This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Abstraction-Free Control Synthesis to Satisfy Temporal Logic Constraints under Sensor Faults and Attacks

Luyao Niu1, Zhouchi Li2, and Andrew Clark3 1Luyao Niu is with the Network Security Lab, Department of Electrical and Computer Engineering, University of Washington, Seattle, WA 98195-2500, [email protected]2Z. Li is with the Department of Electrical and Computer Engineering, Worcester Polytechnic Institute, Worcester, MA 01609, [email protected]3Andrew Clark is with the Electrical and Systems Engineering Department, McKelvey School of Engineering, Washington University in St. Louis, St. Louis, MO 63130, [email protected]This work was supported by the National Science Foundation and the Office of Naval Research via grants CNS-1941670 and N00014-17-S-B001.
Abstract

We study the problem of synthesizing a controller to satisfy a complex task in the presence of sensor faults and attacks. We model the task using Gaussian distribution temporal logic (GDTL), and propose a solution approach that does not rely on computing any finite abstraction to model the system. We decompose the GDTL specification into a sequence of reach-avoid sub-tasks. We develop a class of fault-tolerant finite time convergence control barrier functions (CBFs) to guarantee that a dynamical system reaches a set within finite time almost surely in the presence of malicious attacks. We use the fault-tolerant finite time convergence CBFs to guarantee the satisfaction of ‘reach’ property. We ensure ‘avoid’ part in each sub-task using fault-tolerant zeroing CBFs. These fault-tolerant CBFs formulate a set of linear constraints on the control input for each sub-task. We prove that if the error incurred by system state estimation is bounded by a certain threshold, then our synthesized controller fulfills each reach-avoid sub-task almost surely for any possible sensor fault and attack, and thus the GDTL specification is satisfied with probability one. We demonstrate our proposed approach using a numerical study on the coordination of two wheeled mobile robots.

I Introduction

Cyber-physical systems (CPS) in applications including autonomous vehicles, industrial control systems, and robotics must complete increasingly complex tasks. Temporal logics [1] such as linear temporal logic (LTL) are used to specify tasks because of their concreteness, rigor, and rich expressiveness. To model the process and observation noises that are commonly incurred by CPS, Gaussian distribution temporal logic (GDTL) has been proposed to specify CPS properties involving the uncertainties in the system state [2, 3]. Controller design for CPS under different temporal logic constraints has attracted extensive research attention [4, 5, 2, 3]. There have been two types of methodologies for temporal logic based control synthesis, namely abstraction-based and abstraction-free approaches.

The abstraction-based approaches [6, 7, 8, 2, 3, 4, 5] lift the CPS model from continuous domain to discrete domain by computing a finite abstraction, e.g., finite transition systems [6] and Markov decision processes (MDPs) [7, 8]. A controller is synthesized by applying model checking algorithms on the finite abstraction. In general, the abstraction-based approaches are computationally demanding and suffer from the curse of dimensionality. To mitigate the computation required by the abstraction-based approaches, abstraction-free approaches have been proposed [9, 10, 11, 12, 13, 14, 15]. These approaches eliminate the step of constructing the finite abstractions, and directly compute the controller in the continuous state space.

The previously introduced abstraction-based [6, 7, 8, 2, 3, 4, 5] and abstraction-free approaches [9, 10, 11, 12, 13, 14, 15] assume that there exist no faults or malicious attacks targeting the system sensors. However, CPS have been shown to be vulnerable to sensor faults and malicious attacks [16, 17, 18, 19], making temporal logic based control synthesis more challenging. Sensor faults and malicious attacks can arbitrarily manipulate the sensor measurements. Inaccurate measurements prevent the CPS from correctly evaluating and tracking the satisfaction of temporal logic specifications. Moreover, the faulty or compromised measurements can bias the state estimate and lead to erroneous controller actions, which deviate the system trajectory and violate the temporal logic specification. Although there has been research attention on fault and attack detection [17, 20, 19] and low-level control design under sensor attacks [21], abstraction-free temporal logic based control synthesis under sensor faults and attacks has not been studied.

In this paper, we study the problem of synthesizing a controller for a nonlinear control-affine system to satisfy a GDTL specification in the presence of process and observation noises as well as malicious attacks on sensor measurements. We assume that the adversary has finitely many choices of the sensors to attack. The system estimates the state by employing a set of Extended Kalman Filters (EKFs) associated with each possible set of sensors attacked by the adversary. If there exists a control input that satisfies the GDTL specification for all of the state estimates, then the GDTL formula can be satisfied regardless of the attack. When such control input does not exist, it indicates that the state estimates given by some EKFs conflict with the others. We view two EKFs as conflicting if the divergence between their state estimates exceeds a certain threshold. We then compute a baseline state estimate without using any sensor corresponding to the conflicting EKFs. We compare this baseline with those from the conflicting EKFs, and treat the one that diverges from the baseline as being compromised. We make the following specific contributions:

  • We decompose the GDTL specification into a sequence of reach-avoid sub-tasks with the system state uncertainties being explicitly encoded.

  • We synthesize the controller for each decomposed problem by deriving a set of sufficient conditions. We guarantee the ‘reach’ and ‘avoid’ properties in each sub-task by developing a class of fault-tolerant finite time convergence control barrier functions and applying the fault-tolerant zeroing control barrier functions.

  • We prove that if the errors incurred by the EKFs are bounded by some thresholds, then each decomposed reach-avoid sub-task is satisfied with probability one using our proposed approach, and thus the GDTL specification is satisfied almost surely.

  • We evaluate our proposed approach using a numerical case study on two wheeled mobile robots. We show that our proposed approach satisfies the given GDTL specification, while a baseline fails.

The rest of this paper is organized as follows. Section II reviews related work. Section III presents the system model and reviews necessary background on finite time stability in probability, EKF, and GDTL. Section IV presents our proposed solution approach. A numerical case study is presented in Section V. Section VI concludes the paper.

II Related Work

Abstraction-based approaches have been proposed for control synthesis of CPS in the absence of observation noise and sensor attacks under LTL constraints [4, 6, 7, 8]. When CPS are subject to observation noises, GDTL has been proposed to model the uncertainties [2, 3].

When CPS operate under malicious attacks targeting the system actuators, abstraction-based approaches for control synthesis under LTL constraints have been studied in [22, 23], where the CPS are abstracted as finite stochastic games. For CPS under partially observable environments, partially observable stochastic game is used as the finite abstraction for control synthesis [24]. These works [22, 23, 24] and the present paper focus on different scopes of attacks. In [22, 23, 24], sensor faults or attacks are not considered, and thus the system state is known by the CPS. This paper focuses on sensor faults and attacks. In addition, our proposed approach does not compute a finite abstraction for the CPS, and thus belongs to the class of abstraction-free approaches.

Abstraction-free approaches eliminate the step for computing the finite abstractions to improve the scalability. In [25], the LTL specification is encoded as a set of mixed integer linear constraints. In [10], the authors solve a sequence of stochastic reachability problems via nonlinear PDEs to satisfy a co-safe LTL specification. Alternatively, a mixed continuous-discrete HJB-based formulation is proposed in [11]. In general, solving PDEs and HJB equations are computationally expensive.

In [13, 12, 15], the authors adopt control barrier functions (CBFs), which are originally proposed for safety-critical control synthesis [26, 27], to compute control policies that satisfy LTL specifications for deterministic CPS. In parallel, (control) barrier certificate based verification and synthesis for temporal logic properties have been investigated in [28, 29]. An abstraction-free safe learning scheme is proposed in [30] for systems under adversarial actuator attacks. Our paper differs from these barrier function based works in two aspects. First, we consider the presence of sensor faults and attacks. Second, we develop a class of fault-tolerant finite time convergence CBF, in combined with the recent advancement in fault-tolerant zeroing CBF, to address the presence of noises, faults, and attacks.

III Preliminaries and Problem Formulation

In this section, we first present the system model. We then give some background on GDTL and EKF. We finally present the problem formulation.

III-A System Model

In this subsection, we first give the notations that we will use throughout this paper. We next present the system model.

Notations. We use n\mathbb{R}^{n} to denote the nn-dimensional Euclidean space. Given a vector ana\in\mathbb{R}^{n}, we use supp(a)supp(a) and [a]i[a]_{i} to denote its support and ii-th entry, respectively. For any matrix AA, we denote its trace as tr(A)tr(A). The set of positive semi-definite matrices of dimension n×nn\times n is represented as 𝕊n×n\mathbb{S}^{n\times n}. We denote the set of non-negative real numbers as 0\mathbb{R}_{\geq 0}. A function α:00\alpha:\mathbb{R}_{\geq 0}\rightarrow\mathbb{R}_{\geq 0} belongs to class 𝒦\mathcal{K} if it is continuous, strictly increasing, and α(0)=0\alpha(0)=0. A class 𝒦\mathcal{K} function belongs to class 𝒦\mathcal{K}_{\infty} if α(x)\alpha(x)\rightarrow\infty when xx\rightarrow\infty. We use ()\mathbb{P}(\cdot) to represent the probability of an event.

Consider a nonlinear dynamical system defined as

dxt=(f(xt)+g(xt)ut)dt+σtdWt,dx_{t}=(f(x_{t})+g(x_{t})u_{t})\ dt+\sigma_{t}\ dW_{t}, (1)

where xtnx_{t}\in\mathbb{R}^{n} is the system state, utpu_{t}\in\mathbb{R}^{p} is the control input at time tt, σtn×n\sigma_{t}\in\mathbb{R}^{n\times n}, and WtW_{t} is an nn-dimensional Brownian motion. Functions f:nnf:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n} and g:nn×pg:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n\times p} are locally Lipschitz continuous.

