Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-06-02T06:02:54.569Z Has data issue: false hasContentIssue false

Quaternion-based state-dependent differential Riccati equation for quadrotor drones: Regulation control problem in aerobatic flight

Published online by Cambridge University Press:  17 February 2022

Saeed Rafee Nekoo*
Affiliation:
GRVC Robotics Lab., Departamento de Ingeniería de Sistemas y Automática, Escuela Técnica Superior de Ingeniería, Universidad de Sevilla, Seville, Spain
José Ángel Acosta
Affiliation:
GRVC Robotics Lab., Departamento de Ingeniería de Sistemas y Automática, Escuela Técnica Superior de Ingeniería, Universidad de Sevilla, Seville, Spain
Anibal Ollero
Affiliation:
GRVC Robotics Lab., Departamento de Ingeniería de Sistemas y Automática, Escuela Técnica Superior de Ingeniería, Universidad de Sevilla, Seville, Spain
*
*Corresponding author. E-mail: saerafee@yahoo.com
Rights & Permissions [Opens in a new window]

Abstract

The quaternion is a powerful and common tool to avoid singularity in rotational dynamics in three-dimensional (3D) space. Here it has been particularly used as an alternative to Euler angles and rotation matrix. The application of the quaternion is exercised in quadrotor modeling and control. It changes the dynamics and represents a singularity-free attitude model. Here for the first time (for the best knowledge of authors), the state-dependent differential Riccati equation (SDDRE) control has been implemented on the quaternion-based model of a quadcopter. The proposed control structure is capable of aerobatic flight, and the Pugachev’s Cobra maneuver is chosen to assess the capability of the quaternion-based SDDRE approach. The introduced control simulator is validated by comparison with conventional dynamics based on Euler angles, controlled using a proportional-derivative (PD) controller on a normal regulation flight. The simulator successfully performed the Cobra maneuver and also validated the proposed structure. The more precision in regulation along with lower energy consumption demonstrated the superiority of the introduced approach.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press

1. Introduction

The Euler angles and rotation matrix in three-dimensional (3D) space are vulnerable to singularities, considering dynamics, especially in an aerobatic flight. Here we focus, particularly on multirotor drones and quadcopters. The dynamics of the multirotor drones are usually subjected to hovering assumptions to guarantee a stable flight [Reference Bibik, Narkiewicz, Zasuwa and Żugaj1]. To perform agile maneuvers with sudden changes in attitude, geometric control was introduced that avoids singularities and computes the rotation matrix in another manifold [Reference Fernando, Chandiramani, Lee and Gutierrez2]. Krishna et al. used the geometric control for helicopter trajectory tracking in agile flight regarding the attitude of the system [Reference Krishna, Sen and Kothari3]. The sliding mode was employed as the controller and the error function in the geometric domain constructed the sliding condition to prove the stability. The geometric control was also applied on quadrotor drones subject to wind disturbance and sudden changes in attitude dynamics [Reference Craig, Yeo and Paley4]. A flip is representative of a challenging maneuver that includes passing through a singularity that was addressed by the geometric approach [Reference Nekoo, Acosta and Ollero5]. The $ \pi \left(\textrm{r}\textrm{a}\textrm{d}\right)$ flip in roll angle was considered, sudden motion between two stable roll angles. The change in the manifold and solving singularity cost the user more complicated integration methods [Reference Monforte6]. Variational integration is an effective method [Reference Lee, Leok and McClamroch7]. Improper treatment of the integration may cause numerical drift in the results, especially in big orientation changes.

The quaternion is another representation of the complex numbers in mathematics, with a wide range of usage in theory and application. The focus of this work is to implement the alternative representation of the Euler angles and rotation matrix in 3D space. The application of quaternion in control was reported on aircraft [Reference Terze, Zlatar and Pandža8, Reference Zhou, Zhang, Liu, Zeng and Tian9], orientation and translation control of manipulators [Reference Figueredo, Adorno, Ishihara and Borges10, Reference Savino, Pimenta, Shah and Adorno11], control of autonomous underwater vehicles [Reference Antonelli, Chiaverini, Sarkar and West12Reference de Cerqueira Gava, Jorge, Júnior and Adabo15], helicopter attitude control [Reference Suzuki, Nakazawa, Nonami and Tawara16], etc. Terze et al. used quaternion representation of the rotational dynamics for aircraft simulators and introduced shifting update process to ensure precise integration in long flight simulations [Reference Terze, Zlatar and Pandža8]. Kinematics control of a robotic arm was presented using dual quaternion, aimed to present a robust controller without decoupling of the translational and orientation dynamics [Reference Figueredo, Adorno, Ishihara and Borges10]. Cooke et al. presented a thorough implementation of the quaternion dynamics for flight simulation [Reference Cooke, Zyda, Pratt and McGhee17]. The performance of the quaternion-based control was validated by an omnidirectional intelligent navigator for an underwater platform [Reference Antonelli, Chiaverini, Sarkar and West12]. Suzuki et al. presented the simulation and experimental studies on Lepton-Ex unmanned helicopter using quaternion feedback in control [Reference Suzuki, Nakazawa, Nonami and Tawara16]. Quadrotor and multicopters were also employed quaternions [Reference Islam and Okasha18Reference Nguyen, Nguyen, Prodan and Pereira23]. A study compared three controllers, linear quadratic regulator (LQR), proportional-derivative (PD), and model predictive controller (MPC) for trajectory tracking that revealed the PD and the LQR gained better results in an ideal condition, and the MPC was more accurate in presence of disturbance [Reference Islam and Okasha18]. Quaternion variables in the attitude controller provided an advantage, employment of low-cost sensors due to the high-speed capability of the singularity-free controller [Reference Kang, Sadegh and Urschel20]. Palomo et al. presented an observer-based method based on position-yaw for the ellipsoid method [Reference Oliva-Palomo, Sanchez-Orta, Alazki, Castillo and Muñoz-Vázquez21]. Linear matrix inequalities were used to optimize the feedback of the observer, and experimental results showed successful tracking in high speed maneuver. Euler angles are popular in UAV modeling though they suffer from gimbal lock when two orthogonal axes align and lock together [Reference Chovancová, Fico, Hubinský and Duchoň24]. Another disadvantage of Euler angles is the computational cost by computing so many trigonometric functions [Reference Fresk and Nikolakopoulos25], on the contrary, quaternion mathematics only involves algebraic computations [Reference Zha, Ding, Yu and Wang26]. Sanchez et al. used quaternion dynamics for quadrotor control based on receiving gesture commands [Reference Sanchez, Abaunza and Castillo27], where it was important to cope with the unpredictability of the gesture and quaternion made it more reliable due to insensitivity to singularity.

Performing experiment on aggressive maneuvers requires more safety concerns in addition to solving singularities. Gillula et al. used the Hamilton–Jacobi differential game approach for finding a reachable set for aerobatic maneuvers to guarantee safety [Reference Gillula, Huang, Vitus and Tomlin28]. Considering the constraint of the actuators in the flight experiment is an important issue since, during the flip, it is highly probable to put the rotors in saturation [Reference Cutler and How29]. Here in this work, the simulation of the Cobra maneuver is done and for experimentation, actuator limits and speed of the maneuver must be considered to avoid undesired saturations in the middle of trajectory that could be dangerous.

The state-dependent Riccati equation (SDRE) is a closed-loop optimal nonlinear controller introduced in the 1960s [Reference Pearson30]. The utilization of the quaternion in SDRE was explored in several platforms such as satellite [Reference Parrish and Ridgely31, Reference Romero, Souza and Chagas32], spacecraft [Reference Stansbery and Cloutier33Reference Pukdeboon36], spacecraft in proximity operations [Reference Li, Yue, Chi and Du37], attitude control of a rigid body [Reference Safi, Mortazavi and Dibaji38], and remote sensing CubeSats [Reference Romero and de Souza39]. Here we present the state-dependent differential Riccati equation (SDDRE) control based on quaternion for quadcopter dynamic systems. For the best knowledge of authors, a quaternion-based SDDRE controller has not been used for quadrotors in the literature so far which makes the first contribution of this work. The SDDRE is a finite time controller that penalizes the states (error variables) by a final boundary condition [Reference Cimen40].

The singularity-free attitude control is the principal advantage of the quaternion. The Cobra is a challenging aerobatic flight maneuver, performed by a jet aircraft [Reference Wang, Ma, Chang and Zhang41, Reference Gal-Or and Baumann42]. The motion turns the aircraft vertical (even for pitch angle $ \theta \gt \pi /2$ ) to perform the maneuver along with sudden deceleration; the thrust of the jet engine helps to avoid the system from falling.

A tail-sitter drone is a good choice, with a vertical thrust option, to perform the Cobra maneuver [Reference Xu and Zhang43]. Xu et al. presented iterative learning control for a tail sitter unmanned aerial vertical-take-off-and-landing system for Pugachev’s Cobra maneuver [Reference Xu and Zhang43]. They used the acceleration model that resulted in simple dynamics without system identification. The Cobra was exercised by a quadrotor using adaptive control [Reference Chen and Pérez-Arancibia44]. The quadrotor possessed $ 28(\textrm{g})$ , a small lightweight platform. An adaptive backstepping controller with a modified recursive least-square was employed to control the system.

Contributions: (1) presenting a quaternion-based SDDRE control peculiar to a quadrotor. (2) Implementing Cobra maneuver in pitch angle in a forward flight. The performed Cobra in ref. [Reference Chen and Pérez-Arancibia44] was done in vertical ascending flight. In a forward flight, conducting a Cobra maneuver, the system will lose its thrust ( $ \theta \approx \pi /2$ ) and is subjected to fall. Here in this work, using SDDRE, the Cobra in forward flight is done.

Section 2 describes the preliminaries in quaternion mathematics. Section 3 states the system dynamics details. Section 4 presents the SDDRE control structure. Section 5 expresses the approximate closed-form solution to the SDDRE. Section 6 presents the implementation and method and cascade design. Simulations are reported in Section 7 which includes validation and aerobatic flight, and concluding remarks are summarized in Section 8.

