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

Anytime Solvers for Variational Inequalities:
the (Recursive) Safe Monotone Flows

Ahmed Allibhoy and Jorge Cortés
Abstract

This paper synthesizes anytime algorithms, in the form of continuous-time dynamical systems, to solve monotone variational inequalities. We introduce three algorithms that solve this problem: the projected monotone flow, the safe monotone flow, and the recursive safe monotone flow. The first two systems admit dual interpretations: either as projected dynamical systems or as dynamical systems controlled with a feedback controller synthesized using techniques from safety-critical control. The third flow bypasses the need to solve quadratic programs along the trajectories by incorporating a dynamics whose equilibria precisely correspond to such solutions, and interconnecting the dynamical systems on different time scales. We perform a thorough analysis of the dynamical properties of all three systems. For the safe monotone flow, we show that equilibria correspond exactly with critical points of the original problem, and the constraint set is forward invariant and asymptotically stable. The additional assumption of convexity and monotonicity allows us to derive global stability guarantees, as well as establish the system is contracting when the constraint set is polyhedral. For the recursive safe monotone flow, we use tools from singular perturbation theory for contracting systems to show KKT points are locally exponentially stable and globally attracting, and obtain practical safety guarantees. We illustrate the performance of the flows on a two-player game example and also demonstrate the versatility for interconnection and regulation of dynamical processes of the safe monotone flow in an example of a receding horizon linear quadratic dynamic game.

1 Introduction

Variational inequalities encompass a wide range of problems arising in operations research and engineering applications, including minimizing a function, characterizing Nash equilibria of a game, and seeking saddle points of a function. This paper synthesizes continuous-time flows whose trajectories converge to the solution set of a monotone variational inequality while respecting the constraints at all times.

Our motivation for considering this problem is two-fold. First, iterative algorithms in numerical computing can be interpreted as dynamical systems. This opens the door for the use of controls and system-theoretic tools to characterize their qualitative and quantitative properties, e.g., stability of solutions, convergence rate, and robustness to disturbances. In turn, the availability of such characterization sets the stage for developing sample-data implementations and systematically designing new algorithms equipped with desired properties.

The second motivation stems from problems where the solution to the variational inequality is used to regulate a physical process modeled as a dynamically evolving plant (e.g., providing setpoints, specifying optimization-based controllers, steering the plant toward an optimal steady-state). This type of problem arises in multiple application areas, including power systems, network congestion control, and traffic networks. In these settings, the algorithm used to solve the variational inequality is interconnected with a plant, and thus the resulting coupled system naturally lends itself to system-theoretic analysis and design. We are particularly interested in situations where the problem incorporates constraints which, when violated, would threaten the safe operation of the physical system. In such cases, it is desirable that the algorithm is anytime, meaning that it is guaranteed to return a feasible point even when terminated before it has converged to a solution. The anytime property ensures that the specifications conveyed to the plant remain feasible at all times.

1.1 Related Work

The study of the interplay between continuous-time dynamical systems and variational inequalities has a rich history. In the context of optimization, classical references include  [2, 3, 4, 5]. For general variational inequalities, flows solving them can be obtained through differential inclusions involving monotone set-valued maps, originally introduced in [6]. These systems have been equivalently described as a projected dynamical systems [7] and complementarity systems [8, 9]. Recently, this framework has been extended to settings where monotonicity holds with respect to non-Euclidean norms [10], and when the system evolves on a Riemannian manifold [11]. One limitation of these systems is that they are, in general, discontinuous, which poses challenges both for their theoretical analysis and practical implementation.

The interconnection of algorithms with physical plants has attracted much attention recently. The interconnected system can often be viewed as the coupling of a dynamical system with a set valued operator (cf. [12, 13] for a comprehensive survey of systems having this form). The specific setting where the algorithm optimizes the steady-state of the plant is typically referred to as online feedback optimization[14, 15] and has been studied in continuous time in the context of power systems [16, 17], network congestion control [18], and traffic networks [19], as well as discrete time in [20]. Recently this framework has also been studied in game scenarios [21, 22].

A complementary approach uses extremum seeking control [23], which has been generalized to the setting of games in [24]. Extremum seeking differs from the methods introduced here in that they are typically zeroth-order methods, and do not offer exact stability guarantees. The recent work [25] considers safety guarantees for extremum seeking control for the special case of a single inequality constraint. In fact, the proposed algorithm can be understood as an approximation of the safe gradient flow in [26], which is a precursor for constrained optimization of the algorithms proposed here for constraint sets parameterized by multiple inequality and equalities.

To synthesize our flows, we employ techniques from safety-critical control [27, 28], which refers to the problem of designing a feedback controller that ensures that the state of the system satisfies certain constraints. The problem of ensuring safety is typically formalized by specifying a set of states where the system is said to remain safe, and ensuring the safe set is forward invariant. The work [29] reviews set invariance in control. A popular technique for synthesizing safe controllers uses the concept of control barrier functions (CBFs), see [30, 31, 32] and references therein, to specify optimization-based feedback controllers which “filter” a nominal system to ensure it remains in the safe set. Here, we apply this strategy to synthesize anytime algorithms, viewing the constraint set as a safety set and the monotone operator of the variational inequality as the nominal system. This view has connections to projected dynamical systems, whose relationship with CBF-based control design has recently been explored in [33, 34], and leads to the alternative “projection-based” interpretation of the projected monotone flow and safe monotone flow proposed here.

1.2 Statement of Contributions

We consider the synthesis of continuous-time dynamical systems solving variational inequalities while respecting the constraints at all times. We work primarily with continuous-time systems since we are motivated by online feedback optimization, where the evolution of a physical process determines the functions that define the variational inequality, which in turn regulates the dynamically evolving plant. The methods introduced here can be implemented on digital hardware by discretizing the resulting dynamics.

We discuss three flows that solve this problem. The projected monotone flow is already known, but we provide a novel control-theoretic interpretation as a control system whose closed-loop behavior is as close as possible to the monotone operator while still belonging to the tangent cone of the constraint set. The safe monotone flow can be interpreted either as a control system with a feedback controller synthesized using techniques from safety-critical control or as an approximation of the projected monotone flow. The latter interpretation relies on the novel notion of restricted tangent set, which generalizes the usual concept of tangent cone from variational geometry. We show that equilibria correspond exactly with critical points of the original problem, establish existence and uniqueness of solutions, and characterize the regularity properties of flow. We also show that the constraint set is forward invariant and asymptotically stable, and derive global stability guarantees under the additional assumption of convexity and monotonicity. Our technical analysis relies on a suite of Lyapunov functions to establish stability properties with respect to the constraint set and the whole state space. When the constraint set is polyhedral, we establish that the system is contracting and exponentially stable. Finally, the recursive safe monotone flow bypasses the need for continuously solving quadratic programs along the trajectories by incorporating a dynamics whose equilibria precisely correspond to such solutions, and interconnecting the dynamical systems on different time scales. Using tools from singular perturbation theory for contracting systems, we show that for variational inequalities with polyhedral constraints, the KKT points are locally exponentially stable and globally attracting, and obtain practical safety guarantees. We compare the three flows on a simple two-player game and demonstrate how the safe monotone flow can be interconnected with dynamical processes on an example of a receding horizon linear quadratic dynamic game.

The flows introduced here generalize the safe gradient flow, a continuous-time system proposed in our previous work [26] (in parallel, [20] introduced a discrete-time implementation of a simplified version of it). With respect to the safe gradient flow, our treatment extends the results in three key ways. First, we consider variational inequalities, rather than just constrained optimization problems, making the flows introduced here applicable to a much broader range of problems. Second, with assumptions of monotonicity and convexity of the constraints, we obtain global stability and convergence results, rather than local stability results. Third, the rigorous characterization of the contractivity properties of the safe monotone flow paves the way for its interconnection with other dynamically evolving processes. In fact, the proposed recursive safe monotone flow critically builds on this analysis by leveraging different timescales and singular perturbation theory.

2 Preliminaries

We review here basic notions from variational inequalities, projections, and set invariance. Readers familiar with these concepts can safely skip this section.

2.1 Notation

Let denote the set of real numbers. For cc\in, [c]+=max{0,c}[c]_{+}=\max\{0,c\}. For xnx\in^{n}, xix_{i} denotes the iith component and xix_{-i} denotes all components of xx excluding ii. For v,wnv,w\in^{n}, vwv\leq w (resp. v<wv<w) denotes viwiv_{i}\leq w_{i} (resp. vi<wiv_{i}<w_{i}) for i{1,,n}i\in\{1,\dots,n\}. We let v\left\lVert v\right\rVert denote the Euclidean norm. We write Q0Q\succeq 0 (resp., Q0Q\succ 0) to denote QQ is positive semidefinite (resp., QQ is positive definite). Given Q0Q\succeq 0, let xQ\left\lVert x\right\rVert_{Q} denote the seminorm where xQ=xQx\left\lVert x\right\rVert_{Q}=\sqrt{x^{\top}Qx}. For a symmetric QQ, λmin(Q)\lambda_{\text{min}}(Q) and λmax(Q)\lambda_{\textnormal{max}}{(Q)} denote the minimum and maximum eigenvalues of QQ, resp. Given 𝒞n\mathcal{C}\subset^{n}, the distance of xnx\in^{n} to 𝒞\mathcal{C} is dist(x,𝒞)=infy𝒞xy\textnormal{dist}(x,\mathcal{C})=\inf_{y\in\mathcal{C}}\left\lVert x-y\right\rVert. The projection map onto 𝒞¯\overline{\mathcal{C}} is proj𝒞:n𝒞¯\text{proj}_{\mathcal{C}}:^{n}\rightrightarrows\overline{\mathcal{C}}, where proj𝒞(x)={y𝒞¯xy=dist(x,𝒞)}\textnormal{proj}_{\mathcal{C}}\left(x\right)=\big{\{}y\in\overline{\mathcal{C}}\mid\left\lVert x-y\right\rVert=\textnormal{dist}(x,\mathcal{C})\big{\}}. Given a closed and convex set 𝒞n\mathcal{C}\subset^{n}, the normal cone to 𝒞\mathcal{C} at xnx\in^{n} is N𝒞(x)={dnd(xx)0,x𝒞}N_{\mathcal{C}}(x)=\{d\in^{n}\mid d^{\top}(x^{\prime}-x)\leq 0,\;\forall x^{\prime}\in\mathcal{C}\} and the tangent cone to 𝒞\mathcal{C} at xx is T𝒞(x)={ξndξ0,dN𝒞(x)}T_{\mathcal{C}}(x)=\{\xi\in^{n}\mid d^{\top}\xi\leq 0,\;\forall d\in N_{\mathcal{C}}(x)\}.

Given g:ng:^{n}\to, we denote its gradient by g\nabla g. For g:nmg:^{n}\to^{m}, g(x)x\frac{\partial g(x)}{\partial x} denotes its Jacobian. For I{1,2,,m}I\subset\{1,2,\dots,m\}, we denote by gI(x)x\frac{\partial g_{I}(x)}{\partial x} the matrix whose rows are {gi(x)}iI\{\nabla g_{i}(x)^{\top}\}_{i\in I}. Given a function f:nf:^{n}\to, we say it is directionally differentiable if for all ξn\xi\in^{n}, the following limit exists

f(x;ξ)=limh0+f(x+hξ)f(x)h.f^{\prime}(x;\xi)=\lim_{h\to 0^{+}}\frac{f(x+h\xi)-f(x)}{h}.

Given a vector field 𝒢:nn\mathcal{G}:^{n}\to^{n} and a function V:nV:^{n}\to, the Upper-right Dini derivative of VV along 𝒢\mathcal{G} is

D𝒢+V(x)=lim suph0+1h(V(Φh(x))V(x)),D^{+}_{\mathcal{G}}V(x)=\limsup_{h\to 0^{+}}\frac{1}{h}\left(V(\Phi_{h}(x))-V(x)\right),

where Φh\Phi_{h} is the flow map111 For h0h\in_{\geq 0}, the flow map Φh:nn\Phi_{h}:^{n}\to^{n} is defined by Φh(x0)=x(h)\Phi_{h}(x_{0})=x(h), where x(t)x(t) is the unique trajectory solving x˙=𝒢(x)\dot{x}=\mathcal{G}(x) with x(0)=x0x(0)=x_{0}. corresponding to the system x˙=𝒢(x)\dot{x}=\mathcal{G}(x). When VV is directionally differentiable D𝒢+V(x)=V(x,𝒢(x))D^{+}_{\mathcal{G}}V(x)=V^{\prime}(x,\mathcal{G}(x)) and when VV is differentiable, then D𝒢+V(x)=V(x)𝒢(x)D^{+}_{\mathcal{G}}V(x)=\nabla V(x)^{\top}\mathcal{G}(x).

2.2 Variational Inequalities

Here we review the basic theory of variational inequalities following [35]. Let F:nnF:^{n}\to^{n} be a map and 𝒞n\mathcal{C}\subset^{n} a set of constraints. A variational inequality refers to the problem of finding x𝒞x^{*}\in\mathcal{C} such that

(xx)F(x)0,x𝒞.(x-x^{*})^{\top}F(x^{*})\geq 0,\qquad\forall x\in\mathcal{C}. (1)

We use VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}) to refer to the problem (1) and SOL(F,𝒞)\textnormal{SOL}(F,\mathcal{C}) to denote its set of solutions. Variational inequalities provide a framework to study many different analysis and optimization problems, including

  • Solving the nonlinear equation F(x)=0F(x^{*})=0, which corresponds to VI(F,n)\textnormal{VI}(F,^{n});

  • Minimizing the function f:nf:^{n}\to subject to the constraint that x𝒞x\in\mathcal{C}, which corresponds to VI(f,𝒞)\textnormal{VI}(\nabla f,\mathcal{C});

  • Finding saddle points of the function :n×m\ell:^{n}\times^{m}\to subject to the constraints that x1X1x_{1}\in X_{1} and x2X2x_{2}\in X_{2}, which corresponds to VI([x1;x2],X1×X2)\textnormal{VI}([\nabla_{x_{1}}\ell;-\nabla_{x_{2}}\ell],X_{1}\times X_{2}).

  • Finding the Nash equilibria of a game with NN agents, where the iith agent wants to minimize the cost Ji(xi,xi)J_{i}(x_{i},x_{-i}) subject to the constraint xiXix_{i}\in X_{i}, which corresponds to VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}), where FF is the pseudogradient operator defined by F(x)=(x1J1(x),,xNJN(x))F(x)=(\nabla_{x_{1}}J_{1}(x),\dots,\nabla_{x_{N}}J_{N}(x)), and 𝒞=X1×X2××XN\mathcal{C}=X_{1}\times X_{2}\times\cdots\times X_{N}.