We denote the system output at time tt as ytqy_{t}\in\mathbb{R}^{q}. The output yty_{t} follows an observation process [31] given as:

dyt=(cxt+at)dt+νtdVt,dy_{t}=(cx_{t}+a_{t})\ dt+\nu_{t}\ dV_{t}, (2)

where cq×nc\in\mathbb{R}^{q\times n}, νtq×q\nu_{t}\in\mathbb{R}^{q\times q}, and VtV_{t} is a qq-dimensional Brownian motion. We assume that the system operates in the presence of a malicious adversary. The adversary can inject a signal atqa_{t}\in\mathbb{R}^{q} to manipulate the output yty_{t} with supp(at){1,,q}supp(a_{t})\subseteq\{1,\ldots,q\}. We assume that once the adversary determines the set of sensors to attack, it cannot adjust the support of the attack signal, i.e., the adversary cannot change its target after the attack starts. We assume that there are finitely many choices for the support of ata_{t}, denoted as ={r1,,rm}\mathcal{F}=\{r_{1},\ldots,r_{m}\}, which is also known to the system. In the remainder of this paper, we refer to rir_{i} as a fault pattern. However, the choice supp(at)supp(a_{t}) made by the adversary is not known to the system.

III-B Finite Time Stability in Probability of Stochastic Nonlinear Systems

In this subsection, we introduce background on finite time stability of stochastic nonlinear systems. Consider a stochastic nonlinear system with complete state information given by Eqn. (1). We then give the definition of finite time stability in probability.

Definition 1 (Finite Time Stability in Probability [32]).

The solution of system (1), denoted as x(t;x0)x(t;x_{0}), is said to be finite time stable in probability if system (1) admits a unique solution for any initial state x0nx_{0}\in\mathbb{R}^{n} and the following conditions hold:

  1. 1.

    Finite time attractiveness in probability: For every initial state x0𝟎x_{0}\neq\mathbf{0}, the stochastic settling time τx0=inf{t:x(t;x0)=𝟎}\tau_{x_{0}}=\inf\{t:x(t;x_{0})=\mathbf{0}\} is finite almost surely, i.e., (τx0<)=1\mathbb{P}(\tau_{x_{0}}<\infty)=1.

  2. 2.

    Stability in probability: For every pair of ϵ(0,1)\epsilon\in(0,1) and χ>0\chi>0, there exists some constant vv such that

    (|x(t;x0)|<χ,t0)1ϵ,\mathbb{P}(|x(t;x_{0})|<\chi,~{}\forall t\geq 0)\geq 1-\epsilon, (3)

    for any |x0|<v|x_{0}|<v.

Let VV be a twice continuously differentiable function. Finite time stability in probability is then certified by the following stochastic Lyapunov theorem.

Theorem 1 (Stochastic Lyapunov Theorem [32]).

Assume system (1) admits a unique solution. If there exists a twice continuously differentiable function V:n0V:\mathbb{R}^{n}\rightarrow\mathbb{R}_{\geq 0}, class 𝒦\mathcal{K}_{\infty} functions α1\alpha_{1} and α2\alpha_{2}, positive constants γ>0\gamma>0, and ρ(0,1)\rho\in(0,1) such that

α1(|x|)V(x)α2(|x|),\displaystyle\alpha_{1}(|x|)\leq V(x)\leq\alpha_{2}(|x|), (4a)
Vx(x)(f(x)+g(x)u)+12tr(σ2V(x)x2σ)\displaystyle\frac{\partial V}{\partial x}(x)(f(x)+g(x)u)+\frac{1}{2}tr\left(\sigma^{\top}\frac{\partial^{2}V(x)}{\partial x^{2}}\sigma\right)
+γV(x)ρ0\displaystyle\quad\quad\quad\quad\quad+\gamma\cdot V(x)^{\rho}\leq 0 (4b)

then the solution of system (1) is finite time stable in probability.

III-C Extended Kalman Filter

The system given in Eqn. (1) and (2) employs a set of EKFs [33] to estimate the system state. Each EKF corresponds to a potential fault pattern rir_{i}\in\mathcal{F}. We briefly review the background on EKF in the following.

We denote the state estimate at time tt as x^t,i\hat{x}_{t,i} when fault pattern rir_{i} is chosen by the adversary. Then the dynamics of x^t,i\hat{x}_{t,i} can be represented as

x^t,i=(f(x^t,i)+g(x^t,i)ut)dt+Kt,i(dyt,icix^t,idt),\hat{x}_{t,i}=(f(\hat{x}_{t,i})+g(\hat{x}_{t,i})u_{t})dt+K_{t,i}(dy_{t,i}-c_{i}\hat{x}_{t,i}dt), (5)

where Kt,i=Pt,iciRt,i1K_{t,i}=P_{t,i}c_{i}^{\top}R_{t,i}^{-1}, Rt,i=νt,iνt,iR_{t,i}=\nu_{t,i}\nu_{t,i}^{\top}, and cic_{i} as well as yt,iy_{t,i} are obtained from matrix cc and vector yty_{t} by removing the corresponding rows indexed by rir_{i}. Matrix νt,i\nu_{t,i} is obtained from νt\nu_{t} by removing the rows and columns indexed by rir_{i}. The covariance of xt,ix_{t,i}, denoted as Pt,iP_{t,i}, is the solution to

dPt,idt=At,iPt,i+Pt,iAt,i+QtPt,iciRt,i1cPt,i,\frac{dP_{t,i}}{dt}=A_{t,i}P_{t,i}+P_{t,i}A_{t,i}^{\top}+Q_{t}-P_{t,i}c_{i}^{\top}R_{t,i}^{-1}cP_{t,i}, (6)

where Qt=σtσtQ_{t}=\sigma_{t}\sigma_{t}^{\top} and At,i=f¯x(x^t,u)A_{t,i}=\frac{\partial\bar{f}}{\partial x}(\hat{x}_{t},u) with f¯(x^t,i,u)=f(x^t,i)+g(x^t,i)u\bar{f}(\hat{x}_{t,i},u)=f(\hat{x}_{t,i})+g(\hat{x}_{t,i})u. A belief state bt,in×𝕊n×n𝔹b_{t,i}\in\mathbb{R}^{n}\times\mathbb{S}^{n\times n}\triangleq\mathbb{B} corresponding to fault pattern rir_{i} is defined as bt,i=(x^t,i,Pt,i)b_{t,i}=(\hat{x}_{t,i},P_{t,i}). The evolution of the belief state is described by Eqn. (5) and (6). We make the following assumptions on Eqn. (1) and (2).

Assumption 1.

SDEs in Eqn. (1) and (2) satisfy the following conditions for all fault patterns rir_{i}:

  1. 1.

    The pair [f¯x(x^t,i,u),c][\frac{\partial\bar{f}}{\partial x}(\hat{x}_{t,i},u),c] is uniformly detectable;

  2. 2.

    There exist constants β1\beta_{1} and β2\beta_{2} such that Qtβ1IQ_{t}\geq\beta_{1}I and Rt,iβ2IR_{t,i}\geq\beta_{2}I for all tt;

  3. 3.

    Let

    Δ(xt,x^t,i,u)=f¯(xt,u)f¯(x^t,i,u)f¯x(xtx^t,i,u).\Delta(x_{t},\hat{x}_{t,i},u)=\bar{f}(x_{t},u)-\bar{f}(\hat{x}_{t,i},u)-\frac{\bar{f}}{\partial x}(x_{t}-\hat{x}_{t,i},u).

    Then there exist real numbers kΔk_{\Delta} and ϵΔ\epsilon_{\Delta} such that Δ(xt,x^t,i,u)kΔxtx^t,i22\|\Delta(x_{t},\hat{x}_{t,i},u)\|\leq k_{\Delta}\|x_{t}-\hat{x}_{t,i}\|_{2}^{2} for all xtx_{t} and x^t,i\hat{x}_{t,i} satisfying xtx^t,i2ϵΔ\|x_{t}-\hat{x}_{t,i}\|_{2}\leq\epsilon_{\Delta}.

Given Assumption 1, we have the following result.

Lemma 1 ([33]).

Suppose the conditions of Assumption 1 hold. If there exists ε>0\varepsilon>0 such that if QtεIQ_{t}\leq\varepsilon I and Rt,iεIR_{t,i}\leq\varepsilon I, then for any ζ>0\zeta>0, there exists γ>0\gamma>0 such that

(supt0xtx^t,i2γ)1ζ.\mathbb{P}\left(\sup_{t\geq 0}\|x_{t}-\hat{x}_{t,i}\|_{2}\leq\gamma\right)\geq 1-\zeta. (7)

when the fault pattern is rir_{i}.

III-D Gaussian Distribution Temporal Logic and Problem Formulation

In this subsection, we introduce GDTL [2, 3], which will be used to model the task that needs to be satisfied by system (1)-(2). We then formulate the problem studied in this paper.

The syntax of GDTL is defined as follows:

φ=True|l0|¬φ|φ1φ2|φ1𝒰φ2,\varphi=True|l\leq 0|\neg\varphi|\varphi_{1}\land\varphi_{2}|\varphi_{1}\mathcal{U}\varphi_{2}, (8)

where l:𝔹l:\mathbb{B}\rightarrow\mathbb{R}, and l0l\leq 0 is a predicate.

