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

A Unified Stability Analysis of Safety-Critical Control using Multiple Control Barrier Functions

Matheus F. Reis, José P. Carvalho and A. Pedro Aguiar The authors are with the Research Center for Systems and Technologies (SYSTEC-ARISE), Faculdade de Engenharia, Universidade do Porto, Rua Dr. Roberto Frias, s/n 4200-465, Porto, Portugal {matheus.reis, jpcarvalho, apra}@fe.up.pt
Abstract

Ensuring liveness and safety of autonomous and cyber-physical systems remains a fundamental challenge, particularly when multiple safety constraints are present. This letter advances the theoretical foundations of safety-filter Quadratic Programs (QP) and Control Lyapunov Function (CLF)-Control Barrier Function (CBF) controllers by establishing a unified analytical framework for studying their stability properties. We derive sufficient feasibility conditions for QPs with multiple CBFs and formally characterize the conditions leading to undesirable equilibrium points at possible intersecting safe set boundaries. Additionally, we introduce a stability criterion for equilibrium points, providing a systematic approach to identifying conditions under which they can be destabilized or eliminated. Our analysis extends prior theoretical results, deepening the understanding of the conditions of feasibility and stability of CBF-based safety filters and the CLF-CBF QP framework.

I INTRODUCTION

Due to recent technological advancements, cyber-physical and autonomous systems, including self-driving vehicles and robotics, are becoming increasingly prevalent in our lives. While these systems exhibit higher autonomy levels, the increase in agency also brings more unpredictable and catastrophic failures [2], making it ever more crucial to design controllers that are both performant and provably safe.

The performance of a controller can be associated with its liveness, defined as asymptotic stability towards a set of goal states. On the other hand, safety can be defined as the requirement of forward invariance of the system trajectories with respect to a specified safe set. Combining and guaranteeing the fulfillment of these two requirements is still an open problem. Common approaches are the safety filter architecture [3] and the CLF-CBF minimum-norm controller [1]. In the former, the liveness and safety requirements are decoupled through a performance (nominal) controller that seeks to ensure liveness, while a safety controller modifies the control law to guarantee safety at all times [11], making use of Control Barrier Functions (CBFs) [6] to model the safety requirements. In the latter, safety and liveness are directly combined within an unified control framework [1], using Control Lyapunov Functions (CLFs) and CBFs.

However, it is shown in [8] that due to the conflicting objectives of safety and liveness, the CLF-CBF QP framework can introduce undesirable stable equilibrium points on the safe set boundary. This means that while the framework can guarantee safety, it does not guarantee liveness as the system’s trajectories can get stuck at undesirable equilibrium points, resulting in deadlocks. Research has aimed to mitigate or avoid these deadlocks in both the CLF-CBF framework and CBF-based safety filters.

In [7], a method is proposed to solve the deadlock problem for the safety-filter considering multiple non-convex unsafe regions modeled by CBFs. Undesirable stable equilibria are avoided by using a nonlinear transformation that maps the system state into a new domain called the “ball world”, where all non-convex CBF boundaries are converted into moving nn-spheres and dynamically avoid the state trajectory. In [8], the authors propose a solution based on a rotating CLF, so that equilibrium points are removed in the case of a single convex unsafe set. The idea of modifying the CLF to avoid deadlocks is latter extended in [9], where the theory of linear matrix polynomials was used to compute a CLF that does not cause deadlocks, considering the class of linear systems and quadratic CLFs and CBFs. Finally, in [5], the authors address the deadlock problem by proposing a learning-based framework that adapts the CBF class 𝒦\mathcal{K} function online. Using a probabilistic ensemble neural network (PENN), the framework predicts safety and performance metrics, including deadlock time, which is then minimized to mitigate deadlocks.

This paper builds upon the work [9] and presents a generalized theoretical understanding for the conditions for the formation of undesirable boundary equilibrium points with the CBF-based safety filters and CLF-CBF frameworks for safety-critical control. Particularly:

  • Addresses the stability theory for the closed-loop system of affine nonlinear systems with both the safety-filter QP and CLF-CBF mininum-norm QP using a single, generalized QP framework for both safety-critical controllers.

  • Derives sufficient conditions for the feasibility of the safety-filter QP and CLF-CBF QP with multiple CBFs.

  • Introduces conditions for existence of undesirable equilibrium points that are valid both for the safety-filter QP and for the CLF-CBF QP, considering multiple, possibly overlapping CBF unsafe sets.

  • Derives a novel condition for the stability of boundary equilibrium points of this type with a clear geometric intuition.

II PRELIMINARIES

Notation: Given a matrix An×mA\in\mathbb{R}^{n\times m} or vector vnv\in\mathbb{R}^{n} [A]kn[A]_{k}\in\mathbb{R}^{n} denotes its kk-th column and [v]k[v]_{k}\in\mathbb{R} denotes its kk-th element, while diag{α1,,αr}\text{diag}\{\alpha_{1},\,\cdots,\,\alpha_{r}\} is block matrix with the diagonal stacking of rr scalars or matrices α1,,αr\alpha_{1},\,\cdots,\,\alpha_{r}. Matrix Inn×nI_{n}\in\mathbb{R}^{n\times n} is the n×nn\times n identity matrix. LgfL_{g}f is the Lie derivative of a function ff along a function g:nn×mg:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n\times m}, that is, Lgf=f𝖳gmL_{g}f=\nabla f^{\mathsf{T}}g\in\mathbb{R}^{m}. The inner product between two vectors u,vnu,v\in\mathbb{R}^{n} induced by a positive semidefinite matrix X=X𝖳0X=X^{\mathsf{T}}\geq 0 is given by u,vX=u𝖳Xv\langle u,v\rangle_{X}=u^{\mathsf{T}}X\,v. This inner product induces a norm vX2=v,vX=v𝖳Xv\lVert v\rVert_{X}^{2}=\langle v,v\rangle_{X}=v^{\mathsf{T}}X\,v over n\mathbb{R}^{n}. The standard inner product is then u,v=u,vIn\langle u,v\rangle=\langle u,v\rangle_{I_{n}}, with standard Euclidean norm v2=vIn2\lVert v\rVert^{2}=\lVert v\rVert_{I_{n}}^{2}. The orthogonal complement of a subspace 𝒲\mathcal{W} is denoted by 𝒲\mathcal{W}^{\perp}, with orthogonality dependent on an inner product ,X\langle\cdot,\cdot\rangle_{X}.

II-A Control Lyapunov (CLFs) and Barrier Functions (CBFs)

Consider the nonlinear control affine system

x˙=f(x)+g(x)u\displaystyle\dot{x}=f(x)+g(x)u (1)

where xnx\in\mathbb{R}^{n}, umu\in\mathbb{R}^{m} are the system state and control input, respectively, and f:nnf:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n}, g:nn×mg:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n\times m} are locally Lipschitz.

Definition II.1 (CLFs).

A positive definite function VV is a control Lyapunov function (CLF) for system (1) if it satisfies:

infum[LfV(x)+LgVu]γ(V(x))\displaystyle\inf_{u\in\mathbb{R}^{m}}\left[L_{f}V(x)+L_{g}Vu\right]\leq-\gamma(V(x))

where γ:00\gamma:\mathbb{R}_{\geq 0}\rightarrow\mathbb{R}_{\geq 0} is a class 𝒦\mathcal{K} function [4].

Definition II.1 implies that there exists a set of stabilizing controls that makes the CLF strictly decreasing everywhere outside its global minimum x0nx_{0}\in\mathbb{R}^{n}, given by 𝕂V(x)={um:LfV+LgVuγ(V)}\mathbb{K}_{V}(x)=\{u\!\in\!\mathbb{R}^{m}:L_{f}V+L_{g}Vu\leq-\gamma(V)\}.

Definition II.2 (Safety).

The trajectories of a given system are safe with respect to a set 𝒞\mathcal{C} if 𝒞\mathcal{C} is forward invariant, meaning that for every x(0)𝒞x(0)\in\mathcal{C}, x(t)𝒞x(t)\in\mathcal{C} for all t>0t>0.

Consider NN subsets 𝒞1,,𝒞Nn\mathcal{C}_{1},\dots,\mathcal{C}_{N}\subset\mathbb{R}^{n} defined by the superlevel set of NN continuously differentiable functions hi:nh_{i}:\mathbb{R}^{n}\rightarrow\mathbb{R}, i=1,2,,Ni=1,2,\ldots,N The corresponding ii-th boundary is given by 𝒞i={xn:hi(x)=0}\partial\mathcal{C}_{i}=\{x\in\mathbb{R}^{n}:h_{i}(x)=0\}.

Definition II.3 (CBFs).