Notations: $ {\mathbb{R}}^{n}$ denotes the $ n$ -dimensional Euclidean space, $ {\mathbb{H}}^{n}$ denotes the $ n$ -dimensional Hamilton space, and ${\left(\cdot \right)}^{*}$ performs conjugate transpose. $ {\mathbb{R}}^{n\times m}$ is the set of $ n\times m$ real matrices; $ {\left(\cdot \right)}^{T}$ is the transpose of a matrix or a vector; $\bigotimes$ denotes Kronecker product; $ \textrm{d}\textrm{i}\textrm{a}\textrm{g}\left(\cdot \right)$ means a diagonal matrix; $ {\textbf{I}}_{n\times n}$ and ${\textbf{0}}_{n\times n}$ denote $ n\times n$ identity and zero matrices.

2. Preliminaries: Quaternion mathematics

Here we consider the quaternion definition with real-scalar part $ {q}_{0}$ and vector-imaginary part $ {\textbf{q}}_{\textrm{v}}={q}_{1}\textbf{i}+{q}_{2}\textbf{j}+{q}_{3}\textbf{k}$ , set all together

(1) \begin{align}\textbf{q}=\left[\begin{array}{c}{q}_{0}\\ \\[-7pt] {\textbf{q}}_{\textrm{v}}\end{array}\right]=\left[\begin{array}{c}\textrm{cos}\dfrac{\vartheta }{2}\\ \\[-7pt] \textbf{r}\textrm{sin}\dfrac{\vartheta }{2}\end{array}\right]\!,\end{align}

where $ \textbf{r}\in {\mathbb{R}}^{3}$ is a unit rotation vector and $ \vartheta $ is the corresponding rotation angle about that ref. [Reference Kang, Sadegh and Urschel20]. To solve the ambiguity of the direction of quaternions, $ 0\le {q}_{0}\le 1$ is chosen. A conjugate transpose of the quaternion (1) is presented as

(2) \begin{align}{\textbf{q}}^{\textbf{*}}=\left[\begin{array}{c}{q}_{0}\\ \\[-7pt] -{\textbf{q}}_{\textrm{v}}\end{array}\right].\end{align}

The multiplication product of two arbitrary quaternions $ \textbf{q}$ and $ \textbf{p}$ is defined through Kronecker product

(3) \begin{align} \nonumber\\[-25pt]\textbf{q}\otimes \textbf{p}=\textbf{Q}(\textbf{q})\textbf{p}=\left[\begin{array}{c@{\quad}c@{\quad}c@{\quad}c}{q}_{0} & -{q}_{1} & -{q}_{2} &-{q}_{3}\\ \\[-7pt] {q}_{1} &{q}_{0} &-{q}_{3} & {q}_{2}\\ \\[-7pt] {q}_{2} & {q}_{3} & {q}_{0} & -{q}_{1}\\ \\[-7pt] {q}_{3} & -{q}_{2} & {q}_{1} &{q}_{0}\end{array}\right]\left[\begin{array}{c}{p}_{0}\\ \\[-7pt] {p}_{1}\\ \\[-7pt] {p}_{2}\\ \\[-7pt] {p}_{3}\end{array}\right].\end{align}

A unit quaternion can build a rotation transformation by two multiplications by Kronecker product that could rotate an arbitrary vector $ \textbf{v}$ from the global coordinate to a moving coordinate $ \textbf{q}$ as in the form of $ \textbf{q}\otimes {\left[0,{\textbf{v}}^{T}\right]}^{T}\otimes {\textbf{q}}^{\textrm{*}}$ . So, using definitions (2) and (3), the rotation matrix is built by replacing $ x,y,z$ in turn into $ \textbf{v}$ :

\begin{align*}{\textbf{R}}_{x}(\textbf{q})=\textbf{q}\otimes {\left[\textrm{0,1},\textrm{0,0}\right]}^{T}\otimes {\textbf{q}}^{\textbf{*}},\\[3pt]{\textbf{R}}_{y}(\textbf{q})=\textbf{q}\otimes {\left[\textrm{0,0,1,0}\right]}^{T}\otimes {\textbf{q}}^{\textbf{*}},\\[3pt]{\textbf{R}}_{z}(\textbf{q})=\textbf{q}\otimes {\left[\textrm{0,0,0,1}\right]}^{T}\otimes {\textbf{q}}^{\textbf{*}},\end{align*}

which form [Reference Fresk and Nikolakopoulos25]:

\begin{align*}\textbf{R}(\textbf{q})=\left[{\textbf{R}}_{x}(2:\ 4),{\textbf{R}}_{y}(2:\ 4),{\textbf{R}}_{z}(2:\ 4)\right]=\left[\begin{array}{c@{\quad}c@{\quad}c}{q}_{0}^{2}+{q}_{1}^{2}-{q}_{2}^{2}-{q}_{3}^{2} & 2{q}_{1}{q}_{2}-2{q}_{0}{q}_{3} & 2{q}_{0}{q}_{2}+2{q}_{1}{q}_{3}\\ \\[-7pt] 2{q}_{0}{q}_{3}+2{q}_{1}{q}_{2} &{q}_{0}^{2}-{q}_{1}^{2}+{q}_{2}^{2}-{q}_{3}^{2} & 2{q}_{2}{q}_{3}-2{q}_{0}{q}_{1}\\ \\[-7pt] 2{q}_{1}{q}_{3}-2{q}_{0}{q}_{2} &2{q}_{0}{q}_{1}+2{q}_{2}{q}_{3} & {q}_{0}^{2}-{q}_{1}^{2}-{q}_{2}^{2}+{q}_{3}^{2}\end{array}\right].\end{align*}

where that is $ {\textbf{R}}_{x}(2:\ 4)$ selects three last components of $ {\textbf{R}}_{x}$ .

The angular velocity vector of the quaternion is accessible (supposedly) in local moving coordinate; it is presented by the vector $ {\boldsymbol\omega }\left(\textbf{q},\dot{\textbf{q}}\right)\!:\mathbb{H}\times {\mathbb{R}}^{4}\to {\mathbb{R}}^{3}$ where $ {\boldsymbol\omega }={\left[{\omega }_{1},{\omega }_{2},{\omega }_{3}\right]}^{T}\left(\dfrac{\textrm{r}\textrm{a}\textrm{d}}{\textrm{s}}\right)$ . Then the derivative of the quaternion is found [Reference Diebel45]:

(4) \begin{align}{\dot{\textbf{q}}}_{{\boldsymbol\omega }}=\dfrac{1}{2}\textbf{q}\otimes \left[\begin{array}{c}0\\ \\[-7pt] \boldsymbol\omega \end{array}\right]=\frac{1}{2}\textbf{Q}(\textbf{q})\left[\begin{array}{c}0\\ \\[-7pt] \boldsymbol\omega \end{array}\right]\!,\end{align}

in which $ {\dot{\textbf{q}}}_{{\boldsymbol\omega }}(\textbf{q},{\boldsymbol\omega }):\ \mathbb{H}\times {\mathbb{R}}^{3}\to {\mathbb{R}}^{4}$ and $ \textbf{Q}(\textbf{q})$ has been introduced in Eq. (3).

The relation between the Euler angles $ (\varphi ,\theta ,\psi )$ and quaternions is also needed for the control design [Reference Fresk and Nikolakopoulos25]:

(5) \begin{align}\textbf{q}=\left[\begin{array}{c}\textrm{cos}\dfrac{\varphi }{2}\textrm{cos}\dfrac{\theta }{2}\textrm{cos}\dfrac{\psi }{2}+\textrm{sin}\dfrac{\varphi }{2}\textrm{sin}\dfrac{\theta }{2}\textrm{sin}\dfrac{\psi }{2}\\ \\[-7pt] \textrm{sin}\dfrac{\varphi }{2}\textrm{cos}\dfrac{\theta }{2}\textrm{cos}\dfrac{\psi }{2}-\textrm{cos}\dfrac{\varphi }{2}\textrm{sin}\dfrac{\theta }{2}\textrm{sin}\dfrac{\psi }{2}\\ \\[-7pt] \textrm{cos}\dfrac{\varphi }{2}\textrm{sin}\dfrac{\theta }{2}\textrm{cos}\dfrac{\psi }{2}+\textrm{sin}\dfrac{\varphi }{2}\textrm{cos}\dfrac{\theta }{2}\textrm{sin}\dfrac{\psi }{2}\\ \\[-7pt] \textrm{cos}\dfrac{\varphi }{2}\textrm{cos}\dfrac{\theta }{2}\textrm{sin}\dfrac{\psi }{2}-\textrm{sin}\dfrac{\varphi }{2}\textrm{sin}\dfrac{\theta }{2}\textrm{cos}\dfrac{\psi }{2}\end{array}\right]\!,\end{align}

and also with inverse mapping, one could find:

\begin{align*}\left[\begin{array}{c}\varphi \\ \\[-7pt] \theta \\ \\[-7pt] \psi \end{array}\right]=\left[\begin{array}{c}\textrm{a}\textrm{t}\textrm{a}\textrm{n}2\left(2\left({q}_{0}{q}_{1}+{q}_{2}{q}_{3}\right),{q}_{0}^{2}-{q}_{1}^{2}-{q}_{2}^{2}+{q}_{3}^{2}\right)\\ \\[-7pt] \textrm{a}\textrm{s}\textrm{i}\textrm{n}\left(2\left({q}_{0}{q}_{2}-{q}_{1}{q}_{3}\right)\right)\\ \\[-7pt] \textrm{a}\textrm{t}\textrm{a}\textrm{n}2\left(2\left({q}_{0}{q}_{3}+{q}_{1}{q}_{2}\right),{q}_{0}^{2}+{q}_{1}^{2}-{q}_{2}^{2}-{q}_{3}^{2}\right)\end{array}\right].\end{align*}

3. Dynamics of the system

Consider a “plus-shaped” quadrotor drone with two moving and fixed reference coordinates, body frame denoted by $ \mathcal{B}$ , and global frame marked with $ \mathcal{M}=\left\{X,Y,Z\right\}$ , with respect, see Fig. 1. The position of the moving coordinate is defined through the vector $ {{\boldsymbol\xi }}_{1}(t)={\left[{x}_{\textrm{c}}(t),{y}_{\textrm{c}}(t),{z}_{\textrm{c}}(t)\right]}^{T}\left(\textrm{m}\right).$ The kinematics equation is

(6) \begin{align}{\dot{{\boldsymbol\xi }}}_{1}(t)=\textbf{R}(\textbf{q}(t)){{\boldsymbol\upsilon }}_{1}(t),\end{align}

where $ {{\boldsymbol\upsilon }}_{1}(t)={\left[u(t),v(t),w(t)\right]}^{T}\left(\frac{\textrm{m}}{\textrm{s}}\right)$ , and $ \textbf{R}(\textbf{q}):\mathbb{H}\to {\mathbb{R}}^{3\times 3}$ is the quaternion-based rotation matrix; and $ \left\{\varphi ,\theta ,\psi \right\}\left(\textrm{r}\textrm{a}\textrm{d}\right)$ are Euler angles set in global coordinate. The local angular velocity vector set on the body frame is also named $ {{\boldsymbol\upsilon }}_{2}(t)={\left[p(t),q(t),r(t)\right]}^{T}\left(\frac{\textrm{r}\textrm{a}\textrm{d}}{\textrm{s}}\right)$ .