Let 𝐛=b0b1\mathbf{b}=b^{0}b^{1}\ldots be an infinite sequence of belief states with each bi𝔹b^{i}\in\mathbb{B}. We denote the suffix btbt+1b^{t}b^{t+1}\ldots of 𝐛\mathbf{b} as 𝐛t\mathbf{b}^{t}. Then a GDTL formula is interpreted as follows:

  • 𝐛tTrue\mathbf{b}^{t}\models True

  • 𝐛tl0\mathbf{b}^{t}\models l\leq 0 iff l(bt)0l(b^{t})\leq 0

  • 𝐛t¬φ\mathbf{b}^{t}\models\neg\varphi iff ¬(𝐛tφ)\neg(\mathbf{b}^{t}\models\varphi)

  • 𝐛tφ1φ2\mathbf{b}^{t}\models\varphi_{1}\land\varphi_{2} iff (𝐛tφ1)(𝐛tφ2)(\mathbf{b}^{t}\models\varphi_{1})\land(\mathbf{b}^{t}\models\varphi_{2})

  • 𝐛tφ1𝒰φ2\mathbf{b}^{t}\models\varphi_{1}\mathcal{U}\varphi_{2} iff tt\exists t^{\prime}\geq t s.t. (𝐛tφ2)(𝐛t′′φ1)(\mathbf{b}^{t^{\prime}}\models\varphi_{2})\land(\mathbf{b}^{t^{\prime\prime}}\models\varphi_{1}) for all t′′<tt^{\prime\prime}<t^{\prime}

Let 𝔹l={b:l(b)0}\mathbb{B}_{l}=\{b:l(b)\leq 0\} be the set of belief states such that predicate ll is true and φ\mathcal{L}_{\varphi} be the set of predicates that appear in a GDTL formula φ\varphi. We present the following proposition [3] so as to construct an equivalent deterministic Rabin automaton (DRA) representing GDTL formula φ\varphi.

Proposition 1 ([3]).

Let φ\varphi be a GDTL formula. Let Π~\tilde{\Pi} be an additional finite set of atomic propositions such that |φ|=|Π~||\mathcal{L}_{\varphi}|=|\tilde{\Pi}|. Let e:φΠ~e:\mathcal{L}_{\varphi}\rightarrow\tilde{\Pi} be a bijective map. Then GDTL formula φ\varphi is equivalent to an LTL formula φ~\tilde{\varphi} defined over atomic proposition set ΠΠ~\Pi\cup\tilde{\Pi}, with each predicate ll in φ\mathcal{L}_{\varphi} being replaced by e(l)Π~e(l)\in\tilde{\Pi}.

Leveraging Proposition 1, we can represent a GDTL formula using an equivalent DRA defined as follows:

Definition 2 (Deterministic Rabin Automaton (DRA) [1]).

A Deterministic Rabin Automaton (DRA) is a tuple 𝒜=(𝒬,Σ,δ,q0,F)\mathcal{A}=(\mathcal{Q},\Sigma,\delta,q_{0},F), where 𝒬\mathcal{Q} is a finite set of states, Σ\Sigma is the finite set of alphabet, δ:𝒬×Σ𝒬\delta:\mathcal{Q}\times\Sigma\rightarrow\mathcal{Q} is a finite set of transitions, q0𝒬q_{0}\in\mathcal{Q} is the initial state, and F={(B(1),C(1)),,(B(S),C(S))}F=\{(B(1),C(1)),\ldots,(B(S),C(S))\} is a finite set of Rabin pairs such that B(s),C(s)𝒬B(s),C(s)\subseteq\mathcal{Q} for all s=1,,Ss=1,\ldots,S with SS being a positive integer.

We consider that the system given by Eqn. (1) and (2) needs to satisfy a GDTL specification φ\varphi. We define a labeling function L:𝔹2Π~L:\mathbb{B}\rightarrow 2^{\tilde{\Pi}}, where Π~\tilde{\Pi} is the set of atomic propositions generated by Proposition 1. For any πΠ~\pi\in\tilde{\Pi}, we define π\llbracket\pi\rrbracket as π={b:πL(b)}\llbracket\pi\rrbracket=\{b:\pi\in L(b)\}. Consider Z2Π~Z\in 2^{\tilde{\Pi}}. We define Z\llbracket Z\rrbracket as

Z={𝔹πΠ~π if Z=πZππΠZπ otherwise\llbracket Z\rrbracket=\begin{cases}\mathbb{B}\setminus\cup_{\pi\in\tilde{\Pi}}\llbracket\pi\rrbracket&\mbox{ if }Z=\emptyset\\ \cap_{\pi\in Z}\llbracket\pi\rrbracket\setminus\cup_{\pi\in\Pi\setminus Z}\llbracket\pi\rrbracket&\mbox{ otherwise}\end{cases} (9)

We let 𝐛\mathbf{b} be a trajectory of belief state of infinite length. We now define the trace of 𝐛\mathbf{b} to establish the connection between the satisfaction of φ\varphi to trajectory 𝐛\mathbf{b}.

Definition 3 (Trace of Trajectory [34]).

An infinite sequence Trace(𝐛)=Z0,Z1,,Zi,Trace(\mathbf{b})=Z_{0},Z_{1},\ldots,Z_{i},\ldots, where Zi2Π~Z_{i}\in 2^{\tilde{\Pi}} for all i=0,1,i=0,1,\ldots is a trace of a trajectory 𝐛\mathbf{b} if there exists a sequence t0,t1,,ti,t_{0},t_{1},\ldots,t_{i},\ldots of time instants such that

  1. 1.

    t0=0t_{0}=0

  2. 2.

    tit_{i}\rightarrow\infty as ii\rightarrow\infty

  3. 3.

    ti<ti+1t_{i}<t_{i+1}

  4. 4.

    x^tiZi\hat{x}_{t_{i}}\in\llbracket Z_{i}\rrbracket

  5. 5.

    if ZiZi+1Z_{i}\neq Z_{i+1}, then there exists some ti[ti,ti+1]t_{i}^{\prime}\in[t_{i},t_{i+1}] such that x^tZi\hat{x}_{t}\in\llbracket Z_{i}\rrbracket for all t(ti,ti)t\in(t_{i},t_{i}^{\prime}), x^tZi+1\hat{x}_{t}\in\llbracket Z_{i+1}\rrbracket for all t(ti,ti+1)t\in(t_{i}^{\prime},t_{i+1}), and either x^tiZi\hat{x}_{t_{i}^{\prime}}\in\llbracket Z_{i}\rrbracket or x^tiZi+1\hat{x}_{t_{i}^{\prime}}\in\llbracket Z_{i+1}\rrbracket.

Using the trace of system trajectory, we have that GDTL specification φ\varphi is satisfied if Trace(𝐛)φTrace(\mathbf{b})\models\varphi. The problem investigated in this paper is then formulated as follows:

Problem 1.

Given a parameter ϵ(0,1)\epsilon\in(0,1), compute a control policy μ:{yt:t[0,t)}p\mu:\{y_{t^{\prime}}:t^{\prime}\in[0,t)\}\rightarrow\mathbb{R}^{p} for the system at each time tt mapping from the sequence of outputs {yt:t[0,t)}\{y_{t^{\prime}}:t^{\prime}\in[0,t)\} to a control input utu_{t}, so that the probability of satisfying φ\varphi, denoted as (φ)\mathbb{P}(\varphi), satisfies (φ)1ϵ\mathbb{P}(\varphi)\geq 1-\epsilon for any fault rr\in\mathcal{F} when fault rr occurs.

IV Solution Approach

In this section, we present the proposed solution approach. We first derive a fault-tolerant stochastic finite time convergence CBF. We then decompose the GDTL formula into a sequence of sub-formulae using the DRA representing φ\varphi. We show that the satisfaction probability of GDTL formula φ\varphi can be expressed using that of each sub-formula. We then synthesize a control policy that enables the system to satisfy each sub-formula with certain probability guarantee using fault-tolerant stochastic finite time convergence CBF and fault-tolerant stochastic zeroing CBF. We finally derive the satisfaction probability guarantee for each sub-formula.

IV-A Fault-Tolerant Stochastic Finite Time Convergence CBF

This subsection derives fault-tolerant stochastic finite time convergence CBF. We first define stochastic finite time convergence CBF under complete information in the absence of an adversary. Then we extend it to the incomplete information setting. We finally propose the fault-tolerant stochastic finite time convergence CBF to address the presence of the adversary.

Definition 4 (Stochastic Finite Time Convergence CBF under Complete Information).

Consider the dynamical system given in Eqn. (1) whose solution is denoted as x(t;x0)x(t;x_{0}). Consider a set 𝒞={x:h(x)0}\mathcal{C}=\{x:h(x)\geq 0\} where h:nh:\mathbb{R}^{n}\rightarrow\mathbb{R} is a twice continuously differentiable function. We say hh is a stochastic finite time convergence CBF if the following conditions hold:

  1. 1.

    There exist class 𝒦\mathcal{K}_{\infty} functions α1\alpha_{1} and α2\alpha_{2} such that

    α1(h(x))h(x)α2(h(x))\alpha_{1}(h(x))\leq h(x)\leq\alpha_{2}(h(x)) (10)

    for all xx.

  2. 2.

    There exist control input uu such that

    hx(x)(f(x)+g(x)u)+12tr(σ2h(x)x2σ)+sgn(h(x))|h(x)|0.\frac{\partial h}{\partial x}(x)(f(x)+g(x)u)+\frac{1}{2}tr\left(\sigma^{\top}\frac{\partial^{2}h(x)}{\partial x^{2}}\sigma\right)\\ +sgn(h(x))|h(x)|\geq 0. (11)

Stochastic finite time convergence CBF provides the finite time stability in probability guarantee as given by Definition 1. We formally state this result as follows:

Lemma 2.