Let 𝒞i={xn:hi(x)0}\mathcal{C}_{i}=\{x\in\mathbb{R}^{n}:h_{i}(x)\geq 0\}. Then hi(x)h_{i}(x) is a (zeroing) Control Barrier Function (CBF) for (1) if there exists a locally Lipschitz extended class 𝒦\mathcal{K}_{\infty} function [4] αi:\alpha_{i}:\mathbb{R}\rightarrow\mathbb{R} such that

supum[Lfhi(x)+Lghi(x),u]αi(hi(x))\displaystyle\sup_{u\in\mathbb{R}^{m}}\left[L_{f}h_{i}(x)+\langle L_{g}h_{i}(x),u\rangle\right]\geq-\alpha_{i}(h_{i}(x))

This definition means that there exists a set of safe controls allowing the ii-th CBF to decrease on in the interior of its safe set int(𝒞i)\text{int}(\mathcal{C}_{i}), but not on its boundary 𝒞i\partial\mathcal{C}_{i}, given by 𝕂hi(x)={um:Lfhi+Lghiu+αi(hi)0}\mathbb{K}_{h_{i}}(x)\!=\!\{u\in\mathbb{R}^{m}\!:\!L_{f}h_{i}\!+\!L_{g}h_{i}u\!+\!\alpha_{i}(h_{i})\geq 0\}. The composite safe set associated to all the NN CBFs is simply 𝒞=i=1N𝒞i\mathcal{C}=\bigcap^{N}_{i=1}\mathcal{C}_{i} and the set of controls rendering 𝒞\mathcal{C} forward invariant is the intersection 𝕂h(x)=i=1N𝕂hi(x)\mathbb{K}_{h}(x)=\bigcap^{N}_{i=1}\mathbb{K}_{h_{i}}(x).

II-B QP-Based Safety-Critical Controllers

Consider the closed-loop system for (1)

x˙=fcl(x):=f(x)+g(x)u(x)\displaystyle\dot{x}=f_{cl}(x):=f(x)+g(x)u^{\star}(x) (2)

with a state-feedback control law u(x)u^{\star}(x). A common approach for safety-critical control is the safety-filter algorithm, which minimally modifies a given stabilizing state-feedback control law unom(x)u_{nom}(x) to achieve safety. The safety-filter effectively generates the “closest” safe control u(x)𝕂h(x)u^{\star}(x)\in\mathbb{K}_{h}(x) to the stabilizing control unom(x)u_{nom}(x).

An alternative approach that also makes use of QPs is the minimum-norm CLF-CBF QP by [1]:

u(x),δ(x)\displaystyle u^{\star}(x),\delta^{\star}(x) =argmin(u,δ)12(Δu)𝖳H(x)(Δu)+12pδ2\displaystyle=\operatorname*{argmin}_{(u,\delta)}\frac{1}{2}(\Delta u)^{\mathsf{T}}H(x)(\Delta u)+\frac{1}{2}p\delta^{2} (3)
s.t.\displaystyle s.t.\, LfV+LgV(x),uγ(V)+δ\displaystyle L_{f}V+\langle L_{g}V(x),u\rangle\leq-\gamma(V)+\delta (CLF)
Lfhi+Lghi(x),uαi(hi)\displaystyle L_{f}h_{i}+\langle L_{g}h_{i}(x),u\rangle\geq-\alpha_{i}(h_{i}) (CBFs)

i{1,,N}i\in\{1,\cdots,N\}, H:nn×nH:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n\times n} being a symmetric positive definite matrix function of the state, p>0p>0 and (for now) Δu=u\Delta u=u. The relaxation variable δ\delta in the CLF constraint softens the stabilization objective, aiming to maintain the feasibility of the QP. If feasible, the feedback controller (3) with Δu=u\Delta u=u generates a minimum-norm stabilizing and safe control, guaranteeing local stability of the origin and safety of the closed-loop system trajectories with respect to the composite safe set 𝒞\mathcal{C}.

As pointed out by the works [8, 9, 10], neither the safety-filter nor the CLF-CBF QP (3) (with Δu=u\Delta u=u) can guarantee global stabilization of trajectories towards the origin for the closed-loop system (2), meaning that trajectories could converge towards undesirable equilibrium points. Here, our objective is to study the solutions, existence conditions and stability of equilibrium points of the closed-loop system formed by a generalized QP controller given by (3) with Δu=uunom(x)\Delta u=u-u_{nom}(x), where unomu_{nom} is the stabilizing control law from the safety-filter QP. The structure of controller (3) with Δu=uunom(x)\Delta u=u-u_{nom}(x) allows for the generalization of the safety-filter and the CLF-CBF QPs into a single framework:

  1. 1.

    with H(x)=ImH(x)=I_{m} and V=0V=0 (without CLF), we recover the usual safety-filter QP. In this case, the optimal solution for the slack variable is always δ=0\delta^{\star}=0.

  2. 2.

    with unom(x)=0u_{nom}(x)=0, we recover the usual CLF-CBF QP.

If feasible, the solution of the generalized QP (3) is guaranteed u(x)𝕂h(x)u^{\star}(x)\in\mathbb{K}_{h}(x), and possibly close to the stabilizing set 𝕂V(x)\mathbb{K}_{V}(x). However, due to the slack variable, it is not possible to strictly guarantee that u(x)𝕂V(x)𝕂h(x)xnu^{\star}(x)\in\mathbb{K}_{V}(x)\cap\mathbb{K}_{h}(x)\,\,\forall x\in\mathbb{R}^{n}. That means that safety with respect to 𝒞\mathcal{C} is achieved, but stabilization is (possibly) hampered.

Assumption II.1.

The initial state x(0)nx(0)\in\mathbb{R}^{n} and the origin 0n0\in\mathbb{R}^{n} are contained in the safe set 𝒞\mathcal{C}, that is, hi(x(0))0h_{i}(x(0))\geq 0 and hi(0)0h_{i}(0)\geq 0 for all i{1,,N}i\in\{1,\cdots,N\}.

Theorem 1.

Under Assump. II.1, consider the following assumptions:
(i) There is only one CBF (N=1N=1).
(ii) System (1) is driftless: f(x)=0xnf(x)=0\,\,\,\forall x\in\mathbb{R}^{n}.
(iii) Considering any number of CBFs (N>1N>1),

b2(x)\displaystyle b_{2}(x) ImU𝖳𝒫VGU\displaystyle\in\operatorname{Im}{U^{\mathsf{T}}\mathcal{P}_{V}GU} (4)
b2(x)\displaystyle b_{2}(x) =U𝖳(c1γ(V)GV𝒫Vfnom)α¯\displaystyle=U^{\mathsf{T}}\left(c^{-1}\gamma(V)G\nabla V-\mathcal{P}_{V}f_{nom}\right)-\bar{\alpha}

where c=p1+VG2>0c\!=\!p^{-1}\!+\!\lVert\nabla V\rVert_{G}^{2}>0, G(x)=g(x)H(x)1g(x)𝖳G(x)\!=\!g(x)H(x)^{-1}g(x)^{\mathsf{T}}, fnom(x)=f(x)+g(x)unom(x)f_{nom}(x)\!=\!f(x)\!+\!g(x)u_{nom}(x), 𝒫V=Ic1GVV𝖳\mathcal{P}_{V}=I-c^{-1}G\nabla V\nabla V^{\mathsf{T}}, α¯=[α1(h1)αN(hN)]𝖳\bar{\alpha}\!=\!\begin{bmatrix}\,\alpha_{1}(h_{1})\!\!\!&\!\!\!\cdots\!\!\!&\!\!\!\alpha_{N}(h_{N})\,\end{bmatrix}^{\mathsf{T}} and U(x)=[h1hN]U(x)\!=\!\begin{bmatrix}\,\nabla h_{1}\!\!\!&\!\!\!\cdots\!\!\!&\!\!\!\nabla h_{N}\,\end{bmatrix}.
Then, QP (3) is feasible under Assumptions
(i) or (ii) or (iii).

Proof.