Fig. 1. The definition of the reference coordinates for a sample quadrotor drone.

To find the dynamics of the system, Newton–Euler equation could be used that results in two sets of dynamic equations, the first set is translational:

(7) \begin{align}m{\ddot{{\boldsymbol\xi }}}_{1}(t)={\textbf{R}}_{3}(\textbf{q}(t)){T}_{\textrm{B}}(t)-mg{\textbf{e}}_{3},\end{align}

where $ {\textbf{R}}_{3}(\textbf{q})$ is the last column of $ \textbf{R}(\textbf{q})$ , $ g=9.81\left(\frac{\textrm{m}}{{\textrm{s}}^{2}}\right)$ is the gravity constant, $ m\left(\textrm{k}\textrm{g}\right)$ is the total mass of the drone, and $ {\textbf{e}}_{3}={\left[\textrm{0,0,1}\right]}^{T}$ . The second set is rotational dynamic:

(8) \begin{align}\textbf{J}{\dot{{\boldsymbol\upsilon }}}_{2}(t)={{\boldsymbol\tau }}_{\textrm{B}}(t)-{{\boldsymbol\upsilon }}_{2}(t)\times \textbf{J}{{\boldsymbol\upsilon }}_{2}(t),\end{align}

where $ {{\boldsymbol\tau }}_{\textrm{B}}(t)={\left[{\tau }_{x}(t),{\tau }_{y}(t),{\tau }_{z}(t)\right]}^{T}\left(\textrm{N}\textrm{m}\right)$ is the input torque vector and $ \textbf{J}=\textrm{d}\textrm{i}\textrm{a}\textrm{g}\left({I}_{xx},{I}_{yy},{I}_{zz}\right)\textrm{}\left(\textrm{k}\textrm{g}{\textrm{m}}^{2}\right)$ is the inertia matrix assigned in the body frame.

The state-vector is

\begin{align*}{\left[\textbf{x}(t)\right]}_{13\times 1}={\left[{{\boldsymbol\xi }}_{1}^{T}(t),{\textbf{q}}^{T}(t),{\dot{{\boldsymbol\xi }}}_{1}^{T}(t),{\dot{{\boldsymbol\upsilon }}}_{2}^{T}(t)\right]}^{T},\end{align*}

which provides the state-space representation of the multi-copter using Eqs. (4)–(8) and setting $ {\boldsymbol\omega }(t)={{\boldsymbol\upsilon }}_{2}(t)={\left[p,q,r\right]}^{T}$ in Eq. (4):

(9) \begin{align}\dot{\textbf{x}}(t)=\left[\begin{array}{c}\textbf{R}(\textbf{q}(t)){{\boldsymbol\upsilon }}_{1}(t)\\ \\[-7pt] \dfrac{1}{2}\textbf{Q}(\textbf{q}(t))\left[\begin{array}{c}0\\ \\[-7pt] {{\boldsymbol\upsilon }}_{2}(t)\end{array}\right]\\ \\[-7pt] \dfrac{1}{m}\left({\textbf{R}}_{3}(\textbf{q}(t)){T}_{\textrm{B}}(t)-mg{\textbf{e}}_{3}-\textbf{D}{\dot{{\boldsymbol\xi }}}_{1}(t)\right)\\ \\[-7pt] {\textbf{J}}^{-1}({{\boldsymbol\tau }}_{\textrm{B}}(t)-{{\boldsymbol\upsilon }}_{2}(t)\times \textbf{J}{{\boldsymbol\upsilon }}_{2}(t))\end{array}\right].\end{align}

In state-space Eq. (9), $ \textbf{D}\in {\mathbb{R}}^{3\times 3}$ collects drag and aerodynamics parameters of the quadcopter model and has been added to complete the model.

4. The state-dependent differential Riccati equation controller design

Consider the time-invariant affine-in-control nonlinear system

(10) \begin{align}\dot{\textbf{x}}(t)=\textbf{f}(\textbf{x}(t))+\textbf{g}\left(\textbf{x}(t),\textbf{u}(t)\right),\end{align}

where $ \textbf{x}(t)\in {\mathbb{R}}^{n}$ is the state vector, $ \textbf{u}(t)\in {\mathbb{R}}^{m}$ is the input vector, $ \textbf{f}(\textbf{x}(t)):\ {\mathbb{R}}^{n}\to {\mathbb{R}}^{n}$ and $ \textbf{g}(\textbf{x}(t),\textbf{u}(t)):\ {\mathbb{R}}^{n}\times {\mathbb{R}}^{m}\to {\mathbb{R}}^{n}$ represents the dynamics and they satisfy local Lipschitz condition. $ n$ is the dimension of the state vector and $ m$ is the total number of actuators. The system Eq. (10) is transformed into state-dependent coefficient (SDC) parameterization [Reference Nekoo46]:

(11) \begin{align}\textbf{f}(\textbf{x}(t))=\textbf{A}(\textbf{x}(t))\textbf{x}(t),\end{align}
(12) \begin{align}\textbf{g}(\textbf{x}(t),\textbf{u}(t))=\textbf{B}(\textbf{x}(t))\textbf{u}(t),\\[7pt] \nonumber \end{align}

in which $ \textbf{B}(\textbf{x}(t)):\ {\mathbb{R}}^{n}\to {\mathbb{R}}^{n\times m}$ and $ \textbf{A}(\textbf{x}(t)):\ {\mathbb{R}}^{n}\to {\mathbb{R}}^{n\times n}$ are held. The pair of $ \left\{\textbf{A}(\textbf{x}(t)),\textbf{B}(\textbf{x}(t))\right\}$ is a controllable parameterization of system (10) [Reference Nekoo47].

The cost function of the SDDRE is structured as

(13) \begin{align}J\left(\cdot \right)=\frac{1}{2}{\textbf{x}}^{T}(t)\textbf{F}\textbf{x}(t)+\frac{1}{2}\underset{0}{\overset{{t}_{\textrm{f}}}{\int }}\left[{\textbf{x}}^{T}(t)\textbf{Q}(\textbf{x}(t))\textbf{x}(t)+{\textbf{u}}^{T}(t)\textbf{R}(\textbf{x}(t))\textbf{u}(t)\right]\textrm{d}t,\end{align}

where ${t}_{\textrm{f}}\in {\mathbb{R}}^{+}$ is the final time of the control task, $ \textbf{Q}(\textbf{x}(t)):\ {\mathbb{R}}^{n}\to {\mathbb{R}}^{n\times n}$ , $ \textbf{R}(\textbf{x}(t)){\mathbb{R}}^{n}\to {\mathbb{R}}^{m\times m}$ , and $ \textbf{F}\in {\mathbb{R}}^{n\times n}$ are weighting matrices, for states and inputs in $ t\in [0,{t}_{\textrm{f}})$ and states at the final time $ {t}_{\textrm{f}}$ with respect. The pair of $ \left\{\textbf{A}(\textbf{x}(t)),{\textbf{Q}}^{\frac{1}{2}}(\textbf{x}(t))\right\}$ is an observable parameterization of system (10) where $ {\textbf{Q}}^{\frac{1}{2}}(\textbf{x}(t))$ is the Cholesky decomposition of weighting matrix in (13).

The control law is defined by applying $\frac{\partial \mathcal{H}\left(\textbf{x}(t),\textbf{u}(t),{\boldsymbol\lambda }(t)\right)}{\partial \textbf{u}(t)}=\textbf{0}$ , stationary condition, on Hamiltonian function $\mathcal{H}\left(\textbf{x}(t),\textbf{u}(t),{\boldsymbol\lambda }(t)\right)=J\left(\cdot \right)+{{\boldsymbol\lambda }}^{T}(t)\dot{\textbf{x}}(t)$ :

(14) \begin{align}\textbf{u}(t)=-{\textbf{R}}^{-1}(\textbf{x}(t)){\textbf{B}}^{T}(\textbf{x}(t))\textbf{K}(\textbf{x}(t))\textbf{x}(t),\end{align}

where $ {\boldsymbol\lambda }(t)=\textbf{K}(\textbf{x}(t))\textbf{x}(t)$ is the co-state vector and $\textbf{K}(\textbf{x}(t)):\ {\mathbb{R}}^{n}\to {\mathbb{R}}^{n \times n}$ is the control gain of the control equation, SDDRE.

Using the necessary condition for optimality, $ \frac{\partial \mathcal{H}\left(\textbf{x}(t),\textbf{u}(t),{\boldsymbol\lambda }(t)\right)}{\partial \textbf{x}(t)}=-\dot{{\boldsymbol\lambda }}(t)$ and the derivative of the co-state vector $ \dot{{\boldsymbol\lambda }}=\dot{\textbf{K}}\textbf{x}+\textbf{K}\dot{\textbf{x}}$ , one could find:

\begin{align*}\dot{\textbf{K}}(\textbf{x})+\textbf{K}(\textbf{x})\textbf{A}(\textbf{x})-\textbf{K}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x})\textbf{K}(\textbf{x})+{\textbf{A}}^{T}(\textbf{x})\textbf{K}(\textbf{x})+\textbf{Q}(\textbf{x})+{\boldsymbol\Pi }(\textbf{x})=\textbf{0},\end{align*}

where $ {\boldsymbol\Pi }(\textbf{x})$ collects the derivative terms (see the complete equation in ref. [Reference Nekoo46]), then the SDDRE is represented as [Reference Nekoo48]:

(15) \begin{align}-\dot{\textbf{K}}(\textbf{x})=\textbf{Q}(\textbf{x})-\textbf{K}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x})\textbf{K}(\textbf{x})+\textbf{K}(\textbf{x})\textbf{A}(\textbf{x})+{\textbf{A}}^{T}(\textbf{x})\textbf{K}(\textbf{x}),\end{align}

with $ \textbf{K}\left(\textbf{x}\left({t}_{\textrm{f}}\right)\right)=\textbf{F}$ , a final boundary condition.