Let 𝒞={x:h(x)0}\mathcal{C}=\{x:h(x)\geq 0\}. Suppose the system admits a stochastic finite time convergence CBF h(x)h(x). Let T=inf{t:x(t;x0)𝒞}T=\inf\{t:x(t;x_{0})\in\mathcal{C}\}. Then (T<)=1\mathbb{P}(T<\infty)=1, i.e., the system reaches 𝒞\mathcal{C} within finite time almost surely. Moreover, the system remains in 𝒞\mathcal{C} for all tTt^{\prime}\geq T almost surely.

Proof.

We define a function V(h(x))V(h(x)) as V(h(x))=h(x)2V(h(x))=h(x)^{2}. Using the definition of VV, we have that

Vx(x)(f(x)+g(x)u)+12tr(σ2V(x)x2σ)\displaystyle\frac{\partial V}{\partial x}(x)(f(x)+g(x)u)+\frac{1}{2}tr\left(\sigma^{\top}\frac{\partial^{2}V(x)}{\partial x^{2}}\sigma\right)
=\displaystyle= 2h(x)[hx(x)(f(x)+g(x)u)+12tr(σ2h(x)x2σ)]\displaystyle 2h(x)\left[\frac{\partial h}{\partial x}(x)(f(x)+g(x)u)+\frac{1}{2}tr\left(\sigma^{\top}\frac{\partial^{2}h(x)}{\partial x^{2}}\sigma\right)\right]

If Eqn. (11) holds and x𝒞x\notin\mathcal{C}, we have that

Vx(x)(f(x)+g(x)u)+12tr(σ2V(x)x2σ)2h(x)|V(x)|120.\frac{\partial V}{\partial x}(x)(f(x)+g(x)u)+\frac{1}{2}tr\left(\sigma^{\top}\frac{\partial^{2}V(x)}{\partial x^{2}}\sigma\right)\\ \leq 2h(x)|V(x)|^{\frac{1}{2}}\leq 0.

Note that h(x)<0h(x)<0 when x𝒞x\notin\mathcal{C}. By Theorem 1, we have that (T<)=1\mathbb{P}(T<\infty)=1, where T=inf{t:h(x(t;x0))=0}T=\inf\{t:h(x(t;x_{0}))=0\} is the stochastic settling time. Hence the system reaches 𝒞\mathcal{C} almost surely when x0𝒞x_{0}\notin\mathcal{C}. When x0𝒞x_{0}\in\mathcal{C}, the lemma holds since T=0T=0.

When Eqn. (11) holds, we have that xt𝒞x_{t}\in\mathcal{C} for tTt\geq T and

hx(x)(f(x)+g(x)u)+12tr(σ2h(x)x2σ)h(x).\frac{\partial h}{\partial x}(x)(f(x)+g(x)u)+\frac{1}{2}tr\left(\sigma^{\top}\frac{\partial^{2}h(x)}{\partial x^{2}}\sigma\right)\geq-h(x).

By [35, Thm. 3], we have that (xt𝒞,tT)=1\mathbb{P}(x_{t}\in\mathcal{C},~{}\forall t\geq T)=1. ∎

In the following, we extend Definition 4 and Lemma 2 to the incomplete information setting where the system state is estimated via an EKF. Define

h¯ε=sup{h(x):xx2ε for some xh1({0})}\bar{h}_{\varepsilon}=\sup\{h(x):\|x-x^{\prime}\|_{2}\leq\varepsilon\text{ for some }x^{\prime}\in h^{-1}(\{0\})\}

as the supremum of h(x)h(x) for all xx belonging to the ε\varepsilon-neighborhood of the boundary of 𝒞\mathcal{C}. We then present the following preliminary result [35].

Lemma 3 ([35]).

If xtx^t2ε\|x_{t}-\hat{x}_{t}\|_{2}\leq\varepsilon for all tt and h(x^t)>h¯εh(\hat{x}_{t})>\bar{h}_{\varepsilon} for all tt, then xt𝒞x_{t}\in\mathcal{C} for all tt.

We define h^(x)=h(x)h¯ε\hat{h}(x)=h(x)-\bar{h}_{\varepsilon}. Provided Lemma 3, we are now ready to develop stochastic finite time convergence CBF under incomplete information setting for the system described by (1) and (2) employing EKF for state estimation, assuming that at=𝟎a_{t}=\mathbf{0} for all t0t\geq 0.

Theorem 2.

Let T=inf{t:x(t;x0)𝒞}T=\inf\{t:x(t;x_{0})\in\mathcal{C}\}. Suppose there exists a twice continuously differentiable function h:nh:\mathbb{R}^{n}\rightarrow\mathbb{R}, control input uu, and class 𝒦\mathcal{K} functions α1\alpha_{1}, α2\alpha_{2} such that

α1(h^(x))h(x)α2(h^(x))\displaystyle\alpha_{1}(\hat{h}(x))\leq h(x)\leq\alpha_{2}(\hat{h}(x)) (12a)
hx(x^)(f(x^)+g(x^)u)εhxKc2\displaystyle\frac{\partial h}{\partial x}(\hat{x})(f(\hat{x})+g(\hat{x})u)-\varepsilon\|\frac{\partial h}{\partial x}Kc\|_{2}
+12tr(σtK2hx2Kσ)+sgn(h^(x^))|h^(x^)|0\displaystyle+\frac{1}{2}tr\left(\sigma_{t}^{\top}K^{\top}\frac{\partial^{2}h}{\partial x^{2}}K\sigma\right)+sgn(\hat{h}(\hat{x}))|\hat{h}(\hat{x})|\geq 0 (12b)

then (T<|xtx^t2ε,t)=1\mathbb{P}(T<\infty|\|x_{t}-\hat{x}_{t}\|_{2}\leq\varepsilon,~{}\forall t)=1, i.e., the system reaches 𝒞\mathcal{C} within finite time almost surely.

Proof.

Since we use EKF for state estimation, we have that the state estimate and covariance follow Eqn. (5) and (6). Given the output dynamics of yty_{t}, we have that

dx^t\displaystyle d\hat{x}_{t} =(f(x^t)+g(x^t)ut)dt+Kt(cxtdt+νtdVtcx^tdt)\displaystyle=(f(\hat{x}_{t})+g(\hat{x}_{t})u_{t})d_{t}+K_{t}(cx_{t}dt+\nu_{t}dV_{t}-c\hat{x}_{t}dt)
=(f(x^t)+g(x^t)ut+Ktc(xtx^t))dt+KtνtdVt.\displaystyle=(f(\hat{x}_{t})+g(\hat{x}_{t})u_{t}+K_{t}c(x_{t}-\hat{x}_{t}))dt+K_{t}\nu_{t}dV_{t}.

Let h^(x)=h(x)h¯ε\hat{h}(x)=h(x)-\bar{h}_{\varepsilon}. We have that

dh^t=(hx(x^t)(f(x^t)+g(x^t)ut+Ktc(xtx^t))+12tr(νtKt2hx(x^t)Ktνt))dthxKtνtdVt.d\hat{h}_{t}=\bigg{(}\frac{\partial h}{\partial x}(\hat{x}_{t})\left(f(\hat{x}_{t})+g(\hat{x}_{t})u_{t}+K_{t}c(x_{t}-\hat{x}_{t})\right)\\ +\frac{1}{2}tr\left(\nu_{t}^{\top}K_{t}^{\top}\frac{\partial^{2}h}{\partial x}(\hat{x}_{t})K_{t}\nu_{t}\right)\bigg{)}dt-\frac{\partial h}{\partial x}K_{t}\nu_{t}dV_{t}. (13)

When xtx^t2ε\|x_{t}-\hat{x}_{t}\|_{2}\leq\varepsilon holds, then

hxKtc(xtx^t)hxKtc2xtx^t2εhxKtc2.\frac{\partial h}{\partial x}K_{t}c(x_{t}-\hat{x}_{t})\geq-\left\|\frac{\partial h}{\partial x}K_{t}c\right\|_{2}\|x_{t}-\hat{x}_{t}\|_{2}\geq-\varepsilon\left\|\frac{\partial h}{\partial x}K_{t}c\right\|_{2}.

When Eqn. (12b) holds, we thus have that

hx(x^t)(f(x^t)+g(x^t)ut+Ktc(xtx^t))\displaystyle\frac{\partial h}{\partial x}(\hat{x}_{t})\left(f(\hat{x}_{t})+g(\hat{x}_{t})u_{t}+K_{t}c(x_{t}-\hat{x}_{t})\right)
+12tr(νtKt2hx2(x^t)Ktνt)\displaystyle\quad\quad\quad+\frac{1}{2}tr\left(\nu_{t}^{\top}K_{t}^{\top}\frac{\partial^{2}h}{\partial x^{2}}(\hat{x}_{t})K_{t}\nu_{t}\right)
\displaystyle\geq hx(x^t)(f(x^t)+g(x^t)ut)εhxKtc2\displaystyle\frac{\partial h}{\partial x}(\hat{x}_{t})\left(f(\hat{x}_{t})+g(\hat{x}_{t})u_{t}\right)-\varepsilon\left\|\frac{\partial h}{\partial x}K_{t}c\right\|_{2}
+12tr(νtKt2hx2(x^t)Ktνt)\displaystyle\quad\quad\quad+\frac{1}{2}tr\left(\nu_{t}^{\top}K_{t}^{\top}\frac{\partial^{2}h}{\partial x^{2}}(\hat{x}_{t})K_{t}\nu_{t}\right)
\displaystyle\geq sgn(h^(x^t))|h^(x^t)|.\displaystyle-sgn(\hat{h}(\hat{x}_{t}))|\hat{h}(\hat{x}_{t})|.