The proofs of (i) and (ii) can be found in [1] and [9], respectively. To establish the proof of (iii), we begin by formulating the Lagrangian associated with the QP (3)

(u,δ,λ¯)\displaystyle\mathcal{L}(u,\delta,\bar{\lambda})\! =12(ΔuH2+pδ2)\displaystyle=\!\frac{1}{2}\left(\lVert\Delta u\rVert_{H}^{2}\!+\!p\delta^{2}\right) (5)
+λ0(LfV+LgV,u+γ(V)δ)\displaystyle\!+\lambda_{0}(L_{f}V\!+\!\langle L_{g}V,u\rangle\!+\!\gamma(V)\!-\!\delta)
i=1Nλi(Lfhi+Lghi,u+αi(hi))\displaystyle\!-\sum^{N}_{i=1}\lambda_{i}(L_{f}h_{i}\!+\!\langle L_{g}h_{i},u\rangle\!+\!\alpha_{i}(h_{i}))

where λi0\lambda_{i}\geq 0 and λ=[λ1λN]𝖳0N\lambda\!=\!\begin{bmatrix}\,\lambda_{1}\!\!\!&\!\!\!\cdots\!\!\!&\!\!\!\lambda_{N}\,\end{bmatrix}^{\mathsf{T}}\in\mathbb{R}^{N}_{\geq 0}, λ¯=[λ0λ𝖳]𝖳0N+1\bar{\lambda}\!=\!\begin{bmatrix}\,\lambda_{0}\!\!&\!\!\lambda^{\mathsf{T}}\,\end{bmatrix}^{\mathsf{T}}\in\mathbb{R}^{N+1}_{\geq 0} are vectors of KKT multipliers associated to the optimization problem. Using matrix UU, the stationarity KKT conditions give the following solutions for the QP:

u(x)\displaystyle u^{\star}(x) =unom+H1g𝖳(λ0V+Uλ)\displaystyle=u_{nom}+H^{-1}g^{\mathsf{T}}\left(-\lambda_{0}\nabla V+U\lambda\right) (6)
δ(x)\displaystyle\delta^{\star}(x) =p1λ0\displaystyle=p^{-1}\lambda_{0} (7)

The dual function g(λ¯)=min(u,δ)(u,δ,λ¯)g(\bar{\lambda})=\min_{(u,\delta)}\mathcal{L}(u,\delta,\bar{\lambda}) associated to the QP can be obtained by substituting (6) into (5), yielding the following dual QP:

maxλ¯0N\displaystyle\max_{\bar{\lambda}\in\mathbb{R}^{N}_{\geq 0}} 12λ¯𝖳A(x)λ¯+λ¯𝖳b(x)\displaystyle-\frac{1}{2}\bar{\lambda}^{\mathsf{T}}A(x)\bar{\lambda}+\bar{\lambda}^{\mathsf{T}}b(x) (8)
A(x)\displaystyle A(x) =[cV𝖳GUU𝖳GVU𝖳GU],b(x)=[FVFh]\displaystyle=\begin{bmatrix}c\!\!&\!\!-\nabla V^{\mathsf{T}}GU\\ -U^{\mathsf{T}}G\nabla V\!\!&\!\!U^{\mathsf{T}}GU\end{bmatrix}\,,\,\,b(x)=\begin{bmatrix}F_{V}\\ -F_{h}\end{bmatrix}

with FV=LfnomV+γ(V)F_{V}=L_{f_{nom}}V+\gamma(V), Fh=U𝖳fnom+α¯F_{h}=U^{\mathsf{T}}f_{nom}+\bar{\alpha}. Notice that the dual cost in (8) is bounded from above if and only if b(x)ImA(x)b(x)\in\operatorname{Im}{A(x)}. In that case, the primal QP (3) is feasible. Applying Gauss elimination to the augmented matrix [A(x)|b(x)]\begin{bmatrix}\,A(x)\!\!&\!\!\!\!|\!\!\!\!&\!\!b(x)\,\end{bmatrix} yields

[1c1V𝖳GU|c1FV0U𝖳𝒫VGU|b2(x)]\displaystyle\begin{bmatrix}1\!\!&\!\!-c^{-1}\nabla V^{\mathsf{T}}GU\!\!\!&|&\!\!\!c^{-1}F_{V}\\ 0\!\!&\!\!U^{\mathsf{T}}\mathcal{P}_{V}GU\!\!\!&|&\!\!\!b_{2}(x)\end{bmatrix} (9)

Then, from (9), the condition b2(x)ImU𝖳𝒫VGUb_{2}(x)\in\operatorname{Im}{U^{\mathsf{T}}\mathcal{P}_{V}GU} as stated in (4) is equivalent to b(x)ImA(x)b(x)\in\operatorname{Im}{A(x)}, meaning that the primal QP is feasible under this condition. ∎

Theorem 1(iii) provides a sufficient condition for the feasibility of QP (3): when (4) holds, the corresponding dual QP cost is bounded, which means (3) is feasible. Feasibility condition (4) is of particular importance when assumptions (i) and (ii) fail, giving a sufficient condition for the feasibility of QP (3) in the case when multiple safety objectives are simultaneously required.

III Closed-loop System Analysis

III-A Existence of Equilibrium Points

In this section, we extend a result from [8], regarding the existence of equilibrium points in the safety-filter QP or in the CLF-CBF QP when multiple CBF constraints are present. The results are conditioned to the feasibility of the QP (3): that is, it is assumed that at least one of the conditions of Theorem 1 holds.

Definition III.1 (Equilibrium Manifold).

Let the set 𝒜={a1,,ar}2{1,,N}\mathcal{A}=\{a_{1},\cdots,a_{r}\}\subset 2^{\{1,\cdots,N\}} be a collection of 1rN1\leq r\leq N CBF indexes corresponding to rr CBFs {ha1,,har}\{h_{a_{1}},\cdots,h_{a_{r}}\}. Define the vector field f𝒜:n×0rnf_{\mathcal{A}}:\mathbb{R}^{n}\times\mathbb{R}^{r}_{\geq 0}\rightarrow\mathbb{R}^{n}:

f𝒜(x,λ)\displaystyle f_{\mathcal{A}}(x,\lambda) =fnompγ(V)GV+GU𝒜λ\displaystyle=f_{nom}-p\gamma(V)G\nabla V+GU_{\!\mathcal{A}}\lambda (10)
U𝒜\displaystyle U_{\!\mathcal{A}} =[ha1har]n×r\displaystyle=\begin{bmatrix}\,\nabla h_{a_{1}}\!\!&\!\!\cdots\!\!&\!\!\nabla h_{a_{r}}\,\end{bmatrix}\in\mathbb{R}^{n\times r} (11)

As will be demonstrated in the next sections, (10) and its Jacobian with respect to xx will be of central importance to characterize the existence and stability conditions for the equilibrium points of the closed-loop system.

Theorem 2 (Existence of Equilibrium Points).

Let (2) be the closed-loop system formed by the nonlinear system (1) with controller (3). Let the set 𝒜\mathcal{A} as defined in Definition III.1 representing the indexes of rr overlapping CBF boundaries 𝒞𝒜=i=1r𝒞ai\partial\mathcal{C}_{\mathcal{A}}=\bigcap^{r}_{i=1}\partial\mathcal{C}_{a_{i}}\neq\emptyset. The equilibrium points of (2) come in two distinct types:

𝒞𝒜\displaystyle\mathcal{E}_{\partial\mathcal{C}_{\mathcal{A}}}\! =𝒞𝒜{xn|λ0r s.t. f𝒜(x,λ)=0}\displaystyle=\!\partial\mathcal{C}_{\mathcal{A}}\!\cap\!\{x\in\mathbb{R}^{n}\,|\,\exists\lambda\!\in\!\mathbb{R}^{r}_{\geq 0}\text{ s.t. }f_{\mathcal{A}}(x,\lambda)\!=\!0\} (12)
int(𝒞)\displaystyle\mathcal{E}_{int(\mathcal{C})}\! =int(𝒞){xn|fnom=pγ(V)GV}\displaystyle=\!int(\mathcal{C})\!\cap\!\{x\in\mathbb{R}^{n}\,|\,f_{nom}\!=\!p\gamma(V)G\nabla V\} (13)

where 𝒞𝒜\mathcal{E}_{\partial\mathcal{C}_{\mathcal{A}}} is the set of boundary equilibrium points occurring in 𝒞𝒜\partial\mathcal{C}_{\mathcal{A}} and int(𝒞)\mathcal{E}_{int(\mathcal{C})} is the set of interior equilibrium points. Furthermore, defining the region of the state space where only the a1,,ara_{1},\cdots,a_{r} CBF constraints are simultaneously active as 𝒮𝒜n\mathcal{S}_{\mathcal{A}}\subset\mathbb{R}^{n}, we have 𝒞𝒜𝒮𝒜\mathcal{E}_{\partial\mathcal{C}_{\mathcal{A}}}\subset\mathcal{S}_{\mathcal{A}}.

Proof.

