Hostname: page-component-848d4c4894-sjtt6 Total loading time: 0 Render date: 2024-06-21T06:16:37.914Z Has data issue: false hasContentIssue false

Verification framework for control theory of aircraft

Published online by Cambridge University Press:  10 May 2022

O. A. Jasim
Affiliation:
Department of Automatic Control and Systems Engineering, University of Sheffield, Sheffield, UK
S. M. Veres*
Affiliation:
Department of Automatic Control and Systems Engineering, University of Sheffield, Sheffield, UK
*
*Corresponding author email: s.veres@sheffield.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

A control system verification framework is presented for unmanned aerial vehicles using theorem proving. The framework’s aim is to set out a procedure for proving that the mathematically designed control system of the aircraft satisfies robustness requirements to ensure safe performance under varying environmental conditions. Extensive mathematical derivations, which have formerly been carried out manually, are checked for their correctness on a computer. To illustrate the procedures, a higher-order logic interactive theorem-prover and an automated theorem-prover are utilised to formally verify a nonlinear attitude control system of a generic multi-rotor UAV over a stability domain within the dynamical state space of the drone. Further benefits of the procedures are that some of the resulting methods can be implemented onboard the aircraft to detect when its controller breaches its flight envelope limits due to severe weather conditions or actuator/sensor malfunction. Such a detection procedure can be used to advise the remote pilot, or an onboard intelligent agent, to decide on some alterations of the planned flight path or to perform emergency landing.

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 (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Royal Aeronautical Society

Nomenclature

b

torque constant of the rotor

l

aerodynamic force constant of the rotor

r

position (translational) vector of x, y, z in the world frame

u

control input vector

$\textbf{u}_\textbf{d}$

represents uncertainty and disturbances

q

quaternions elements

I

symmetric and positive-definite inertia matrix

$\hat{I}$

estimated matrix of the inertia matrix I

T

time sub-domain

${F_{B}}$

total force of thrusters

V

Lyapunov function

Q

symmetric positive-definite matrix

$K_{\omega},K_{q}$

positive-definite diagonal gain matrices

Greek Symbol

$\tau$

torque vector

$\tau_d$

disturbance torque vector

$\phi$

roll of the aircraft

$\theta$

pitch of the aircraft

$\psi$

yaw of the aircraft

$\omega$

angular velocity vector

$\dot{\omega}$

angular acceleration vector

${{\xi}}$

error vector

$\dot{{\xi}}$

error rate vector

${{\eta}}$

error and error rate vector

$\Omega$

rotor angular velocity

$\ell$

length from the centre of mass of the multi-rotor to the rotor

$\gamma$

known positive constant bound

$\delta$

positive constant

$\varepsilon$

positive constant

$\lambda_{\min}, \lambda_{\max}$

lower and upper constant bound, respectively

1.0 Introduction

Control law design for aircraft normally combines control engineering knowledge with tests of stability and smoothness of control responses under disturbances. This often takes the form of an iterative process of control law refinements accompanied with wind tunnel and in flight tests. Given a particular open loop dynamical model, control engineering relies on mathematical theory that enables the design of a flight controller. When the flight envelope is defined, it introduces numerical values for the design, which need to be carried through derivations and proofs of stability and acceptable handling within the flight envelope. In this paper manual derivations are replaced by theorem proving methods for robust control performance. The procedures presented go beyond algebraic computation and use higher-order logic, including handling of functionals, operators, stability and levels of smoothness measures. The formal approach of proofs on a computer has two advantages: avoiding mistakes of manual mathematical derivations and fast robustness calculations against parameters of the flight envelope required.

The approach taken in this paper fits into one of the three stages of formally verifiable controller design as outlined in [Reference Jasim and Veres1], where robust control theory verification is followed by verification of the software used for implementation. A third stage is to prove that the quantisation affects on a digital controller do not significantly affect the results in stage one of the controller theory used. In implementations of aviation software, verification is often followed by redundancy-based safety analysis for critical sensors and actuators using voting principles, the use and effect of these lies outside the scope of this paper.

Realtime code-verification normally consists systematically checking the correctness of the encoded controller such as in [Reference Feron2, Reference Jobredeaux3] and [Reference Wang4]. This paper focuses on verification of the control principles before it is implemented in realtime code. An interactive theorem prover (ITP) is used for performance verification under bounded nominal ambient conditions. Also monitoring of stability and performance are checked in flight using an automated theorem prover (ATP) for excessive conditions, including some sensor or actuator failures.

There have been prior initiatives on verification of safety-critical and cyber-physical systems. Such are the European Integrated Tool Chain for Model-based Design of Cyber-Physical Systems (INTO-CPS) [5], where a Functional Mockup Interface (FMI) has been developed for integrating the formal verification of Cyber-Physical Systems using the PVS (Prototype Verification System) [Reference Owre, Rushby and Shankar6] theorem prover with model-based software to co-simulate these systems. This approach integrates simulated models in model-based tools such as Modelica [Reference Tiller7], Simulink/Matlab [8] or 20-sim [9] with the FMI interface to verify the control system according to the required specifications using formal methods.

The FMI is also implemented in [Reference Zeyda, Ouy, Foster and Cavalcanti10] using Isabelle/Unifying Theory of Programming (UTP) [11], where Modelica is used to model the control system of a train, which is encoded in Isabelle/UTP within an FMI framework for co-simulation. Another project is the ERATO Metamathematics for Systems Design (MMSD) [Reference Team12], where a framework is developed to use formal methods to verify automotive-related applications. Other projects are conducted by the verification team at the NASA Langley Research Center [13] such as [Reference Denman and Muñoz14, Reference Muáoz15] and [Reference Munoz, Dutle, Narkawicz and Upchurch16]. Though these approaches use formal methods to verify control systems, the derivations of control laws are not covered by verification before implementing them. In addition, there is an absence of onboard real-time stability monitoring of these systems by formal methods.