Finally, according to Lemma 2, we have that time T=inf{t:x(t;x0)𝒞}T=\inf\{t:x(t;x_{0})\in\mathcal{C}\} is finite with probability (T<|xtx^t2ε,t)=1\mathbb{P}(T<\infty|\|x_{t}-\hat{x}_{t}\|_{2}\leq\varepsilon,~{}\forall t)=1. ∎

We are now ready to present fault-tolerant stochastic finite time convergence CBF for the system given by (1)-(2) as follows. The system utilizes mm EKFs, with each associated with one attack pattern rir_{i}\in\mathcal{F}, where i=1,,mi=1,\ldots,m.

Lemma 4.

Let h¯εi=sup{h(x):xx2εi for some xh1({0})}\bar{h}_{\varepsilon_{i}}=\sup\{h(x):\|x-x^{\prime}\|_{2}\leq\varepsilon_{i}\text{ for some }x^{\prime}\in h^{-1}(\{0\})\} and h^i(x)=h(x)h¯εi\hat{h}_{i}(x)=h(x)-\bar{h}_{\varepsilon_{i}}. Let T=inf{t:x(t;x0)𝒞}T=\inf\{t:x(t;x_{0})\in\mathcal{C}\}. Suppose ε1,,εm\varepsilon_{1},\ldots,\varepsilon_{m} and θij\theta_{ij} are chosen such that there exists ϱ>0\varrho>0, making any XtXt(ϱ)X_{t}^{\prime}\subseteq X_{t}(\varrho) satisfying x^t,ix^t,jθij\|\hat{x}_{t,i}-\hat{x}_{t,j}\|\leq\theta_{ij} for all i,jXti,j\in X_{t}^{\prime} there exists uu satisfying

Λi(x^t,i)u>0,iXt\Lambda_{i}(\hat{x}_{t,i})u>0,~{}\forall i\in X_{t}^{\prime} (14)

where Λi(x^t,i)=hix(x^t,i)g(x^t,i)\Lambda_{i}(\hat{x}_{t,i})=\frac{\partial h_{i}}{\partial x}(\hat{x}_{t,i})g(\hat{x}_{t,i}), Xt(ϱ)={i:h^i(x^t,i)<ϱ}X_{t}(\varrho)=\{i:\hat{h}_{i}(\hat{x}_{t,i})<\varrho\}, then

(T<|x^t,ix^t,i,j2θij/2,j,x^t,ixt2εi,t,ri)=1.\mathbb{P}(T<\infty|\|\hat{x}_{t,i}-\hat{x}_{t,i,j}\|_{2}\leq\theta_{ij}/2,~{}\forall j,\\ \|\hat{x}_{t,i}-x_{t}\|_{2}\leq\varepsilon_{i},~{}\forall t,\forall r_{i}\in\mathcal{F})=1.

The proof of Lemma 4 is omitted due to space constraint. Functions h1,,hmh_{1},\ldots,h_{m} satisfying conditions 1) and 2) in Lemma 4 are fault tolerant stochastic finite time convergence CBFs.

IV-B Decomposition of GDTL Formula φ\varphi and Satisfaction of the Decomposed Sub-formulae

Given the GDTL formula, we construct the DRA representing it, as defined in Definition 2. We then pick an accepting run of the automaton 𝒜\mathcal{A}, denoted as η\eta. We rewrite η\eta into the prefix suffix form η=q1,,qn,(ηsuff)ω\eta=q_{1},\ldots,q_{n},(\eta_{suff})^{\omega}. We next decompose the accepting run η\eta into a sequence of sub-formulae, denoted as ψ1,,ψN\psi_{1},\ldots,\psi_{N}. Here

ψj=Φj𝒰(ϕjΦj+1),\psi_{j}=\Phi_{j}\mathcal{U}\Box(\phi_{j}\land\Phi_{j+1}), (15)

where Φj\Phi_{j} is the input word for self-loop transition δ(qj,Φj)=qj\delta(q_{j},\Phi_{j})=q_{j}, and ϕj\phi_{j} is the input word corresponding to the transition from state qjq_{j} to qj+1q_{j+1}. That is, sub-formula ψj\psi_{j} models a one-step transition from state qjq_{j} to qj+1q_{j+1} along accepting run η\eta.

We next synthesize the control policy to satisfy each sub-formula leveraging fault-tolerant finite time convergence control barrier functions in Section IV-A and fault-tolerant zeroing control barrier functions proposed in [36]. We then derive the satisfaction probability of each sub-formula, which is used to bound the probability of satisfying φ\varphi.

We let functions hjh^{j} and djd^{j} be given as

{b:hj(b)0}={b:(L(b)ϕjΦj+1)(L(b)Φj)}\displaystyle\{b:h^{j}(b)\geq 0\}=\{b:(L(b)\models\phi_{j}\land\Phi_{j+1})\lor(L(b)\models\Phi_{j})\}
{b:dj(b)0}={b:L(b)Φj+1}.\displaystyle\{b:d^{j}(b)\geq 0\}=\{b:L(b)\models\Phi_{j+1}\}.

Then sub-formula ψj\psi_{j} indicates that the system trajectory needs to guarantee hj(b)0h^{j}(b)\geq 0 for all [tj,tj+1][t_{j},t_{j+1}] and dj(b)0d^{j}(b)\geq 0 for some t[tj,tj+1]t^{\prime}\in[t_{j},t_{j+1}], where tjt_{j} and tj+1t_{j+1} are some time instants satisfying tj+1tj0t_{j+1}\geq t_{j}\geq 0. We now define

h¯εj=sup{hj(b):bb2ε\displaystyle\bar{h}^{j}_{\varepsilon}=\sup\{h^{j}(b):\|b-b^{\prime}\|_{2}\leq\varepsilon
 for some b(hj)1({0})}\displaystyle\quad\quad\quad\text{ for some }b^{\prime}\in(h^{j})^{-1}(\{0\})\} (16)
d¯εj=sup{dj(b):bb2ε\displaystyle\bar{d}^{j}_{\varepsilon}=\sup\{d^{j}(b):\|b-b^{\prime}\|_{2}\leq\varepsilon
 for some b(dj)1({0})}\displaystyle\quad\quad\quad\text{ for some }b^{\prime}\in(d^{j})^{-1}(\{0\})\} (17)
h^j(b)=hj(b)h¯εj,d^j(b)=dj(b)d¯εj.\displaystyle\hat{h}^{j}(b)=h^{j}(b)-\bar{h}^{j}_{\varepsilon},\quad\quad\hat{d}^{j}(b)=d^{j}(b)-\bar{d}^{j}_{\varepsilon}. (18)

We next present an algorithm for synthesizing the control policy so that the probability of satisfying sub-formula ψj\psi_{j} is lower bounded in Algorithm 1. In line 3, we compute Xt(ϱ)={i:h^ij(bt,i)<ϱ1,d^ij(bt,i)<ϱ2}X_{t}(\varrho)=\{i:\hat{h}_{i}^{j}(b_{t,i})<\varrho_{1},\hat{d}_{i}^{j}(b_{t,i})<\varrho_{2}\}, where ϱ=[ϱ1,ϱ2]\varrho=[\varrho_{1},\varrho_{2}]^{\top} and ϱ1,ϱ2>0\varrho_{1},\varrho_{2}>0 are constants. In line 4, we first compute the set of control inputs Ωi\Omega_{i} and Γi\Gamma_{i}, where Ωi\Omega_{i} ensures that hj(bt,i)0h^{j}(b_{t,i})\geq 0 for all t[tj,tj+1]t\in[t_{j},t_{j+1}], and Γi\Gamma_{i} ensures that there exists t[tj,tj+1]t^{\prime}\in[t_{j},t_{j+1}] such that dj(x)0d^{j}(x)\geq 0, provided that the fault pattern is rir_{i}. If iXt(ϱ)(ΩiΓi)\cap_{i\in X_{t}(\varrho)}(\Omega_{i}\cap\Gamma_{i})\neq\emptyset, then any control input uiXt(ϱ)(ΩiΓi)u\in\cap_{i\in X_{t}(\varrho)}(\Omega_{i}\cap\Gamma_{i}) guarantees the satisfaction of sub-formula ψj\psi_{j} regardless of the fault pattern. If such uu does not exist, then line 5-11 aims to identify the fault pattern by comparing the state estimate when fault pattern rir_{i} is removed. Line 17 computes the control input by solving a quadratic program. If no feasible control input can be found, we will repeat the procedure for another accepting run ηη\eta^{\prime}\neq\eta.

Theorem 3.

Let bt,i=(x^t,i,Pt,i)b_{t,i}=(\hat{x}_{t,i},P_{t,i}). We define

Λij(bt,i,u)=hijx(bt,i)g(x^t,i)u\displaystyle\Lambda_{i}^{j}(b_{t,i},u)=\frac{\partial h_{i}^{j}}{\partial x}(b_{t,i})g(\hat{x}_{t,i})u
+tr[hijP(bt,i)(gx(x^t,i)uPt,i+Pt,i(gx(x^t,i)u))],\displaystyle~{}+tr\left[\frac{\partial h_{i}^{j}}{\partial P}(b_{t,i})\left(\frac{\partial g}{\partial x}(\hat{x}_{t,i})uP_{t,i}+P_{t,i}(\frac{\partial g}{\partial x}(\hat{x}_{t,i})u)^{\top}\right)\right],
Ξij(bt,i,u)=dijx(bt,i)g(x^t,i)u\displaystyle\Xi_{i}^{j}(b_{t,i},u)=\frac{\partial d_{i}^{j}}{\partial x}(b_{t,i})g(\hat{x}_{t,i})u
+tr[dijP(bt,i)(gx(x^t,i)uPt,i+Pt,i(gx(x^t,i)u))].\displaystyle~{}+tr\left[\frac{\partial d_{i}^{j}}{\partial P}(b_{t,i})\left(\frac{\partial g}{\partial x}(\hat{x}_{t,i})uP_{t,i}+P_{t,i}(\frac{\partial g}{\partial x}(\hat{x}_{t,i})u)^{\top}\right)\right].