The map F:nnF:^{n}\to^{n} is monotone if (x1x2)(F(x1)F(x2))0(x_{1}-x_{2})^{\top}(F(x_{1})-F(x_{2}))\geq 0, for all x1,x2nx_{1},x_{2}\in^{n}, and FF is μ\mu-strongly monotone if there exists μ>0\mu>0 such that (x1x2)(F(x1)F(x2))μx1x22(x_{1}-x_{2})^{\top}(F(x_{1})-F(x_{2}))\geq\mu\left\lVert x_{1}-x_{2}\right\rVert^{2}, for all x1,x2nx_{1},x_{2}\in^{n}. When FF is a gradient map, i.e., F=fF=\nabla f for some function f:nf:^{n}\to, then monotonicity (resp. μ\mu-strong monotonicity) is equivalent to convexity (resp. strong convexity) of ff. When FF is monotone and 𝒞\mathcal{C} is convex, VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}) is a monotone variational inequality.

In order to provide a characterization of the solution set SOL(F,𝒞)\textnormal{SOL}(F,\mathcal{C}), we need to introduce a more explicit description of the set of constraints. Suppose that g:nmg:^{n}\to^{m} and h:nkh:^{n}\to^{k} are continuously differentiable and 𝒞\mathcal{C} is described by inequality constraints and affine equality constraints,

𝒞={xng(x)0,h(x)=Hxch=0},\mathcal{C}=\{x\in^{n}\mid g(x)\leq 0,\;h(x)=Hx-c_{h}=0\}, (2)

where Hk×nH\in^{k\times n} and chkc_{h}\in^{k}. For xnx\in^{n}, we denote the active constraint, inactive constraint, and constraint violation sets, resp., as

I0(x)\displaystyle I_{0}(x) ={i[1,m]gi(x)=0},\displaystyle=\{i\in[1,m]\mid g_{i}(x)=0\},
I(x)\displaystyle I_{-}(x) ={i[1,m]gi(x)<0},\displaystyle=\{i\in[1,m]\mid g_{i}(x)<0\},
I+(x)\displaystyle I_{+}(x) ={i[1,m]gi(x)>0}.\displaystyle=\{i\in[1,m]\mid g_{i}(x)>0\}.

The problem (1) satisfies the constant-rank condition at x𝒞x\in\mathcal{C} if there exists an open set UU containing xx such that for all II0(x)I\subset I_{0}(x), the rank of {gi(y)}iI{hj(y)}j=1k\{\nabla g_{i}(y)\}_{i\in I}\cup\{\nabla h_{j}(y)\}_{j=1}^{k} remains constant for all yUy\in U. The problem (1) satisfies the Mangasarian-Fromovitz Constraint Qualification (MFCQ) condition at x𝒞x\in\mathcal{C} if {hj(x)}i=1k\{\nabla h_{j}(x)\}_{i=1}^{k} are linearly independent, and there exists ξn\xi\in^{n} such that gi(x)ξ<0\nabla g_{i}(x)^{\top}\xi<0 for all iI0(x)i\in I_{0}(x), and hj(x)ξ=0\nabla h_{j}(x)^{\top}\xi=0 for all j{1,,k}j\in\{1,\dots,k\}. If MFCQ holds at x𝒞x^{*}\in\mathcal{C}, then, if xx^{*} satisfies (1), there exists (u,v)m×k(u^{*},v^{*})\in^{m}\times^{k} such that

F(x)+i=1muigi(x)+j=1kvjhj(x)\displaystyle F(x^{*})+\sum_{i=1}^{m}u^{*}_{i}\nabla g_{i}(x^{*})+\sum_{j=1}^{k}v_{j}^{*}\nabla h_{j}(x^{*}) =0\displaystyle=0 (3a)
g(x)\displaystyle g(x^{*}) 0\displaystyle\leq 0 (3b)
h(x)\displaystyle h(x^{*}) =0\displaystyle=0 (3c)
u\displaystyle u^{*} 0\displaystyle\geq 0 (3d)
(u)g(x)\displaystyle(u^{*})^{\top}g(x^{*}) =0.\displaystyle=0. (3e)

Equations (3) are called the KKT conditions. A point (x,u,v)(x^{*},u^{*},v^{*}) satisfying them is a KKT triple and the pair (u,v)(u^{*},v^{*}) is a Lagrange multiplier. For monotone variational inequalities, when MFCQ holds at xx^{*}, then the KKT conditions are both necessary and sufficient for xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}). When FF is monotone SOL(F,𝒞)\textnormal{SOL}(F,\mathcal{C}) is closed and convex. If FF is additionally μ\mu-strongly monotone, then the set of solutions is a singleton.

2.3 Controller Synthesis for Set Invariance

We review here notions from the theory of set invariance for control systems following [29] and discuss methods for synthesizing feedback controllers that ensure it. Consider a control-affine system

x˙\displaystyle\dot{x} =(x,ν)\displaystyle=\mathcal{F}(x,\nu) (4)
=F0(x)+i=1rνiFi(x),\displaystyle=F_{0}(x)+\sum_{i=1}^{r}\nu_{i}F_{i}(x),

with Lipschitz-continuous vector fields Fi:nnF_{i}:^{n}\to^{n}, for i{0,,r}i\in\{0,\dots,r\}, and a set 𝒰r\mathcal{U}\subset^{r} of valid control inputs ν\nu. Let 𝒞n\mathcal{C}\subset^{n} be a constraint set of the form (2) to which we want to restrict the evolution of the system. We consider the problem of designing a feedback controller κ:n𝒰\kappa:^{n}\to\mathcal{U} such that 𝒞\mathcal{C} is forward invariant with respect to the closed-loop dynamics x˙=(x,κ(x))\dot{x}=\mathcal{F}(x,\kappa(x)). In applications, 𝒞\mathcal{C} often corresponds to the set of states for which the system can operate safely. For this reason, we refer to 𝒞\mathcal{C} as the safety set, and call the system safe under a controller κ\kappa if 𝒞\mathcal{C} is forward invariant. A controller ensuring safety is safeguarding. We discuss two optimization-based strategies for synthesizing safeguarding controllers.

2.3.1 Safeguarding Control via Projection

The first strategy ensures the closed-loop dynamics lie in the tangent cone of the safety set. Whenever MFCQ holds at x𝒞x\in\mathcal{C}, the tangent cone can conveniently be expressed as, cf. [36, Theorem 6.31], T𝒞(x)={ξn|h(x)xξ=0,gI0(x)xξ0}T_{\mathcal{C}}(x)=\{\xi\in^{n}\left|\frac{\partial h(x)}{\partial x}\xi=0,\frac{\partial g_{I_{0}}(x)}{\partial x}\xi\leq 0\right.\}. We then define the set-valued map Kproj:n𝒰K_{\text{proj}}:^{n}\rightrightarrows\mathcal{U} which characterizes the set of inputs that ensure the system remains inside the safety set. The set has the form,

Kproj(x)={μ𝒰|\displaystyle K_{\text{proj}}(x)\!=\!\Big{\{}\mu\in\mathcal{U}\;\Big{|}\; DF0+gi(x)+=1rμDF+gi(x)0,iI(x),\displaystyle D^{+}_{F_{0}}g_{i}(x)\!+\!\sum_{\ell=1}^{r}\mu_{\ell}D^{+}_{F_{\ell}}g_{i}(x)\leq 0,\,i\in I(x),
DF0+hj(x)+=1rμDF+hj(x)=0,j=1,,k}.\displaystyle D^{+}_{F_{0}}h_{j}(x)+\sum_{\ell=1}^{r}\mu_{\ell}D^{+}_{F_{\ell}}h_{j}(x)=0,\,j=1,\dots,k\Big{\}}.

As we in the following result, the feedback k:𝒞𝒰k:\mathcal{C}\to\mathcal{U} such that k(x)Kproj(x)k(x)\in K_{\text{proj}}(x) for x𝒞x\in\mathcal{C} renders 𝒞\mathcal{C} forward invariant. The proof is omitted for brevity, though it follows directly from Nagumo’s Theorem [37, Theorem 2].

Lemma 2.1 (Projection-based Safeguarding Feedback).

Consider the system (4) with safety set 𝒞\mathcal{C} and suppose that Kproj(x)K_{\text{proj}}(x)\neq\emptyset for all x𝒞x\in\mathcal{C}. Then, the feedback controller κ:𝒞𝒰\kappa:\mathcal{C}\to\mathcal{U} is safeguarding if κ(x)Kproj(x)\kappa(x)\in K_{\text{proj}}(x) for all x𝒞x\in\mathcal{C} and the closed-loop system x˙=(x,κ(x))\dot{x}=\mathcal{F}(x,\kappa(x)) admits a unique solution for all initial conditions.

To synthesize a safeguarding controller, we propose a strategy where κ(x)\kappa(x) at each x𝒞x\in\mathcal{C} is expressed as the solution to a mathematical program. Because Kproj(x)K_{\text{proj}}(x) is defined in terms of affine constraints on the control input ν\nu, we can readily express a feedback satisfying the hypotheses of Lemma 2.1 in the form of a mathematical program,

κ(x)argminνKproj(x)J(x,ν),\kappa(x)\in\underset{\nu\in K_{\text{proj}}(x)}{\text{argmin}}J(x,\nu), (5)

for an appropriate choice of cost function J:𝒞×𝒰J:\mathcal{C}\times\mathcal{U}\to. In general, care must be taken to ensure that the set KprojK_{\text{proj}} is nonempty and that the controller κ\kappa in (5) satisfies appropriate regularity conditions to ensure existence and uniqueness for solutions of the resulting closed-loop dynamics. Even if these properties hold, the approach has several limitations: the controller is ill-defined for initial conditions lying outside the safety set and the closed-loop system in general is nonsmooth.

2.3.2 Safeguarding Control via Control Barrier Functions

The second strategy for synthesizing safeguarding controllers addresses the limitations of projection-based methods. The approach relies on the notion of a vector control barrier functions [27, 26]. Given a set X𝒞X\supset\mathcal{C} and set of valid control inputs 𝒰m\mathcal{U}\subset^{m}, we say the pair (g,h):n×km(g,h):^{n}\times^{k}\to^{m} is a (m,k)(m,k)-vector control barrier function (VCBF) for 𝒞\mathcal{C} on XX relative to 𝒰\mathcal{U} if there exists α>0\alpha>0 such that the map Kcbf,α:n𝒰K_{\textnormal{cbf},\alpha}:^{n}\rightrightarrows\mathcal{U} given by

Kcbf,α(x)={μ𝒰|\displaystyle K_{\textnormal{cbf},\alpha}(x)\!=\!\Big{\{}\mu\in\mathcal{U}\;\Big{|}\; DF0+gi(x)+=1rμDF+gi(x)+αgi(x)0,\displaystyle D^{+}_{F_{0}}g_{i}(x)\!+\!\sum_{\ell=1}^{r}\mu_{\ell}D^{+}_{F_{\ell}}g_{i}(x)\!+\!\alpha g_{i}(x)\leq 0,
DF0+hj(x)+=1rμDF+hj(x)+αhj(x)=0,\displaystyle D^{+}_{F_{0}}h_{j}(x)+\sum_{\ell=1}^{r}\mu_{\ell}D^{+}_{F_{\ell}}h_{j}(x)+\alpha h_{j}(x)=0,
1im, 1jk},\displaystyle 1\leq i\leq m,\;1\leq j\leq k\Big{\}},

takes nonempty values for all xXx\in X. Similar to the previous strategy, the set Kcbf,αK_{\textnormal{cbf},\alpha} characterizes the set of inputs which ensure that the state remains inside the safe set. Unlike the previous strategy, this assurance is implemented gradually: the parameter α\alpha corresponds to how tolerant we are of increasing the value of inactive constraints, with small values of α\alpha corresponding to more aggressive bounds on the growth rate. To see this, note that if gi(x)<0g_{i}(x)<0 then the left-hand sides of the constraints parameterizing Kcbf,αK_{\textnormal{cbf},\alpha} decrease as α\alpha increases, making the constraint easier to satisfy. For α=\alpha=\infty, if gi(x)<0g_{i}(x)<0, then the left-hand side of the corresponding constraint becomes -\infty, thus the only nontrivial constraints correspond to ii where gi(x)=0g_{i}(x)=0, and for x𝒞x\in\mathcal{C} the set corresponds to KprojK_{\text{proj}}.

When m=1m=1 and k=0k=0, an (m,k)(m,k)-vector control barrier function is equivalent to the usual notion of a control barrier function [27, Definition 2]. The generalization provided by VCBFs allows us to consider a broader class of safety sets.

Lemma 2.2 (VCBF-based Safeguarding Feedback).

Consider the system (4) with safety set 𝒞\mathcal{C} and suppose (g,h)(g,h) is a vector control barrier function for 𝒞\mathcal{C} on XX relative to 𝒰\mathcal{U} and MFCQ holds on 𝒞\mathcal{C}. Then, the feedback controller κ:X𝒰\kappa:X\to\mathcal{U} is safeguarding and ensures asymptotic stability of 𝒞\mathcal{C} on XX if κ(x)Kcbf,α(x)\kappa(x)\in K_{\textnormal{cbf},\alpha}(x) for all xXx\in X and the closed-loop system x˙=(x,κ(x))\dot{x}=\mathcal{F}(x,\kappa(x)) admits a unique solution for all initial conditions.

The proof of the Lemma 2.2 is omitted for space, but the result follows by Nagumo’s Theorem [37, Theorem 2] since one can show that the closed-loop dynamics are contained in the tangent cone. To synthesize a safeguarding feedback controller, one can pursue a design using a similar approach to Section 2.3.1. Given a cost function J:X×𝒰J:X\times\mathcal{U}\to, we let κ(x)\kappa(x) solve the following mathematical program:

κ(x)argminνKcbf,α(x)J(x,ν).\kappa(x)\in\underset{\nu\in K_{\textnormal{cbf},\alpha}(x)}{\text{argmin}}J(x,\nu). (6)

Similarly to the case of projection-based safeguarding feedback control, care must be taken to verify the existence and uniqueness of solutions to the closed-loop system, as well as to handle situations where (6) does not have unique solutions. If these properties hold, then the control design addresses the challenges of projection-based methods. In particular, we can ensure that a controller of the form (6) is well-defined outside the safety set and results in closed-loop system with continuous solutions, under mild conditions which we discuss in the following sections.