There have been several studies conducted to formally verify control systems. In [Reference Hasan and Ahmad17], a framework for reasoning about the transfer function and steady state errors in feedback control systems using the higher-order-logic (HOL) Light [Reference Harrison18] theorem prover is proposed. Adnan et al. [Reference Rashid and Hasan19] formalised the theory of Laplace transforms using an ITP to analyse linear control systems. In [Reference Ma and Chen20], formal verification of coordinate transformations is proposed for aircraft control systems using the Coq [Reference Huet and Paulin-Mohring21] theorem prover. An autonomous vehicle controller is formalised and proven in [Reference Domenici, Fagiolini and Palmieri22] using the PVS theorem prover to illustrate its correctness in terms of stability of the Jacobian matrix and then co-simulated using the PVSio environment in the PVS prover. Denman et al. [Reference Denman, Zaki, Tahar and Rodrigues23] presented a method for the implementation of a formal Nichols plot analysis of a flight control system using the MetiTarski ATP [Reference Paulson24] with the aim of stability verification. Araiza-Illan et al. [Reference Araiza-Illan, Eder and Richards25] proposed a framework for automatically translating the system block diagrams modelled in Simulink into the Why3 [Reference Filliâtre and Paskevich26] prover for verification. Verifying hybrid control systems using differential dynamic logic can be carried out in the KeYmaera prover [Reference Platzer27]. In [Reference Aréchiga, Loos, Platzer and Krogh28, Reference Aréchiga and Krogh29] abstractions of the closed-loop behaviour have been constructed for a class of controllers using KeYmaera.

The framework proposed here is different from these approaches as we are verifying the correctness of the manually derived control law of an aviation system at the design stage before the encoding and testing/simulation step, which is followed by real-time monitoring of stability by an ATP onboard the craft. To the best of our knowledge this initial verification step has not been carried out for robust control systems at this level of HOL-complexity in any prior work.

The novelty of this paper is that it demonstrates that the presented combination of interactive theorem provers are suitable for verifying the correctness of robust control theory for a prescribed flight envelope of a rotary wing craft. Although prior work suggested this may be a possibility, this is a first evidence of this kind. This involves formal stability analysis to guarantee system robustness under variable ambient conditions. Also a method of continuous stability monitoring is proposed through an onboard automated theorem prover. The latter methods are implemented in Isabelle [Reference Nipkow, Paulson and Wenzel30] and MetiTarski [Reference Paulson24]. The associated HOL codes have been made available online for use by the aviation community.

Our verification framework has been applied to a control scheme of a generic quad-copter presented in [Reference Jasim and Veres31], which consists of a nonlinear attitude controller to deal with modelling uncertainty and external disturbances. The design is based on the well-known dynamic inversion technique [Reference Spong, Hutchinson and Vidyasagar32, Reference Sciavicco and Siciliano33]. Lyapunov’s method [Reference Slotine and Li34] has been used as part of the attitude controller design and to analyse the system to ensure its stability. We have applied our verification scheme to the attitude control system, including all definitions, assumptions, formal derivations and performance proofs. Lyapunov stability of the control system is formalised and proven in Isabelle/HOL.

We have implemented the onboard stability analysis in the MetiTarski ATP. This system is suitable to check whether the aircraft is in its permitted flight envelope and it can inform the pilot or an autonomous agent to perform some emergency action.

The paper is structured as follow: Section 2 describes theorem proving tools, where Isabelle and MetiTarski provers are introduced; Section 3 presents the proposed verification framework; Section 4 demonstrates the applicability of the framework to a multi-rotor control system and describes onboard stability monitoring; Section 5 concludes the work.

2.0 Theorem proving

Computer-based theorem proving relies on a set computational tools in some logical system that can be used to prove the soundness and correctness of mathematical arguments. There are two main types of theorem proving: Automated Theorem Proving (ATP), which proves mathematical statements automatically without human interaction, and Interactive Theorem Proving (ITP), which automates steps of formal proofs by aid of a developer guiding the process of proof. The automated steps rely on mathematical logic and automated reasoning techniques. In this paper an ITP will be used for the verification framework proposed. ITPs are proof assistants, which formally define and prove mathematical theorems. Therefore, their user can implement a mathematical theory by defining assumptions and some valid logical statements to start with. Then the ITP procedure is used to prove a sequence of statements, relying on available formal theories, and also by using existing logical methods and techniques or some external resources, such as ATP tools.

The distinction between ITP and ATP systems is not only that ATP systems fully automate proofs but that ATPs tend to have restricted expressivity. Unlike ITPs, they cannot handle the highest order mathematical theories. Instead, they tend to be used to prove less complex mathematical formulae containing inequalities over real numbers, quantified variables, complex algebra and special functions. On the other hand, ITP systems have the ability to aid formalising and proving mathematical theories involving higher order logic, techniques of abstract mathematics, including the concept of convergence, continuity and linear and nonlinear operators to represent dynamical systems.

To verify complex mathematical theories, ATPs can be utilised locally in an ITP to prove a step in the proof by adding their packages to the ITP software. A large number of theorem provers are listed on the “System on TPTP” [Reference Sutcliffe and Suttner35] web site, including the most powerful ATPs that can be used to prove mathematical statements automatically. A good example of a useful ATP, from a control engineering point of view, is MetiTarski [Reference Paulson24], which is a first-order logic (FOL) prover designed to work over the field of real numbers on decidable problems.

Model checking is one of the well-known tools that can be applied to engineering models and to formally verify and validate their correctness [Reference Clarke, Grumberg and Peled36]. Systems are modelled as finite state transition systems, or timed automata, and their properties are expressed, for instance, in propositional temporal logic. Checking of system properties is reduced to a graph search and exhaustive exploration of all possible abstracted states. However, model checkers are not which structures to infinite dimensional state space properties, and they cannot be used to prove the correctness of robust control systems. In contrast, ITPs can be employed to verify infinite dimensional state space systems and other abstract mathematical constructs.

2.1 Isabelle/HOL

Isabelle is an interactive higher-order-logic (HOL) theorem prover, a kind of proof assistant, based on automated reasoning techniques and logical calculus. It is written using ML [Reference Milner37], a functional programming language, in which rules are presented as propositions and the proofs are structured by combining rules using the $\lambda$ -calculus. Isabelle provides the ability to express mathematical theorems of control engineering in formulae of formal logic and prove them using the automated derivation tools provided. When a proof is completed in Isabelle, one can be certain that the mathematical theorem at hand is valid.

Isabelle/HOL is one of Isabelle’s platforms with quantifiers over domains of variables and formal semantics. Isabelle/HOL has a proof-language called Isar, which structures the proofs in a human readable and understandable mathematical formal text for both users and computers. The mathematical formulae can be formalised and proven in Isar with the aid of Isabelle’s logical tools.