Substituting (6) in (2) and using the definition of G(x)G(x) yields the following closed-loop dynamics:

fcl(x)=fnom(x)+G(x)(λ0V+Uλ)\displaystyle f_{cl}(x)=f_{nom}(x)+G(x)\left(-\lambda_{0}\nabla V+U\lambda\right) (14)

At an equilibrium point xex_{e}\in\mathcal{E}, fcl(xe)=0f_{cl}(x_{e})=0. Applying this condition to (2) yields

fnom(xe)=G(xe)(λ0V(xe)U(xe)λ)\displaystyle f_{nom}(x_{e})=G(x_{e})\left(\lambda_{0}\nabla V(x_{e})-U(x_{e})\lambda\right) (15)

Case 1. Consider the region of the state space where the CLF constraint is inactive: LfclV+γ(V)δ<0L_{f_{cl}}V+\gamma(V)-\delta^{\star}<0. Following the exact same steps of Case 1 of Theorem 2 on [9], we conclude that no equilibrium points can occur at this region. Case 2. Consider the region where CLF constraint is active: LfclV+γ(V)=δL_{f_{cl}}V+\gamma(V)=\delta^{\star}. In the case of the safety-filter, since V=0V=0, the CLF constraint becomes simply 0δ0\leq\delta and since the optimal solution for the relaxation variable is always δ=0\delta^{\star}=0, the CLF constraint is always active. At an equilibrium point xex_{e}\in\mathcal{E}, LfclV(xe)=0L_{f_{cl}}V(x_{e})=0. Therefore, using (7), γ(V(xe))=δ(xe)=p1λ0\gamma(V(x_{e}))=\delta^{\star}(x_{e})=p^{-1}\lambda_{0}. Then, at any equilibrium point xex_{e}\in\mathcal{E}, the multiplier associated to the CLF constraint is λ0(xe)=pγ(V(xe))0\lambda_{0}(x_{e})=p\gamma(V(x_{e}))\geq 0. Therefore, equation (15) yields

fnom(xe)=G(xe)(pγ(V(xe))V(xe)U(xe)λ)\displaystyle f_{nom}(x_{e})\!=\!G(x_{e})\left(p\gamma(V(x_{e}))\nabla V(x_{e})-U(x_{e})\lambda\right) (16)

For the safety-filter, V=0V=0 and V=0\nabla V=0 are valid in (16). For the next cases, the CLF constraint is assumed to be active.
Case 3. Consider the region where exactly rNr\leq N CBF constraints are simultaneously active. Their corresponding indexes are denoted by the set 𝒜={a1,,ar}\mathcal{A}=\{a_{1},\cdots,a_{r}\} as defined previously. Therefore, Lfclhi+αi(hi)=0L_{f_{cl}}h_{i}+\alpha_{i}(h_{i})=0, for i𝒜i\in\mathcal{A}. At an equilibrium point xex_{e} occurring in this region, Lfclhi(xe)=0L_{f_{cl}}h_{i}(x_{e})=0, implying that ha1(xe)==har(xe)=0h_{a_{1}}(x_{e})=\cdots=h_{a_{r}}(x_{e})=0. Therefore, xex_{e} must lie at the boundary intersection associated to the rr active CBFs, that is, xe𝒞𝒜x_{e}\in\partial\mathcal{C}_{\mathcal{A}}. The conclusion is that boundary equilibrium points at 𝒞𝒜\partial\mathcal{C}_{\mathcal{A}} can only occur at 𝒮𝒜\mathcal{S}_{\mathcal{A}}, that is, the region where only the CBF constraints corresponding to ha1,,harh_{a_{1}},\cdots,h_{a_{r}} are simultaneously active. Since in this case the remaining CBF constraints are all inactive, λi=0i𝒜\lambda_{i}=0\,\,\forall i\notin\mathcal{A}, and (16) reduces to

fnom(xe)=G(xe)(pγ(V)V(xe)U𝒜(xe)λa)\displaystyle f_{nom}(x_{e})=G(x_{e})\left(p\gamma(V)\nabla V(x_{e})-U_{\mathcal{A}}(x_{e})\lambda_{a}\right) (17)

where U𝒜U_{\mathcal{A}} is given by (11) and λa0r\lambda_{a}\in\mathbb{R}^{r}_{\geq 0} is a vector of appropriate size with the non-negative corresponding KKT multipliers associated to the active CBF constraints. Notice that (17) is equivalent to f𝒜(x,λa)=0f_{\mathcal{A}}(x,\lambda_{a})=0 as defined in (10). Thus, in this case, the equilibrium point is at the boundary intersection 𝒞𝒜\partial\mathcal{C}_{\mathcal{A}} and satisfies f𝒜(xe,λa)=0f_{\mathcal{A}}(x_{e},\lambda_{a})=0 for some λa0r\lambda_{a}\in\mathbb{R}^{r}_{\geq 0}, demonstrating (12).
Case 4. Consider the region where all CBF constraints are inactive: Lfclhi+αi(hi)>0L_{f_{cl}}h_{i}+\alpha_{i}(h_{i})>0, i=1,,Ni=1,\cdots,N. Following the same steps of Case 4 of Theorem 2 on [9], we conclude that equilibrium points occurring in this region must lie in the interior of the safe set, that is, xeint(𝒞)x_{e}\in int(\mathcal{C}). Additionally, (16) must be satisfied with λ=0\lambda=0, which means that f(xe)=pγ(V(xe))G(xe)V(xe)f(x_{e})=p\gamma(V(x_{e}))G(x_{e})\nabla V(x_{e}). This demonstrates (13). ∎

The work [10] has proposed a system transformation that removes certain types of interior equilibrium points from the closed-loop system with the CLF-CBF QP controller. We conjecture that a similar transformation could be performed in the case of the safety-filter. Thereby, in the remaining of the paper, we focus on the stability properties for boundary equilibrium points.

III-B Stability of Boundary Equilibrium Points

Lemma 1 (Closed-Loop Jacobian).

Let (2) be the closed-loop system formed by the nonlinear system (1) with controller (3), and let 𝒜\mathcal{A} from Definition III.1 represent the indexes of 1rn1\leq r\leq n active CBF constraints. For a boundary equilibrium point xe𝒞𝒜x_{e}\in\mathcal{E}_{\partial\mathcal{C}_{\mathcal{A}}} with full rank and U𝒜𝖳(xe)g(xe)0U_{\mathcal{A}}^{\mathsf{T}}(x_{e})g(x_{e})\neq 0, the Jacobian matrix Jcl(xe)n×nJ_{cl}(x_{e})\in\mathbb{R}^{n\times n} of the closed-loop system (2) computed at xex_{e} is given by

Jfcl(xe)\displaystyle J_{f_{cl}}(x_{e})\! =𝒫U𝒜(𝒫VJ𝒜(xe,λ¯e)c1γ(V)GVV𝖳)\displaystyle=\!\mathcal{P}_{U_{\mathcal{A}}}\!\left(\mathcal{P}_{V}J_{\mathcal{A}}(x_{e},\bar{\lambda}_{e})\!-\!c^{-1}\gamma^{\prime}(V)G\nabla V\nabla V^{\mathsf{T}}\right)
𝒫VGU𝒜SA1ΛU𝒜𝖳\displaystyle\qquad\qquad\qquad-\mathcal{P}_{V}GU_{\mathcal{A}}S_{A}^{-1}\Lambda^{\prime}U_{\mathcal{A}}^{\mathsf{T}} (18)

where λ¯e=[pγ(V(xe))λe𝖳]𝖳\bar{\lambda}_{e}=\begin{bmatrix}\,p\gamma(V(x_{e}))&\lambda_{e}^{\mathsf{T}}\,\end{bmatrix}^{\mathsf{T}}, and λe0r\lambda_{e}\in\mathbb{R}^{r}_{\geq 0} is the vector with the rr corresponding KKT multipliers at xex_{e} for the rr active CBF constraints, 𝒫U𝒜=In𝒫VGU𝒜SA1U𝒜𝖳\mathcal{P}_{U_{\mathcal{A}}}\!=\!I_{n}\!-\!\mathcal{P}_{V}GU_{\mathcal{A}}\,S_{A}^{-1}U^{\mathsf{T}}_{\mathcal{A}} with SA=U𝒜𝖳𝒫VGU𝒜S_{A}\!\!=\!U^{\mathsf{T}}_{\mathcal{A}}\mathcal{P}_{V}GU_{\mathcal{A}}, Λ=diag{αa1(ha1),,αar(har)}>0\Lambda^{\prime}\!\!=\!\text{diag}\{\alpha^{\prime}_{a_{1}}\!(h_{a_{1}}),\!\cdots\!,\alpha^{\prime}_{a_{r}}\!(h_{a_{r}})\}\!>\!0 and

J𝒜(x,λ¯a)\displaystyle J_{\mathcal{A}}(x,\bar{\lambda}_{a})\! =fnomxλ0(GV)x+i𝒜(Ghi)xλi\displaystyle=\!\frac{\partial f_{nom}}{\partial x}\!-\!\lambda_{0}\frac{\partial(G\nabla V)}{\partial x}\!+\!\sum_{i\in\mathcal{A}}\frac{\partial(G\nabla h_{i})}{\partial x}\lambda_{i} (19)

with λ¯a=[λ0λa𝖳]𝖳\bar{\lambda}_{a}=\begin{bmatrix}\,\lambda_{0}&\lambda_{a}^{\mathsf{T}}\,\end{bmatrix}^{\mathsf{T}}.