3 Problem Formulation

Consider a variational inequality VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}) defined by a continuously differentiable map F:nnF:^{n}\to^{n} and a convex set 𝒞\mathcal{C} of the form (2), where g:nmg:^{n}\to^{m} is continuously differentiable. We assume throughout that SOL(F,𝒞)\textnormal{SOL}(F,\mathcal{C})\neq\emptyset. Our goal is to synthesize a solver for this problem in the form of a continuous-time dynamical system.

Problem 1 (Anytime solver of variational inequality).

Design a dynamical system, x˙=𝒢(x)\dot{x}=\mathcal{G}(x), which is well defined on a set XX containing 𝒞\mathcal{C} such that

  1. (i)

    Trajectories of the system converge to SOL(F,𝒞)\textnormal{SOL}(F,\mathcal{C});

  2. (ii)

    𝒞\mathcal{C} is forward invariant;

  3. (iii)

    𝒞\mathcal{C} is asymptotically stable as a set.

Item (i) ensures that the dynamical system can be viewed as a solver of the problem (1): solutions can be obtained by simulating system trajectories and taking the limit as tt\to\infty of x(t)x(t). Furthermore, the solver can be implemented on computational platforms by discretizing the dynamics. Item (ii) ensures that this solver is anytime, meaning that even if terminated early, it is guaranteed to return a feasible solution provided the initial condition is feasible. Item (iii) accounts for infeasible initial conditions, and ensures asymptotic safety. Both the expression of the algorithm in the form of a continuous-time dynamical system and the anytime property are particularly useful for real-time applications, where the algorithm might be interconnected with other physical processes – e.g., when the algorithm output is used to regulate a physical plant and constraints of the optimization problem ensure the safe operation of the plant.

In the following, we introduce three dynamics to solve Problem 1. The first, synthesized using the technique outlined in Section 2.3.1, is the projected monotone flow. This dynamics is already well-known but we reinterpret it here through the lens of control theory. The next two, synthesized using the technique outlined in Section 2.3.2, are the safe monotone flow and the recursive safe monotone flow. Both dynamics are entirely novel.

4 Projected Monotone Flow

Here, we discuss our first solution to Problem 1, in the form of the projected monotone flow. We show that the system can be viewed in two equivalent ways: either as a control system with a feedback controller designed using the strategy outlined in Section 2.3.1, or as a projected dynamical system. In fact, this system admits many other equivalent descriptions, for example in terms of monotone differential inclusions, or complementarity systems [8, 9, 37] (see [13] for a modern survey on the relationships between systems in various formalisms). Its properties have been extensively studied [7] and techniques for computational implementation have been considered in [38]. However, we focus here on the control-based and projection-based forms. In the coming sections we describe in detail the derivation of each implementation, show their equivalence, and discuss the properties of the resulting flow regarding safety and stability.

4.1 Control-Based Implementation

Our design strategy originates from the observation that, when FF is monotone, the system x˙=F(x)\dot{x}=-F(x) finds solutions to the unconstrained variational inequality VI(F,n)\textnormal{VI}(F,^{n}). However, trajectories flowing along this dynamics might leave the constraint set 𝒞\mathcal{C}. This leads us to consider the control-affine system:

x˙\displaystyle\dot{x} =(x,u,v)\displaystyle=\mathcal{F}(x,u,v) (7)
=F(x)i=1muigi(x)j=1kvjhj(x).\displaystyle=-F(x)-\sum_{i=1}^{m}u_{i}\nabla g_{i}(x)-\sum_{j=1}^{k}v_{j}\nabla h_{j}(x).

Here, we have augmented the system with inputs from the admissible set 𝒰=0m×k\mathcal{U}=^{m}_{\geq 0}\times^{k} to modify the flow of the original drift F-F to account for the constraints in a way that ensures that the solutions to (7) stay inside of or approach 𝒞\mathcal{C}. The idea is that if the constraint gi(x)0g_{i}(x)\leq 0 is in danger of being violated, the corresponding input uiu_{i} can be increased to ensure trajectories continue to satisfy it. Likewise, the input vjv_{j} can be increased or decreased to ensure the corresponding constraint hj(x)=0h_{j}(x)=0 is satisfied along trajectories.

Our design proceeds by thinking of 𝒞\mathcal{C} as a safety set for the system and using the approach outlined in Section 2.3.1 to synthesize a safeguarding feedback controller (u,v)=κ(x)(u,v)=\kappa(x). Assuming that MFCQ holds for all x𝒞x\in\mathcal{C}, the map Kproj:n0m×kK_{\text{proj}}:^{n}\rightrightarrows^{m}_{\geq 0}\times^{k} takes the form

Kproj(x)={(u,v)0m×k|\displaystyle K_{\text{proj}}(x)=\Big{\{}(u,v)\in^{m}_{\geq 0}\times^{k}\;\Big{|} gI0xF(x)gI0xgxugI0xhxv0,\displaystyle-\frac{\partial g_{I_{0}}}{\partial x}F(x)-\frac{\partial g_{I_{0}}}{\partial x}\frac{\partial g}{\partial x}^{\top}u-\frac{\partial g_{I_{0}}}{\partial x}\frac{\partial h}{\partial x}^{\top}v\leq 0, (8)
hxF(x)hxgxuhxhxv=0}.\displaystyle-\frac{\partial h}{\partial x}F(x)-\frac{\partial h}{\partial x}\frac{\partial g}{\partial x}^{\top}u-\frac{\partial h}{\partial x}\frac{\partial h}{\partial x}^{\top}v=0\Big{\}}.

The following result states that the set of admissible controls is nonempty. We omit its proof for space reasons, but note that it readily follows from Farka’s Lemma [39].

Lemma 4.1.

(Projection onto Tangent Cone is Feasible) If x𝒞x\in\mathcal{C} and MFCQ holds at xx, then Kproj(x)K_{\text{proj}}(x)\neq\emptyset.

We then use the feedback controller

κ(x)argmin(u,v)Kproj(x)J(x,u,v),\kappa(x)\in\underset{(u,v)\in K_{\text{proj}}(x)}{\text{argmin}}J(x,u,v), (9)

where we set the objective function to be

J(x,u,v)=12i=1muigi(x)+j=1kvjhj(x)2.J(x,u,v)=\frac{1}{2}\bigg{\|}\sum_{i=1}^{m}u_{i}\nabla g_{i}(x)+\sum_{j=1}^{k}v_{j}\nabla h_{j}(x)\bigg{\|}^{2}. (10)

This function measures the magnitude of the “modification” of the drift term in (7). Thus, the QP-based controller (9) has the interpretation, at each xx, of finding the control input such that the closed-loop system dynamics are as close as possible to F(x)-F(x), while still being in T𝒞(x)T_{\mathcal{C}}(x). In general, the program given by (9) does not have unique solutions. Despite this, we show below that the closed-loop dynamics of (7) is well defined regardless of which solution to (5) is chosen. We refer to it as the projected monotone flow.

4.2 Projection-Based Implementation

The second implementation of the projected monotone flow consists of projecting F(x)-F(x) onto the tangent cone of the constraint set. In general, the tangent cone does not have a representation that allows us to compute the projection easily. However, when the appropriate constraint qualification condition holds, the tangent cone admits a convenient parameterization which allows for the projection to be implemented as a quadratic program. Let x𝒞x\in\mathcal{C} and suppose that MFCQ holds at xx. It follows that the tangent cone can be parameterized222 The right hand side of (11) is called the linearization cone of 𝒞\mathcal{C}, and denoted as T𝒞lin(x)T^{\text{lin}}_{\mathcal{C}}(x). In general, T𝒞lin(x)T𝒞(x)T^{\text{lin}}_{\mathcal{C}}(x)\subset T_{\mathcal{C}}(x), and we say that Abadie’s Constraint Qualification (ACQ) [40] holds at x𝒞x\in\mathcal{C} if the reverse inclusion holds, i.e., T𝒞lin(x)=T𝒞(x)T^{\text{lin}}_{\mathcal{C}}(x)=T_{\mathcal{C}}(x). MFCQ is a stronger condition that implies ACQ. Many of the results we state in the paper can be generalized to hold under the weaker assumption of ACQ, however for simplicity, we avoid a detailed technical discussion of constraint qualifications. as

T𝒞(x)={ξn|h(x)xξ=0,gI0(x)xξ0}.\displaystyle T_{\mathcal{C}}(x)=\bigg{\{}\xi\in^{n}\;\Big{|}\;\frac{\partial h(x)}{\partial x}\xi=0,\frac{\partial g_{I_{0}}(x)}{\partial x}\xi\leq 0\bigg{\}}. (11)

The projection-based implementation of the projected monotone flow takes then the following form:

x˙\displaystyle\dot{x} =projT𝒞(x)(F(x))\displaystyle=\textnormal{proj}_{T_{\mathcal{C}}(x)}\left(-F(x)\right) (12)
=argminξn12ξ+F(x)2\displaystyle=\underset{\xi\in^{n}}{\text{argmin}}\quad\frac{1}{2}\left\lVert\xi+F(x)\right\rVert^{2}
subject togI0(x)xξ0,h(x)xξ=0.\displaystyle\quad\;\text{subject to}\quad\frac{\partial g_{I_{0}}(x)}{\partial x}\xi\leq 0,\frac{\partial h(x)}{\partial x}\xi=0.

The projection onto the tangent ensures that 𝒞\mathcal{C} is forward invariant, by Nagumo’s Theorem [29, Theorem 3.1].

4.3 Properties of Projected Monotone Flow

Here, we lay out the properties of the projected monotone flow. We begin by establishing the equivalence between the control- and projection-based implementations. We then discuss existence and uniqueness of solutions, and finally the stability and safety properties of the dynamics.

4.3.1 Equivalence of Control-Based and Projection-Based Implementations

Equivalence follows directly from the properties of the tangent cone, as we show next.

Proposition 4.2.

(Equivalence of Control-Based and Projected-Based Implementations) Assume MFCQ holds at x𝒞x\in\mathcal{C} and let (u,v)(u,v) be any solution to (9). Then, (x,u,v)=projT𝒞(x)(F(x))\mathcal{F}(x,u,v)=\textnormal{proj}_{T_{\mathcal{C}}(x)}\left(-F(x)\right).

Proof.

Let (u,v)(u,v) be any solution to (9) and ξ=projT𝒞(x)(F(x))\xi=\textnormal{proj}_{T_{\mathcal{C}}(x)}\left(-F(x)\right). Then (x,u,v)T𝒞(x)\mathcal{F}(x,u,v)\in T_{\mathcal{C}}(x), so it follows immediately by optimality of ξ\xi that

ξ+F(x)2(x,u,v)+F(x)2.\left\lVert\xi+F(x)\right\rVert^{2}\leq\left\lVert\mathcal{F}(x,u,v)+F(x)\right\rVert^{2}.

Next, because ξ\xi is given by a projection, there exists wN𝒞(x)w\in N_{\mathcal{C}}(x) such that ξ+F(x)+w=0\xi+F(x)+w=0, see e.g., [8, Corollary 2]. If MFCQ holds at x𝒞x\in\mathcal{C}, by [36, Theorem 6.14], there exists (u¯,v¯)(\bar{u},\bar{v}) such that ww can be written as

w=i=1mu¯igi(x)+j=1kv¯jhj(x),u¯0,u¯g(x)=0.\displaystyle w=\sum_{i=1}^{m}\bar{u}_{i}\nabla g_{i}(x)+\sum_{j=1}^{k}\bar{v}_{j}\nabla h_{j}(x),\qquad\bar{u}\geq 0,\quad\bar{u}^{\top}g(x)=0.

Combining this expression with the fact that ξ=F(x)wT𝒞(x)\xi=-F(x)-w\in T_{\mathcal{C}}(x) and using the parameterization of the tangent cone in (11), we deduce that (u¯,v¯)Kproj(x)(\bar{u},\bar{v})\in K_{\text{proj}}(x). By optimality of (u,v)(u,v),

ξ+F(x)2\displaystyle\left\lVert\xi+F(x)\right\rVert^{2} =i=1mu¯igi(x)+j=1kv¯jhj(x)2\displaystyle=\bigg{\|}\sum_{i=1}^{m}\bar{u}_{i}\nabla g_{i}(x)+\sum_{j=1}^{k}\bar{v}_{j}\nabla h_{j}(x)\bigg{\|}^{2}
i=1muigi(x)+j=1kvjhj(x)2\displaystyle\geq\bigg{\|}\sum_{i=1}^{m}u_{i}\nabla g_{i}(x)+\sum_{j=1}^{k}v_{j}\nabla h_{j}(x)\bigg{\|}^{2}
=(x,u,v)+F(x)2.\displaystyle=\left\lVert\mathcal{F}(x,u,v)+F(x)\right\rVert^{2}.

But since the projection onto the tangent cone must be unique, we conclude ξ=(x,u,v)\xi=\mathcal{F}(x,u,v). ∎

The value of Proposition 4.2 stems from showing that safety-critical control can be used to design algorithms that solve variational inequalities. Though the control strategy pursued in Section 4.1 results in a known flow, this sets up the basis for employing other design strategies from safety-critical control to yield novel methods, as we show later.

4.3.2 Existence and Uniqueness of Solutions

The projected monotone flow is discontinuous, and hence one must consider notions of solutions beyond the classical ones, see e.g., [41]. Here, we consider Carathéodory solutions, which are absolutely continuous functions that satisfy (12) almost everywhere. The existence and uniqueness of solutions for all initial conditions follows readily from [37, Chapter 3.2, Theorem 1(i)] or [13, Theorem 5.8].

4.3.3 Safety and Stability of Projected Monotone Flow

We now show that the projected monotone flow is safe, meaning that the constraint set 𝒞\mathcal{C} is forward invariant, and the solution set SOL(F,𝒞)\textnormal{SOL}(F,\mathcal{C}) is stable. Forward invariance of 𝒞\mathcal{C} follows directly from Nagumo’s Theorem. The equilibria of the projected monotone flow correspond to solutions to VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}). Finally, stability of a solution xx^{*} can be certified using the Lyapunov function

V(x)=12xx2,V(x)=\frac{1}{2}\left\lVert x-x^{*}\right\rVert^{2},

as a consequence of [37, Chapter 3.2, Theorem 1(ii)] or [42, Theorems 1 and 2]. These properties are summarized next.