5. An approximate closed-form solution to the SDDRE

The quadrotor dynamics in Eq. (9) must be controlled by the SDDRE approach. The extended linearization model of (9) is so-called state-dependent coefficient parameterization (or apparent linearization) [Reference Jayaram and Tadi49]. To solve the SDDRE (15) and to find the control gain, several methods could be used such as backward integration (BI) [Reference Korayem and Nekoo50], state-transition matrix [Reference Mathavaraj and Padhi51], and Lyapunov-based approach [Reference Heydari and Balakrishnan52]. The BI imposes a two-round solution that might not be proper for systems that need frequent changes in initial and final conditions, nonrepetitive systems. The STM approach is not computationally robust when the final time is long [Reference Korayem and Nekoo53], then the Lyapunov-based approach has been selected for this work. It works with both positive and negative solutions to the Riccati equation and delivers an approximate closed-form answer, in just one round. This work uses Lyapunov-based method via negative root to the related SDRE, $ {\textbf{K}}_{\textrm{s}\textrm{s}}^{-}(\textbf{x})$ to find the symmetric positive-definite solution to the SDDRE, $ \textbf{K}(\textbf{x}(t))$ [Reference Korayem and Nekoo50]. Subtracting (15) from

(16) \begin{align}{\textbf{A}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})+{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\textbf{A}(\textbf{x})+\textbf{Q}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})=\textbf{0},\end{align}

generates

(17) \begin{align}-\dot{\textbf{K}}(\textbf{x})&=\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]\textbf{A}(\textbf{x})+{\textbf{A}}^{T}(\textbf{x})\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right] \nonumber\\[3pt]&\quad -\textbf{K}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x})\textbf{K}(\textbf{x})\nonumber\\[3pt]&\quad +{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x}).\end{align}

Adding and subtracting $ \textbf{K}\textbf{B}{\textbf{R}}^{-1}{\textbf{B}}^{T}{\textbf{K}}_{\textrm{s}\textrm{s}}$ , $ {\textbf{K}}_{\textrm{s}\textrm{s}}\textbf{B}{\textbf{R}}^{-1}{\textbf{B}}^{T}\textbf{K}$ and $ {\textbf{K}}_{\textrm{s}\textrm{s}}\textbf{B}{\textbf{R}}^{-1}{\textbf{B}}^{T}{\textbf{K}}_{\textrm{s}\textrm{s}}$ to (17) result in:

(18) \begin{align}-\dot{\textbf{K}}(\textbf{x})&={\textbf{A}}^{T}(\textbf{x})\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]+\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]\textbf{A}(\textbf{x})\nonumber\\&\quad -\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\nonumber\\&\quad -{\left[\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]}^{T}\left[\textbf{K}\right(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x}\left)\right]\nonumber\\&\quad -\textbf{K}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x})\textbf{K}(\textbf{x})\nonumber\\&\quad +\textbf{K}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\nonumber\\&\quad +{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x})\textbf{K}(\textbf{x})\nonumber\\&\quad -{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x}).\end{align}

Rewriting (18):

\begin{align*}-\dot{\textbf{K}}(\textbf{x})&=\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]\textbf{A}(\textbf{x})+{\textbf{A}}^{T}(\textbf{x})\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right] \\&\quad -\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\\&\quad -{\left[\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]}^{T}\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]\\&\quad -\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x})\left[\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x})\right]\!,\end{align*}

and introducing a new variable

\begin{align*}{\textbf{P}}^{-1}(\textbf{x})=\textbf{K}(\textbf{x})-{\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x}),\end{align*}

and expressing the closed-loop matrix of the system

\begin{align*}{\textbf{A}}_{\text{cl}}(\textbf{x})=\textbf{A}(\textbf{x})-\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{K}}_{\textrm{s}\textrm{s}}(\textbf{x}),\end{align*}

one could present

(19) \begin{align}-\dot{\textbf{K}}(\textbf{x})={\textbf{P}}^{-1}(\textbf{x}){\textbf{A}}_{\text{cl}}(\textbf{x})+{\textbf{A}}_{\text{cl}}^{T}(\textbf{x}){\textbf{P}}^{-1}(\textbf{x})-{\textbf{P}}^{-1}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}){\textbf{P}}^{-1}(\textbf{x})\end{align}

Regarding that $ \frac{\textrm{d}}{\textrm{d}t}\left({\textbf{P}}^{-1}(\textbf{x})\right)=-{\textbf{P}}^{-1}(\textbf{x})\dot{\textbf{P}}(\textbf{x}){\textbf{P}}^{-1}(\textbf{x})$ , Eq. (19) should be changed to

\begin{align*}{\textbf{P}}^{-1}(\textbf{x})\dot{\textbf{P}}(\textbf{x}){\textbf{P}}^{-1}(\textbf{x})={\textbf{P}}^{-1}(\textbf{x}){\textbf{A}}_{\text{cl}}(\textbf{x})+{\textbf{A}}_{\text{cl}}^{T}(\textbf{x}){\textbf{P}}^{-1}(\textbf{x})-{\textbf{P}}^{-1}(\textbf{x})\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{}(\textbf{x}){\textbf{P}}^{-1}(\textbf{x}),\end{align*}

and consequently, results in a state-dependent differential Lyapunov Eq. [Reference Korayem and Nekoo50]:

(20) \begin{align}\dot{\textbf{P}}(\textbf{x})={\textbf{A}}_{\text{cl}}(\textbf{x})\textbf{P}(\textbf{x})+\textbf{P}(\textbf{x}){\textbf{A}}_{\text{cl}}^{T}(\textbf{x})-\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x}),\end{align}

with a final boundary condition $ \textbf{P}\left({t}_{\text{f}}\right)={\left[\textbf{F}-{\textbf{K}}_{\text{ss}}(\textbf{x}(t))\right]}^{-1}$ . A solution to (20) is

(21) \begin{align}\textbf{P}(\textbf{x})=\textbf{E}(\textbf{x})+\textrm{e}\textrm{x}\textrm{p}\left\{{\textbf{A}}_{\text{cl}}(\textbf{x})\left(t-{t}_{\text{f}}\right)\right\}\left[\textbf{P}\left({t}_{\text{f}}\right)-\textbf{E}(\textbf{x})\right]\textrm{e}\textrm{x}\textrm{p}\left\{{\textbf{A}}_{\text{cl}}^{T}(\textbf{x})\left(t-{t}_{\text{f}}\right)\right\},\end{align}

in which $ \textbf{E}(\textbf{x}(t))$ is an answer to a state-dependent algebraic Lyapunov equation:

(22) \begin{align}\textbf{E}(\textbf{x}){\textbf{A}}_{\text{cl}}^{T}(\textbf{x})+{\textbf{A}}_{\text{cl}}(\textbf{x})\textbf{E}(\textbf{x})-\textbf{B}(\textbf{x}){\textbf{R}}^{-1}(\textbf{x}){\textbf{B}}^{T}(\textbf{x})=\textbf{0}.\end{align}

Proof of (21) could be checked by the substitution of Eq. (21) into (20):

(23) \begin{align}&{\dot{\textbf{E}}}({\textbf{x}}) + {{\textrm{d}\left( {{{\textbf{A}}_{\textrm{cl}}}({\textbf{x}})\left( {t - {t_{\textrm{f}}}} \right)} \right)} \over {\textrm{d}t}}{\textrm{exp}}\{ {{\textbf{A}}_{\textrm{cl}}}({\textbf{x}})(t - {t_\textrm{f}})\} [{\textbf{P}}({t_\textrm{f}}) - {\textbf{E}}({\textbf{x}})]{\textrm{exp}}\{ {\textbf{A}}_{\textrm{cl}}^T({\textbf{x}})(t - {t_\textrm{f}})\}\nonumber\\[3pt]&\quad - {\textrm{exp}}\{ {{\textbf{A}}_{\textrm{cl}}}({\textbf{x}})(t - {t_\textrm{f}})\} {\dot{\textbf{E}}}({\textbf{x}}){\textrm{exp}}\{ {\textbf{A}}_{\textrm{cl}}^T({\textbf{x}})(t - {t_\textrm{f}})\}\nonumber\\[3pt] &\quad + {\textrm{exp}}\{ {{\textbf{A}}_{\textrm{cl}}}({\textbf{x}})(t - {t_\textrm{f}})\} [{\textbf{P}}({t_\textrm{f}}) - {\textbf{E}}({\textbf{x}})]{\textrm{exp}}\{ {\textbf{A}}_{\textrm{cl}}^T({\textbf{x}})(t - {t_\textrm{f}})\} {{\textrm{d}\left( {{\textbf{A}}_{\textrm{cl}}^T({\textbf{x}})\left( {t - {t_\textrm{f}}} \right)} \right)} \over {\textrm{d}t}}\nonumber\\[3pt] &= {\textbf{E}}({\textbf{x}}){\textbf{A}}_{\textrm{cl}}^T({\textbf{x}}) + {\textrm{exp}}\{ {{\textbf{A}}_{\textrm{cl}}}({\textbf{x}})(t - {t_\textrm{f}})\} [{\textbf{P}}({t_\textrm{f}}) - {\textbf{E}}({\textbf{x}})]{\textrm{exp}}\{ {\textbf{A}}_{\textrm{cl}}^T({\textbf{x}})(t - {t_\textrm{f}})\} {\textbf{A}}_{\textrm{cl}}^T({\textbf{x}}) + {{\textbf{A}}_{\textrm{cl}}}({\textbf{x}}){\textbf{E}}({\textbf{x}})\nonumber\\[3pt] &\quad + {{\textbf{A}}_{\textrm{cl}}}({\textbf{x}}){\textrm{exp}}\{ {{\textbf{A}}_{\textrm{cl}}}({\textbf{x}})(t - {t_\textrm{f}})\} [{\textbf{P}}({t_\textrm{f}}) - {\textbf{E}}({\textbf{x}})]{\textrm{exp}}\{ {\textbf{A}}_{\textrm{cl}}^T({\textbf{x}})(t - {t_\textrm{f}})\} - {\textbf{B}}({\textbf{x}}){{\textbf{R}}^{ - 1}}({\textbf{x}}){{\textbf{B}}^T}({\textbf{x}}).\end{align}