Examples of such tools are the simplifier, which performs operation and reasoning on equations; the classical reasoner that carries out long chains of reasoning procedures; and algebraic decision procedures using advanced pattern matching by the package sledgehammer for automatically finding the proofs based on already-proven theorems in Isabelle’s library. The latter also calls external FOL (first-order-logic) provers (ATPs) and Satisfiability Modulo Theories (SMT) solvers [Reference Barrett and Tinelli38]. Isabelle has been chosen in this work for its rich logical and automation techniques since it has a large library, which contains most of the mathematical theories that are useful in control systems theory proofs.

The use of Isabelle is justified as advanced robust and adaptive control theory heavily uses concepts of convergence in limits, differentiation and integration, complex analysis, abstract vector spaces, spaces of functionals and nonlinear operators, stability theory, and formal measures of control performance. Isabelle also supports other logical systems such as Hoare logics for systems verification [Reference Nipkow39] in addition to the ability to generate executable codes of the proven statements. There is a wide range of syntax in Isabelle/HOL, the most common and useful Isabelle/HOL expressions and symbols are summarised in Table 1.

Table 1. Isabelle/HOL symbols and expressions

2.2 MetiTarski

MetiTarski is a FOL automated theorem prover, which is designed as a combination of the resolution prover Metis [Reference Hurd40] and a decision procedure tool QEPCAD [Reference Brown41] to conduct proofs of theorems stated algebraically over the field of real numbers. It can be utilised to prove universally quantified inequalities of transcendental and other functions as well, for instance ln, log, exp, etc. MetiTarski reduces problems to decidable ones by replacing some functions with their upper and lower bounds. It can be made to perform proofs automatically with the aid of three reasoners: QEPCAD, Z3 [Reference De Moura and Bjørner42] and Mathematica. It can, for instance, be used to check the stability of an unmanned aerial vehicle (UAV) during flight, due to its ability to prove inequalities, which occur in the theory of dynamical stability, at a reasonable speed.

3.0 A Control theory verification framework

A framework is introduced here to perform the verification steps of the control theory that was applied at the stage of aircraft design. Following an engineer’s initial control design of an aircraft, a thorough process can be started to verify the correctness of the control system using this framework. Such a computerised verification process is able to deal with stability proofs of aircraft dynamics under actuators constraints. It is able confirm or it can make flight envelope specifications more precise. In addition, its methods can be extended to onboard stability monitoring of aircraft to enhance the decision making process for aircraft safety. The framework is intended as a complementary work to conventional verification processes such as software and control code verification. It is not to be mixed up with control system code verification, which concerns itself with the correctness of implementation in code and also with numerical inaccuracies due to digital processing, sensor signal processing, control surface and engine actuator errors.

An example framework is outlined in Fig. 1. It consists of two stages:

  1. (i) The ITP represented by the Isabelle/HOL is used to prove the mathematical derivation of the designed control system and its stability analysis.

  2. (ii) The ATP represented by the MetiTarski prover is deployed to continuously check aircraft stability during the flight.

Figure 1. UAVs verification framework.

To perform the first stage, the aircraft’s dynamical parameters need to be included in the Isabelle/HOL prover to carry out the verification process. The work starts with formalising the aircraft’s dynamics in the HOL syntax of the Isabelle prover. The aircraft’s formal dynamical theory, which need to be included in an ITP representation, contains the following:

  1. (a) the dynamical equations of motion,

  2. (b) the coordinate system in the rigid body frame,

  3. (c) the transformations between the world and body frames,

  4. (d) the controller design and its stability analysis.

Other needed concepts are also formalised in HOL such as definitions of the time domain, signals, real vectors and matrices with their properties, etc.

In a second stage, the aircraft’s stability analysis is to be formalised in the FOL syntax of the MetiTarski prover for verification of stability conditions onboard the craft. The derivative of the Lyapunov function is formalised in order to check that the Lyapunov stability criterion is satisfied for the current state. If the criterion is not satisfied that can be indicative of the aircraft being outside its nominal stability domain, i.e. it violates its flight envelope.

4.0 A Case study: multi-rotor verification

This section illustrates the verification stages within framework in some detail. The stages in Fig. 1 are considered in separate subsections. The Isabelle and MetiTarski codes of these demonstrations are downloadable by aerospace engineers from an online repository [Reference Jasim and Veres43].

4.1 Verification using the Isabelle/HOL prover

The attitude controller of a generic multi-rotor UAV, as proposed in [Reference Jasim and Veres31], is considered here. The aircraft’s attitude dynamics are controlled by a robust nonlinear controller that takes into account the dynamical modelling uncertainties and payload disturbances. Figure 2 illustrates the attitude controller.

Figure 2. The inner control loop of the multi-rotor controller [Reference Jasim and Veres31] for attitude control. q is the measured and $q_r$ is the desired attitude, $q_e$ is its error in quaternions, $\xi$ is angular error vector, $\omega, \omega_r$ are the angular velocity and reference, $K_q, K_\omega$ are control gains, $I,\tilde{I}$ are the inertia matrix and its estimate. $\Gamma(\omega)$ is a cross-product matrix for the Coriolis forces such that $\Gamma(\omega) I \omega = \omega \times I \omega$ is satisfied and $u_d$ is torque disturbance of the attitude.

Simulation cannot be used to fully prove that the UAV’s control system is robust for all possible flight conditions within its flight envelope. To ensure correctness of the controller robustness stated in [Reference Jasim and Veres31], the controller design’s derivations and stability analysis are to be verified using the Isabelle/HOL prover. The Isabelle/HOL software system is applied for this purpose due to its rich library of mathematical theorems, which are required to perform the UAV’s control system verification.

The verification process using Isabelle/HOL is illustrated in Fig. 3, which consists of two parts: formalisation (HOL Platform) and proof procedures (Proof Tools). The first part starts by formalising the multi-rotor UAV system into the Isabelle/HOL syntax such as the coordinate system, rotational dynamics, time-domain functions, proposed assumptions and the aircraft’s stability analysis. The implementation of the control design and aircraft dynamics includes a series of Isabelle items such as definition, lemma and theorem.

Figure 3. Formalising and proving UAV’s controller in Isabelle/HOL theorem prover.

In this demo some new lemmas were needed to be stated and proven, which did not yet exist in Isabelle as that library is still under development. The formalisation also required importing some proven mathematical theories and lemmas from the prover’s library, which were used in formalising the control system equations with the controller’s assumptions and definitions.