Theorem 4.3.

(Safety and Stability Properties of Projected Monotone Flow) Let 𝒞\mathcal{C} be convex and suppose MFCQ holds everywhere on 𝒞\mathcal{C}. The following hold for the projected monotone flow:

  1. (i)

    𝒞\mathcal{C} is forward invariant;

  2. (ii)

    xx^{*} is an equilibrium of the projected monotone flow if and only if xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C});

  3. (iii)

    If xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) and FF is monotone, then xx^{*} is globally Lyapunov stable relative to 𝒞\mathcal{C};

  4. (iv)

    If FF is μ\mu-strongly monotone, then the projected monotone flow is contracting at rate μ\mu. In particular, the unique solution xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) is globally exponentially stable relative to 𝒞\mathcal{C}.

5 Safe Monotone Flow

In this section, we discuss a second solution to Problem 1, which results in an entirely novel flow, termed safe monotone flow. Similar to the projected monotone flow, this system admits two equivalent implementations: either as a control-system with a safeguarding feedback controller or as a projected dynamical system. Because of its continuity properties, the system can be solved using standard ODE discretization schemes, cf. [26, Remark 5.6], but we leave the formal analysis of this for future work.

5.1 Control-Based Implementation

We start with the control system (7) with the admissible control set 𝒰=0m×k\mathcal{U}=^{m}_{\geq 0}\times^{k}, viewing 𝒞\mathcal{C} as a safety set, and design a safeguarding controller. We synthesize this controller using the function (g,h)(g,h) as a VCBF, following the approach outlined in Section 2.3.2.

Letting α>0\alpha>0 be a parameter, the set of control inputs ensuring safety is given by

Kcbf,α(x)={(u,v)0m×k|\displaystyle K_{\textnormal{cbf},\alpha}(x)=\Big{\{}(u,v)\in_{\geq 0}^{m}\times^{k}\;\Big{|} gxF(x)gxgxugxhxvαg(x)\displaystyle-\frac{\partial g}{\partial x}F(x)-\frac{\partial g}{\partial x}\frac{\partial g}{\partial x}^{\top}u-\frac{\partial g}{\partial x}\frac{\partial h}{\partial x}^{\top}v\leq-\alpha g(x)
hxF(x)hxgxuhxhxv=αh(x)}.\displaystyle-\frac{\partial h}{\partial x}F(x)-\frac{\partial h}{\partial x}\frac{\partial g}{\partial x}^{\top}u-\frac{\partial h}{\partial x}\frac{\partial h}{\partial x}^{\top}v=-\alpha h(x)\Big{\}}.

The next result shows that this set is nonempty on an open set containing 𝒞\mathcal{C}.

Lemma 5.1 (Vector Control Barrier Function for (7)).

Assume MFCQ holds for all x𝒞x\in\mathcal{C}. Then there exists an open set X𝒞X\supset\mathcal{C} on which ϕ=(g,h)\phi=(g,h) is a vector-control barrier function of (7) for 𝒞\mathcal{C}, on XX, relative to ×k0m{}^{m}_{\geq 0}\times^{k}.

The proof of this result is identical to [26, Lemma 4.1] and we omit it for brevity. By Lemma 5.1, the feedback controller (u,v)=κ(x)(u,v)=\kappa(x) where

κ(x)argmin(u,v)Kcbf,α(x)J(x,u,v),\kappa(x)\in\underset{(u,v)\in K_{\textnormal{cbf},\alpha}(x)}{\text{argmin}}J(x,u,v), (13)

and JJ is given by (10), is well defined on XX. This controller has the same interpretation as before: determining the control input belonging to Kcbf,α(x)K_{\textnormal{cbf},\alpha}(x) such that the closed-loop system dynamics are as close as possible to F(x)-F(x). Similar to the case with projection-based methods, the problem (6) does not necessarily have unique solutions. However, we show below that the closed-loop system is well-defined regardless of which solution is chosen. We refer to it as the safe monotone flow with safety parameter α\alpha, denoted x˙=𝒢α(x)\dot{x}=\mathcal{G}_{\alpha}(x).

5.2 Projection-Based Implementation

Here we describe the implementation of the safe monotone flow as a projected dynamical system. Similar to the projected monotone flow, the projected system is obtained by projecting F(x)-F(x) onto a set-valued map. However, because the projection onto the tangent cone is in general discontinuous as a function of the state, we replace the tangent cone with the α\alpha-restricted tangent set (here α>0\alpha>0), denoted T𝒞(α)T_{\mathcal{C}}^{(\alpha)}, defined as

T𝒞(α)(x)={ξn|g(x)xξαg(x),h(x)xξ=αh(x)}.\displaystyle T^{(\alpha)}_{\mathcal{C}}(x)=\Big{\{}\xi\in^{n}\;\Big{|}\;\frac{\partial g(x)}{\partial x}\xi\leq-\alpha g(x),\frac{\partial h(x)}{\partial x}\xi=-\alpha h(x)\Big{\}}. (14)

Figure 1 illustrates this definition. This set can be interpreted as an approximation of the usual tangent cone, but differs in several key ways. First, the restricted tangent set is not a cone, meaning that vectors in T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x) cannot be scaled arbitrarily: in certain direction, the magnitude of vectors in T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x) is restricted. An important property of T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x) is that, even though the tangent cone is undefined for x𝒞x\not\in\mathcal{C}, this is not the case for the restricted tangent set. In fact, it can be shown that T𝒞(α)T_{\mathcal{C}}^{(\alpha)} takes nonempty values on an open set containing 𝒞\mathcal{C}. This property allows for the safe monotone flow to be well-defined for infeasible initial conditions. The next result states properties of the α\alpha-restricted tangent set.

Refer to caption
Refer to caption
Figure 1: Illustration of the notion of tangent cone, and α\alpha-restricted tangent set. The gray-shaded region represents the set 𝒞\mathcal{C}. The colored regions depict either type of set, which consists of vectors centered various points xix_{i}. The dashed border indicates directions in which the magnitude of vectors in the set are unbounded. (a) The α\alpha-restricted tangent set. Note that the set is well-defined at x2𝒞x_{2}\not\in\mathcal{C}, however because the region does not overlap with the point x2x_{2}, the set T𝒞(α)(x2)T^{(\alpha)}_{\mathcal{C}}(x_{2}) does not contain any zero vectors, and all vectors point strictly toward the feasible set. (b) The tangent cone. Note that the tangent cone is not well defined at points outside 𝒞\mathcal{C}.
Proposition 5.2.

(Properties of α\alpha-Restricted Tangent Set) Assume MFCQ holds for all x𝒞x\in\mathcal{C}. The set-valued map T𝒞(α):nnT_{\mathcal{C}}^{(\alpha)}:^{n}\rightrightarrows^{n} satisfies:

  1. (i)

    T𝒞(α)(x)T_{\mathcal{C}}^{(\alpha)}(x) is convex for all xnx\in^{n};

  2. (ii)

    For any fixed x𝒞x\in\mathcal{C}, the set T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x) satisfies MFCQ at all ξT𝒞(α)(x)\xi\in T^{(\alpha)}_{\mathcal{C}}(x).

  3. (iii)

    There exists an open set XX containing 𝒞\mathcal{C} such that T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x)\neq\emptyset for all xXx\in X;

  4. (iv)

    If x𝒞x\in\mathcal{C}, then T𝒞(α)T𝒞(x)T_{\mathcal{C}}^{(\alpha)}\subset T_{\mathcal{C}}(x).

Proof.

We observe that (i) follows from the fact that the constraints characterizing TC(α)(x)T_{C}^{(\alpha)}(x) are affine in the variable ξ\xi. We prove (ii) using the same strategy as [26, Lemma 4.5], which we sketch here. If MFCQ holds at x𝒞x\in\mathcal{C}, then the inequalities defining (14) satisfy Slater’s condition [43, Chapter 5.2.3] at xx and therefore MFCQ holds for all ξT𝒞(α)(x)\xi\in T^{(\alpha)}_{\mathcal{C}}(x). To show (iii), we note that Slater’s condition implies that the affine constraints parameterizing T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x) are regular [44, Theorem 2], meaning that the system remains feasible with respect to perturbations. Since T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x) is nonempty for all x𝒞x\in\mathcal{C}, it follows that there exists an open set XX containing 𝒞\mathcal{C} such that T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x) is nonempty for all xXx\in X. Finally, (iv) follows from the definition of the tangent cone. ∎

Using the α\alpha-restricted tangent set, we can define the projected dynamical system

x˙\displaystyle\dot{x} =projT𝒞(α)(x)(F(x))\displaystyle=\textnormal{proj}_{T^{(\alpha)}_{\mathcal{C}}(x)}\left(-F(x)\right) (15)
=argminξn12ξ+F(x)2\displaystyle=\underset{\xi\in^{n}}{\text{argmin}}\quad\frac{1}{2}\left\lVert\xi+F(x)\right\rVert^{2}
subject tog(x)xξαg(x)\displaystyle\quad\text{subject to}\quad\frac{\partial g(x)}{\partial x}\xi\leq-\alpha g(x)
h(x)xξ=αh(x).\displaystyle\quad\phantom{subjectto}\quad\frac{\partial h(x)}{\partial x}\xi=-\alpha h(x).

Similar to the projected monotone flow, the projection operation ensures that the trajectories of the system remain in the safety set. However, as we show next, the advantages of projecting onto the restricted tangent cone is that the system is well defined for infeasible initial conditions, and trajectories of the system are smooth.

5.3 Properties of Safe Monotone Flow

We now discuss the properties of the safe monotone flow. We begin by establishing the equivalence of the control-based and projection-based implementations. Next, we discuss its stability and safety properties.

5.3.1 Equivalence of Control-Based and Projection-Based Implementations

We establish here that the control-based and projection-based implementations of the safe monotone flow are equivalent. The next result states that the closed-loop dynamics resulting from the implementation of (6) over (7) is equivalent to the projection onto T𝒞(α)(x)T^{(\alpha)}_{\mathcal{C}}(x). The structure of the proof mirrors that of Proposition 4.2.

Proposition 5.3.

(Equivalence of Control-Based and Projection-Based Implementations) Assume MFCQ holds for everywhere on 𝒞\mathcal{C} and let XnX\subset^{n} be an open set containing 𝒞\mathcal{C} on which Kcbf,αK_{\textnormal{cbf},\alpha} takes nonempty values. Let (u,v)(u,v) be any solution to (13) at xXx\in X (note that 𝒢α(x)=(x,u,v)\mathcal{G}_{\alpha}(x)=\mathcal{F}(x,u,v)). Then, 𝒢α(x)=projT𝒞(α)(x)(F(x))\mathcal{G}_{\alpha}(x)=\textnormal{proj}_{T^{(\alpha)}_{\mathcal{C}}(x)}\left(-F(x)\right).

Proof.

Let (u,v)(u,v) be any solution to (13) and ξ=projT𝒞(α)(x)(F(x))\xi=\textnormal{proj}_{T^{(\alpha)}_{\mathcal{C}}(x)}\left(-F(x)\right). Then (x,u,v)T𝒞(α)(x)\mathcal{F}(x,u,v)\in T^{(\alpha)}_{\mathcal{C}}(x), so it follows immediately by optimality of ξ\xi that

ξ+F(x)2(x,u,v)+F(x)2.\left\lVert\xi+F(x)\right\rVert^{2}\leq\left\lVert\mathcal{F}(x,u,v)+F(x)\right\rVert^{2}.

Next, because ξ\xi is given by a projection, there exists wNT(ξ)w\in N_{T}(\xi), where T=T𝒞(α)(x)T=T^{(\alpha)}_{\mathcal{C}}(x) such that ξ+F(x)+w=0\xi+F(x)+w=0, see e.g., [8, Corollary 2], and where

w=i=1mu¯igi(x)+j=1kv¯jhj(x),u¯0,u¯(g(x)+αg(x))=0.\displaystyle w=\sum_{i=1}^{m}\bar{u}_{i}\nabla g_{i}(x)+\sum_{j=1}^{k}\bar{v}_{j}\nabla h_{j}(x),\quad\bar{u}\geq 0,\qquad\bar{u}^{\top}(\nabla g(x)^{\top}+\alpha g(x))=0.

Combining this expression with the fact that ξ=F(x)wT𝒞(α)(x)\xi=-F(x)-w\in T^{(\alpha)}_{\mathcal{C}}(x) and using the definition of the α\alpha-restricted tangent cone, we deduce that (u¯,v¯)Kcbf,α(x)(\bar{u},\bar{v})\in K_{\textnormal{cbf},\alpha}(x). By optimality of (u,v)(u,v), we have

ξ+F(x)2\displaystyle\left\lVert\xi+F(x)\right\rVert^{2} =i=1mu¯igi(x)+j=1kv¯jhj(x)2\displaystyle=\bigg{\|}\sum_{i=1}^{m}\bar{u}_{i}\nabla g_{i}(x)+\sum_{j=1}^{k}\bar{v}_{j}\nabla h_{j}(x)\bigg{\|}^{2}
i=1muigi(x)+j=1kvjhj(x)2\displaystyle\geq\bigg{\|}\sum_{i=1}^{m}u_{i}\nabla g_{i}(x)+\sum_{j=1}^{k}v_{j}\nabla h_{j}(x)\bigg{\|}^{2}
=(x,u,v)+F(x)2.\displaystyle=\left\lVert\mathcal{F}(x,u,v)+F(x)\right\rVert^{2}.

But since the projection onto the α\alpha-restricted tangent set is unique, we conclude ξ=(x,u,v)\xi=\mathcal{F}(x,u,v). ∎

5.3.2 Existence and Uniqueness of Solutions

We now discuss conditions for the existence and uniqueness of solutions of the safe monotone flow.

Proposition 5.4.

(Existence and Uniqueness of Solutions to Safe Monotone Flow) Assume MFCQ and the constant-rank condition hold on 𝒞\mathcal{C} for all x𝒞x\in\mathcal{C} and let XX be the open set containing 𝒞\mathcal{C} in Proposition 5.2(iii). Then

  1. (i)

    For all x0𝒞x_{0}\in\mathcal{C}, there exists a unique solution x:0nx:_{\geq 0}\to^{n} to the safe monotone flow with x(0)=x0x(0)=x_{0}.

  2. (ii)

    For all x0Xx_{0}\in X, there exists a unique solution x:[0,tf]nx:[0,t_{f}]\to^{n} such that x(0)=x0x(0)=x_{0}. Furthermore, the solution can be extended so that either tf=t_{f}=\infty or x(t)Xx(t)\to\partial X as ttft\to t_{f}.