Remark III.1.

The Jacobian J𝒜J_{\mathcal{A}} as defined in (19) is not the same as Jf𝒜J_{f_{\mathcal{A}}}, the Jacobian of f𝒜f_{\mathcal{A}} in (10): they differ precisely by a factor of pγ(V)VV𝖳p\gamma^{\prime}(V)\nabla V\nabla V^{\mathsf{T}}.

Proof.

This demonstration is a direct continuation of Case 3 from the proof of Theorem 2. Using the complementary slackness conditions, V˙+LgV,u=δ\dot{V}+\langle L_{g}V,u^{\star}\rangle=\delta^{\star}, h˙i+Lghi,u=0i𝒜\dot{h}_{i}+\langle L_{g}h_{i},u^{\star}\rangle=0\,\,\forall i\in\mathcal{A}. Substituting (6)-(7) these expressions and using the fact that λi=0\lambda_{i}=0 for all i𝒜i\notin\mathcal{A} yields the following system:

[cV𝖳GU𝒜U𝒜𝖳GVU𝒜𝖳GU𝒜]Aa(x)[λ0λa]λ¯a=[FVFha]ba(x)\displaystyle\underbrace{\begin{bmatrix}c\!&\!-\nabla V^{\mathsf{T}}GU_{\mathcal{A}}\\ -U_{\mathcal{A}}^{\mathsf{T}}G\nabla V\!&\!U_{\mathcal{A}}^{\mathsf{T}}GU_{\mathcal{A}}\end{bmatrix}}_{A_{a}(x)}\underbrace{\begin{bmatrix}\lambda_{0}\\ \lambda_{a}\end{bmatrix}}_{\bar{\lambda}_{a}}\!=\!\underbrace{\begin{bmatrix}F_{V}\\ -F_{h_{a}}\end{bmatrix}}_{b_{a}(x)} (20)

where λa0r\lambda_{a}\in\mathbb{R}^{r}_{\geq 0}, Fha=U𝒜𝖳fnom+α¯arF_{h_{a}}\!=U_{\mathcal{A}}^{\mathsf{T}}\,f_{nom}+\bar{\alpha}_{a}\in\mathbb{R}^{r} and α¯a=[αa1(ha1)αar(har)]𝖳\bar{\alpha}_{a}\!=\!\begin{bmatrix}\,\alpha_{a_{1}}(h_{a_{1}})\!\!&\!\!\cdots\!\!&\!\!\alpha_{a_{r}}(h_{a_{r}})\,\end{bmatrix}^{\mathsf{T}}. Here, Aa(x)(r+1)×(r+1)A_{a}(x)\in\mathbb{R}^{(r+1)\times(r+1)} and ba(x)r+1b_{a}(x)\in\mathbb{R}^{r+1} are essentially reduced versions of matrices A(x)A(x) and b(x)b(x) from (8), with fewer rows and columns (since not all constraints are active). In particular, the set 𝒮𝒜\mathcal{S}_{\mathcal{A}} where the CLF and only the CBF constraints from 𝒜\mathcal{A} are active is given by

𝒮𝒜={xn|\displaystyle\mathcal{S}_{\mathcal{A}}=\{x\in\mathbb{R}^{n}\,|\, fcl(x)=fnomλ0GV+GU𝒜λa}\displaystyle f_{cl}(x)\!=\!f_{nom}\!-\!\lambda_{0}G\nabla V\!+\!GU_{\mathcal{A}}\lambda_{a}\} (21)

where λ0,λa\lambda_{0},\lambda_{a} are the positive solutions of (20). Using the known formula for inversion of block matrices, since c>0c>0, Aa(x)A_{a}(x) is invertible if its Schur complement SA=U𝒜𝖳𝒫VGU𝒜r×rS_{A}=U^{\mathsf{T}}_{\mathcal{A}}\mathcal{P}_{V}GU_{\mathcal{A}}\in\mathbb{R}^{r\times r} is invertible. Since 𝒫V(x)>0x\mathcal{P}_{V}(x)>0\,\forall x, SA1S_{A}^{-1} exists if U𝒜U_{\mathcal{A}} has full column rank and if U𝒜𝖳(x)g(x)0U^{\mathsf{T}}_{\mathcal{A}}(x)g(x)\neq 0. Under these assumptions, a formula for Aa1(x)A_{a}^{-1}(x) is

Aa1(x)=[c1000]+1c2[V𝖳GU𝒜cIr]SA1[V𝖳GU𝒜cIr]𝖳\displaystyle A_{a}^{-1}(x)\!=\!\!\begin{bmatrix}c^{-1}\!\!&\!\!0\\ 0\!\!&\!\!0\end{bmatrix}\!+\!\frac{1}{c^{2}}\!\!\begin{bmatrix}\nabla V^{\mathsf{T}}GU_{\mathcal{A}}\\ c\,I_{r}\end{bmatrix}\!S_{A}^{-1}\!\!\begin{bmatrix}\nabla V^{\mathsf{T}}GU_{\mathcal{A}}\\ c\,I_{r}\end{bmatrix}^{\mathsf{T}}\!\!\! (22)

where the dimensions of vectors and matrices are conformable for matrix addition and multiplication. Then, the KKT multipliers λ0\lambda_{0}, λa\lambda_{a} can be found by solving (20).

Taking the derivative of (20) with respect to the kk-th state component xkx_{k} and solving for kλ¯a\partial_{k}\bar{\lambda}_{a} yields

kλ¯a(x)\displaystyle\partial_{k}\bar{\lambda}_{a}(x) =Aa1(kbakAaλ¯a)\displaystyle=A_{a}^{-1}\left(\partial_{k}b_{a}-\partial_{k}A_{a}\bar{\lambda}_{a}\right) (23)

Defining U¯=[VU𝒜]n×(r+1)\bar{U}=\begin{bmatrix}\,-\nabla V\!\!&\!\!U_{\mathcal{A}}\,\end{bmatrix}\in\mathbb{R}^{n\times(r+1)}, the partial derivatives of Aa(x)A_{a}(x) and ba(x)b_{a}(x) are:

kAa(x)\displaystyle\partial_{k}A_{a}(x)\! =k(U¯𝖳GU¯)\displaystyle=\!\partial_{k}(\bar{U}^{\mathsf{T}}G\bar{U}) (24)
kba(x)\displaystyle\partial_{k}b_{a}(x)\! =k(U¯𝖳fnom)Λ¯[U¯𝖳]k\displaystyle=\!-\partial_{k}(\bar{U}^{\mathsf{T}}f_{nom})\!-\!\bar{\Lambda}^{\prime}[\bar{U}^{\mathsf{T}}]_{k} (25)

where Λ¯=diag{γ(V),αa1(ha1),,αar(har)}\bar{\Lambda}^{\prime}=\text{diag}\{\gamma^{\prime}(V),\alpha^{\prime}_{a_{1}}(h_{a_{1}}),\,\cdots,\,\alpha^{\prime}_{a_{r}}(h_{a_{r}})\}. From (19), notice that [J𝒜]k=kfnom+k(GU¯)λ¯a[J_{\mathcal{A}}]_{k}=\partial_{k}f_{nom}+\partial_{k}(G\bar{U})\bar{\lambda}_{a}. Using this fact, combining equations (24)-(25) to compute the term kbakAaλ¯a\partial_{k}b_{a}-\partial_{k}A_{a}\bar{\lambda}_{a} in (23), left multiplying it by Aa1A_{a}^{-1} and using the fact that fcl(x)=fnom+GU¯λ¯af_{cl}(x)=f_{nom}+G\bar{U}\bar{\lambda}_{a} yields

kλ¯a(x)\displaystyle\partial_{k}\bar{\lambda}_{a}(x)\! =Aa1(U¯𝖳[J𝒜]k+Λ¯[U¯𝖳]k+(kU¯)𝖳fcl)\displaystyle=\!-A_{a}^{-1}\!\left(\bar{U}^{\mathsf{T}}[J_{\mathcal{A}}]_{k}\!+\!\bar{\Lambda}^{\prime}[\bar{U}^{\mathsf{T}}]_{k}\!+\!(\partial_{k}\bar{U})^{\mathsf{T}}f_{cl}\right) (26)

where the fact that [J𝒜]k=k(fnom+GU¯λ¯a)[J_{\mathcal{A}}]_{k}=\partial_{k}(f_{nom}+G\bar{U}\bar{\lambda}_{a}) was used. Then, taking the derivative of the closed-loop system dynamics (2) with λi=0i𝒜\lambda_{i}=0\,\,\forall i\notin\mathcal{A} yields