From (22) we have

(24) \begin{align}{\textbf{B}}({\textbf{x}}){{\textbf{R}}^{ - 1}}({\textbf{x}}){{\textbf{B}}^T}({\textbf{x}}) = {{\textbf{A}}_{{\textrm{cl}}}}({\textbf{x}}){\textbf{E}}({\textbf{x}}) + {\textbf{E}}({\textbf{x}}){\textbf{A}}_{{\textrm{cl}}}^T({\textbf{x}}).\end{align}

The algebraic Lyapunov Eq. (22) results in ${\dot{\textbf{E}}}({\textbf{x}}) = \textbf{0}$ . Regarding frozen computation at each simulation time-step, we neglect the time derivative of ${{\textbf{A}}_{{\textrm{cl}}}}({\textbf{x}})$ and as a result, ${{{\textrm{d}}\left( {{{\textbf{A}}_{{\textrm{cl}}}}({\textbf{x}})\left( {t - {t_{\textrm{f}}}} \right)} \right)} \over {{\textrm{d}}t}} = {{\textbf{A}}_{{\textrm{cl}}}}({\textbf{x}}) + \underbrace {{\textrm{ }}{{{\dot{\textbf{A}}}}_{{\textrm{cl}}}}({\textbf{x}})\left( {t - {t_{\textrm{f}}}} \right)}_{\textbf{0}}$ . Substituting (24) into (23), mathematical manipulation cancels all terms and shows that the solution (21) holds for Eq. (20). Since we neglected ${{\dot{\textbf{A}}}_{{\textrm{cl}}}}({\textbf{x}})\left( {t - {t_{\textrm{f}}}} \right) \approx \textbf{0}$ in the derivative $\frac{{{\textrm{d}}\left( {{{\textbf{A}}_{{\textrm{cl}}}}({\textbf{x}})\left( {t - {t_{\textrm{f}}}} \right)} \right)}}{{{\textrm{d}}t}}$ and considered ${\dot{\textbf{E}}}({\textbf{x}}) = \textbf{0}$ , this approach is so-called a closed-form approximate solution.

The positive gain of the SDDRE (15) is regarded as ${\textbf{K}}\left( {{\textbf{x}}(t)} \right) = {{\textbf{K}}_{{\textrm{ss}}}}\left( {{\textbf{x}}(t)} \right) + {{\textbf{P}}^{ - 1}}\left( {{\textbf{x}}(t)} \right),$ in which ${{\textbf{K}}_{{\textrm{ss}}}}\left( {{\textbf{x}}(t)} \right)$ could be a negative definite ${\textbf{K}}_{{\textrm{ss}}}^ - \left( {{\textbf{x}}(t)} \right)$ or positive definite ${\textbf{K}}_{{\textrm{ss}}}^ + \left( {{\textbf{x}}(t)} \right)$ solution to the SDRE (16). It works with both of them.

The details of the positive and negative roots of the SDRE are reported in Sections 3.1.1 and 3.3.2 of Ref. [Reference Korayem and Nekoo50]. The negative root is computationally more robust than the positive one, and it has been used here in this work. Finally, notice that the negative definite solution to the SDRE (16) is the positive definite answer to

\begin{align*}- {\textbf{K}}_{\textrm{n}}^ + ({\textbf{x}}){\textbf{A}}({\textbf{x}}) - {{\textbf{A}}^T}({\textbf{x}}){\textbf{K}}_{\textrm{n}}^ + ({\textbf{x}}) - {\textbf{K}}_{\textrm{n}}^ + ({\textbf{x}}){\textbf{B}}({\textbf{x}}){{\textbf{R}}^{ - 1}}({\textbf{x}}){{\textbf{B}}^T}({\textbf{x}}){\textbf{K}}_{\textrm{n}}^ + ({\textbf{x}}) + {\textbf{Q}}({\textbf{x}}) = \textbf{0},\end{align*}

with consideration of ${\textbf{K}}_{{\textrm{ss}}}^ - \left( {{\textbf{x}}(t)} \right) = - {\textbf{K}}_{\textrm{n}}^ + \left( {{\textbf{x}}(t)} \right)$ .

6. Implementation of the SDDRE on quaternion-based dynamics

The state-space equation of the system (9) must be represented by SDC forms (11) and (12). On the one hand, the dimension of the state vector (13 states) is not compatible with the cascade approach, commonly used in quadrotor control (12 states) [Reference Nekoo, Acosta and Ollero5]. On the other hand, the separation of rotational and translational dynamics was reported helpful in the control design due to different speeds of them, slow and fast dynamic [Reference Nekoo, Acosta, Gomez-Tamm and Ollero54]. So, here we propose two subcontrollers for the translation and rotational dynamics, connected through cascade design. For the translation dynamic, the set of SDC parameterization is

\begin{align*}{{\textbf{A}}_\textrm{t}}( {{\textbf{x}}(t)}) = \left[ {\begin{array}{c@{\quad}c}{{\textbf{0}_{3 \times 3}}} {}& {{\textbf{R}}({{\textbf{q}}(t)})}\\ \\[-7pt] {{\textbf{0}_{3 \times 3}}}& {}{ - \dfrac{{\textbf{D}}}{m}}\end{array}} \right]\!,{{\textbf{B}}_\textrm{t}} = \left[ {\begin{array}{*{20}{c}}{{\textbf{0}_{3 \times 3}}}\\ \\[-7pt] {\dfrac{{{{\textbf{I}}_{3 \times 3}}}}{m}}\end{array}} \right]\!,\end{align*}

in which hovering condition has been assumed to find ${\dot {\boldsymbol\xi} _1} = \underbrace {{\textbf{R}}({\textbf{q}})}_{\textbf{I}}{\boldsymbol\upsilon _1} \approx {\boldsymbol\upsilon _1}$ , to make the factorization possible, and “t” stands for translation. For the orientation section, we neglect the scalar part of the quaternion, the first column, and the row of ${\textbf{Q}}({{\textbf{q}}(t)})$ in (9), then the SDC parameterization is

\begin{align*}{{\textbf{A}}_{\textrm{o}}}( {{\textbf{x}}(t)}) = \left[ {\begin{array}{c@{\quad}c}{{\textbf{0}_{3 \times 3}}}& {{{\left[ {{\textbf{Q}}\left( {2:\ 4,2:\ 4} \right)} \right]}_{3 \times 3}}}\\ \\[-7pt] {{\textbf{0}_{3 \times 3}}} &{ - {{\textbf{J}}^{ - 1}}{{\left[ {{{{\boldsymbol\upsilon }}_2} \times {\textbf{I}}} \right]}_{3 \times 3}}}\end{array}} \right]\!,{{\textbf{B}}_{\textrm{o}}} = \left[ \begin{array}{c}{{\textbf{0}_{3 \times 3}}}\\ \\[-7pt] {{{\textbf{J}}^{ - 1}}}\end{array} \right]\!,\end{align*}

where ${\textbf{Q}}\left( {2:\ 4,2:\ 4} \right)$ collects the second to fourth components of ${\textbf{Q}}({{\textbf{q}}(t)})$ in columns and rows, and “o” stands for orientation. Based on (14), the translation control law for regulation to the desired condition rather than zero is

(25) \begin{align}{{\textbf{u}}_\textrm{t}}(t) = - {\textbf{R}}_\textrm{t}^{ - 1}\left( {{\textbf{x}}(t)} \right){\textbf{B}}_\textrm{t}^T{{\textbf{K}}_\textrm{t}}\left( {{\textbf{x}}(t)} \right)\left[ {\begin{array}{*{20}{c}}{{{{\boldsymbol\xi }}_1}(t) - {{{\boldsymbol\xi }}_{1,{\textrm{des}}}}}\\ \\[-7pt] {{{{\dot{\boldsymbol{\xi}}}}_1}(t) - {{{\dot{\boldsymbol{\xi}}}}_{1,{\textrm{des}}}}}\end{array}} \right]\!,\end{align}

where ${{\textbf{R}}_{}}$ , ${{\textbf{Q}}_\textrm{t}}$ , ${{\textbf{F}}_\textrm{t}}$ , and ${{\textbf{K}}_\textrm{t}}$ are the corresponding matrices (with proper dimension) for the translation and Riccati equation and “des” defines the desired condition.

The computation of input law (25) is based on a fully actuated system, which is not possible in reality. So, to transform the ${\left[ {{{\textbf{u}}_{\textrm{t}}}(t)} \right]_{3 \times 1}}$ to the scalar thrust input considering the gravity as well, we use cascade design [Reference Nekoo, Acosta and Ollero55]:

\begin{align*}{T_{\textrm{B}}}(t) = m\left( {{{\left[ {{{\textbf{R}}_3}({{\textbf{q}}(t)})} \right]}_1}{u_{\textrm{t},1}}(t) + {{\left[ {{{\textbf{R}}_3}({{\textbf{q}}(t)})} \right]}_2}{u_{\textrm{t},2}}(t) + {{\left[ {{{\textbf{R}}_3}({{\textbf{q}}(t)})} \right]}_3}\left( {{u_{\textrm{t},3}}(t) + g} \right)} \right),\end{align*}

where that is ${\left[ {{{\textbf{R}}_3}({{\textbf{q}}(t)})} \right]_1}$ is the first component of ${{\textbf{R}}_3}({{\textbf{q}}(t)})$ . The cascade design delivers the necessary desired Euler angles

(26) \begin{align}{\theta _{{\textrm{des}}}}(t) = {\textrm{ta}}{{\textrm{n}}^{ - 1}}\left( {\frac{{{u_{\textrm{t},1}}{\textrm{cos}}{\psi _{{\textrm{des}}}} + {u_{\textrm{t},2}}{\textrm{sin}}{\psi _{{\textrm{des}}}}}}{{{u_{\textrm{t},3}} + g}}} \right),\end{align}
(27) \begin{align}{\phi _{{\textrm{des}}}}(t) = {\textrm{si}}{{\textrm{n}}^{ - 1}}\left( {\frac{{{u_{\textrm{t},1}}{\textrm{sin}}{\psi _{{\textrm{des}}}} - {u_{\textrm{t},2}}{\textrm{cos}}{\psi _{{\textrm{des}}}}}}{{\sqrt {u_{\textrm{t},1}^2 + u_{\textrm{t},2}^2 + {{\left( {{u_{\textrm{t},3}} + g} \right)}^2}} }}} \right),\\ \nonumber\end{align}

in which ${\psi _{{\textrm{des}}}}$ could possess an arbitrary value.