Proof.

We first note that the program (15) satisfies the General Strong Second-Order Sufficient Condition (cf. [45]) and Slater’s condition at xXx\in X. Because the objective function and constraints of (15) are twice continuously differentiable, we can apply [45, Theorem 3.6] to conclude that 𝒢α\mathcal{G}_{\alpha} is locally Lipschitz at xx. Therefore, 𝒢α\mathcal{G}_{\alpha} is also lower semicontinuous and by [37, Chapter 2, Theorem 1] there exists for all x0Xx_{0}\in X a solution x:[0,tf]nx:[0,t_{f}]\to^{n} for some tf>0t_{f}>0 with x(0)=x0x(0)=x_{0}. Furthermore, either tf=t_{f}=\infty or x(t)Xx(t)\to\partial X as ttft\to t_{f}. Uniqueness of solutions holds by local Lipschitznes and (ii) follows.

To show (i), we note that 𝒢α(x)T𝒞(x)\mathcal{G}_{\alpha}(x)\in T_{\mathcal{C}}(x), and by [29, Theorem 3.1], for any solution with x(0)𝒞x(0)\in\mathcal{C}, we have that x(t)𝒞x(t)\in\mathcal{C} for all t0t\geq 0 on the interval on which the solution exists. Since 𝒞int(X)\mathcal{C}\subset\text{int}(X), solutions beginning in 𝒞\mathcal{C} cannot approach X\partial X, and exist for all time. ∎

5.3.3 Safety of Safe Monotone Flow

Here we establish the safety properties of the safe monotone flow. We begin by characterizing optimality conditions for the closed-loop dynamics.

Lemma 5.5.

(Optimality Conditions for Closed-loop Dynamics) For xnx\in^{n}, consider the equations

ξ+F(x)+g(x)xu+h(x)xv\displaystyle\xi+F(x)+\frac{\partial g(x)}{\partial x}^{\top}u+\frac{\partial h(x)}{\partial x}^{\top}v =0,\displaystyle=0, (16a)
g(x)xξ+αg(x)\displaystyle\frac{\partial g(x)}{\partial x}\xi+\alpha g(x) 0,\displaystyle\leq 0, (16b)
h(x)xξ+αh(x)\displaystyle\frac{\partial h(x)}{\partial x}\xi+\alpha h(x) =0,\displaystyle=0, (16c)
u\displaystyle u 0,\displaystyle\geq 0, (16d)
u(g(x)xξ+αg(x))\displaystyle u^{\top}\bigg{(}\frac{\partial g(x)}{\partial x}\xi+\alpha g(x)\bigg{)} =0,\displaystyle=0, (16e)

in (ξ,u,v)(\xi,u,v). Let Λα:n0m×k\Lambda_{\alpha}:^{n}\rightrightarrows^{m}_{\geq 0}\times^{k} be

Λα(x)={(u,v)ξ such that (ξ,u,v) solves (16)}.\Lambda_{\alpha}(x)=\{(u,v)\mid\exists\xi\text{ such that }(\xi,u,v)\text{ solves~{}\eqref{eq:KKT_proj}}\}.

Assume MFCQ holds everywhere on 𝒞\mathcal{C}. Then, there exists an open set X𝒞X\supset\mathcal{C} such that, if xXx\in X, then Λα(x)\Lambda_{\alpha}(x)\neq\emptyset. If (ξ,u,v)(\xi,u,v) solves (16), then 𝒢α(x)=ξ\mathcal{G}_{\alpha}(x)=\xi and (u,v)(u,v) solves (13).

Proof.

Let F~(x,ξ)=F(x)+ξ\tilde{F}(x,\xi)=F(x)+\xi. Then ξ=projT𝒞(α)(x)(F(x))\xi=\textnormal{proj}_{T^{(\alpha)}_{\mathcal{C}}(x)}\left(-F(x)\right) is a solution to the monotone variational inequality VI(F~(x,),T𝒞(α)(x))\textnormal{VI}(\tilde{F}(x,\cdot),T^{(\alpha)}_{\mathcal{C}}(x)), parameterized by xx. Since MFCQ holds at all ξT𝒞(α)(x)\xi\in T^{(\alpha)}_{\mathcal{C}}(x) by Proposition 5.2(iii), we can use the KKT conditions to characterize 𝒢α(x)\mathcal{G}_{\alpha}(x), which correspond to (16). Further, by Proposition 5.2(iv), solutions to (16) exist on an open set XX containing 𝒞\mathcal{C}. Since F~\tilde{F} is strongly monotone with respect to ξ\xi, the solution to VI(F~(x,),T𝒞(α)(x))\textnormal{VI}(\tilde{F}(x,\cdot),T^{(\alpha)}_{\mathcal{C}}(x)) is unique, proving the result. ∎

We rely on the optimality conditions in Lemma 5.5 to establish the following result characterizing the equilibria and safety properties of the safe monotone flow.

Theorem 5.6.

(Equilibria and Safety of Safe Monotone Flow) Let α>0\alpha>0, 𝒞\mathcal{C} be convex, and suppose MFCQ and the constant rank condition holds everywhere on 𝒞\mathcal{C}. The following hold for the safe monotone flow:

  1. (i)

    𝒞\mathcal{C} is forward invariant and asymptotically stable on XX;

  2. (ii)

    xx^{*} is an equilibrium if and only if xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C});

Proof.

To show (i), note that by Proposition 5.3, for all xXx\in X there exists (u(x),v(x))Kcbf,α(x)(u(x),v(x))\in K_{\textnormal{cbf},\alpha}(x) such that 𝒢α(x)=(x,u(x),v(x))\mathcal{G}_{\alpha}(x)=\mathcal{F}(x,u(x),v(x)). Given the existence and uniqueness of solutions of the closed-loop system, cf. Propositions 5.4, the result follows from Lemma 2.2 since ϕ(x)=(g(x),h(x))\phi(x)=(g(x),h(x)) is a VCBF. Statement (ii) follows from the observation that, if 𝒢α(x)=0\mathcal{G}_{\alpha}(x^{*})=0, by Lemma 5.5, there exists (u,v)(u^{*},v^{*}) such that (0,u,v)(0,u^{*},v^{*}) solves (16), which holds if and only if (x,u,v)(x^{*},u^{*},v^{*}) solves (3). ∎

5.3.4 Stability of Safe Monotone Flow

Here we characterize the stability properties of the safe monotone flow. We begin by establishing conditions for stability relative to 𝒞\mathcal{C}.

Theorem 5.7.

(Stability of Safe Monotone Flow Relative to 𝒞\mathcal{C}) Assume MFCQ holds everywhere on 𝒞\mathcal{C}. Then

  1. (i)

    If xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) and FF is monotone, then xx^{*} is globally Lyapunov stable relative to 𝒞\mathcal{C};

  2. (ii)

    If xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) and FF is μ\mu-strongly monotone, then xx^{*} is globally asymptotically stable relative to 𝒞\mathcal{C}.

Before proving Theorem 5.7, we provide several intermediate results. Our strategy relies on fixing xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) and considering the candidate Lyapunov function

V(x)=12xx2V~(x)1α2infξT𝒞(α)(x){ξF(x)+12ξ2}W(x).\displaystyle V(x)\!=\!\underbrace{\frac{1}{2}\left\lVert x-x^{*}\right\rVert^{2}}_{\tilde{V}(x)}\!-\frac{1}{\alpha^{2}}\!\underbrace{\inf_{\xi\in T^{(\alpha)}_{\mathcal{C}}(x)}\!\{\xi^{\top}F(x)\!+\!\frac{1}{2}\left\lVert\xi\right\rVert^{2}\!\}}_{W(x)}\!. (17)

We first present bounds on the Dini derivative of V~\tilde{V} along 𝒢α\mathcal{G}_{\alpha}. The result is proven in Appendix A.

Lemma 5.8 (Dini Derivative of V~\tilde{V}).

Assume MFCQ holds everywhere on 𝒞\mathcal{C}. For xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}), let (u,v)(u^{*},v^{*}) be Lagrange multipliers corresponding to xx^{*}. For xXx\in X and (u,v)Λα(x)(u,v)\in\Lambda_{\alpha}(x), then

D𝒢α+V~(x)μxx2(uu)(g(x)g(x))(vv)h(x),\displaystyle D^{+}_{\mathcal{G}_{\alpha}}\tilde{V}(x)\leq-\mu\left\lVert x-x^{*}\right\rVert^{2}-(u-u^{*})^{\top}(g(x)-g(x^{*}))-(v-v^{*})^{\top}h(x),

if FF is μ\mu-strongly monotone (the inequality holds with μ=0\mu=0 if F is monotone instead).

We now move on to characterizing properties of WW. The following result is proven in Appendix B.

Lemma 5.9 (Properties of WW).

Assume MFCQ holds everywhere on 𝒞\mathcal{C}. Define the matrix-valued function Q:X×0mn×nQ:X\times^{m}_{\geq 0}\to^{n\times n} by

Q(x,u)=12(F(x)x+F(x)x)+i=1mui2gi(x).Q(x,u)=\frac{1}{2}\bigg{(}\frac{\partial F(x)}{\partial x}+\frac{\partial F(x)}{\partial x}^{\top}\bigg{)}+\sum_{i=1}^{m}u_{i}\nabla^{2}g_{i}(x).

Then, for all xXx\in X and (u,v)Λα(x)(u,v)\in\Lambda_{\alpha}(x),

W(x)=12𝒢α(x)2+αug(x)+αvh(x)W(x)=-\frac{1}{2}\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}+\alpha u^{\top}g(x)+\alpha v^{\top}h(x) (18)

and

D𝒢α+W(x)\displaystyle D^{+}_{\mathcal{G}_{\alpha}}W(x) 𝒢α(x)Q(x,u)2α2ug(x)α2vh(x).\displaystyle\geq\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}_{Q(x,u)}-\alpha^{2}u^{\top}g(x)-\alpha^{2}v^{\top}h(x). (19)

We are now ready to prove Theorem 5.7.

Proof of Theorem 5.7.

Let xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) and suppose the hypotheses of (i) hold. Consider the function V:XV:X\to defined by (17). We show that VV is a (strict) Lyapunov function when FF is (μ\mu-strongly) monotone. Let x𝒞x\in\mathcal{C} and (u,v)Λα(x)(u,v)\in\Lambda_{\alpha}(x). Then, using (16d),

αug(x)+vh(x)0,\alpha u^{\top}g(x)+v^{\top}h(x)\leq 0,

so by examining the expression in (18) we see that W(x)0W(x)\leq 0 for all x𝒞x\in\mathcal{C} with equality if and only if xSOL(F,𝒞)x\in\textnormal{SOL}(F,\mathcal{C}). Thus VV is positive definite with respect to xx^{*}. Next,

D𝒢α+V(x)=D𝒢α+V~(x)1α2D𝒢α+W(x),D^{+}_{\mathcal{G}_{\alpha}}V(x)=D^{+}_{\mathcal{G}_{\alpha}}\tilde{V}(x)-\frac{1}{\alpha^{2}}D^{+}_{\mathcal{G}_{\alpha}}W(x),

and by Lemmas 5.8 and 5.9,

D𝒢α+V(x)1α2𝒢α(x)Q(x,u)2+(u)g(x)+(v)h(x)+ug(x)+vh(x).\displaystyle D^{+}_{\mathcal{G}_{\alpha}}V(x)\leq-\frac{1}{\alpha^{2}}\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}_{Q(x,u)}+(u^{*})^{\top}g(x)+(v^{*})^{\top}h(x)+u^{\top}g(x^{*})+v^{\top}h(x^{*}).

Since u0u\geq 0 and x𝒞x^{*}\in\mathcal{C}, we have g(x)0g(x^{*})\leq 0 and h(x)=0h(x^{*})=0, and therefore

ug(x)+vh(x)0.u^{\top}g(x^{*})+v^{\top}h(x^{*})\leq 0.

Similarly, since u0u^{*}\geq 0, and x𝒞x\in\mathcal{C}, we have g(x)0g(x)\leq 0 and h(x)=0h(x)=0, and therefore

(u)g(x)+(v)h(x)0.(u^{*})^{\top}g(x)+(v^{*})^{\top}h(x)\leq 0.

Finally, since FF is monotone and gg is convex, it follows that Q(x,u)Q(x,u) is positive semi-definite, and therefore D𝒢α+V(x)0D^{+}_{\mathcal{G}_{\alpha}}V(x)\leq 0. To show (ii), we can use the same reasoning above to show that D𝒢α+V(x)μxx2D^{+}_{\mathcal{G}_{\alpha}}V(x)\leq-\mu\left\lVert x-x^{*}\right\rVert^{2}. ∎

Next, we discuss stability with respect to the entire state space, which ensures the safe monotone flow can be used to solve VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}) even for infeasible initial conditions.

Theorem 5.10.

(Stability of Safe Monotone Flow with Respect to n) Assume MFCQ and the constant-rank condition holds on 𝒞\mathcal{C}. Then

  1. (i)

    If xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) and FF is monotone, then xx^{*} is globally Lyapunov stable;

  2. (ii)

    If xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) and FF is μ\mu-strongly monotone, then xx^{*} is globally asymptotically stable.

To prove Theorem 5.10, we can no longer rely on the Lyapunov function VV defined in (17) because it is no longer positive definite and may take negative values for x𝒞x\not\in\mathcal{C}. Instead, we consider the new candidate Lyapunov function

Vϵ(x)=V~(x)+[1α2W(x)]++δϵ(x)V_{\epsilon}(x)=\tilde{V}(x)+\bigg{[}-\frac{1}{\alpha^{2}}W(x)\bigg{]}_{+}+\delta_{\epsilon}(x) (20)

where ϵ>0\epsilon>0 and δϵ\delta_{\epsilon} is the penalty function given by

δϵ(x)=1ϵi=1m[gi(x)]++1ϵj=1k|hj(x)|.\delta_{\epsilon}(x)=\frac{1}{\epsilon}\sum_{i=1}^{m}[g_{i}(x)]_{+}+\frac{1}{\epsilon}\sum_{j=1}^{k}|h_{j}(x)|.

Before proceeding to the proof of Theorem 5.10, we provide a bound for the Dini derivative of δϵ\delta_{\epsilon} along 𝒢α\mathcal{G}_{\alpha}, which is proven in Appendix C.