The theories implemented under an HOL formalism will be described here to illustrate the proof procedures. The $HOL.thy$ is at the core of HOL formalism, which includes definitions of real numbers (real.thy), functions (fun.thy), sets (set.thy), etc. These are necessary in most formal procedures. The $Quaternions.thy$ can be used for quaternion definitions, operations and also to represent aircraft attitude. A multi-variable analysis package $Analysis.thy$ can be exploited for function operations over real numbers. The Finite_Cartesian_Product.thy and Inner_Product.thy can be utilised for definitions and operations over real vectors, L2_norm.thy and Norm_Arth.thy are used for real vector norms and their operations, respectively.

A time sub-domain has been defined as $T= \{t | \ t \in [0,\infty\}\}$ and relied on in definitions of sets of vector signals. For instance, the concept of a function of time (i.e. signal) over $[0,\infty)$ and its mapping to another signal is an operator. Formal nonlinear operators have been defined within Isabelle to express unknown dynamics, also used in a previous work [Reference Jasim and Veres1], to map signals to signals by dynamical operators. These nonlinear operators are very important in descriptions and proofs of controller robustness. It is worth mentioning that in the Isabelle/HOL it was possible to define such operations using the “locale” syntax [Reference Nipkow, Paulson and Wenzel30].

The mentioned theory $Analysis.thy$ includes definitions of real vectors (Finite_Cartesian_Product.thy), vector norms (L2_norm.thy, Norm_Arth.thy) and their operations. These theories enable the definitions of the aircraft’s three-dimensional rotation vectors and their norms including the torque, angular velocity and acceleration vectors, where each component of a vector is represented as a continuous time-domain function for signals. The continuous functions defined in $Fun.thy$ , $Function\_Algebras.thy$ and $Topological\_Spaces.thy$ theories have been heavily relied on.

Matrices are formalised in $Matrix.thy$ , and their operations and derivations performed in $Analysis.thy$ , $Finite\_Cartesian\_Product.thy$ and $Real\_$ $Vector\_Spaces.thy$ . The rate change of the multi-rotor’s attitudes, i.e. velocities and accelerations, are formalised by time derivatives in the theories $Deriv.thy$ and $derivative.thy$ .

Multi-rotor controller design requires several robustness assumptions, which need inequalities over the field of real-numbers. Fortunately, such inequalities have been defined in the Isabelle prover in the theory $Orderings.thy$ . This is an important feature for any robust control design.

The second part “Theorem Proving” in Fig. 1 is an interactive process between the designer/engineer and the automated proving tools that the Isabelle prover supports. The role of designer/engineer is to exploit the ITP step-by-step the achieve a proof, where the system is not able to find a proof automatically, by simplifying the statement into several steps. Each step should be proven using the provided automated tools before moving to the next one, otherwise the prover will not pass the statement. Examples of the automated tools supported in Isabelle are: CVC4, Z3, SPASS, E prover, Remote_Vampire and SMT solvers. In addition, Isabelle has its own automatic proving tools such as auto, simp, blast, etc. Most of the control systems verified in this work required an interaction with the prover due to the design complexity making it not possible for the prover to solve them automatically.

The Isabelle code used in the case study is too long to be stated here, only the relevant definitions and proofs are shown, while the complete code can be found in an online repository [Reference Jasim and Veres43]. Formalisation and proofs are illustrated in the remainder of this section.

The multi-rotor attitude dynamics in [Reference Jasim and Veres31],

(1) \begin{equation} I \dot{\boldsymbol{\omega}} + \Gamma(\boldsymbol{\omega}) I \boldsymbol{\omega} + \boldsymbol{\tau}_d = \boldsymbol{\tau},\end{equation}

where $I \in \Re^{3 \times 3}$ is a symmetric and positive-definite inertia matrix of the craft about its mass centre. $\boldsymbol{\tau}_d(t)=[\tau_{d \phi}(t) \ \tau_{d \theta}(t) \ \tau_{d \psi}(t)]^{T} \in \Re^{3}$ are unknown disturbances torques with $\phi,\theta$ and $ \psi$ being the roll, pitch and yaw angles, respectively. $\boldsymbol{\tau}(t) = [\tau_{\phi}(t) \ \tau_{\theta}(t) \ \tau_{\psi}(t) ]^{T} \in \Re^{3} $ is the torque vector of the onboard controller in the body frame which produces the multi-rotor motion. These can be formalised by the following code in Isabelle/HOL:

and the torque vector $\boldsymbol{\tau}$ in [Reference Jasim and Veres31],

(2) \begin{equation} \boldsymbol{\tau} =\begin{bmatrix}\tau_{\phi} \\ \tau_{\theta} \\ \tau_{\psi}\end{bmatrix}=\begin{bmatrix}\ell l(\Omega^{2}_{2} - \Omega^{2}_{4}) \\\ell l(-\Omega^{2}_{1} + \Omega^{2}_{3}) \\b(-\Omega^{2}_{1} + \Omega^{2}_{2} - \Omega^{2}_{3} + \Omega^{2}_{4})\end{bmatrix},\end{equation}

is defined by bounding all rotor angular velocities $\Omega_i$ with their maximum value $\Omega_{\max}$ as

where $\ell$ is the length from the centre of mass of the multi-rotor to the rotor, and l and b are the aerodynamic force and torque constants of the rotors. The control input $\textbf{u}$ [Reference Jasim and Veres31],

(3) \begin{equation}\textbf{u} = \dot{\boldsymbol{\omega}}_{r} + K_{\boldsymbol{\omega}}\dot{\boldsymbol{\xi}} + K_{\textbf{q}}\boldsymbol{\xi},\end{equation}

and the control torque [Reference Jasim and Veres31],

(4) \begin{equation}\boldsymbol{\tau} = \hat{I} \textbf{u} + \textbf{u}_d + \hat{\textbf{c}}(\boldsymbol{\omega}),\end{equation}

are defined in the prover through the following code:

where $\hat{I}$ is an estimated matrix of the inertia matrix I of the craft, $\textbf{u}$ represents a new input vector to be designed later on in Equation (3), $\boldsymbol{\omega}(t)$ is the angular velocities vector, $\dot{\boldsymbol{\omega}}(t)$ and $\dot{\boldsymbol{\omega}}_{r}(t)$ are the angular acceleration and reference angular acceleration vectors, respectively. $K_{\omega}=diag[k_{\omega_{1}} k_{\omega_{2}} k_{\omega_{3}}] \in \Re^{3 \times 3}$ is a positive-definite diagonal gain matrix setting the error gains in feedback and $K_{q}=diag[k_{q_{1}} k_{q_{2}} k_{q_{3}}]\in \Re^{3 \times 3}$ is a positive-definite diagonal gain matrix. $[q_{e_{1}} \ q_{e_{2}} \ q_{e_{3}}]^{T}$ is a term defined to reduce the the dimensions of quaternions. $\hat{\textbf{c}}(\boldsymbol{\omega})$ is an estimate of $\textbf{c}(\boldsymbol{\omega})= \Gamma(\boldsymbol{\omega}) I \boldsymbol{\omega}$ as based on $\hat{I}$ and measured $\boldsymbol{\omega}$ . The additional term $\textbf{u}_d$ is added to render the effects of uncertainty and disturbances in order to guarantee robustness against these effects; $\textbf{u}_d$ will be defined later, to counter these, in Equation (14). The derivative of $\dot{\boldsymbol{\omega}}$ [Reference Jasim and Veres31] is

(5) \begin{equation} \begin{split}\dot{\boldsymbol{\omega}} &= I^{-1}\hat{I} \textbf{u} + I^{-1} \textbf{u}_d + I^{-1} [\boldsymbol{\Delta} (\boldsymbol{\omega}) - \boldsymbol{\tau}_d] \\&= \textbf{u} + (I^{-1} \hat{I} - \mathbb{I}) \textbf{u} + I^{-1} \textbf{u}_d + I^{-1}[\boldsymbol{\Delta} (\boldsymbol{\omega}) - \boldsymbol{\tau}_d] \\& = \textbf{u} + I^{-1} \textbf{u}_d - \textbf{y} \\\end{split}\end{equation}

where

(6) \begin{equation} \begin{split}\textbf{y} &=[\mathbb{I} - I^{-1}\hat{I}] \textbf{u} -I^{-1}[\boldsymbol{\Delta} (\boldsymbol{\omega}) - \boldsymbol{\tau}_d] \ \ , \ \ \boldsymbol{\Delta}(\boldsymbol{\omega})=\hat{\textbf{c}}(\boldsymbol{\omega})- \textbf{c}(\boldsymbol{\omega})\end{split}\end{equation}

is formalised in Isabelle/HOL based on the $att\_dyms$ , $cont\_u$ and $cont\_law$ (see the the repository [Reference Jasim and Veres43] for the details). The closed-loop error dynamics [Reference Jasim and Veres31] is

(7) \begin{equation} \begin{split}\dot{\boldsymbol{\eta}} &= A\boldsymbol{\eta} + G[\textbf{y} - I^{-1} \textbf{u}_d]\end{split}\end{equation}

where $\boldsymbol{\eta} = [\boldsymbol{\xi}^T \ \ \dot{\boldsymbol{\xi}}^T]^{T} \in \Re^{6 \times 1}$ and

(8) \begin{equation} A =\begin{bmatrix}0^{3 \times 3}\;\;\; & \mathbb{I}^{3 \times 3} \\-K_{q}^{3 \times 3}\;\;\; & -K_{\omega}^{3 \times 3}\end{bmatrix}, \ G =\begin{bmatrix}0^{3 \times 3} \\ \mathbb{I}^{3 \times 3}\end{bmatrix},\end{equation}

are implemented as

The $set\_of\_definitions$ in the above code are definitions, which rely on many available, pre-defined definitions in Isabelle. The flight envelope assumptions proposed in [Reference Jasim and Veres31] are as follows.

Assumption 1,

(9) \begin{equation}\sup (\| \dot{\boldsymbol{\omega}}_{r}\|) < \alpha.\end{equation}

Assumption 2,

(10) \begin{equation}\lambda_{\min} \le \| I^{-1}\| \le \lambda_{\max}.\end{equation}
(11) \begin{equation} \| \mathbb{I} - I^{-1}\hat{I}\| \le \delta \le 1,\end{equation}

where $\lambda_{\min}>0$ and $\lambda_{\max}>0$ a lower and upper bound, respectively; $\delta>0$ is a constant.

Assumption 3 ,

(12) \begin{equation}\| \boldsymbol{\tau}_d \| \le \gamma,\end{equation}

where $\gamma>0$ is a known upper constant bound. The above assumptions are formalised in Isabelle/HOL as follows:

Stability analysis of the attitude controller as stated in Equations (38)–(47) [Reference Jasim and Veres31] is implemented in Isabelle/HOL using a set of definitions in (definition), several lemmas in (lemma) and short theorems in terms of (theorem). This structure of using several lemmas and theorems during the proof is due to the fact that the reasoning system of the theorem prover cannot handle long proofs with many assumptions, i.e. the Isabelle system is unable to prove statements, which have many equations, if they are formalised in only one lemma or theorem style. However, stability analysis starts by defining the candidate Lyapunov function V [Reference Jasim and Veres31],

(13) \begin{equation} V(\boldsymbol{\eta}) = \boldsymbol{\eta}^{T} Q \boldsymbol{\eta} >0 \ , \ \forall \boldsymbol{\eta} \ne 0\end{equation}

which is formalised as a definition in Isabelle/HOL:

where $Q \in \Re^{6 \times 6}$ is a symmetric positive-definite matrix. Taking the candidate Lyapunov function V, the time derivative of Lyapunov function is derived and the derivations in Equations (39)–(41) in [Reference Jasim and Veres31] are proven symbolically and detailed in the online repository. An outline is shown below:

The term $\textbf{u}_\textbf{d}$ [Reference Jasim and Veres31],

(14) \begin{equation}\textbf{u}_d =\begin{cases}\dfrac{ \zeta(\boldsymbol{\eta},t)}{\|G^{T}Q \boldsymbol{\eta}\|} G^{T}Q \boldsymbol{\eta}, &\ \ \ if \ \ \|G^{T}Q \boldsymbol{\eta}\| \ge \mu\\\dfrac{ \zeta(\boldsymbol{\eta},t)}{\mu} G^{T}Q \boldsymbol{\eta},&\ \ \ if \ \ \|G^{T}Q \boldsymbol{\eta}\| < \mu.\end{cases}\end{equation}

is defined then as the derivative in Equation (43) [Reference Jasim and Veres31]. This is performed using the Cauchy-Schwartz inequality (see “theorem $Eq\_43$ ” in the repository). Based on the $\zeta(\boldsymbol{\eta},t)$ definition, $\varepsilon > 0$ , $\lambda_{\min} >0$ ,

(15) \begin{equation} \zeta(\boldsymbol{\eta},t) \ge \dfrac{\varepsilon}{\lambda_{\min}},\end{equation}

and the upper bound of the norm of $\textbf{y}$ derived in Equation (45) (31) (see “theorem $Eq\_45$ ” in the repository), $\zeta(\boldsymbol{\eta},t)$ in (15) is obtained (see “theorem $Eq\_46$ ” in the repository). The terms $\textbf{u}_\textbf{d}$ and $\zeta(\boldsymbol{\eta},t)$ are implemented in the prover as “ $u_{d}\_def$ ” and “ $zeta\_def$ ” respectively:

Note that the short arrow $\rightarrow$ in the code refers to implies, while the longer $ \longrightarrow$ refers to convergence in HOL.

