpha$ shows this advantage. Finally, in the light of Lemmas 1–5, we propose an API-based finite trajectory tracking control algorithm for underactuated airships without command filters or dynamic surfaces. Figures 11–12 illustrate the finite-time convergence of position errors to a preassigned residual set, validating the theoretical predictions in Remark 4 and Theorem 2.
6.0 Conclusion
This brief presented a novel finite-time velocity-free trajectory tracking control algorithm for an underactuated airship under the condition of modeling imperfections and environmental disturbances. First, a coordinate transformation was conducted to address the underactuated problem, which make the presented approach can be extended easily to a wide range of second-order mechanical systems. Second, an FTO was built to form a output-feedback control structure with disturbance estimation and attenuation ability. Finally, we blended the backstepping technique and API method into a Lyapunov design, which successfully guaranteed the finite-time convergence of the position errors $x_{e}$ and $y_{e}$ into a preassigned residual set around zero. The control design did not cover any analytically formidable calculation, filters, or self-tuning mechanisms (e.g. FLSs or NNs), leading to a structurally simple control attribute. Our future work will focus on extending this approach to a stratospheric airship with actuator faults.
Acknowledgment
This work was supported in part by the National Natural Science Foundation of China (Nos. 51205253 and 51906141).
Conflicts interest
There is no conflict of interest.