Lemma 5.11 (Dini Derivative of δϵ\delta_{\epsilon}).

For all xXx\in X and ξn\xi\in^{n}, δϵ\delta_{\epsilon} is directionally differentiable at xx along the direction ξ\xi. In particular,

D𝒢α+δϵ(x)αϵiI+(x)gi(x)αϵjIh(x)|hj(x)|,D^{+}_{\mathcal{G}_{\alpha}}\delta_{\epsilon}(x)\leq-\frac{\alpha}{\epsilon}\sum_{i\in I_{+}(x)}g_{i}(x)-\frac{\alpha}{\epsilon}\sum_{j\in I_{h}(x)}\left|h_{j}(x)\right|, (21)

where Ih(x)={j[1,k]hj(x)0}I_{h}(x)=\{j\in[1,k]\mid h_{j}(x)\neq 0\}.

We are now ready to prove Theorem 5.10.

Proof of Theorem 5.10.

We begin by showing (i). Let xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}). Note that, from the optimality conditions (16), Λα(x)\Lambda_{\alpha}(x^{*}) corresponds to the set of Lagrange multipliers of the solution xx^{*} to VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}). Because MFCQ holds at xx^{*}, it follows that Λα(x)\Lambda_{\alpha}(x^{*}) is bounded. Thus, it is possible to choose ϵ>0\epsilon>0 small enough so that

αϵ>sup(u,v)Λα(x){[uv]}.\frac{\alpha}{\epsilon}>\sup_{(u^{*},v^{*})\in\Lambda_{\alpha}(x^{*})}\bigg{\{}\left\lVert\begin{bmatrix}u^{*}\\ v^{*}\end{bmatrix}\right\rVert_{\infty}\bigg{\}}. (22)

Next, we observe that it follows immediately from the definition (20) that VϵV_{\epsilon} is positive definite with respect to xx^{*}. Finally, we compute D𝒢α+Vϵ(x)D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x) and show that it is negative semidefinite. Let xXx\in X. We consider three cases: W(x)<0W(x)<0, W(x)>0W(x)>0, and W(x)=0W(x)=0. In the case where W(x)<0W(x)<0,

D𝒢α+Vϵ=D𝒢α+V~(x)1α2D𝒢α+W(x)+D𝒢α+δϵ(x).D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}=D^{+}_{\mathcal{G}_{\alpha}}\tilde{V}(x)-\frac{1}{\alpha^{2}}D^{+}_{\mathcal{G}_{\alpha}}W(x)+D^{+}_{\mathcal{G}_{\alpha}}\delta_{\epsilon}(x).

Combining the bounds in Lemmas 5.85.9, and 5.11,

D𝒢α+Vϵ(x)1α2𝒢α(x)Q(x,u)2+iI+(x)(uiαϵ)gi(x)+jIh(x)(vjαϵ)|hj(x)|.\displaystyle D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x)\leq-\frac{1}{\alpha^{2}}\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}_{Q(x,u)}+\sum_{i\in I_{+}(x)}\left(u_{i}^{*}-\frac{\alpha}{\epsilon}\right)g_{i}(x)+\sum_{j\in I_{h}(x)}\left(v_{j}^{*}-\frac{\alpha}{\epsilon}\right)\left|h_{j}(x)\right|. (23)

Since ϵ\epsilon satisfies (22), it follows that D𝒢α+Vϵ(x)0D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x)\leq 0.

For the case where W(x)>0W(x)>0, we rearrange (18) to write

ug(x)+vh(x)=1αW(x)+12α𝒢(x)2>12α𝒢(x)2.u^{\top}g(x)+v^{\top}h(x)=\frac{1}{\alpha}W(x)+\frac{1}{2\alpha}\left\lVert\mathcal{G}(x)\right\rVert^{2}>\frac{1}{2\alpha}\left\lVert\mathcal{G}(x)\right\rVert^{2}.

Then, we have

D𝒢α+Vϵ(x)\displaystyle D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x) =D𝒢α+V~(x)+D𝒢α+δϵ(x)\displaystyle=D^{+}_{\mathcal{G}_{\alpha}}\tilde{V}(x)+D^{+}_{\mathcal{G}_{\alpha}}\delta_{\epsilon}(x) (24)
(uu)g(x)(vv)h(x)αϵiI+(x)gi(x)αϵjIh(x)|hj(x)|\displaystyle\leq-(u-u^{*})^{\top}g(x)-(v-v^{*})^{\top}h(x)-\frac{\alpha}{\epsilon}\sum_{i\in I_{+}(x)}g_{i}(x)-\frac{\alpha}{\epsilon}\sum_{j\in I_{h}(x)}\left|h_{j}(x)\right|
12α𝒢α(x)2+iI+(x)(uiαϵ)gi(x)+jIh(x)(vjαϵ)|hj(x)|,\displaystyle\leq-\frac{1}{2\alpha}\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}+\sum_{i\in I_{+}(x)}\left(u_{i}^{*}-\frac{\alpha}{\epsilon}\right)g_{i}(x)+\sum_{j\in I_{h}(x)}\left(v_{j}^{*}-\frac{\alpha}{\epsilon}\right)\left|h_{j}(x)\right|,

and D𝒢α+Vϵ(x)0D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x)\leq 0. In the case where W(x)=0W(x)=0,

D𝒢α+Vϵ(x)=D𝒢α+V~(x)+1α2[D𝒢α+W(x)]++D𝒢α+δϵ(x),D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x)=D^{+}_{\mathcal{G}_{\alpha}}\tilde{V}(x)+\frac{1}{\alpha^{2}}[-D^{+}_{\mathcal{G}_{\alpha}}W(x)]_{+}+D^{+}_{\mathcal{G}_{\alpha}}\delta_{\epsilon}(x),

which leads us to two subcases: (a) D𝒢α+W(x)<0D^{+}_{\mathcal{G}_{\alpha}}W(x)<0 and (b) D𝒢α+W(x)0D^{+}_{\mathcal{G}_{\alpha}}W(x)\geq 0. In subcase (a), D𝒢α+Vϵ(x)D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x) satisfies the bound in (23) and, therefore, D𝒢α+Vϵ(x)0D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x)\leq 0. In subcase (b),

ug(x)+vh(x)=12α𝒢α(x)2u^{\top}g(x)+v^{\top}h(x)=\frac{1}{2\alpha}\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}

and therefore D𝒢α+Vϵ(x)D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x) satisfies the bound in (24), so D𝒢α+Vϵ(x)0D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x)\leq 0. To prove (ii), we can use the same arguments above to show in each case D𝒢α+Vϵ(x)μxx2D^{+}_{\mathcal{G}_{\alpha}}V_{\epsilon}(x)\leq-\mu\left\lVert x-x^{*}\right\rVert^{2}. ∎

We conclude this section by discussing the contraction properties of the safe monotone flow. Contraction refers to the property that any two trajectories of the system approach each other exponentially (cf. [46, 47] for a precise definition), and implies exponential stability of an equilibrium. We show that, for sufficiently large α\alpha, the safe monotone flow system is contracting provided FF is globally Lipschitz and the constraint set 𝒞\mathcal{C} is polyhedral. Our analysis relies relies on the following result, which follows as a direct consequence of [48, Lemma 2.1].

Lemma 5.12 (​​[48, Lemma 2.1]).

Consider the following quadratic program

min(u,v)0m×k12[uv]Q~2+c[uv]+p,\underset{(u,v)\in^{m}_{\geq 0}\times^{k}}{\textnormal{min}}\;\frac{1}{2}\left\lVert\begin{bmatrix}u\\ v\end{bmatrix}\right\rVert_{\tilde{Q}}^{2}+c^{\top}\begin{bmatrix}u\\ v\end{bmatrix}+p, (25)

where Q~0\tilde{Q}\succeq 0. Then (u,v)(u^{*},v^{*}) solves (25) if and only if it is a solution to the linear program

min(u,v)0m×k(Q~[uv]+c)[uv].\underset{(u,v)\in^{m}_{\geq 0}\times^{k}}{\textnormal{min}}\left(\tilde{Q}\begin{bmatrix}u^{*}\\ v^{*}\end{bmatrix}+c\right)^{\top}\begin{bmatrix}u\\ v\end{bmatrix}. (26)

We now show that the safe monotone flow is contracting.

Theorem 5.13.

(Contraction and Exponential Stability of Safe Monotone Flow) Let FF be μ\mu-strongly monotone and globally Lipschitz with constant F\ell_{F} and 𝒞\mathcal{C} a polyhedral set defined by (2) with g(x)=Gxcgg(x)=Gx-c_{g} and h(x)=Hxchh(x)=Hx-c_{h}. If

α>F24μ,\alpha>\frac{\ell_{F}^{2}}{4\mu}, (27)

then the safe monotone flow is contracting with rate c=μF24αc=\mu-\frac{\ell_{F}^{2}}{4\alpha}. In particular, the unique solution xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) is globally exponentially stable.

Proof.

We claim that if the assumptions hold, then

(xy)(𝒢α(x)𝒢α(y))cxy2,(x-y)^{\top}(\mathcal{G}_{\alpha}(x)-\mathcal{G}_{\alpha}(y))\leq-c\left\lVert x-y\right\rVert^{2}, (28)

in which case the system is contracting by [46, Theorem 31], and exponential stability of xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) follows as a consequence. To show the claim, from (16a), note that

𝒢α(x)=F(x)GuxHvx.\mathcal{G}_{\alpha}(x)=-F(x)-G^{\top}u_{x}-H^{\top}v_{x}.

for any (ux,vx)Λα(x)(u_{x},v_{x})\in\Lambda_{\alpha}(x). Let then x,yXx,y\in X and (ux,vx)Λα(x)(u_{x},v_{x})\in\Lambda_{\alpha}(x) and (uy,vy)Λα(y)(u_{y},v_{y})\in\Lambda_{\alpha}(y). Then, using the strong monotonicity of FF,

(xy)(𝒢α(x)𝒢α(y))\displaystyle(x-y)^{\top}(\mathcal{G}_{\alpha}(x)-\mathcal{G}_{\alpha}(y)) =(xy)(F(x)F(y))(xy)[GH][uxuyvxvy]\displaystyle=-(x-y)^{\top}(F(x)-F(y))-(x-y)^{\top}\begin{bmatrix}G^{\top}&H^{\top}\end{bmatrix}\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix} (29)
μxy2[uxuyvxvy][G(xy)H(xy)].\displaystyle\leq-\mu\left\lVert x-y\right\rVert^{2}-\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}^{\top}\begin{bmatrix}G(x-y)\\ H(x-y)\end{bmatrix}.

Next, let J~(x;u,v)=infξnL(x;ξ,u,v)\tilde{J}(x;u,v)=-\inf_{\xi\in^{n}}L(x;\xi,u,v), where LL is the Lagrangian of (15), defined as

L(x;ξ,u,v)=ξF(x)+12ξ2+i=1mui(gi(x)ξ+αgi(x))+i=1kvi(hi(x)ξ+αhi(x)),\displaystyle L(x;\xi,u,v)=\xi^{\top}F(x)+\frac{1}{2}\left\lVert\xi\right\rVert^{2}+\sum_{i=1}^{m}u_{i}(\nabla g_{i}(x)^{\top}\xi+\alpha g_{i}(x))+\sum_{i=1}^{k}v_{i}(\nabla h_{i}(x)^{\top}\xi+\alpha h_{i}(x)), (30)

and Q~(m+k)×(m+k)\tilde{Q}\in^{(m+k)\times(m+k)} is given by

Q~=[GGGHHGHH].\tilde{Q}=\begin{bmatrix}GG^{\top}&GH^{\top}\\ HG^{\top}&HH^{\top}\end{bmatrix}. (31)

For xnx\in^{n}, LL is minimized when ξ=F(x)GuHv\xi=-F(x)-G^{\top}u-H^{\top}v, and therefore

J~(x;u,v)=12[uv]Q~2+[GF(x)α(Gxcg)HF(x)α(Hxch)][uv]+12F(x)2.\displaystyle\tilde{J}(x;u,v)=\frac{1}{2}\left\lVert\begin{bmatrix}u\\ v\end{bmatrix}\right\rVert^{2}_{\tilde{Q}}\!\!+\!\begin{bmatrix}GF(x)\!-\!\alpha(Gx-c_{g})\\ HF(x)\!-\!\alpha(Hx-c_{h})\end{bmatrix}^{\top}\!\!\begin{bmatrix}u\\ v\end{bmatrix}+\frac{1}{2}\left\lVert F(x)\right\rVert^{2}. (32)

If (ux,vx)Λα(x)(u_{x},v_{x})\in\Lambda_{\alpha}(x), then (ux,vx)(u_{x},v_{x}) is a solution to the program

min(u,v)0m×kJ~(x,u,v),\min_{(u,v)\in^{m}_{\geq 0}\times^{k}}\tilde{J}(x,u,v),

which is the Lagrangian dual333By convention, the Lagrangian dual problem is a maximization problem (cf. [43, Chapter 5]). However, the minus sign in the definition of J~\tilde{J} ensures that here it is a minimization. The reason for this sign convention is to make the notation simpler. of (15). By Lemma 5.12, (ux,vx)(u_{x},v_{x}) is also a solution to the linear program,

min(u,v)0m×k(Q~[uxvx]+[GF(x)α(Gxcg)HF(x)α(Hxch)])[uv].\underset{(u,v)\in^{m}_{\geq 0}\times^{k}}{\textnormal{min}}\left(\tilde{Q}\begin{bmatrix}u_{x}\\ v_{x}\end{bmatrix}+\begin{bmatrix}GF(x)-\alpha(Gx-c_{g})\\ HF(x)-\alpha(Hx-c_{h})\end{bmatrix}\right)^{\top}\begin{bmatrix}u\\ v\end{bmatrix}.

Since (uy,vy)(u_{y},v_{y}) is also feasible for the previous linear program, by optimality of (ux,vx)(u_{x},v_{x}),

\displaystyle- [uxuyvxvy][GxcgHxch]1α[uxvx]Q~2+1α[uyvy]Q~[uxvx]1α[uxuyvxvy][GF(x)HF(x)].\displaystyle\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}^{\top}\begin{bmatrix}Gx-c_{g}\\ Hx-c_{h}\end{bmatrix}\leq-\frac{1}{\alpha}\left\lVert\begin{bmatrix}u_{x}\\ v_{x}\end{bmatrix}\right\rVert^{2}_{\tilde{Q}}+\frac{1}{\alpha}\begin{bmatrix}u_{y}\\ v_{y}\end{bmatrix}^{\top}\tilde{Q}\begin{bmatrix}u_{x}\\ v_{x}\end{bmatrix}-\frac{1}{\alpha}\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}^{\top}\begin{bmatrix}GF(x)\\ HF(x)\end{bmatrix}.