Finally, based on all the above definitions and assumptions, it has been verified that the proposed control system is asymptotically stable since the time derivative of Lyapunov function [Reference Jasim and Veres31],

(16) \begin{equation}\dot{V}(\boldsymbol{\eta}) = - \boldsymbol{\eta}^{T} P \boldsymbol{\eta} + 2 \boldsymbol{\eta}^{T}QG (\textbf{y}- I^{-1} \dfrac{\zeta(\boldsymbol{\eta},t)}{\|G^{T}Q \boldsymbol{\eta}\|} G^{T}Q \boldsymbol{\eta}) <0,\end{equation}
(17) \begin{equation}\dot{V}(\boldsymbol{\eta}) = - \boldsymbol{\eta}^{T} P \boldsymbol{\eta} + 2 \boldsymbol{\eta}^{T}Q G(\textbf{y}- I^{-1} \dfrac{ \zeta(\boldsymbol{\eta},t)}{\mu} G^{T}Q \boldsymbol{\eta} ) <0.\end{equation}

is strictly negative for $\forall \boldsymbol{\eta} \ne 0$ where $A^{T}Q + QA = -P$ . It has also been proven that the tracking error converges to zero as the time converges to infinity ( $\|\boldsymbol{\eta}\| \longrightarrow 0$ ). The code below illustrates the symbolic proof in the Isabelle theorem prover.

Figure 4. Onboard verification framework of UAVs.

4.2 Onboard verification for a safe flight using MetiTarski

The control system of the UAV can be designed, simulated and verified at the modelling/design stage. The verified controller can then be converted into code and implemented on the autopilot system, which controls the aircraft trajectory. The aircraft can be exposed to gusts of wind, and the payload may also vary. Both may cause, in the extremes, unstable flight when the flight envelope is violated. The autopilot should provide a warning when the aircraft has entered an unstable region of its dynamical states as that may cause a crash or damage. This monitoring of the occurrence of instability can be implemented in an ATP tool, for instance the MetiTarski prover. The MetiTarski ATP has been chosen here to verify the controller stability of the aircraft due to its ability to deal with algebraic inequalities over numerical real numbers quickly. Unlike the previous verification stage using Isabelle, MetiTarski proves inequalities automatically, without the need for any interaction with a pilot or engineer.

MetiTarski can be implemented on the autopilot’s electronic board such as PixHawk [44] or Navio2 [45] Raspberry Pi and many others. Interaction between the autopilot and MetiTarski is easy to create in practice. A possible system architecture is illustrated in Fig. 4. The applicability of this approach is demonstrated here in simulation by interfacing the Simulink/Matlab model with MetiTarski and testing the stability of the control system. The multi-rotor model used was based on the realistic MathWorks Simulink model from [Reference Horton46].

For the stability analysis stated in Equations (16) and (17), the time derivative of the Lyapunov function needs to be tested to check whether it is negative definite or not. If it is not negative definite, then this indicates that the control system is out of its stability region, hence the autopilot can pass a warning message to the pilot to take some suitable action or perform emergency landing.

The verification process starts by formalising the stability Equations (16) and (17) into a FOL syntax. The parameters of these equations are then passed from the autopilot system to the MetiTarski prover via an interface. The test is conducted in the MetiTarski prover to test the condition ( $\dot{V}(\boldsymbol{\eta}) < 0$ ). The stability equations, Equations (16) and (17), were simplified using symbolic computations in Matlab before formalising them into FOL syntax. The parameters included in the stability equations were passed from Simulink/Matlab to MetiTarski to perform the monitoring. The following code illustrates stability checking in the MetiTarski prover as based on Equation (16) (the full code can be found in the online repository [Reference Jasim and Veres43]:

and for Equation (17),

The system entering an unstable region has been demonstrated in Simulink- Matlab by applying an external force to the nonlinear multi-rotor’s dynamics to simulate a strong gust of wind. The control system tried to overcome this dynamical change and it failed to keep the system stable because the external force was high and continuous. Without stability monitoring, this could have lead to the craft crashing.

The approach has been implemented not only to check Lyapunov stability but also to check if the limits of the system’s robustness parameters have been exceeded. This has been achieved by checking whether the maximum value of one or all of the rotor angular velocities $0<\Omega_i<\Omega_{\max}$ have been exceeded. This can be known from checking the torques for $\|\tau\| < \tau_{\max}$ , where $\tau_{\max}$ can be obtained from the known maximum angular velocity of each propeller $\Omega_{\max}$ and Equation (2). In the simulation this meant that the craft was unable to compensate against the external wind and caused the craft unable to follow its prescribed flight path. For more details, see our previous work [Reference Jasim and Veres47]. In such a situation the pilot, or autonomous agent, may decide not to insist on the planned flight path and make the most of the developing situation to save the craft by replanning.

All that has been demonstrated in simulation can also be implemented on a real UAV. This can be achieved as follows. First, implement MetiTarski on the autopilot’s electronic board with the stability and verification framework. Second, implement the interface between the autopilot and MetiTarski prover. Finally, implement the emergency decision framework inside the autopilot code as in Fig. 4.

The steps of a single cycle of stability testing, in a Simulink free onboard environment, can be implemented as follows:

  • An interface is created for the autopilot which enables communication with the MetiTarski prover.

  • The required parameters are communicated by the autopilot to the MetiTarski.

  • The interface formalises the stability condition in FOL for MetiTarski.

  • MetiTarski tests stability and communicates the result to the autopilot through the interface.

  • The autopilot informs the human pilot or the onboard decision making software.