kfcl(x)=[J𝒜]k+GU¯kλ¯a\displaystyle\partial_{k}f_{cl}(x)=[J_{\mathcal{A}}]_{k}+G\bar{U}\partial_{k}\bar{\lambda}_{a} (27)

which by using the expression for kλ¯a\partial_{k}\bar{\lambda}_{a} in (26) yields an expression for the kk-th column of the closed-loop Jacobian matrix Jfcl(x)J_{f_{cl}}(x) at 𝒮𝒜\mathcal{S}_{\mathcal{A}}. Consider an equilibrium point xe𝒞𝒜x_{e}\in\mathcal{E}_{\partial\mathcal{C}_{\mathcal{A}}}. Since fcl(xe)=0f_{cl}(x_{e})=0 by definition, the last term on the right-hand side of (23) vanishes. Then, substituting kλ¯a(xe)\partial_{k}\bar{\lambda}_{a}(x_{e}) in (27) and using the fact that λ0(xe)=pγ(V(xe))\lambda_{0}(x_{e})=p\gamma(V(x_{e})) yields

kfcl(xe)\displaystyle\partial_{k}f_{cl}(x_{e}) =(IGU¯Aa1U¯𝖳)[J𝒜(xe,λ¯e)]k\displaystyle=(I\!-\!G\bar{U}A_{a}^{-1}\bar{U}^{\mathsf{T}})[J_{\mathcal{A}}(x_{e},\bar{\lambda}_{e})]_{k} (28)
GU¯Aa1Λ¯[U¯𝖳]k\displaystyle\qquad\qquad\qquad-G\bar{U}A_{a}^{-1}\bar{\Lambda}^{\prime}[\bar{U}^{\mathsf{T}}]_{k}

On the assumptions of the theorem, namely a full rank U𝒜(xe)U_{\mathcal{A}}(x_{e}) and U𝒜𝖳(xe)g(xe)0U^{\mathsf{T}}_{\mathcal{A}}(x_{e})g(x_{e})\neq 0, equation (22) can be used to simplify the following expressions at (28):

(IGU¯Aa1U¯𝖳)\displaystyle(I-G\bar{U}A_{a}^{-1}\bar{U}^{\mathsf{T}}) =𝒫U𝒜𝒫V\displaystyle=\mathcal{P}_{U_{\mathcal{A}}}\mathcal{P}_{V} (29)
GU¯Aa1Λ¯[U¯𝖳]k\displaystyle G\bar{U}A_{a}^{-1}\bar{\Lambda}^{\prime}[\bar{U}^{\mathsf{T}}]_{k} =c1γ(V)𝒫U𝒜GV[V]k\displaystyle=c^{-1}\gamma^{\prime}(V)\mathcal{P}_{U_{\mathcal{A}}}G\nabla V[\nabla V]_{k}
+𝒫VGU𝒜SA1Λ[U𝒜𝖳]k\displaystyle+\mathcal{P}_{V}GU_{\mathcal{A}}S_{A}^{-1}\Lambda^{\prime}[U_{\mathcal{A}}^{\mathsf{T}}]_{k} (30)

Substituting (29)-(30) in (28) yields

kfcl(xe)\displaystyle\partial_{k}f_{cl}(x_{e})\! =𝒫U𝒜(𝒫VJ𝒜(xe,λ¯e)c1γ(V)GV[V]k)\displaystyle=\!\mathcal{P}_{U_{\mathcal{A}}}\!\left(\mathcal{P}_{V}J_{\mathcal{A}}(x_{e},\bar{\lambda}_{e})\!-\!c^{-1}\gamma^{\prime}(V)G\nabla V[\nabla V]_{k}\right)
𝒫VGU𝒜SA1Λ[U𝒜]k\displaystyle\qquad\qquad\qquad-\mathcal{P}_{V}GU_{\mathcal{A}}S_{A}^{-1}\Lambda^{\prime}[U_{\mathcal{A}}]_{k} (31)

which is simply the kk-th column of the closed-loop Jacobian matrix Jfcl(xe)J_{f_{cl}}(x_{e}) computed at xex_{e}. ∎

Remark III.2.

Notice that U𝒜𝖳𝒫U𝒜=U𝒜𝖳SASA1U𝒜𝖳=0U_{\mathcal{A}}^{\mathsf{T}}\mathcal{P}_{U_{\mathcal{A}}}=U_{\mathcal{A}}^{\mathsf{T}}-S_{A}S_{A}^{-1}U_{\mathcal{A}}^{\mathsf{T}}=0, therefore showing that 𝒫U𝒜\mathcal{P}_{U_{\mathcal{A}}} is a generalized projection matrix for the orthogonal complement of the column space of U𝒜U_{\mathcal{A}}. From (18), that means that U𝒜𝖳(xe)Jfcl(xe)=ΛU𝒜𝖳(xe)U_{\mathcal{A}}^{\mathsf{T}}(x_{e})J_{f_{cl}}(x_{e})=-\Lambda^{\prime}U_{\mathcal{A}}^{\mathsf{T}}(x_{e}). Therefore, the gradients ha1(xe),,har(xe)\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e}) are left eigenvectors of Jfcl(xe)J_{f_{cl}}(x_{e}) with corresponding negative eigenvalues αa1(ha1),,αar(har)<0-\alpha_{a_{1}}^{\prime}(h_{a_{1}}),\cdots,-\alpha_{a_{r}}^{\prime}(h_{a_{r}})<0. In particular, this implies that if an equilibrium point xex_{e} on the conditions of Theorem 2 occurs at the intersection of exactly r=nr=n barrier boundaries, xex_{e} must be stable.

Lemma 2.

Let Xn×nX\in\mathbb{R}^{n\times n} be a symmetric positive definite matrix defining an inner product ,X\langle\cdot,\cdot\rangle_{X} over n\mathbb{R}^{n}, and 𝒵={z1,,zr}\mathcal{Z}=\{z_{1},\cdots,z_{r}\} be a set of rr linearly independent vectors. Additionally, let 𝒱={v1,,vnr}\mathcal{V}=\{v_{1},\cdots,v_{n-r}\} be a basis for the orthogonal complement of 𝒵\mathcal{Z} with respect to ,X\langle\cdot,\cdot\rangle_{X}. Then, 𝒵𝒱\mathcal{Z}\cup\mathcal{V} is a basis for n\mathbb{R}^{n}.

Proof.

Recall that since the span of 𝒱\mathcal{V} is the orthogonal complement of 𝒵\mathcal{Z} with respect to ,X\langle\cdot,\cdot\rangle_{X}, vi,zjX=0i,j\langle v_{i},z_{j}\rangle_{X}=0\,\,\forall i,j. Let i=1raizi+i=1nrbivi=0\sum^{r}_{i=1}a_{i}z_{i}\!+\!\sum^{n-r}_{i=1}b_{i}v_{i}\!=\!0 be the equation for deciding linear independence. Taking the inner product ,X\langle\cdot,\cdot\rangle_{X} with zj𝒵z_{j}\in\mathcal{Z} yields i=1raizj,ziX=0\sum^{r}_{i=1}a_{i}\langle z_{j},z_{i}\rangle_{X}\!=\!0. Since the vectors from 𝒵\mathcal{Z} are linearly independent, the only solution is ai=0,i=1,,ra_{i}=0,\,i=1,\cdots,r. Similarly, taking the inner product ,X\langle\cdot,\cdot\rangle_{X} with vj𝒱v_{j}\in\mathcal{V} yields i=1nrbivj,viX=0\sum^{n-r}_{i=1}b_{i}\langle v_{j},v_{i}\rangle_{X}=0. Since the vectors from 𝒱\mathcal{V} are also linearly independent (𝒱\mathcal{V} is a basis), the only solution is bi=0,i=1,,nrb_{i}=0,\,i=1,\cdots,n-r. Therefore, the set 𝒵𝒱\mathcal{Z}\cup\mathcal{V} must be composed of nn linearly independent vectors, constituting a basis for n\mathbb{R}^{n}. ∎

Proposition 1.

Let 𝒲={w1,,wn1}\mathcal{W}=\{w_{1},\cdots,w_{n-1}\} be a basis for {GV}\{G\nabla V\}^{\perp}. Then, with BV=[Vw1wn1]n×nB_{V}=\begin{bmatrix}\,\nabla V\!\!&\!\!w_{1}\!\!&\!\!\cdots\!\!&\!\!w_{n-1}\,\end{bmatrix}\in\mathbb{R}^{n\times n}, the following formula holds:

𝒫VJ𝒜c1γ(V)GVV𝖳\displaystyle\mathcal{P}_{V}J_{\mathcal{A}}\!-\!c^{-1}\gamma^{\prime}(V)G\nabla V\nabla V^{\mathsf{T}} =(BV𝖳)1DBV𝖳Jf𝒜\displaystyle=(B_{V}^{\mathsf{T}})^{-1}DB_{V}^{\mathsf{T}}J_{f_{\mathcal{A}}} (32)

where D=diag{p1c1,In1}>0D=\text{diag}\{p^{-1}c^{-1},I_{n-1}\}>0.