By a similar line of reasoning,

\displaystyle- [uyuxvyvx][GycgHych]1α[uyvy]Q~2+1α[uxvx]Q~[uyvy]1α[uyuxvyvx][GF(y)HF(y)].\displaystyle\begin{bmatrix}u_{y}-u_{x}\\ v_{y}-v_{x}\end{bmatrix}^{\top}\begin{bmatrix}Gy-c_{g}\\ Hy-c_{h}\end{bmatrix}\leq-\frac{1}{\alpha}\left\lVert\begin{bmatrix}u_{y}\\ v_{y}\end{bmatrix}\right\rVert^{2}_{\tilde{Q}}+\frac{1}{\alpha}\begin{bmatrix}u_{x}\\ v_{x}\end{bmatrix}^{\top}\tilde{Q}\begin{bmatrix}u_{y}\\ v_{y}\end{bmatrix}-\frac{1}{\alpha}\begin{bmatrix}u_{y}-u_{x}\\ v_{y}-v_{x}\end{bmatrix}^{\top}\begin{bmatrix}GF(y)\\ HF(y)\end{bmatrix}.

Combining the previous two expressions, we obtain

[uxuyvxvy][G(xy)H(xy)]\displaystyle-\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}^{\top}\begin{bmatrix}G(x-y)\\ H(x-y)\end{bmatrix} 1α[uxuyvxvy][G(F(x)F(y))H(F(x)F(y))]1α[uxuyvxvy]Q~2\displaystyle\leq-\frac{1}{\alpha}\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}^{\top}\begin{bmatrix}G(F(x)-F(y))\\ H(F(x)-F(y))\end{bmatrix}-\frac{1}{\alpha}\left\lVert\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}\right\rVert^{2}_{\tilde{Q}}
Fα[GH][uxuyvxvy]xy1α[GH][uxuyvxvy]2,\displaystyle\leq\frac{\ell_{F}}{\alpha}\left\lVert\begin{bmatrix}G\\ H\end{bmatrix}^{\top}\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}\right\rVert\left\lVert x-y\right\rVert-\frac{1}{\alpha}\left\lVert\begin{bmatrix}G\\ H\end{bmatrix}^{\top}\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}\right\rVert^{2}\!\!,

where we used (u,v)Q~=M(u,v)\|(u,v)\|_{\tilde{Q}}=\left\lVert M^{\top}(u,v)\right\rVert, with M=[G;H]M=[G;H]. For any ϵ>0\epsilon>0, by Young’s Inequality [49, pp. 140] it follows that

[uxuyvxvy]\displaystyle-\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}^{\top} [G(xy)H(xy)]F2ϵαxy2(1αϵF2α)[GH][uxuyvxvy]2.\displaystyle\begin{bmatrix}G(x-y)\\ H(x-y)\end{bmatrix}\leq\frac{\ell_{F}}{2\epsilon\alpha}\left\lVert x-y\right\rVert^{2}-\Big{(}\frac{1}{\alpha}-\frac{\epsilon\ell_{F}}{2\alpha}\Big{)}\left\lVert\begin{bmatrix}G\\ H\end{bmatrix}^{\top}\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}\right\rVert^{2}\!\!.

Substituting into (29), we obtain

(xy)\displaystyle(x-y)^{\top} (𝒢α(x)𝒢α(y))(μF2ϵα)xy2(1αϵF2α)[GH][uxuyvxvy]2.\displaystyle(\mathcal{G}_{\alpha}(x)-\mathcal{G}_{\alpha}(y))\leq-\Big{(}\mu-\frac{\ell_{F}}{2\epsilon\alpha}\Big{)}\left\lVert x-y\right\rVert^{2}-\Big{(}\frac{1}{\alpha}-\frac{\epsilon\ell_{F}}{2\alpha}\Big{)}\left\lVert\begin{bmatrix}G\\ H\end{bmatrix}^{\top}\begin{bmatrix}u_{x}-u_{y}\\ v_{x}-v_{y}\end{bmatrix}\right\rVert^{2}.

Therefore, (28) holds with c=μF2ϵαc=\mu-\frac{\ell_{F}}{2\epsilon\alpha} if ϵ\epsilon satisfies

F2αμϵ2F.\frac{\ell_{F}}{2\alpha\mu}\leq\epsilon\leq\frac{2}{\ell_{F}}.

Such ϵ\epsilon can be chosen if α>F24μ\alpha>\frac{\ell_{F}^{2}}{4\mu}, which corresponds to the condition (27). The optimal estimate of the contraction rate is c=μF24αc=\mu-\frac{\ell_{F}^{2}}{4\alpha}. ∎

Remark 5.14 (Connection with Safe Gradient Flow).

The safe monotone flow is a generalization of the safe gradient flow introduced in [26]. The latter was originally studied in the context of nonconvex optimization and, similar to the case of the safe monotone flow, enjoys safety of the feasible set and correspondence between equilibria and critical points. Further, under certain constraint qualifications, the local stability of equilibria relative to the constraint set under the safe gradient flow can be established using the objective function as a Lyapunov function. Because we are working with variational inequalities, FF may not correspond to the gradient of a scalar-valued function, so the Lyapunov functions used in [26] are not directly applicable. However, the assumption of convexity and monotonicity here allows us to construct novel Lyapunov functions to obtain global stability results. The interested reader can check [26, Section VI] for a comparison of the safe gradient flow with other continuous-time methods for optimization. \bullet

6 Recursive Safe Monotone Flow

A drawback of the projected and safe monotone flows is that, in order to implement them, one needs to solve either the quadratic programs (9) or (13) at each time along the trajectory of the system. As a third algorithmic solution to Problem 1, in this section we introduce the recursive safe monotone flow which gets around this limitation by incorporating a dynamics whose equilibria correspond to the solutions of the quadratic program. We begin by showing how to derive the dynamics for general constraint sets 𝒞\mathcal{C} by interconnecting two systems on multiple time scales. Next, we use the theory of singular perturbations of contracting flows to obtain stability guarantees in the case where 𝒞\mathcal{C} is polyhedral, and show that trajectories of the recursive safe monotone flow track those of the safe monotone flow. The latter property enables us to formalize a notion of “practical safety” that the recursive safe monotone flow satisfies.

6.1 Construction of the Dynamics

We discuss here the construction of the recursive safe monotone flow. The starting point for our derivation is the control-affine system (7). The safe monotone flow consists of this system with a feedback controller specified by the quadratic program (13). Rather than solving this program exactly, the approach we take is to replace it with a monotone variational inequality parameterized by the state. For fixed xXx\in X, we can solve solve this inequality, and hence obtain the feedback κ(x)\kappa(x), using the safe monotone flow corresponding to this problem. Coupling this flow with the control system (7) yields the recursive safe monotone flow. The reason for using the term “recursive” is that the system can be interpreted as coupling one instance of the safe monotone flow (corresponding to the problem (1)), with another instance of the safe monotone flow (corresponding to the problem (6)).

In this section we carry out this strategy in mathematically precise terms. We rely on the following result, which provides an alternative characterization of the CBF-QP (13).

Lemma 6.1.

(Alternative Characterization of Safe Feedback) For xnx\in^{n}, consider the optimization

minimize(u,v)0m×k\displaystyle\underset{(u,v)\in_{\geq 0}^{m}\times^{k}}{\textnormal{minimize}} 12i=1muigi(x)+j=1kvjhj(x)2\displaystyle\frac{1}{2}\bigg{\|}\sum_{i=1}^{m}u_{i}\nabla g_{i}(x)+\sum_{j=1}^{k}v_{j}\nabla h_{j}(x)\bigg{\|}^{2} (33)
+u(gxF(x)αg(x))+v(hxF(x)αh(x)).\displaystyle+u^{\top}\bigg{(}\frac{\partial g}{\partial x}F(x)-\alpha g(x)\bigg{)}+v^{\top}\bigg{(}\frac{\partial h}{\partial x}F(x)-\alpha h(x)\bigg{)}.

If (u,v)(u,v) is a solution to (33), then (u,v)(u,v) is a solution to (13).

Proof.

Note that the constraints of (33) satisfy MFCQ for all (u,v)0m×k(u,v)\in^{m}_{\geq 0}\times^{k}. Since the objective function in (33) is convex in (u,v)(u,v), one can see that necessary and sufficient conditions for optimality are given by a KKT system that, after some manipulation, takes the form

gxgxugxhxvgxF(x)αg(x)\displaystyle-\frac{\partial g}{\partial x}\frac{\partial g}{\partial x}^{\top}u-\frac{\partial g}{\partial x}\frac{\partial h}{\partial x}^{\top}v-\frac{\partial g}{\partial x}F(x)-\alpha g(x) 0\displaystyle\leq 0
hxgxuhxhxvhxF(x)αh(x)\displaystyle-\frac{\partial h}{\partial x}\frac{\partial g}{\partial x}^{\top}u-\frac{\partial h}{\partial x}\frac{\partial h}{\partial x}^{\top}v-\frac{\partial h}{\partial x}F(x)-\alpha h(x) =0\displaystyle=0
u\displaystyle u 0\displaystyle\geq 0
u(gxgxugxhxvgxF(x)αg(x))\displaystyle u^{\top}\bigg{(}-\frac{\partial g}{\partial x}\frac{\partial g}{\partial x}^{\top}u-\frac{\partial g}{\partial x}\frac{\partial h}{\partial x}^{\top}v-\frac{\partial g}{\partial x}F(x)-\alpha g(x)\bigg{)} =0.\displaystyle=0.

It follows immediately that if (u,v)(u,v) satisfies the above equations, then (u,v)Kcbf,α(x)(u,v)\in K_{\textnormal{cbf},\alpha}(x). ∎

The rationale for considering (33), rather than working with (13) directly, is that the constraints of (33) are independent of xx, which will be important for reasons we show next. Being a constrained optimization problem, (33) can be expressed in terms of a variational inequality (parameterized by xnx\in^{n}). Formally, let F~(x,u,v)\tilde{F}(x,u,v) be given by

F~(x,u,v)=[gx(x,u,v)αg(x)hx(x,u,v)αh(x)],\displaystyle\tilde{F}(x,u,v)=\begin{bmatrix}-\frac{\partial g}{\partial x}\mathcal{F}(x,u,v)-\alpha g(x)\\ -\frac{\partial h}{\partial x}\mathcal{F}(x,u,v)-\alpha h(x)\end{bmatrix}, (34)

where \mathcal{F} is given by (7) and let 𝒞~=0m×k\tilde{\mathcal{C}}=^{m}_{\geq 0}\times^{k}, which we parameterize as

𝒞~={(u,v)m×ku0}.\tilde{\mathcal{C}}=\{(u,v)\in^{m}\times^{k}\mid u\geq 0\}. (35)

The optimization problem (33) corresponds to the variational inequality VI(F~(x,,),𝒞~)\textnormal{VI}(\tilde{F}(x,\cdot,\cdot),\tilde{\mathcal{C}}).

Our next step is to write down the safe monotone flow with safety parameter β>0\beta>0 corresponding to the variational inequality VI(F~(x,),𝒞~)\textnormal{VI}(\tilde{F}(x,\cdot),\tilde{\mathcal{C}}). Note that the β\beta-restricted tangent set (14) of 𝒞~\tilde{\mathcal{C}} is T𝒞~(β)(u,v)={(ξu,ξv)m×k|ξuβu}.T^{(\beta)}_{\tilde{\mathcal{C}}}(u,v)=\{(\xi_{u},\xi_{v})\in^{m}\times^{k}\;|\;\xi_{u}\geq-\beta u\}. The projection onto T𝒞~(β)(u,v)T^{(\beta)}_{\tilde{\mathcal{C}}}(u,v) has the following closed-form solution

projT𝒞~(β)(u,v)([ab])=[max{βu,a}b].\textnormal{proj}_{T^{(\beta)}_{\tilde{\mathcal{C}}}(u,v)}\left(\begin{bmatrix}a\\ b\end{bmatrix}\right)=\begin{bmatrix}\max\{-\beta u,a\}\\ b\end{bmatrix}.

Using this expression and applying Proposition 5.3, we write the safe monotone flow corresponding to the variational inequality VI(F~(x,),𝒞~)\textnormal{VI}(\tilde{F}(x,\cdot),\tilde{\mathcal{C}}) as

u˙\displaystyle\dot{u} =max{βu,g(x)x(x,u,v)+αg(x)}\displaystyle=\max\bigg{\{}-\beta u,\frac{\partial g(x)}{\partial x}\mathcal{F}(x,u,v)+\alpha g(x)\bigg{\}} (36)
v˙\displaystyle\dot{v} =h(x)x(x,u,v)+αh(x).\displaystyle=\frac{\partial h(x)}{\partial x}\mathcal{F}(x,u,v)+\alpha h(x).

Under certain assumptions, which we formalize in the sequel, for a fixed xx, trajectories of (36) converge to solutions of the QP (13).

This discussion suggests a system solving the original variational inequality VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}) can be obtained by coupling (36) with the dynamics (7) as follows:

x˙\displaystyle\dot{x} =(x,u,v)\displaystyle=\mathcal{F}(x,u,v) (37a)
τu˙\displaystyle\tau\dot{u} =max{βu,g(x)x(x,u,v)+αg(x)}\displaystyle=\max\bigg{\{}-\beta u,\frac{\partial g(x)}{\partial x}\mathcal{F}(x,u,v)+\alpha g(x)\bigg{\}} (37b)
τv˙\displaystyle\tau\dot{v} =h(x)x(x,u,v)+αh(x).\displaystyle=\frac{\partial h(x)}{\partial x}\mathcal{F}(x,u,v)+\alpha h(x). (37c)

We refer to the system (37) as the recursive safe monotone flow. The parameter τ\tau characterizes the separation of timescales between the system (37a) and (37b)-(37c). The interpretation of the dynamics is that, when τ\tau is sufficiently small, (37b)-(37c) evolve on a much faster timescale and rapidly approach the solution set of (13). The system on the slower timescale (37a) then approximates the safe monotone flow. We formalize this analysis next.

Remark 6.2.