Using these methods the autopilot can send warning messages to the pilot, or in case of an autonomous mission, to an onboard intelligent software agent, to decide on a suitable counter manoeuvre [Reference Lusk, Glaab, Glaab and Beard48].

5.0 Conclusions

A novel verification framework has been introduced for safety-critical control systems by applying the power of higher-order-logic-based interactive theorem provers and first-order logic-based automated theorem provers to verify the control system of unmanned aerial vehicles and to ensure aircraft control system stability and performance. The framework relies on two stages, the first is for verifying the design of the control system and its stability and the second is onboard monitoring of the aircraft’s stability to ensure flight safety.

The framework has been demonstrated on a robust attitude controller of a generic multi-rotor UAV to verify the correctness of the design and stability analysis in addition to onboard monitoring of the conditions of its dynamical stability while the aircraft is flying. The aircraft’s attitudes have been controlled by a nonlinear robust controller, which was designed to take into account dynamical uncertainty and external disturbances.

The methods needed in the verification stages go significantly beyond symbolic computation of inequalities for the Lyapunov theory, which can be based on first order logic (FOL). To prove robust control theory, higher order logic (HOL) concepts are needed such as quantification over sets of functions and operators as used in Isabelle HOL. These techniques were not found in prior literature on aviation control systems.

The paper illustrates the applicability and the power of interactive theorem provers relying on HOL. The methodological approach is supported by working code in an online repository [Reference Jasim and Veres43]. This has been an initial effort that may encourage the use of such methods for control system verification in aviation and for safety-critical systems in general in the future.