Suppose ε1,,εm\varepsilon_{1},\ldots,\varepsilon_{m} and θij\theta_{ij} are chosen such that there exists ϱ\varrho making bt,ibt,kθik\|b_{t,i}-b_{t,k}\|\leq\theta_{ik} holds for all i,kXti,k\in X_{t}^{\prime} and for any XtXt(ϱ)X_{t}^{\prime}\subseteq X_{t}(\varrho). If there exists some uu satisfying

Λi(bt,i,u)>0,Ξi(bt,i,u)>0,iXt.\Lambda_{i}(b_{t,i},u)>0,\quad\Xi_{i}(b_{t,i},u)>0,~{}\forall i\in X_{t}^{\prime}. (19)

then

(ψj|bt,ibt,i,k2θik/2,j,bt,ibt2εi,t,ri)=1.\mathbb{P}(\psi_{j}|\|b_{t,i}-b_{t,i,k}\|_{2}\leq\theta_{ik}/2,~{}\forall j,\\ \|b_{t,i}-b_{t}\|_{2}\leq\varepsilon_{i},~{}\forall t,\forall r_{i}\in\mathcal{F})=1.
Algorithm 1 Control synthesis for sub-formula ψj\psi_{j}.
1:Input: Sub-formula ψj\psi_{j}, system dynamics (1)-(2), parameters ε1,,εm\varepsilon_{1},\ldots,\varepsilon_{m}, θij\theta_{ij} where i<ji<j, and ϱ1,ϱ2>0\varrho_{1},\varrho_{2}>0.
2:Output: control input utu_{t}
3:Compute Xt(ϱ)={i:h^ij(bt,i)<ϱ1,d^ij(bt,i)<ϱ2}X_{t}(\varrho)=\{i:\hat{h}_{i}^{j}(b_{t,i})<\varrho_{1},\hat{d}_{i}^{j}(b_{t,i})<\varrho_{2}\}, where ϱ=[ϱ1,ϱ2]\varrho=[\varrho_{1},\varrho_{2}]^{\top}
4:For each iXt(ϱ)i\in X_{t}(\varrho), compute
Ωi={u:hijx(f(x^t,i)+g(x^t,i)u)εhijxKt,ic2\displaystyle\Omega_{i}=\Big{\{}u:\frac{\partial h_{i}^{j}}{\partial x}(f(\hat{x}_{t,i})+g(\hat{x}_{t,i})u)-\varepsilon\|\frac{\partial h_{i}^{j}}{\partial x}K_{t,i}c\|_{2}
+12tr(νt,iKt,i2hijx2Kt,iνt,i)h^(bt,i)}\displaystyle+\frac{1}{2}tr\left(\nu_{t,i}^{\top}K_{t,i}^{\top}\frac{\partial^{2}h_{i}^{j}}{\partial x^{2}}K_{t,i}\nu_{t,i}\right)\geq-\hat{h}(b_{t,i})\Big{\}} (20)
Γi={u:dijx(f(x^t,i)+g(x^t,i)u)εdijxKt,ic2\displaystyle\Gamma_{i}=\Big{\{}u:\frac{\partial d_{i}^{j}}{\partial x}(f(\hat{x}_{t,i})+g(\hat{x}_{t,i})u)-\varepsilon\|\frac{\partial d_{i}^{j}}{\partial x}K_{t,i}c\|_{2}
+12tr(νt,iKt,i2dijx2Kt,iνt,i)\displaystyle+\frac{1}{2}tr\left(\nu_{t,i}^{\top}K_{t,i}^{\top}\frac{\partial^{2}d_{i}^{j}}{\partial x^{2}}K_{t,i}\nu_{t,i}\right)
sgn(d^ij(bt,i))|d^(bt,i)|}\displaystyle\quad\quad\geq-sgn(\hat{d}_{i}^{j}(b_{t,i}))|\hat{d}(b_{t,i})|\Big{\}} (21)
5:while iXt(ϱ)(ΩiΓi)=\cap_{i\in X_{t}(\varrho)}(\Omega_{i}\cap\Gamma_{i})=\emptyset and Xt(ϱ)X_{t}(\varrho)\neq\emptyset do
6:     for i,k{1,,m}i,k\in\{1,\ldots,m\} do
7:         if x^t,ix^t,k2>θik\|\hat{x}_{t,i}-\hat{x}_{t,k}\|_{2}>\theta_{ik} and x^t,ix^t,i,k2>θik/2\|\hat{x}_{t,i}-\hat{x}_{t,i,k}\|_{2}>\theta_{ik}/2 then
8:              Update Xt(ϱ)Xt(ϱ){i}X_{t}(\varrho)\leftarrow X_{t}(\varrho)\setminus\{i\}
9:         end if
10:     end for
11:end while
12:while iXt(ϱ)(ΩiΓi)=\cap_{i\in X_{t}(\varrho)}(\Omega_{i}\cap\Gamma_{i})=\emptyset and Xt(ϱ)X_{t}(\varrho)\neq\emptyset do
13:     kargmax{yt,kckx^t,k:kXt(ϱ)}k^{*}\leftarrow\arg\!\max\{y_{t,k}-c_{k}\hat{x}_{t,k}:k\in X_{t}(\varrho)\}
14:     Update Xt(ϱ)Xt(ϱ){k}X_{t}(\varrho)\leftarrow X_{t}(\varrho)\setminus\{k^{*}\}
15:end while
16:if iXt(ϱ)(ΩiΓi)\cap_{i\in X_{t}(\varrho)}(\Omega_{i}\cap\Gamma_{i})\neq\emptyset then
17:     Solve utu_{t} as ut=argminutiXt(ϱ)(ΩiΓi)utMutu_{t}=\underset{u_{t}\in\cap_{i\in X_{t}(\varrho)}(\Omega_{i}\cap\Gamma_{i})}{\arg\min}{u_{t}}^{\top}Mu_{t}, where MM is positive definite
18:end if
19:Return utu_{t}
Proof.

We show that if bt,ibt2εi\|b_{t,i}-b_{t}\|_{2}\leq\varepsilon_{i} and bt,ibt,i,k2θik/2\|b_{t,i}-b_{t,i,k}\|_{2}\leq\theta_{ik}/2 for all tt, then utΩiΓiu_{t}\in\Omega_{i}\cap\Gamma_{i} holds when h^ij(bt,i)<ϱ1\hat{h}_{i}^{j}(b_{t,i})<\varrho_{1} and d^ij(bt,i)<ϱ2\hat{d}_{i}^{j}(b_{t,i})<\varrho_{2}.

Suppose that the fault pattern r=rir=r_{i}. At time tt, suppose that h^ij(bt,i)<ϱ1\hat{h}_{i}^{j}(b_{t,i})<\varrho_{1} and d^ij(bt,i)<ϱ2\hat{d}_{i}^{j}(b_{t,i})<\varrho_{2} hold. In addition, suppose that bt,ibt,i,k2θik/2\|b_{t,i}-b_{t,i,k}\|_{2}\leq\theta_{ik}/2. We divide our discussion into three cases.

Case I. We first consider bt,ibt,k2θik\|b_{t,i}-b_{t,k}\|_{2}\leq\theta_{ik} for all i,kXt(ϱ)i,k\in X_{t}(\varrho). Note that Ωi\Omega_{i} and Γi\Gamma_{i} can be represented as

Ωi={u:Λi(bt,i,u)ωi},Γi={u:Ξi(bt,i,u)ιi},\Omega_{i}=\{u:\Lambda_{i}(b_{t,i},u)\geq\omega_{i}\},\quad\Gamma_{i}=\{u:\Xi_{i}(b_{t,i},u)\geq\iota_{i}\},

where ωi,ιi\omega_{i},\iota_{i}\in\mathbb{R} are some constants. When bt,ibt,k2θik\|b_{t,i}-b_{t,k}\|_{2}\leq\theta_{ik} for all i,kXt(ϱ)i,k\in X_{t}(\varrho) holds, by condition 1), we then have that there exists utu_{t} satisfying

Λi(bt,i,ut)>0,Ξi(bt,i,ut)>0,iXt.\Lambda_{i}(b_{t,i},u_{t})>0,\quad\Xi_{i}(b_{t,i},u_{t})>0,~{}\forall i\in X_{t}^{\prime}.

indicating that utiXt(ϱ)(ΩiΓi)u_{t}\in\cap_{i\in X_{t}(\varrho)}(\Omega_{i}\cap\Gamma_{i}).

Case II. We next consider that bt,ibt,k2θik\|b_{t,i}-b_{t,k}\|_{2}\leq\theta_{ik} for all kXt(ϱ)k\in X_{t}(\varrho), but there exists some z,kXt(ϱ){i}z,k\in X_{t}(\varrho)\setminus\{i\} such that bt,kbt,z>θzk\|b_{t,k}-b_{t,z}\|>\theta_{zk}. In this case, by the iteration between line 5-11 in Algorithm 1, ΩkΓk\Omega_{k}\cap\Gamma_{k} is removed. By the hypothesis that bt,ibt,i,k2θik\|b_{t,i}-b_{t,i,k}\|_{2}\leq\theta_{ik} for all kXt(ϱ)k\in X_{t}(\varrho), we have that the updated Xt(ϱ)X_{t}(\varrho) now reduces to Case I, and thus our previous analysis can be applied.