The orientation control law is also presented as

(28) \begin{align}{{\textbf{u}}_{\textrm{o}}}(t) = - {\textbf{R}}_{\textrm{o}}^{ - 1}\left( {{\textbf{x}}(t)} \right){\textbf{B}}_{\textrm{o}}^T{{\textbf{K}}_{\textrm{o}}}\left( {{\textbf{x}}(t)} \right)\left[ {\begin{array}{*{20}{c}}{\mathcal{F}{{\textbf{e}}_{\textbf{q}}}(t)}\\ \\[-7pt] {{{{\boldsymbol\upsilon }}_2}(t) - {{{\boldsymbol\upsilon }}_{2,{\textrm{des}}}}}\end{array}} \right]\!,\end{align}

where ${{\textbf{e}}_{\textbf{q}}}(t)$ is quaternion error and

\begin{align*}\mathcal{F} = \left[ {\begin{array}{c@{\quad}c@{\quad}c@{\quad}c}0 &0& 0& 0\\ \\[-7pt]0& 1 &0& 0\\ \\[-7pt] 0& 0& 1 &0\\ \\[-7pt] 0 &0 &0& 1\end{array}} \right].\end{align*}

The quaternion error in (28) is defined as

\begin{align*}{{\textbf{e}}_{\textbf{q}}}(t) = - {{\textbf{q}}_{{\textrm{des}}}}(t) \otimes {{\textbf{q}}^*}(t) = - {\textbf{Q}}\left( {{{\textbf{q}}_{{\textrm{des}}}}(t)} \right){{\textbf{q}}^*}(t),\end{align*}

where ${{\textbf{q}}_{{\textrm{des}}}}(t)$ is defined by substituting (26), (27), and ${\psi _{{\textrm{des}}}}$ into Eq. (5).

7. Simulations

7.1. Validation

To validate the proposed controller design and the simulator, a comparison has been done with the PD controller and a conventional sliding mode control (SMC). We consider a system based on the Euler angles in rotational dynamics, controlled by a simple PD to have it as a reference. Then the SDDRE controller is implemented on the quaternion-based dynamics to check the performance and also to validate the correctness of the implementation. The SMC has been frequently used in UAV control [Reference Xu and Ozguner56], in the context of robustness or combined with other techniques [Reference Abro, Zulkifli, Asirvadam and Ali57]. Based on that, the SMC is selected for comparison to add the more detailed result.

The mass of the system is $m = 0.468\left( {{\textrm{kg}}} \right)$ , the drag coefficient matrix is ${\textbf{D}} = {\textrm{diag}}\left( {0.25,0.25,0.25} \right)\left( {\frac{{{\textrm{kg}}}}{{\textrm{s}}}} \right)$ , the components of the inertia matrix are ${I_{xx}} = 4.856 \times {10^{ - 3}}\left( {{\textrm{kg}}{{\textrm{m}}^2}} \right)$ , ${I_{yy}} = 4.856 \times {10^{ - 3}}\left( {{\textrm{kg}}{{\textrm{m}}^2}} \right)$ , and ${I_{zz}} = 8.801 \times {10^{ - 3}}\left( {{\textrm{kg}}{{\textrm{m}}^2}} \right)$ . The time of the simulation has been set ${t_\textrm{f}} = 10\left( {\textrm{s}} \right)$ , and the drone flies from zero coordinate to the desired position in Cartesian coordinate $\left( { - 2,3,1.5} \right)\left( {\textrm{m}} \right)$ . All of the initial conditions are set zero including position, velocity, Euler angles, and angular velocity. The initial condition of the quaternions is found by substituting $\left( {\phi \left( 0 \right),\theta \left( 0 \right),\psi \left( 0 \right)} \right)$ into (5) to reach ${\textbf{q}}\left( 0 \right) = {\left[ {1,0,0,0} \right]^T}$ .

The PD control gains are set as ${{\textbf{K}}_{\textrm{P,t}}} = {{\textbf{I}}_{3 \times 3}}$ , ${{\textbf{K}}_{\textrm{D,t}}} = 2 \times {{\textbf{I}}_{3 \times 3}}$ , ${{\textbf{K}}_{\textrm{P,o}}} = {{\textbf{I}}_{3 \times 3}}$ , and ${{\textbf{K}}_{\textrm{D,o}}} = 0.5 \times {{\textbf{I}}_{3 \times 3}}$ . The SDDRE controller gains are selected as follows: ${{\textbf{R}}_\textrm{t}} = {{\textbf{I}}_{3 \times 3}}$ , ${{\textbf{Q}}_\textrm{t}} = {\textrm{diag}}\left( {0.1,0.1,0.1,0.2,0.2,0.2} \right)$ , ${{\textbf{F}}_\textrm{t}} = 100 \times {{\textbf{Q}}_\textrm{t}}$ , ${{\textbf{R}}_{\textrm{o}}} = {{\textbf{I}}_{3 \times 3}}$ , ${{\textbf{Q}}_{\textrm{o}}} = {\textrm{diag}}\left( {2,2,2,0,0,0} \right)$ , ${{\textbf{F}}_\textrm{o}} = 10 \times {{\textbf{Q}}_\textrm{o}}$ .

Two separate SMC controllers are considered for translation and orientation parts, consistent with Section 6. The sliding surface is $\textbf{s}_{i}(\textbf{X})=\dot{\tilde{\textbf{X}}}_{i}+\boldsymbol\Lambda_{i}\tilde{\textbf{X}}_{i}$ for $i = \left\{ {\textrm{t,o}} \right\}$ (translation and orientation) where $\tilde{\textbf{X}}_{i}=\boldsymbol\xi_{i}-\boldsymbol\xi_{i,\textrm{des}}$ and ${{{\boldsymbol\Lambda }}_i}$ is a strictly positive constant matrix. The control law is also in the form of ${{\textbf{u}}_i} = {\textbf{B}}_{i,{\textrm{SMC}}}^{ - 1}({\textbf{x}})\left( {{{{\ddot{\boldsymbol\xi}}}_{i,{\textrm{des}}}} - {{\textbf{f}}_{i,{\textrm{SMC}}}}({\textbf{x}}) - {{\textbf{K}}_{i,{\textrm{SMC}}}}\tanh \left( {\frac{{{{\textbf{s}}_i}({\textbf{x}})}}{\sigma }} \right)} \right)$ , where ${{\textbf{B}}_{\textrm{t},{\textrm{SMC}}}} = 1/m{{\textbf{I}}_{3 \times 3}}$ , ${{\textbf{B}}_{\textrm{o},{\textrm{SMC}}}}({\textbf{x}}) = {\textbf{J}}_\textrm{c}^{ - 1}\left( {{{{\boldsymbol\xi }}_2}} \right)$ , ${{\textbf{f}}_{\textrm{t},{\textrm{SMC}}}}({\textbf{x}}) = - \frac{1}{m}{\textbf{D}}{{\dot{\boldsymbol{\xi}}}_1}(t)$ , ${{\textbf{f}}_{\textrm{o},{\textrm{SMC}}}}({\textbf{x}}) = - {{\textbf{J}}^{ - 1}}\left( {{{{\boldsymbol\xi }}_2}} \right){\textbf{C}}\left( {{{{\boldsymbol\xi }}_2},{{{\dot{\boldsymbol{\xi}}}}_2}} \right){{\dot{\boldsymbol{\xi}}}_2}$ , and ${{\textbf{K}}_{i,{\textrm{SMC}}}}$ is the correction gain of the SMC. More details on conventional dynamics, such as ${{\textbf{J}}_\textrm{c}}\left( {{{{\boldsymbol\xi }}_2}} \right) = {{\textbf{W}}^T}\left( {{{{\boldsymbol\xi }}_2}} \right){\textbf{IW}}\left( {{{{\boldsymbol\xi }}_2}} \right)$ , can be found in ref. [Reference Nekoo, Acosta, Gomez-Tamm and Ollero54]. To avoid chattering in SMC, $\tanh \left( {\frac{{{{\textbf{s}}_i}({\textbf{x}})}}{\sigma }} \right)$ is used where $\sigma = 0.2$ in this simulation. The SMC control parameters are selected as ${{{\boldsymbol\Lambda }}_\textrm{t}} = {\textrm{diag}}\left( {0.5,0.5,1} \right)$ , ${{{\boldsymbol\Lambda }}_\textrm{o}} = 1.5\times{{\textbf{I}}_{3 \times 3}}$ , ${{\textbf{K}}_{\textrm{t},{\textrm{SMC}}}} = {\textrm{diag}}\left( {5,5,1} \right)$ , ${{\textbf{K}}_{\textrm{o},{\textrm{SMC}}}} = 2.5\times{{\textbf{I}}_{3 \times 3}}$ , and desired accelerations are set zero, ${{\ddot{\boldsymbol\xi}}_{i,{\textrm{des}}}} = \textbf{0}$ .

Simulating the system, the results are found in the following. The position variables of the drone are illustrated in Fig. 2 to Fig. 4 with respect. The roll and pitch angles of the multi-copter are plotted in Fig. 5 and Fig. 6 with respect. The input norm of the quadrotor is illustrated in Fig. 7 and the configuration and trajectories of the drones are shown in Fig. 8.

Fig. 2. x-axis regulation of the system, comparison with PD and SMC.

Fig. 3. y-axis regulation of the system, comparison with PD and SMC.

Fig. 4. z-axis regulation of the system, comparison with PD and SMC.

Fig. 5. The roll angle of the system, comparison with PD and SMC.

Fig. 6. The pitch angle of the system, comparison with PD and SMC.

Fig. 7. The input norm of the inputs of the system, comparison with PD and SMC.

Fig. 8. The configuration and trajectories of the quadrotor drones with PD, SMC, and SDDRE controllers.

The error of the regulation with PD controller was gained higher than the other two and the error of the SDDRE was obtained the least, see Table I. Since the norm of the inputs (representative of the energy consumption) of the SDDRE is less than the PD and SMC, the performance of the proposed system is satisfactory. The results also confirm the validity of the quaternion-based dynamics and also the control implementation.

Table I. Comparison of PD, SMC, and SDDRE controller.

Validation with previous work: To confirm the correctness of the quaternion-based dynamics and the SDDRE controller, an existing model will be employed for comparison. Xiong and Zhang used a global fast terminal sliding mode controller (TSMC) for quadrotor regulation and also compared the results with conventional SMC [Reference Xiong and Zhang58]. Here the parameters of the system are substituted into the quaternion model and the SDDRE controls the model. The results are similar to the ones in Fig. 2 of ref. [Reference Xiong and Zhang58] presented in this section, in Fig. 9. The regulation of translation control is quite like the TSMC and regulated to desired values around 2s, without overshoot. The controller parameters are similar to Sections 7.2.