Proof.

By Lemma 2, the set V={V,w1,,wn1}\mathcal{B}_{V}=\{\nabla V,w_{1},\cdots,w_{n-1}\} is a basis for n\mathbb{R}^{n}. Therefore, the square matrix BV=[Vw1wn1]B_{V}=\begin{bmatrix}\,\nabla V\!\!&\!\!w_{1}\!\!&\!\!\cdots\!\!&\!\!w_{n-1}\,\end{bmatrix} is full rank. Left-multiplying matrix 𝒫VJ𝒜c1γ(V)GVV𝖳\mathcal{P}_{V}J_{\mathcal{A}}\!-\!c^{-1}\gamma^{\prime}(V)G\nabla V\nabla V^{\mathsf{T}} by V𝖳\nabla V^{\mathsf{T}} and wi𝖳w_{i}^{\mathsf{T}} and carrying out the algebraic simplifications due to 1c1VG2=(pc)11-c^{-1}\lVert\nabla V\rVert_{G}^{2}=(pc)^{-1} and wi𝖳GV=0w_{i}^{\mathsf{T}}G\nabla V=0 yields p1c1V𝖳Jf𝒜p^{-1}c^{-1}\nabla V^{\mathsf{T}}J_{f_{\mathcal{A}}} and wi𝖳Jf𝒜w_{i}^{\mathsf{T}}J_{f_{\mathcal{A}}}, respectively, i=1,,n1i=1,\cdots,n-1. Combining these nn equations in matrix form yields (32) with both sides left-multiplied by BV𝖳B_{V}^{\mathsf{T}}. ∎

Next, we demonstrate the main result of this work, a theorem for the stability of boundary equilibrium points at 𝒞𝒜\mathcal{E}_{\partial\mathcal{C}_{\mathcal{A}}}, occurring at the intersection of exactly r<nr<n barrier boundaries.

Theorem 3 (Stability of Equilibrium Points).

Consider the same assumptions of Lemma 1, let xe𝒞𝒜x_{e}\in\mathcal{E}_{\partial\mathcal{C}_{\mathcal{A}}} be a boundary equilibrium point of the closed-loop system, and r=|𝒜|<nr=|\mathcal{A}|<n. If there exists v{ha1(xe),,har(xe)}v\in\{\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e})\}^{\perp} (with the standard inner product ,\langle\cdot,\cdot\rangle) such that

v𝖳Jf𝒜(xe,λe)v>0\displaystyle v^{\mathsf{T}}J_{f_{\mathcal{A}}}(x_{e},\lambda_{e})v>0 (33)

then xex_{e} is unstable. Otherwise, it is stable. In (33), Jf𝒜J_{f_{\mathcal{A}}} is the Jacobian of the vector field (10) with respect to xx.

Proof.

Consider a boundary equilibrium point xe𝒞𝒜x_{e}\in\mathcal{E}_{\partial\mathcal{C}_{\mathcal{A}}} with U𝒜(xe)0U_{\mathcal{A}}(x_{e})\neq 0. The first order Taylor series approximation of the closed-loop system on a neighborhood of xex_{e} is x˙Jcl(xe)Δx\dot{x}\approx J_{cl}(x_{e})\Delta x with Δx=(xxe)\Delta x=(x-x_{e}) being a disturbance vector around the equilibrium point. Since the CBF gradients are linearly independent at xex_{e} by assumption, Δx\Delta x can be written using a basis {ha1(xe),,har(xe),v1,,vn1(xe)}\{\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e}),v_{1},\cdots,v_{n-1}(x_{e})\}, where v1,,vnrv_{1},\cdots,v_{n-r} are fixed basis vectors for {ha1(xe),,har(xe)}\{\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e})\}^{\perp}. Therefore, vj𝖳U𝒜(xe)=0jv_{j}^{\mathsf{T}}U_{\mathcal{A}}(x_{e})=0\,\,\forall j by construction. One can write x=xe+U𝒜(xe)a+vx=x_{e}+U_{\mathcal{A}}(x_{e})\,a+v, where vv is a linear combination of the viv_{i}, i=1,,nri=1,\cdots,n-r and ara\in\mathbb{R}^{r} is a vector of coordinates. The time derivative of xx in this new basis is x˙=U𝒜(xe)a˙+v˙\dot{x}=U_{\mathcal{A}}(x_{e})\,\dot{a}+\dot{v}. Left-multiplying this equation and Jfcl(xe)ΔxJ_{f_{cl}}(x_{e})\Delta x by U𝒜𝖳(xe)U_{\mathcal{A}}^{\mathsf{T}}(x_{e}), and using the expression (18) for the closed-loop Jacobian at xex_{e} and the fact that U𝒜𝖳v=0U_{\mathcal{A}}^{\mathsf{T}}v=0 yields

U𝒜𝖳x˙\displaystyle U_{\mathcal{A}}^{\mathsf{T}}\dot{x} =U𝒜𝖳(U𝒜a˙+v˙)=U𝒜𝖳U𝒜a˙\displaystyle=U_{\mathcal{A}}^{\mathsf{T}}(U_{\mathcal{A}}\,\dot{a}+\dot{v})=U_{\mathcal{A}}^{\mathsf{T}}U_{\mathcal{A}}\,\dot{a} (34)
U𝒜𝖳Jfcl(xe)Δx\displaystyle U_{\mathcal{A}}^{\mathsf{T}}J_{f_{cl}}(x_{e})\Delta x =ΛU𝒜𝖳(U𝒜a+v)=ΛU𝒜𝖳U𝒜a\displaystyle=-\Lambda^{\prime}U_{\mathcal{A}}^{\mathsf{T}}(U_{\mathcal{A}}a+v)=-\Lambda^{\prime}U_{\mathcal{A}}^{\mathsf{T}}U_{\mathcal{A}}a (35)

Comparing (34)-(35) yields a˙=(U𝒜𝖳U𝒜)1ΛU𝒜𝖳U𝒜ar\dot{a}=-(U_{\mathcal{A}}^{\mathsf{T}}U_{\mathcal{A}})^{-1}\Lambda^{\prime}U_{\mathcal{A}}^{\mathsf{T}}U_{\mathcal{A}}\,a\in\mathbb{R}^{r}. Since Λ>0\Lambda^{\prime}>0, this subsystem is asymptotically stable, which means that the column space of U𝒜U_{\mathcal{A}} is contained in the stable subspace associated to xex_{e}. Using these equations to solve for the dynamics of vv and letting a0a\rightarrow 0, one can conclude that the stability of xex_{e} is fully determined by the subsystem v˙=Jfcl(xe)v\dot{v}=J_{f_{cl}}(x_{e})v.

The corresponding Lyapunov equation for Jcl(xe)J_{cl}(x_{e}) is

Y\displaystyle Y =Jcl(xe)𝖳X+XJcl(xe)\displaystyle=J_{cl}(x_{e})^{\mathsf{T}}X+XJ_{cl}(x_{e}) (36)

with X=U𝒜ΛaU𝒜𝖳+WΛwW𝖳X=U_{\mathcal{A}}\Lambda_{a}U_{\mathcal{A}}^{\mathsf{T}}+W\Lambda_{w}W^{\mathsf{T}}, where Λa,Λw>0\Lambda_{a},\Lambda_{w}>0 are diagonal matrices and the column space of WW is the orthogonal complement of {ha1(xe),,har(xe)}\{\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e})\} with an inner product induced by 𝒫VG>0\mathcal{P}_{V}G>0, that is, U𝒜𝖳𝒫VGW=0U_{\mathcal{A}}^{\mathsf{T}}\mathcal{P}_{V}GW=0. This means that X>0X>0. Using (18) and Proposition 1, notice that XJfcl(xe)v=WΛwW𝖳(BV𝖳)1DBV𝖳Jf𝒜vXJ_{f_{cl}}(x_{e})v=W\Lambda_{w}W^{\mathsf{T}}(B_{V}^{\mathsf{T}})^{-1}DB_{V}^{\mathsf{T}}J_{f_{\mathcal{A}}}v, where again vv is an arbitrary vector in {ha1(xe),,har(xe)}\{\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e})\}^{\perp}. Define the Lyapunov candidate V(v)=v𝖳Xv>0V(v)=v^{\mathsf{T}}Xv>0. Taking its time derivative and using the dynamics of vv yields