We finally suppose that bt,ibt,k2>θik\|b_{t,i}-b_{t,k}\|_{2}>\theta_{ik} for some kXt(ϱ)k\in X_{t}(\varrho). We then have that

θik\displaystyle\theta_{ik} bt,ibt,i,k+bt,i,kbt,k2\displaystyle\leq\|b_{t,i}-b_{t,i,k}+b_{t,i,k}-b_{t,k}\|_{2}
bt,ibt,i,k2+bt,i,kbt,k2\displaystyle\leq\|b_{t,i}-b_{t,i,k}\|_{2}+\|b_{t,i,k}-b_{t,k}\|_{2}
θik2+bt,i,kbt,k2\displaystyle\leq\frac{\theta_{ik}}{2}+\|b_{t,i,k}-b_{t,k}\|_{2}

where the second inequality holds by triangle inequality, and the last inequality holds by the hypothesis that bt,ibt,i,k2θik/2\|b_{t,i}-b_{t,i,k}\|_{2}\leq\theta_{ik}/2. In this case, bt,i,kbt,k2θik/2\|b_{t,i,k}-b_{t,k}\|_{2}\geq\theta_{ik}/2, indicating that {k}\{k\} is removed by line 8 of Algorithm 1. Then our previous analysis becomes applicable once all such kk are removed. Lastly, using Theorem 2 and [36], we have that the control input uu satisfies sub-formula ψj\psi_{j} almost surely. ∎

We conclude this section by quantifying the probability of satisfying GDTL formula φ\varphi as follows.

Theorem 4.

If the conditions in Theorem 3 are satisfied, then the probability of satisfying GDTL specification φ\varphi satisfies

(φ|bt,ibt,i,k2θik/2,j,bt,ibt2εi,t,ri)=1.\mathbb{P}(\varphi|\|b_{t,i}-b_{t,i,k}\|_{2}\leq\theta_{ik}/2,~{}\forall j,\\ \|b_{t,i}-b_{t}\|_{2}\leq\varepsilon_{i},~{}\forall t,\forall r_{i}\in\mathcal{F})=1. (23)
Proof.

We denote the probability that ψj,,ψj\psi_{j},\ldots,\psi_{j^{\prime}} occur in order as (ψj,,ψj)\mathbb{P}(\psi_{j},\ldots,\psi_{j^{\prime}}), where jjj\leq j^{\prime}. We also denote the event that bt,ibt,i,k2θik/2\|b_{t,i}-b_{t,i,k}\|_{2}\leq\theta_{ik}/2 holds for all jj and bt,ibt2εi\|b_{t,i}-b_{t}\|_{2}\leq\varepsilon_{i} holds for all tt under fault rir_{i} as Υi\Upsilon_{i}. By Bayes’ theorem, we have that

(φ|Υi,ri)(ψ1ψN|Υi,ri)\displaystyle\mathbb{P}(\varphi|\Upsilon_{i},\forall r_{i}\in\mathcal{F})\geq\mathbb{P}(\psi_{1}\ldots\psi_{N}|\Upsilon_{i},\forall r_{i}\in\mathcal{F})
=\displaystyle= (ψN|ψ1ψN1,Υi,ri)\displaystyle\mathbb{P}(\psi_{N}|\psi_{1}\ldots\psi_{N-1},\Upsilon_{i},~{}\forall r_{i}\in\mathcal{F})
(ψN1|ψ1ψN2,Υi,ri)\displaystyle\quad\cdot\mathbb{P}(\psi_{N-1}|\psi_{1}\ldots\psi_{N-2},\Upsilon_{i},~{}\forall r_{i}\in\mathcal{F})\ldots
(ψ1|Υi,ri)(Υi,ri).\displaystyle\quad\cdot\mathbb{P}(\psi_{1}|\Upsilon_{i},~{}\forall r_{i}\in\mathcal{F})\mathbb{P}(\Upsilon_{i},~{}\forall r_{i}\in\mathcal{F}).

By our hypothesis, we have that (Υi,ri)=1\mathbb{P}(\Upsilon_{i},~{}\forall r_{i}\in\mathcal{F})=1. Using Theorem 3, we have that (ψj|ψ1ψj1,Υi,ri)=1\mathbb{P}(\psi_{j}|\psi_{1}\ldots\psi_{j-1},\Upsilon_{i},~{}\forall r_{i}\in\mathcal{F})=1 holds for all j=1,,Nj=1,\ldots,N. Therefore, we have that Eqn. (23) holds, completing the proof. ∎

Based on Theorem 4, we have that if the EKFs guarantee that (Υi)1ϵ\mathbb{P}(\Upsilon_{i})\geq 1-\epsilon for all rir_{i}\in\mathcal{F}, then the probability of satisfying GDTL specification φ\varphi can be bounded as (φ)1ϵ\mathbb{P}(\varphi)\geq 1-\epsilon, as desired by Problem 1. The accuracy of EKFs can be obtained via Lemma 1.

V Case Study

In this section, we present a case study on the coordination of two wheeled mobile robots (WMRs) in a 2-D domain. Each WMR follows dynamics:

[[x˙t]1[x˙t]2z˙t]=[coszt0sinzt001][vtωt]+wt\begin{bmatrix}[\dot{x}_{t}]_{1}\\ [\dot{x}_{t}]_{2}\\ \dot{z}_{t}\end{bmatrix}=\begin{bmatrix}\cos z_{t}&0\\ \sin z_{t}&0\\ 0&1\end{bmatrix}\begin{bmatrix}v_{t}\\ \omega_{t}\end{bmatrix}+w_{t} (24)

where [xt]1[x_{t}]_{1} and [xt]2[x_{t}]_{2} model the position of the robot at time tt, ztz_{t} is the orientation of the robot at time tt, ut=[vt,ωt]u_{t}=[v_{t},\omega_{t}]^{\top} is the control input consisting of the linear and angular velocities, and ww is a zero-mean Gaussian process noise. Following [37], we apply feedback linearization to Eqn. (24) and use 77 sensors yt7y_{t}\in\mathbb{R}^{7} to measure the location and velocities for each WMR. In particular, we let [yt]1[y_{t}]_{1} and [yt]2[y_{t}]_{2} measure horizontal position [xt]1[x_{t}]_{1}, and use [yt]3[y_{t}]_{3} and [yt]4[y_{t}]_{4} to measure vertical position [xt]2[x_{t}]_{2}. Outputs [yt]5[y_{t}]_{5} and [yt]6[y_{t}]_{6} measure the velocities of the WMR. The orientation of the WMR is measured by [yt]7[y_{t}]_{7}.

The WMRs need to coordinate to satisfy a GDTL specification φ\varphi given as φ=i=13φ~i\varphi=\land_{i=1}^{3}\tilde{\varphi}_{i}, where φ~1=(dest1adest1b)dest2\tilde{\varphi}_{1}=\Diamond(dest1a\land\Diamond dest1b)\land\Diamond dest2, φ~2=(¬Obs)\tilde{\varphi}_{2}=\Box(\neg Obs), and φ~3=(tr(P)0.9)\tilde{\varphi}_{3}=\Box(tr(P)\leq 0.9). Specification φ\varphi requires the WMRs to reach a set of destinations, denoted as dest1a,dest1bdest1a,dest1b, and dest2dest2, in order, and to avoid the unsafe region obsobs. In addition, the WMRs must maintain an uncertainty that is less than 0.90.9. There exists a malicious adversary that manipulates the outputs of [yt]2[y_{t}]_{2} and [yt]4[y_{t}]_{4} to bias the WMRs’ location estimates.

Refer to caption
Figure 1: The trajectories of two WMRs generated using our proposed approach and a baseline are plotted in black and blue colors, respectively. The baseline applies a CBF-based control policy with state estimate being calculated using all sensors. Our proposed approach satisfies the given GDTL specification φ\varphi, whereas the baseline violates it.

We compare our proposed approach with a baseline approach [15]. The baseline approach adopts the state-of-the-art abstraction-free CBF-based control synthesis using the state estimate obtained by all sensor measurements.

In the following, we demonstrate our proposed approach. We generate the DRA representing GDTL specification φ\varphi. We then pick an accepting run η\eta and decompose φ\varphi into a sequence of sub-formulae using η\eta. In this case study, accepting run η\eta requires the WMRs to reach destinations dest1adest1a, dest1bdest1b, and dest2dest2 in this order, while guaranteeing φ~2\tilde{\varphi}_{2} and φ~3\tilde{\varphi}_{3} to be always satisfied.

We present the trajectories for both robots using our proposed approach and the baseline in Fig. 1 using lines in black and blue colors, respectively. The trajectories of WMR 11 and 22 are presented using solid and dashed lines, respectively. We observe that the trajectories of the robots using our proposed approach never enter the unsafe region obsobs and reach their assigned destinations. In addition, our proposed approach yields tr(P)=0.030.9tr(P)=0.03\leq 0.9 and thus satisfies specification φ3\varphi_{3}. However, the trajectories generated by the baseline approach enter the the unsafe region obsobs and thus violate specification φ2\varphi_{2}. Moreover, WMR 11 is biased by the sensor attack and fails to reach its destination dest1adest1a using the baseline approach. To summarize, our proposed approach guarantees that WMR 11 and 22 can coordinate to satisfy GDTL formula φ\varphi in the presence of sensor attacks.