References

Jasim, O.A. and Veres, S.M. Towards formal proofs of feedback control theory, 2017 21st International Conference on System Theory, Control and Computing (ICSTCC), 2017, pp 4348. doi: 10.1109/ICSTCC.2017.8107009.CrossRefGoogle Scholar
Feron, E. From control systems to control software, IEEE Control Syst. Mag., 2010, 30, (6), pp 5071. doi: 10.1109/MCS.2010.938196.Google Scholar
Jobredeaux, R.J. Formal Verification of Control Software, PhD Thesis, Georgia Institute of Technology, 2015.Google Scholar
Wang, T. Credible Autocoding of Control Software, PhD Thesis, Georgia Institute of Technology, 2015.Google Scholar
INTO-CPS, [Accessed: 16 May 2021] (May 2021). URL http://projects.au.dk/into-cps/ Google Scholar
Owre, S., Rushby, J.M. and Shankar, N. PVS: A prototype verification system, in: Kapur, D. (Ed.), Automated Deduction—CADE-11, Springer, 1992, Berlin, Heidelberg, pp 748752.Google Scholar
Tiller, M. Introduction to Physical Modeling with Modelica, vol. 615, Springer Science & Business Media, 2012, New York.Google Scholar
M. Inc., Simulink/Matlab, [Accessed: 16 May 2021] (May 2021). URL https://uk.mathworks.com/ Google Scholar
20-sim, [Accessed: 16 May 2021] (May 2021). URL https://www.20sim.com/ CrossRefGoogle Scholar
Zeyda, F., Ouy, J., Foster, S. and Cavalcanti, A. Formalising cosimulation models, International Conference on Software Engineering and Formal Methods, Springer, 2017, Berlin, Heidelberg, pp 453468.CrossRefGoogle Scholar
Isabelle/UTP, [Accessed: 16 May 2021] (May 2021). URL https://www-users.cs.york.ac.uk/simonf/utp-isabelle/ Google Scholar
Team, ERATO, ERATO MMSD, [Accessed: 16 May 2021] (May 2021). URL http://www.jst.go.jp/erato/hasuo/en/ Google Scholar
NASA Langley Formal Methods Research Program, [Accessed: 16 May 2021] (May 2021). URL https://shemesh.larc.nasa.gov/fm/index.html Google Scholar
Denman, W. and Muñoz, C. Automated real proving in PVS via MetiTarski, International Symposium on Formal Methods, Springer, 2014, pp 194199.CrossRefGoogle Scholar
Muáoz, C.A. Formal methods in air traffic management: The case of unmanned aircraft systems (invited lecture), International Colloquium on Theoretical Aspects of Computing, Springer, 2015, pp 5862.CrossRefGoogle Scholar
Munoz, C.A., Dutle, A., Narkawicz, A. and Upchurch, J. Unmanned aircraft systems in the national airspace system: A formal methods perspective, ACM SIGLOG News, 2016, 3, (3), pp 6776.CrossRefGoogle Scholar
Hasan, O. and Ahmad, M. Formal analysis of steady state errors in feedback control systems using hol-light, 2013 Design, Automation Test in Europe Conference Exhibition (DATE), 2013, pp 14231426. doi: 10.7873/DATE.2013.290.CrossRefGoogle Scholar
Harrison, J. HOL Light: A tutorial introduction, in: Srivas, M. and Camilleri, A. (Eds), Formal Methods in Computer-Aided Design, Springer, 1996, Berlin, Heidelberg, pp 265269.Google Scholar
Rashid, A. and Hasan, O. Formal analysis of linear control systems using theorem proving, in: Duan, Z. and Ong, L. (Eds), Formal Methods and Software Engineering, Springer International Publishing, 2017, Cham, pp 345361.Google Scholar
Ma, Z. and Chen, G. Formal derivation and verification of coordinate transformations in theorem prover Coq, 2017 International Conference on Dependable Systems and Their Applications (DSA), 2017, pp 127136. doi: 10.1109/DSA.2017.29.CrossRefGoogle Scholar
Huet, G.K.G. and Paulin-Mohring, C. The coq proof assistant: A tutorial, Tech Rep 178 [Online]. Available: http://cs.swan.ac.uk/csoliver/ok-sat-library/OKplatform/ExternalSources/sources/Coq/Tutorial.pdf, National Institute of Research in Information and Automation (INRIA), 2009.Google Scholar
Domenici, A., Fagiolini, A. and Palmieri, M. Integrated simulation and formal verification of a simple autonomous vehicle, in: Cerone, A. and Roveri, M. (Eds), Software Engineering and Formal Methods, Springer International Publishing, 2018, Cham, pp 300314.Google Scholar
Denman, W., Zaki, M.H., Tahar, S. and Rodrigues, L. Towards flight control verification using automated theorem proving, in: Bobaru, M., Havelund, K., Holzmann, G.J. and Joshi, R. (Eds), NASA Formal Methods, Springer, 2011, Berlin, Heidelberg, pp 89100.Google Scholar
Paulson, L.C. Metitarski: Past and future, International Conference on Interactive Theorem Proving, Springer, 2012, pp 110.CrossRefGoogle Scholar
Araiza-Illan, D., Eder, K. and Richards, A. Formal verification of control systems’ properties with theorem proving, 2014 UKACC International Conference on Control (CONTROL), 2014, pp 244249. doi: 10.1109/CONTROL.2014.6915147.CrossRefGoogle Scholar
Filliâtre, J.-C. and Paskevich, A. Why3 — where programs meet provers, in: Felleisen, M. and Gardner, P. (Eds), Programming Languages and Systems, Springer, 2013, Berlin, Heidelberg, pp 125128.Google Scholar
Platzer, A. Differential dynamic logic for hybrid systems, J. Autom. Reason., 2008, 41, (2), pp 143189. doi: 10.1007/s10817-008-9103-8. URL https://doi.org/10.1007/s10817-008-9103-8 CrossRefGoogle Scholar
Aréchiga, N., Loos, S.M., Platzer, A. and Krogh, B.H. Using theorem provers to guarantee closed-loop system properties, 2012 American Control Conference (ACC), IEEE, 2012, pp 35733580. doi: 10.1109/ACC.2012.6315388.Google Scholar
Aréchiga, N. and Krogh, B. Using verified control envelopes for safe controller design, 2014 American Control Conference, IEEE, 2014, pp 29182923. doi: 10.1109/ACC.2014.6859307.CrossRefGoogle Scholar
Nipkow, T., Paulson, L.C. and Wenzel, M. Isabelle/HOL: A Proof Assistant for Higher-Order Logic, vol. 2283, Springer Science and Business Media, 2002, Berlin, Heidelberg. doi: 10.1007/3-540-45949-9.CrossRefGoogle Scholar
Jasim, O.A. and Veres, S.M. A robust controller for multi rotor UAV, Aerospace Sci. Technol., 2020, 105, 106010.CrossRefGoogle Scholar
Spong, M.W., Hutchinson, S. and Vidyasagar, M. Robot Modeling and Control, vol. 3, Wiley, 2006, New York.Google Scholar
Sciavicco, L. and Siciliano, B. Modelling and Control of Robot Manipulators , Springer Science and Business Media, 2012, Berlin, Heidelberg. doi: 10.1007/978-1-4471-0449-0.Google Scholar
Slotine, J.-J.E., Li, W. Applied Nonlinear Control, vol. 199, Prentice-Hall, 1991, Englewood Cliffs, New Jersey.Google Scholar
Sutcliffe, Geoff and Suttner, Christian, System on TPTP, [Accessed: 16 May 2021]. (2001). URL http://www.tptp.org/cgi-bin/SystemOnTPTP Google Scholar
Clarke, E.M., Grumberg, O. and Peled, D. Model Checking, MIT Press, 1999, London.Google Scholar
Milner, R. A theory of type polymorphism in programming, J. Comput. Syst. Sci., 1978 17, (3), pp 348375.CrossRefGoogle Scholar
Barrett, C. and Tinelli, C. Satisfiability modulo theories, in: Handbook of Model Checking, Springer, 2018, pp 305343.CrossRefGoogle Scholar
Nipkow, T. Hoare logics in Isabelle/HOL, in: Proof and System-Reliability, Springer, 2002, Berlin, Heidelberg, pp 341367.CrossRefGoogle Scholar
Hurd, J. First-order proof tactics in higher-order logic theorem provers, Design and Application of Strategies/Tactics in Higher Order Logics, number NASA/CP-2003-212448 in NASA Tech Rep, 2003, pp 5668.Google Scholar
Brown, C.W. QEPCAD B: A program for computing with semi-algebraic sets using cads, ACM SIGSAM Bull., 37, (4), 2003, pp 97108.CrossRefGoogle Scholar
De Moura, L. and Bjørner, N. Z3: An efficient SMT solver, Tools and Algorithms for the Construction and Analysis of Systems, 2008, Berlin, Heidelberg, pp 337340.CrossRefGoogle Scholar
Jasim, O.A. and Veres, S.M. Isabelle and MetiTraski Code for Controller Verification of an Attitude Controller for Multirotor UAV, [Accessed: 14 May 2021] (May 2021). URL https://github.com/ojrobotics/fvuav Google Scholar
PixHawk, [Accessed: 16 May 2021] (2021). URL http://pixhawk.org/ Google Scholar
Emlid. Navio2, [Accessed: 16 May 2021] (May 2021). URL https://emlid.com/navio/ Google Scholar
Horton, B. Modelling simulation and control of a quadcopter, MATLAB Academic Conference. Australia and New Zealand, 2016, pp 4–14, [Accessed: 16 May 2021]. URL https://uk.mathworks.com/videos/modelling-simulation-and-control-of-a-quadcopter-122872.html Google Scholar
Jasim, O.A. and Veres, S.M. Nonlinear attitude control design and verification for a safe flight of a small-scale unmanned helicopter, 2019 6th International Conference on Control, Decision and Information Technologies (CoDIT), 2019, pp 16521657. doi: 10.1109/CoDIT.2019.8820310.CrossRefGoogle Scholar
Lusk, P.C., Glaab, P.C., Glaab, L.J. and Beard, R.W. Safe2ditch: Emergency landing for small unmanned aircraft systems, J. Aerospace Inf. Syst., 2019, pp 327339.CrossRefGoogle Scholar
Figure 0

Table 1. Isabelle/HOL symbols and expressions

Figure 1

Figure 1. UAVs verification framework.

Figure 2

Figure 2. The inner control loop of the multi-rotor controller [31] for attitude control. q is the measured and $q_r$ is the desired attitude, $q_e$ is its error in quaternions, $\xi$ is angular error vector, $\omega, \omega_r$ are the angular velocity and reference, $K_q, K_\omega$ are control gains, $I,\tilde{I}$ are the inertia matrix and its estimate. $\Gamma(\omega)$ is a cross-product matrix for the Coriolis forces such that $\Gamma(\omega) I \omega = \omega \times I \omega$ is satisfied and $u_d$ is torque disturbance of the attitude.

Figure 3

Figure 3. Formalising and proving UAV’s controller in Isabelle/HOL theorem prover.

Figure 4

Figure 4. Onboard verification framework of UAVs.