Fig. 9. The validation results of the quaternion-based SDDRE with previous work in ref. [Reference Xiong and Zhang58].

It should be noted that since the loop is closed, it cannot be said that the dynamics are validated; however, observing the behavior of the system in comparison with the SMC, the closed-loop system is validated.

7.2. Cobra maneuver

The motivation of using quaternions and avoiding Euler angles in rotational dynamics are to gain a singularity-free controller, and as a result, obtain agile flight and aerobatic maneuvers. One of the hardest positions in quadrotor control in $\pi /2\left( {{\textrm{rad}}} \right)$ rotation either in roll or pitch angles. The Cobra maneuver is famous for a jet aircraft to perform aerobatic shows or in combat for sudden brake, etc. For the quadrotors, it is the first time that this motion is simulated in a forward flight; the Cobra in ascending motion was reported [Reference Chen and Pérez-Arancibia44]. That is a challenge since, in $\phi = \pi /2$ or $\theta = \pi /2$ , there is no thrust to compensate the gravity; for an aircraft, a jet engine supports the gravity. This might cause a crash or fall for the multirotor. To perform the Cobra maneuver, ${\theta _{{\textrm{des}}}} = \pi /2$ is imposed in $t \in \left[ {1,1.35} \right]\left( {\textrm{s}} \right)$ instead of Eq. (26). After that, the multirotor drone tries to recover the stability and regulate to final condition. The simulation time is ${t_\textbf{f}} = 10\left( {\textrm{s}} \right),$ and the parameters of the control are ${{\textbf{R}}_\textbf{t}} = {{\textbf{I}}_{3 \times 3}}$ , ${{\textbf{Q}}_\textbf{t}} = {\textrm{diag}}\left( {1,1,1,0.5,0.5,0.5} \right)$ , ${{\textbf{F}}_\textbf{t}} = 10 \times {{\textbf{Q}}_{}}$ , ${{\textbf{R}}_\textbf{o}} = {{\textbf{I}}_{3 \times 3}}$ , ${{\textbf{Q}}_\textbf{o}} = {\textrm{diag}}\left( {2,2,2,0,0,0} \right)$ , ${{\textbf{F}}_\textbf{o}} = 10 \times {{\textbf{Q}}_\textbf{o}}$ . The start point of the regulation is set at zero along with other initial conditions; the endpoint is chosen $\left( {5,1,1.25} \right)\left( {\textrm{m}} \right)$ . Simulating the drone, the error is found $7.75\left( {{\textrm{mm}}} \right)$ and the system successfully performed the maneuver. The position variables and attitude ones are shown in Fig. 10 and Fig. 11 with respect. The trajectory and configuration of the system are demonstrated in Fig. 12. The quaternions are plotted in Fig. 13. The input thrust and input torque signals are presented in Fig. 14 and Fig. 15 with respect.

Fig. 10. The position variables of the system in aerobatic maneuver.

Fig. 11. The Euler angles of the drone in aerobatic maneuver.

Fig. 12. The Cobra maneuver via the SDDRE and quaternion dynamics.

Fig. 13. The quaternions.

Fig. 14. The input thrust of the quadrotor.

Fig. 15. The input torque signals of the drone.

8. Conclusions

This work investigated the quaternion-based control design using the SDDRE to control a quadrotor in aerobatic flight. The Euler angles are vulnerable to big changes in attitude and rotational dynamics, specifically, the rotation matrix. The diagonal components of ${\textbf{R}}\left( {\phi ,\theta ,\psi } \right)$ become zero for either of $\phi ,\theta ,\psi $ at $\pi /2$ . This means the omission of thrust in flight and unstable conditions. Specifically, the controllability pair will be unsatisfied. To solve this problem, quaternion representation has been used to have a singular-free attitude control and rotation matrix. The introduced model has been validated through a comparison with the conventional Euler dynamics controlled by a PD input law. To show the application of the proposed method, a challenging maneuver, Cobra, has been simulated in the forward flight, successfully controlled. The Cobra maneuver has put the drone in a position without thrust to compensate for the gravity; however, this approach generated a stable motion to reduce the fall in that condition.

Acknowledgment

This work is supported by the HYFLIERS project (HYbrid FLying-rolling with-snakE-aRm robot for contact inspection) funded by the European Commission H2020 Programme under grant agreement ID: 779411 (https://cordis.europa.eu/project/rcn/213049). The work has also been partially funded by the European Commission H2020 Programme under AERIAL-CORE project 871479.

Author Contributions

SRN: writing initial draft, review and editing, conceptualization, methodology, validation, and investigation. JAA: review and editing, supervision, and investigation. AO: project administration, funding acquisition, review and editing, and investigation.

Supplementary Material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0263574722000091.