VI Conclusion

In this paper, we studied the problem of abstraction-free synthesis for a control-affine nonlinear system to satisfy a Gaussian distribution temporal logic specification in the presence of process and observation noises as well as sensor faults and attacks. We proposed a solution approach that decomposes the given specification into a sequence of reach-avoid tasks that needs to be satisfied by the system. We developed the sufficient conditions for the controller so that each decomposed sub-formula can be satisfied by the system in the presence of sensor faults and attacks using our proposed fault-tolerant finite time convergence control barrier functions, along with the fault-tolerant zeroing control barrier functions. We proved that our synthesized controller guarantees almost sure satisfaction for each decomposed sub-formula, and thus ensures that the system satisfies the Gaussian distribution temporal logic specification with probability one, provided that the extended Kalman filters used by the system ensures certain accuracy bounds. We presented a numerical case study on the coordination of two wheeled mobile robots to demonstrate our proposed approach.

References

  • [1] C. Baier, J.-P. Katoen, and K. G. Larsen, Principles of Model Checking.   MIT Press, 2008.
  • [2] C.-I. Vasile, K. Leahy, E. Cristofalo, A. Jones, M. Schwager, and C. Belta, “Control in belief space with temporal logic specifications,” in 55th IEEE Conference on Decision and Control (CDC).   IEEE, 2016, pp. 7419–7424.
  • [3] K. Leahy, E. Cristofalo, C.-I. Vasile, A. Jones, E. Montijano, M. Schwager, and C. Belta, “Control in belief space with temporal logic specifications using vision-based localization,” The International Journal of Robotics Research, vol. 38, no. 6, pp. 702–722, 2019.
  • [4] H. Kress-Gazit, G. E. Fainekos, and G. J. Pappas, “Temporal-logic-based reactive mission and motion planning,” IEEE Transactions on Robotics, vol. 25, no. 6, pp. 1370–1381, 2009.
  • [5] S. Coogan, E. A. Gol, M. Arcak, and C. Belta, “Traffic network control from temporal logic specifications,” IEEE Transactions on Control of Network Systems, vol. 3, no. 2, pp. 162–172, 2015.
  • [6] G. E. Fainekos, A. Girard, H. Kress-Gazit, and G. J. Pappas, “Temporal logic motion planning for dynamic robots,” Automatica, vol. 45, no. 2, pp. 343–352, 2009.
  • [7] X. Ding, S. L. Smith, C. Belta, and D. Rus, “Optimal control of Markov decision processes with linear temporal logic constraints,” IEEE Transactions on Automatic Control, vol. 59, no. 5, pp. 1244–1257, 2014.
  • [8] T. Wongpiromsarn, U. Topcu, and R. M. Murray, “Receding horizon temporal logic planning for dynamical systems,” in 48th IEEE Conference on Decision and Control (CDC).   IEEE, 2009, pp. 5997–6004.
  • [9] E. M. Wolff, U. Topcu, and R. M. Murray, “Robust control of uncertain Markov decision processes with temporal logic specifications,” in 51st Conference on Decision and Control.   IEEE, 2012, pp. 3372–3379.
  • [10] M. B. Horowitz, E. M. Wolff, and R. M. Murray, “A compositional approach to stochastic optimal control with co-safe temporal logic specifications,” in IEEE/RSJ International Conference on Intelligent Robots and Systems.   IEEE, 2014, pp. 1466–1473.
  • [11] I. Papusha, J. Fu, U. Topcu, and R. M. Murray, “Automata theory meets approximate dynamic programming: Optimal control with temporal logic constraints,” in 55th IEEE Conference on Decision and Control (CDC).   IEEE, 2016, pp. 434–440.
  • [12] M. Srinivasan, S. Coogan, and M. Egerstedt, “Control of multi-agent systems with finite time control barrier certificates and temporal logic,” in IEEE 57th Conference on Decision and Control (CDC).   IEEE, 2018, pp. 1991–1996.
  • [13] L. Lindemann and D. V. Dimarogonas, “Control barrier functions for signal temporal logic tasks,” IEEE control systems letters, vol. 3, no. 1, pp. 96–101, 2018.
  • [14] G. Yang, C. Belta, and R. Tron, “Continuous-time signal temporal logic planning with control barrier functions,” in 2020 American Control Conference (ACC), 2020, pp. 4612–4618.
  • [15] L. Niu and A. Clark, “Control barrier functions for abstraction-free control synthesis under temporal logic constraints,” in 59th IEEE Conference on Decision and Control.   IEEE, 2020, pp. 816–823.
  • [16] Y. Liu, P. Ning, and M. K. Reiter, “False data injection attacks against state estimation in electric power grids,” ACM Transactions on Information and System Security, vol. 14, no. 1, p. 13, 2011.
  • [17] Y. Mo and B. Sinopoli, “False data injection attacks in control systems,” in 1st workshop on Secure Control Systems, 2010, pp. 1–6.
  • [18] A. A. Cárdenas, S. Amin, and S. Sastry, “Research challenges for the security of control systems.” in Proceedings of the 3rd Conference on Hot Topics in Security, vol. 5.   USENIX Association, 2008, p. 15.
  • [19] H. Fawzi, P. Tabuada, and S. Diggavi, “Secure estimation and control for cyber-physical systems under adversarial attacks,” IEEE Transactions on Automatic Control, vol. 59, no. 6, pp. 1454–1467, 2014.
  • [20] Y. Shoukry, P. Nuzzo, A. Puggelli, A. L. Sangiovanni-Vincentelli, S. A. Seshia, and P. Tabuada, “Secure state estimation for cyber-physical systems under sensor attacks: A satisfiability modulo theory approach,” IEEE Transactions on Automatic Control, vol. 62, no. 10, pp. 4917–4932, 2017.
  • [21] A. Clark and L. Niu, “Linear quadratic Gaussian control under false data injection attacks,” in Annual American Control Conference (ACC).   IEEE, 2018, pp. 5737–5743.
  • [22] L. Niu and A. Clark, “Optimal secure control with linear temporal logic constraints,” IEEE Transactions on Automatic Control, vol. 65, no. 6, pp. 2434–2449, 2019.
  • [23] L. Niu, J. Fu, and A. Clark, “Optimal minimum violation control synthesis of cyber-physical systems under attacks,” IEEE Transactions on Automatic Control, vol. 66, no. 3, pp. 995–1008, 2020.
  • [24] B. Ramasubramanian, L. Niu, A. Clark, L. Bushnell, and R. Poovendran, “Secure control in partially observable environments to satisfy LTL specifications,” IEEE Transactions on Automatic Control, vol. 66, no. 12, pp. 5665–5679, 2020.
  • [25] E. M. Wolff, U. Topcu, and R. M. Murray, “Optimization-based trajectory generation with linear temporal logic specifications,” in International Conference on Robotics and Automation (ICRA).   IEEE, 2014, pp. 5319–5325.
  • [26] L. Wang, A. D. Ames, and M. Egerstedt, “Safety barrier certificates for collisions-free multirobot systems,” IEEE Transactions on Robotics, vol. 33, no. 3, pp. 661–674, 2017.
  • [27] A. D. Ames, X. Xu, J. W. Grizzle, and P. Tabuada, “Control barrier function based quadratic programs for safety critical systems,” IEEE Transactions on Automatic Control, vol. 62, no. 8, pp. 3861–3876, 2016.
  • [28] P. Jagtap, S. Soudjani, and M. Zamani, “Formal synthesis of stochastic systems via control barrier certificates,” IEEE Transactions on Automatic Control, vol. 66, no. 7, pp. 3097–3110, 2020.
  • [29] M. Anand, P. Jagtapt, and M. Zamani, “Verification of switched stochastic systems via barrier certificates,” in IEEE 58th Conference on Decision and Control (CDC).   IEEE, 2019, pp. 4373–4378.
  • [30] C. Sun and K. G. Vamvoudakis, “Continuous-time safe learning with temporal logic constraints in adversarial environments,” in American Control Conference (ACC).   IEEE, 2020, pp. 4786–4791.
  • [31] M. Fujisaki, G. Kallianpur, and H. Kunita, “Stochastic differential equations for the non linear filtering problem,” Osaka Journal of Mathematics, vol. 9, no. 1, pp. 19–40, 1972.
  • [32] J. Yin, S. Khoo, Z. Man, and X. Yu, “Finite-time stability and instability of stochastic nonlinear systems,” Automatica, vol. 47, no. 12, pp. 2671–2677, 2011.
  • [33] K. Reif, S. Gunther, E. Yaz, and R. Unbehauen, “Stochastic stability of the continuous-time extended Kalman filter,” IEE Proceedings-Control Theory and Applications, vol. 147, no. 1, pp. 45–52, 2000.
  • [34] T. Wongpiromsarn, U. Topcu, and A. Lamperski, “Automata theory meets barrier certificates: Temporal logic verification of nonlinear systems,” IEEE Transactions on Automatic Control, vol. 61, no. 11, pp. 3344–3355, 2015.
  • [35] A. Clark, “Control barrier functions for stochastic systems,” Automatica, vol. 130, p. 109688, 2021.
  • [36] A. Clark, Z. Li, and H. Zhang, “Control barrier functions for safe CPS under sensor faults and attacks,” in 59th IEEE Conference on Decision and Control (CDC).   IEEE, 2020, pp. 796–803.
  • [37] Z. Chen, L. Li, and X. Huang, “Building an autonomous lane keeping simulator using real-world data and end-to-end learning,” IEEE Intelligent Transportation Systems Magazine, 2018.