V˙\displaystyle\dot{V} =v𝖳(Jfcl𝖳X+XJfcl)v\displaystyle=v^{\mathsf{T}}\left(J_{f_{cl}}^{\mathsf{T}}X+XJ_{f_{cl}}\right)v
=2v𝖳WΛwW𝖳(BV𝖳)1DBV𝖳Jf𝒜v\displaystyle=2v^{\mathsf{T}}W\Lambda_{w}W^{\mathsf{T}}(B_{V}^{\mathsf{T}})^{-1}DB_{V}^{\mathsf{T}}J_{f_{\mathcal{A}}}v (37)

Since (BV𝖳)1DBV𝖳(B_{V}^{\mathsf{T}})^{-1}DB_{V}^{\mathsf{T}} is similar to D>0D>0, its eigenvalues are p1c1>0p^{-1}c^{-1}>0 and ones. Since {ha1(xe),,har(xe)}\{\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e})\}^{\perp} is contained in the column space of WΛwW𝖳W\Lambda_{w}W^{\mathsf{T}}, given any v{ha1(xe),,har(xe)}v\in\{\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e})\}^{\perp}, it is always possible to choose WΛwW𝖳0W\Lambda_{w}W^{\mathsf{T}}\geq 0 such that vv is one of its eigenvectors with an associated positive eigenvalue σ>0\sigma>0. Therefore, with this choice for Λw>0\Lambda_{w}>0 and WW, (37) becomes V˙=2σv𝖳Jf𝒜v\dot{V}=2\sigma v^{\mathsf{T}}J_{f_{\mathcal{A}}}v. Hence, by Chetaev’s instability theorem, if there exists v{ha1(xe),,har(xe)}v\in\{\nabla h_{a_{1}}(x_{e}),\cdots,\nabla h_{a_{r}}(x_{e})\}^{\perp} such that v𝖳Jf𝒜v>0v^{\mathsf{T}}J_{f_{\mathcal{A}}}v>0 holds, V˙>0\dot{V}>0 and xex_{e} is an unstable equilibrium point. Otherwise, V˙0\dot{V}\leq 0 and xex_{e} is stable. ∎

Theorems 2 and 3 show that the existence conditions and stability properties of boundary equilibrium points are completely determined by the vector field f𝒜f_{\mathcal{A}} and its state derivatives as defined in (10). For both frameworks for safety-critical control considered in this work, the stability of boundary equilibrium points depends on the state derivatives of the system dynamics ff and gg and on the Hessians of the active CBFs, Hha1,,HharH_{h_{a_{1}}},\cdots,H_{h_{a_{r}}}. Particularly, for the safety filter QP, it also depends on the Jacobian of the nominal controller Junom(x)J_{u_{nom}}(x), and for the CLF-CBF QP, it also depends on the Hessian matrix of the CLF, HVH_{V}.

Refer to caption
Figure 1: Example of a asymptotically stable equilibrium point occurring at the intersection of two quadratic CBF boundaries.

In Fig. 1, we show the simulation result of a safety-critical control task with the CLF-CBF QP-based controller. 111 The code repository used for producing these results is publicly available at https://github.com/CaipirUltron/CompatibleCLFCBF/tree/mydevel.. Here, unom=0u_{nom}=0 and H(x)=I3H(x)=I_{3} in (3), and the proposed system in 3\mathbb{R}^{3} is x˙1=u12u3\dot{x}_{1}=u_{1}-2u_{3}, x˙2=u2\dot{x}_{2}=u_{2}, x˙3=2u1+u3\dot{x}_{3}=-2u_{1}+u_{3}. The CBFs are two quadratic functions h1h_{1} and h2h_{2} centered at the points (1,0,3)(-1,0,3) and (1,0,3)(1,0,3), respectively. Their boundaries are the red ellipsoids in Fig. 1, with the union of their interiors constituting the unsafe set which the system trajectories should avoid. The CLF VV is also a quadratic centered on the origin, and its level set is shown in blue, at an asymptotically stable equilibrium point xex_{e} at the top of the boundary intersection 𝒞𝒜=𝒞1𝒞2\partial\mathcal{C}_{\mathcal{A}}=\partial\mathcal{C}_{1}\cap\partial\mathcal{C}_{2}, shown in Fig. 1 by the dashed black circle. A trajectory converging to xex_{e} is shown, and the normalized gradients h1(xe)\nabla h_{1}(x_{e}), h2(xe)\nabla h_{2}(x_{e}) are shown as the two red vectors pointing up at the equilibrium point. At xex_{e}, since f𝒜(xe,λe)=0f_{\mathcal{A}}(x_{e},\lambda_{e})=0 for some λe=[λe1λe2]𝖳02\lambda_{e}=\begin{bmatrix}\,\lambda_{e_{1}}\!&\!\lambda_{e_{2}}\,\end{bmatrix}^{\mathsf{T}}\in\mathbb{R}^{2}_{\geq 0}, f(xe)=pγ(V(xe))V(xe)G(xe)(λe1h1(xe)+λe2h2(xe))f(x_{e})=p\gamma(V(x_{e}))\nabla V(x_{e})-G(x_{e})(\lambda_{e_{1}}\nabla h_{1}(x_{e})+\lambda_{e_{2}}\nabla h_{2}(x_{e})). Since the system is driftless, f(x)=0xf(x)=0\,\,\forall x. Furthermore, G(xe)G(x_{e}) is a positive definite matrix. Then, the following holds: pγ(V(xe))V(xe)=λe1h1+λe2h2p\gamma(V(x_{e}))\nabla V(x_{e})=\lambda_{e_{1}}\nabla h_{1}+\lambda_{e_{2}}\nabla h_{2}, meaning that the gradient of the CLF is a conical combination of the gradients of the active CBFs at xex_{e}. That is precisely the case at Fig. 1. Furthermore, due to the fact that the system is driftless with a full rank G(xe)G(x_{e}), carrying out the needed simplifications at v𝖳Jf𝒜(xe,λe)vv^{\mathsf{T}}J_{f_{\mathcal{A}}}(x_{e},\lambda_{e})v from (33), one can conclude that the stability of xex_{e} is determined by λe1Hh1+λe2Hh2pγ(V(xe))HV\lambda_{e_{1}}H_{h_{1}}\!+\!\lambda_{e_{2}}H_{h_{2}}\!-\!p\gamma(V(x_{e}))H_{V}, that is, essentially by a difference of curvatures between the CBFs and the CLF at xex_{e}, extending the result in [8].

References

  • [1] Aaron D Ames, Jessy W Grizzle, and Paulo Tabuada. Control barrier function based quadratic programs with application to adaptive cruise control. In 53rd IEEE Conference on Decision and Control, pages 6271–6278. IEEE, 2014.
  • [2] David ”davidad” Dalrymple et al. Towards guaranteed safe ai: A framework for ensuring robust and reliable ai systems, 2024.
  • [3] Kai-Chieh Hsu, Haimin Hu, and Jaime F. Fisac. The safety filter: A unified view of safety-critical control in autonomous systems. Annual Review of Control, Robotics, and Autonomous Systems, 7(Volume 7, 2024):47–72, 2024.
  • [4] Hassan K Khalil. Nonlinear systems; 3rd ed. Prentice-Hall, Upper Saddle River, NJ, 2002.
  • [5] Taekyung Kim, Robin Inho Kee, and Dimitra Panagou. Learning to refine input constrained control barrier functions via uncertainty-aware online parameter adaptation, 2025.
  • [6] G. Notomista, S. F. Ruf, and M. Egerstedt. Persistification of robotic tasks using control barrier functions. IEEE Robotics and Automation Letters, 3(2):758–763, April 2018.
  • [7] Gennaro Notomista and Matteo Saveriano. Safety of dynamical systems with multiple non-convex unsafe sets using control barrier functions. IEEE Control Systems Letters, 6:1136–1141, 2022.
  • [8] M. F. Reis, A. P. Aguiar, and P. Tabuada. Control barrier function-based quadratic programs introduce undesirable asymptotically stable equilibria. IEEE Control Systems Letters, 5(2):731–736, 2021.
  • [9] Matheus F. Reis and A. Pedro Aguiar. On the stability of undesirable equilibria in the quadratic program framework for safety-critical control, 2024.
  • [10] Xiao Tan and Dimos V Dimarogonas. On the undesired equilibria induced by control barrier function based quadratic programs. Automatica, 159:111359, 2024.
  • [11] Kim P. Wabersich, Andrew J. Taylor, Jason J. Choi, Koushil Sreenath, Claire J. Tomlin, Aaron D. Ames, and Melanie N. Zeilinger. Data-driven safety filters: Hamilton-jacobi reachability, control barrier functions, and predictive methods for uncertain systems. IEEE Control Systems Magazine, 43(5):137–177, 2023.