References

Bibik, P., Narkiewicz, J., Zasuwa, M. and Żugaj, M., “Quadrotor Dynamics and Control for Precise Handling,” In: Innovative Simulation Systems, Studies in Systems, Decision and Control (A. Nawrat and K. Jędrasiak, eds.) (Springer International Publishing, 2016) pp. 335351. ISBN: 978-3-319-21118-3.CrossRefGoogle Scholar
Fernando, T., Chandiramani, J., Lee, T. and Gutierrez, H., “Robust Adaptive Geometric Tracking Controls on SO (3) with an Application to the Attitude Dynamics of a Quadrotor UAV,” 50th IEEE Conference on Decision and Control and European Control Conference, Orlando, FL, USA (2011) pp. 73807385.Google Scholar
Krishna, A. B., Sen, A. and Kothari, M., “Robust Geometric Control of a Helicopter Using Sliding Mode Control,International Conference on Unmanned Aircraft Systems, Athens, Greece (2020) pp. 737743.Google Scholar
Craig, W. S., Yeo, D. W. and Paley, D. A., “Geometric Control of a Quadrotor in Wind with Flow Sensing and Thrust Constraints: Attitude and Position Control,” AIAA Scitech 2019 Forum, San Diego, CA (2019) p. 1192.Google Scholar
Nekoo, S. R., Acosta, J. Á. and Ollero, A., “Geometric control using the state-dependent Riccati equation: application to aerial-acrobatic maneuvers,” Int. J. Control, 1–13 (2021). https://doi.org/10.1080/00207179.2021.1881165 CrossRefGoogle Scholar
Monforte, J. C., Geometric, Control and Numerical Aspects of Nonholonomic Systems (Springer-Verlag, Berlin Heidelberg, 2004).Google Scholar
Lee, T., Leok, M. and McClamroch, N. H., Global Formulations of Lagrangian and Hamiltonian Dynamics on Manifolds, vol. 13 (Springer, Cham, 2017) p. 31.Google Scholar
Terze, Z., Zlatar, D. and Pandža, V., “Aircraft attitude reconstruction via novel quaternion-integration procedure,” Aerospace Sci. Technol. 97, 105617 (2020).CrossRefGoogle Scholar
Zhou, Z., Zhang, Q., Liu, Q., Zeng, Q. and Tian, X., “Adaptive quaternion particle filter using generalized likelihood ratio test for aircraft attitude estimation in the presence of anomalous measurement,” Meas. Sci. Technol. 32(4), 045004 (2021).CrossRefGoogle Scholar
Figueredo, L. F. C., Adorno, B. V., Ishihara, J. Y. and Borges, G. A., “Robust Kinematic Control of Manipulator Robots Using Dual Quaternion Representation,” 2013 IEEE International Conference on Robotics and Automation, Karlsruhe, Germany (2013) pp. 1949–1955.Google Scholar
Savino, H. J., Pimenta, L. C. A., Shah, J. A. and Adorno, B. V., “Pose consensus based on dual quaternion algebra with application to decentralized formation control of mobile manipulators,” J. Franklin Inst. 357(1), 142178 (2020).CrossRefGoogle Scholar
Antonelli, G., Chiaverini, S., Sarkar, N. and West, M., “Adaptive Control of an Autonomous Underwater Vehicle: Experimental Results on ODIN,” IEEE Trans. Control Syst. Technol. 9(5), 756765 (2001).CrossRefGoogle Scholar
Liu, X., Zhang, M., Chen, J. and Yin, B., “Trajectory tracking with quaternion-based attitude representation for autonomous underwater vehicle based on terminal sliding mode control,” Appl. Ocean Res. 104, 102342 (2020).CrossRefGoogle Scholar
Rodriguez, J., Castañeda, H. and Gordillo, J. L., “Lagrange modeling and navigation based on quaternion for controlling a micro AUV under perturbations,” Rob. Auton. Syst. 124, 103408 (2020).CrossRefGoogle Scholar
de Cerqueira Gava, P. D., Jorge, V. A. M., Júnior, C. L. N. and Adabo, G. J., “AUV Cruising Auto Pilot for a Long Straight Confined Underwater Tunnel,” 2020 IEEE International Systems Conference (SysCon), Montreal, QC, Canada (2020)pp. 18.Google Scholar
Suzuki, S., Nakazawa, D., Nonami, K. and Tawara, M., “Attitude control of small electric helicopter by using quaternion feedback,” J. Syst. Des. Dyn. 5(2), 231247 (2011).Google Scholar
Cooke, J. M., Zyda, M. J., Pratt, D. R. and McGhee, R. B., “NPSNET: Flight simulation dynamic modeling using quaternions,” Presence Teleoperators Virtual Environ. 1(4), 404420 (1992).CrossRefGoogle Scholar
Islam, M. and Okasha, M., “A Comparative Study of PD, LQR and MPC on Quadrotor Using Quaternion Approach,” 2019 7th International Conference on Mechatronics Engineering (ICOM), Putrajaya, Malaysia (2019) pp. 16.Google Scholar
Islam, M., Okasha, M. and Sulaeman, E., “A model predictive control (MPC) approach on unit quaternion orientation based quadrotor for trajectory tracking,” Int. J. Control Autom. Syst. 17(19), 28192832 (2019).CrossRefGoogle Scholar
Kang, J.-W., Sadegh, N. and Urschel, C., “Quaternion Based Nonlinear Trajectory Control of Quadrotors with Guaranteed Stability,” American Control Conference, Denver, CO, USA (2020) pp. 38343839.Google Scholar
Oliva-Palomo, F., Sanchez-Orta, A., Alazki, H., Castillo, P. and Muñoz-Vázquez, A.-J., “Robust global observer position-yaw control based on ellipsoid method for quadrotors,” Mech. Syst. Signal Process. 158, 107721 (2021).CrossRefGoogle Scholar
Najafi, E., Lopes, G. A. D. and Babuška, R., “Balancing a legged robot using state-dependent Riccati equation control,” IFAC Proc. 47(3), 21772182 (2014).Google Scholar
Nguyen, H. T., Nguyen, N. T., Prodan, I. and Pereira, F. L., “Trajectory tracking for a multicopter under a quaternion representation,” IFAC-PapersOnLine 53(2), 57315736 (2020).CrossRefGoogle Scholar
Chovancová, A., Fico, T., Hubinský, P. and Duchoň, F., “Comparison of various quaternion-based control methods applied to quadrotor with disturbance observer and position estimator,” Rob. Auton. Syst. 79, 8798 (2016).CrossRefGoogle Scholar
Fresk, E. and Nikolakopoulos, G., “Full Quaternion Based Attitude Control for a Quadrotor,” European Control Conference, Zürich, Switzerland (2013) pp. 38643869.Google Scholar
Zha, C., Ding, X., Yu, Y., and Wang, X., “Quaternion-based nonlinear trajectory tracking control of a quadrotor unmanned aerial vehicle,” Chin. J. Mech. Eng. 30(1), 7792 (2017).CrossRefGoogle Scholar
Sanchez, L. F., Abaunza, H. and Castillo, P., “User–robot interaction for safe navigation of a quadrotor,” Robotica 38(12),21892203 (2020).CrossRefGoogle Scholar
Gillula, J. H., Huang, H., Vitus, M. P. and Tomlin, C. J., “Design of Guaranteed Safe Maneuvers Using Reachable Sets: Autonomous Quadrotor Aerobatics in Theory and Practice,” 2010 IEEE International Conference on Robotics and Automation, Anchorage, Alaska, USA (2010) pp. 16491654.Google Scholar
Cutler, M. and How, J., “Actuator constrained Trajectory Generation and Control for Variable-Pitch Quadrotors,” AIAA Guidance, Navigation, and Control Conference, Minneapolis, MN (2012) p. 4777.Google Scholar
Pearson, J. D., “Approximation methods in optimal control I. Sub-optimal control,” Int. J. Electron. 13(5), 453469 (1962).Google Scholar
Parrish, D. K. and Ridgely, D. B., “Attitude Control of a Satellite Using the SDRE Method,” Proceedings of the American Control Conference, Albuquerque, New Mexico (1997) pp. 942946.Google Scholar
Romero, A. G., Souza, L. C. G. and Chagas, R. A., “Application of the SDRE Technique in the Satellite Attitude and Orbit Control System with Nonlinear Dynamics,” 2018 SpaceOps Conference, Marseille, France (2018) p. 2536.Google Scholar
Stansbery, D. T. and Cloutier, J. R., “Position and Attitude Control of a Spacecraft Using The State-Dependent Riccati Equation Technique,” Proceedings of the American Control Conference, Chicago, Illinois (2000) pp. 1867–1871.Google Scholar
Luo, W. and Chu, Y.-C., “Attitude Control Using the SDRE Technique,” 7th International Conference on Control, Automation, Robotics and Vision, Singapore (2002) pp. 12811286.Google Scholar
Pukdeboon, C. and Zinober, A. S., “Optimal sliding mode controllers for spacecraft attitude manoeuvres,” IFAC Proc. Vol. 42(6), 173178 (2009).CrossRefGoogle Scholar
Pukdeboon, C., “Robust optimal output feedback sliding mode control for spacecraft attitude tracking maneuvers,” Int. J. Pure Appl. Math. 79(1), 1127 (2012).Google Scholar
Li, P., Yue, X., Chi, X. and Du, C., “Optimal Relative Attitude Tracking Control for Spacecraft Proximity Operation,”25th Chinese Control and Decision Conference, Taiyuan, China (2013) pp. 45824587.Google Scholar
Safi, M., Mortazavi, M. and Dibaji, S. M., “Global stabilization of attitude dynamics: SDRE-based control designs,” AUT J. Model. Simul. 50(2), 203210 (2018).Google Scholar
Romero, A. G. and de Souza, L. C. G., “State-dependent Riccati equation controller using Java in remote sensing CubeSats,” J. Appl. Remote Sens. 13(3), 032509 (2019).CrossRefGoogle Scholar
Cimen, T., “Survey of state-dependent Riccati equation in nonlinear optimal feedback control synthesis,” J. Guidance Control Dyn. 35(4), 10251047 (2012).CrossRefGoogle Scholar
Wang, N., Ma, R., Chang, X. and Zhang, L., “Numerical virtual flight simulation of quasi-cobra maneuver of a fighter aircraft,” J. Aircraft 58(1), 138152 (2021).CrossRefGoogle Scholar
Gal-Or, B. and Baumann, D. D., “Mathematical phenomenology for thrust-vectoring-induced agility comparisons,”J. Aircraft. 30(2), 248255 (1993).CrossRefGoogle Scholar
Xu, W. and Zhang, F., “Learning Pugachev’s Cobra maneuver for tail-sitter UAVs using acceleration model,” IEEE Rob. Autom. Lett. 5(2), 34523459 (2020).CrossRefGoogle Scholar
Chen, Y. and Pérez-Arancibia, N. O., “Adaptive Control of Aerobatic Quadrotor Maneuvers in the Presence of Propeller-Aerodynamic-Coefficient and Torque-Latency Time-Variations,” 2019 International Conference on Robotics and Automation (ICRA), Montreal, Canada (2019) pp. 64476453.Google Scholar
Diebel, J., “Representing attitude: Euler angles, unit quaternions, and rotation vectors,” Matrix 58(15), 135 (2006).Google Scholar
Nekoo, S. R., “Tutorial and review on the state-dependent Riccati equation,” J. Appl. Nonlinear Dyn. 8(2), 109166 (2019).CrossRefGoogle Scholar
Nekoo, S. R., “A PDE breach to the SDRE,” Asian J. Control 22(2), 667676 (2020).CrossRefGoogle Scholar
Nekoo, S. R., “Digital implementation of a continuous-time nonlinear optimal controller: An experimental study with real-time computations,” ISA Trans. 101, 346357 (2020).CrossRefGoogle ScholarPubMed
Jayaram, A. and Tadi, M., “Synchronization of chaotic systems based on SDRE method,” Chaos Solitons Fractals 28(3), 707715 (2006).CrossRefGoogle Scholar
Korayem, M. H. and Nekoo, S. R., “Finite-time state-dependent Riccati equation for time-varying nonaffine systems: Rigid and flexible joint manipulator control,” ISA Trans. 54, 125144 (2015).CrossRefGoogle ScholarPubMed
Mathavaraj, S. and Padhi, R., “Finite-Time LQR and SDRE for Satellite Formation Flying,” In: Satellite Formation Flying (Springer, Singapore, 2021) pp. 99109. ISBN: 978-981-15-9631-5.CrossRefGoogle Scholar
Heydari, A. and Balakrishnan, S. N., “Approximate Closed-Form Solutions to Finite-Horizon Optimal Control of Nonlinear Systems,” American Control Conference, Fairmont Queen Elizabeth, Montréal, Canada (2012) pp. 26572662.Google Scholar
Korayem, M. H. and Nekoo, S. R., “Nonlinear Optimal Control via Finite Time Horizon State-Dependent Riccati Equation,” Second RSI/ISM International Conference on Robotics and Mechatronics, Tehran, Iran (2014) pp. 878883.Google Scholar
Nekoo, S. R., Acosta, J. Á., Gomez-Tamm, A. E. and Ollero, A., “Optimized Thrust Allocation of Variable-Pitch Propellers Quadrotor Control: A Comparative Study on Flip Maneuver,” 2019 Workshop on Research, Education and Development of Unmanned Aerial Systems (RED UAS), Cranfield, UK (2019) pp. 8695.Google Scholar
Nekoo, S. R., Acosta, J. Á. and Ollero, A., “Fully Coupled Six-DoF Nonlinear Suboptimal Control of a Quadrotor: Application to Variable-Pitch Rotor Design,” Iberian Robotics Conference, Porto, Portugal (2019) pp. 7283.Google Scholar
Xu, R. and Ozguner, U., “Sliding Mode Control of a Quadrotor Helicopter,” Proceedings of the 45th IEEE Conference on Decision and Control, San Diego, CA, USA (2006) pp. 49574962.Google Scholar
Abro, G. E. M., Zulkifli, S. A., Asirvadam, V. S. and Ali, Z. A., “Model-Free-based Single-Dimension Fuzzy SMC Design for Underactuated Quadrotor UAV,” In: Actuators (H. Wang, ed.) (MDPI, Switzerland, 2021) p. 191.CrossRefGoogle Scholar
Xiong, J.-J. and Zhang, G.-B., “Global fast dynamic terminal sliding mode control for a quadrotor UAV,” ISA Trans. 66, 233240 (2017).CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. The definition of the reference coordinates for a sample quadrotor drone.

Figure 1

Fig. 2. x-axis regulation of the system, comparison with PD and SMC.

Figure 2

Fig. 3. y-axis regulation of the system, comparison with PD and SMC.

Figure 3

Fig. 4. z-axis regulation of the system, comparison with PD and SMC.

Figure 4

Fig. 5. The roll angle of the system, comparison with PD and SMC.

Figure 5

Fig. 6. The pitch angle of the system, comparison with PD and SMC.

Figure 6

Fig. 7. The input norm of the inputs of the system, comparison with PD and SMC.

Figure 7

Fig. 8. The configuration and trajectories of the quadrotor drones with PD, SMC, and SDDRE controllers.

Figure 8

Table I. Comparison of PD, SMC, and SDDRE controller.

Figure 9

Fig. 9. The validation results of the quaternion-based SDDRE with previous work in ref. [58].

Figure 10

Fig. 10. The position variables of the system in aerobatic maneuver.

Figure 11

Fig. 11. The Euler angles of the drone in aerobatic maneuver.

Figure 12

Fig. 12. The Cobra maneuver via the SDDRE and quaternion dynamics.

Figure 13

Fig. 13. The quaternions.

Figure 14

Fig. 14. The input thrust of the quadrotor.

Figure 15

Fig. 15. The input torque signals of the drone.

Supplementary material: File

Nekoo et al. supplementary material

Nekoo et al. supplementary material

Download Nekoo et al. supplementary material(File)
File 6.5 MB