(Relationship to Primal-Dual Flows) This recursive safe monotone flow is closely related to primal-dual flows, which seek a solution to (3) by flowing along a monotone operator in the primal variable, and performing a gradient ascent in the dual variables:

x˙\displaystyle\dot{x} =F(x)g(x)xuh(x)xv\displaystyle=-F(x)-\frac{\partial g(x)}{\partial x}^{\top}u-\frac{\partial h(x)}{\partial x}^{\top}v (38a)
u˙\displaystyle\dot{u} =projT0m(u)(g(x))\displaystyle=\textnormal{proj}_{T_{{}^{m}_{\geq 0}}(u)}\left(g(x)\right) (38b)
v˙\displaystyle\dot{v} =h(x)\displaystyle=h(x) (38c)

This method has been well-studied classically [2, 50], and has attracted recent attention due to its suitability for distributed implementation on network problems [51, 52, 53, 54]. One drawback of (38) is that it does not ensure that the primal variable remains in the feasible set 𝒞\mathcal{C}. The recursive monotone flow can be interpreted as a primal-dual flow which overcomes this limitation, and provides practical safety guarantees, by modifying the dual-ascent term. One can show that the right hand side of (37) converges exactly to the right hand side of (38) by letting β=1ϵ2\beta=\frac{1}{\epsilon^{2}}, α=1ϵ\alpha=\frac{1}{\epsilon}, and τ=1ϵ\tau=\frac{1}{\epsilon}, and taking the limit as ϵ0\epsilon\to 0. Future work will study further connections of the recursive safe monotone flow with primal-dual flows. \bullet

6.2 Stability of Recursive Safe Monotone Flow

To prove stability of the system (37), we rely on results from contraction theory [46]. Specifically, we derive conditions on the time-scale separation τ\tau that ensures that (37) is contracting and, as a consequence, globally attractive and locally exponentially stable. Throughout the section, we assume the following assumption holds.

Assumption 1.

(Strong Monotonicity, Lipschitzness, and Polyhedral Constraints) The following holds:

  1. (i)

    FF is μ\mu-strongly monotone and F\ell_{F}-Lipschitz;

  2. (ii)

    𝒞\mathcal{C} is a polyhedral set defined by (2) with g(x)=Gxcgg(x)=Gx-c_{g} and h(x)=Hxchh(x)=Hx-c_{h}, and the matrix [G,H][G^{\top},H^{\top}]^{\top} has full rank444 This assumption is slightly stronger than the Linear Independence Contraint Qualification (LICQ) assumption. LICQ requires linear independence of the gradients of only the active constraints, whereas here we require linear independence of the gradients of all constraints..

Next, we show that it is possible to tune the parameters β\beta so the system (36) is contracting, uniformly in xx.

Lemma 6.3.

(Contractivity of (36)) Under Assumption 1, if β>14λmax(Q~)λmin(Q~),\beta>\frac{1}{4}\frac{\lambda_{\textnormal{max}}(\tilde{Q})}{\lambda_{\textnormal{min}}(\tilde{Q})}, then the system (36) is contracting with rate c¯=λmin(Q~)λmax(Q~)4β\bar{c}=\lambda_{\textnormal{min}}(\tilde{Q})-\frac{\lambda_{\textnormal{max}}(\tilde{Q})}{4\beta} uniformly in xx.

Proof.

By Assumption 1, Q~0\tilde{Q}\succ 0 so one can show F~\tilde{F} is (i) λmin(Q~)\lambda_{\textnormal{min}}(\tilde{Q})-strongly monotone in (u,v)(u,v) uniformly in xx and (ii) Q~\|\tilde{Q}\|-Lipschitz in (u,v)(u,v) uniformly in xx. By Theorem 5.13, if β>Q~24λmin(Q~)\beta>\frac{\|\tilde{Q}\|^{2}}{4\lambda_{\textnormal{min}}(\tilde{Q})}, the system (36) is uniformly contracting. The result follows by observing that Q~2=λmax(Q~)\|\tilde{Q}\|^{2}=\lambda_{\textnormal{max}}(\tilde{Q}). ∎

We now characterize the contraction and stability properties of the recursive safe monotone flow.

Theorem 6.4.

(Contractivity of Recursive Safe Monotone Flow) Assume FF is μ\mu-strongly monotone and F\ell_{F} globally Lipschitz, and α\alpha satisfies (27). Under Assumption 1 and β\beta chosen as in Lemma 6.3, then

  1. (i)

    the unique KKT triple, (x,u,v)(x^{*},u^{*},v^{*}), corresponding to VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}) is the only equilibrium of (37).

Moreover, for all ϵ>0\epsilon>0, there exists τ>0\tau^{*}>0, such that for all 0<τ<τ0<\tau<\tau^{*},

  1. (ii)

    the system (37) is contracting on the set

    𝒵ϵ={(x,u,v)\displaystyle\mathcal{Z}_{\epsilon}=\big{\{}(x,u,v)\in X×m×k(u,v)κ(x)ϵ},\displaystyle X\times^{m}\times^{k}\mid\left\lVert(u,v)-\kappa(x)\right\rVert\leq\epsilon\big{\}},

    and every solution of (37) eventually enters 𝒵ϵ\mathcal{Z}_{\epsilon} in finite time. In particular, there exists a class 𝒦\mathcal{K}\mathcal{L} function β:0×0\beta:_{\geq 0}\times_{\geq 0}\to such that for every solution (x(t),u(t),v(t))(x(t),u(t),v(t))

    (u(x(t)),v(x(t)))κ(x(t))β((u(x(0)),v(x(0)))κ(x(0)),t);\displaystyle\left\lVert\big{(}u(x(t)),v(x(t))\big{)}-\kappa(x(t))\right\rVert\leq\beta\big{(}\left\lVert\big{(}u(x(0)),v(x(0))\big{)}-\kappa(x(0))\right\rVert,t\big{)};
  2. (iii)

    the unique KKT triple (x,u,v)(x^{*},u^{*},v^{*}) is locally exponentially stable and globally attracting.

Proof.

We begin with (i). By direct examination of (37), we see that the equilibria correspond exactly with triples satisfying (3). Since the matrix Q~\tilde{Q} has full rank, the gradients of all the constraints are linearly independent, and hence MFCQ holds on 𝒞\mathcal{C}. Since FF is μ\mu-strongly monotone, the solution xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}) is unique and there exists a unique Lagrange multiplier (u,v)(u^{*},v^{*}) such that (x,u,v)(x^{*},u^{*},v^{*}) satisfies (3).

To show (ii), we verify that all hypotheses in [55, Theorem 4] hold. First, note that the map x(x,u,v)x\mapsto\mathcal{F}(x,u,v) is F\ell_{F}-Lipschitz in xx uniformly in (u,v)(u,v), and [G;H]\left\lVert[G;H]\right\rVert-Lipschitz in (u,v)(u,v), uniformly in xx. Let \mathcal{H} denote the righthand side of (36). Because \mathcal{H} is piecewise affine in (u,v)(u,v) and FF globally Lipschitz, there exists constants ,x\ell_{\mathcal{H},x}, ,u,v>0\ell_{\mathcal{H},u,v}>0, such that \mathcal{H} is ,x\ell_{\mathcal{H},x}-Lipschitz in xx uniformly in (u,v)(u,v) and ,u,v\ell_{\mathcal{H},u,v}-Lipschitz in (u,v)(u,v) uniformly in xx. By Lemma 6.3, there exists c¯>0\bar{c}>0 such that (36) is c¯\bar{c}-contracting, uniformly in xx. Finally, we note that the reduced system corresponding to (37) is x˙=𝒢α(x)\dot{x}=\mathcal{G}_{\alpha}(x), which is contracting by Theorem 5.13. Thus all the hypotheses of [55, Theorem 4] hold and (ii) follows. Finally (iii) follows from combining (i) and (ii). ∎

6.3 Safety of Recursive Safe Monotone Flow

Here we discuss the safety properties of the recursive safe monotone flow. In general, even if the initial condition belongs to 𝒞\mathcal{C}, i.e., x(0)𝒞x(0)\in\mathcal{C}, it is not guaranteed that solutions of the system (37) satisfy x(t)𝒞x(t)\in\mathcal{C} for t>0t>0. However, under appropriate conditions, we can show that the system is “practically safe”, in the sense that x(t)x(t) remains in a slightly expanded form of the original constraint set 𝒞\mathcal{C}.

Theorem 6.5.

(Practical Safety of Recursive Safe Monotone Flow) Assume FF is μ\mu-strongly monotone and F\ell_{F} globally Lipschitz, and α\alpha satisfies (27). Under Assumption 1 and β\beta chosen as in Lemma 6.3, then for all ϵ>0\epsilon>0, there exists δ>0\delta>0 and τ\tau^{*} such that, if 0<τ<τ0<\tau<\tau^{*}, any solution to (37) with

  • x(0)𝒞x(0)\in\mathcal{C};

  • (u(0),v(0))κ(x(0))δ\left\lVert(u(0),v(0))-\kappa(x(0))\right\rVert\leq\delta;

satisfies x(t)𝒞ϵx(t)\in\mathcal{C}_{\epsilon} for all t0t\geq 0, where 𝒞ϵ={xng(x)ϵ,|h(x)|ϵ}\mathcal{C}_{\epsilon}=\{x\in^{n}\mid g(x)\leq\epsilon,\left|h(x)\right|\leq\epsilon\}.

To prove Theorem 6.5, we rely on the notion of input-to-state safety. Consider the system

x˙=𝒢α(x)i=1nei(t)gi(x)j=1mdj(t)hj(x).\dot{x}=\mathcal{G}_{\alpha}(x)-\sum_{i=1}^{n}e_{i}(t)\nabla g_{i}(x)-\sum_{j=1}^{m}d_{j}(t)\nabla h_{j}(x). (39)

This system can be interpreted as the safe monotone flow perturbed by a disturbance (e(t),d(t))(e(t),d(t)). The set 𝒞\mathcal{C} is input-to-state safe (ISSf) with respect to (39), with gain γ\gamma, if there exists a class 𝒦\mathcal{K} function γ\gamma such that, if γ((e,d))<ϵ\gamma(\left\lVert(e,d)\right\rVert_{\infty})<\epsilon, then 𝒞ϵ\mathcal{C}_{\epsilon} is forward invariant under (39). This notion of input-to-state safety is a slight generalization of the standard definition, cf. [56], to the case where the safe set is parameterized by multiple equality and inequality constraints. We show next that (39) is ISSf.

Lemma 6.6 (Perturbed Safe Monotone Flow is ISSf).

Under Assumption 1, the set 𝒞\mathcal{C} is input-to-state safe with respect to (39) with gain γ(r)=λmax(Q~)αr\gamma(r)=\frac{\lambda_{\textnormal{max}}(\tilde{Q})}{\alpha}r, where Q~\tilde{Q} is defined in (31)

Proof.

For i{1,,m}i\in\{1,\dots,m\}, under (39)

g˙i(x)\displaystyle\dot{g}_{i}(x) =Gi(𝒢α(x)i=1nei(t)gi(x)j=1mdj(t)hj(x))\displaystyle=G_{i}^{\top}\big{(}\mathcal{G}_{\alpha}(x)-\sum_{i=1}^{n}e_{i}(t)\nabla g_{i}(x)-\sum_{j=1}^{m}d_{j}(t)\nabla h_{j}(x)\big{)}
αgi(x)Gi(i=1nei(t)gi(x)+j=1mdj(t)hj(x))\displaystyle\leq-\alpha g_{i}(x)-G^{\top}_{i}\bigg{(}\sum_{i=1}^{n}e_{i}(t)\nabla g_{i}(x)+\sum_{j=1}^{m}d_{j}(t)\nabla h_{j}(x)\bigg{)}
αgi(x)+λmax(Q~)(e(t),d(t)),\displaystyle\leq-\alpha g_{i}(x)+\lambda_{\textnormal{max}}(\tilde{Q})\left\lVert(e(t),d(t))\right\rVert,

where GiG^{\top}_{i} is the iith row of GG. It follows from [56, Theorem 1] that the set 𝒞gi={xnGix(cg)i0}\mathcal{C}_{g_{i}}=\{x\in^{n}\mid G_{i}^{\top}x-(c_{g})_{i}\leq 0\} is input-to-state safe with gain γ\gamma with respect to (37).

For j{1,,k}j\in\{1,\dots,k\}, under (39),

h˙\displaystyle\dot{h} (x)j=Hj(𝒢α(x)i=1nei(t)gi(x)j=1mdj(t)hj(x))\displaystyle{}_{j}(x)=H_{j}^{\top}\big{(}\mathcal{G}_{\alpha}(x)-\sum_{i=1}^{n}e_{i}(t)\nabla g_{i}(x)-\sum_{j=1}^{m}d_{j}(t)\nabla h_{j}(x)\big{)}
=αhj(x)Hj(i=1nei(t)gi(x)+j=1mdj(t)hj(x)),\displaystyle=-\alpha h_{j}(x)-H_{j}^{\top}\bigg{(}\sum_{i=1}^{n}e_{i}(t)\nabla g_{i}(x)+\sum_{j=1}^{m}d_{j}(t)\nabla h_{j}(x)\bigg{)},

where HjH^{\top}_{j} is the jjth row of HH. It follows that

h˙j(x)\displaystyle\dot{h}_{j}(x) αhj(x)+λmax(Q~)(e(t),d(t)),\displaystyle\leq-\alpha h_{j}(x)+\lambda_{\textnormal{max}}(\tilde{Q})\left\lVert(e(t),d(t))\right\rVert,
h˙j(x)\displaystyle\dot{h}_{j}(x) αhj(x)λmax(Q~)(e(t),d(t)).\displaystyle\geq-\alpha h_{j}(x)-\lambda_{\textnormal{max}}(\tilde{Q})\left\lVert(e(t),d(t))\right\rVert.

Thus, by [56, Theorem 1], the sets 𝒞hj={xnHjx(ch)j0}\mathcal{C}^{-}_{h_{j}}=\{x\in^{n}\mid H_{j}^{\top}x-(c_{h})_{j}\leq 0\}, and 𝒞hj+={xnHjx(ch)j0}\mathcal{C}^{+}_{h_{j}}=\{x\in^{n}\mid H_{j}^{\top}x-(c_{h})_{j}\geq 0\} are also input-to-state safe with gain γ\gamma with respect to (37). Finally, input-to-state safety of 𝒞\mathcal{C} follows from the fact that

𝒞=(i=1m𝒞gi)(j=1k(𝒞hj+𝒞hj)).\mathcal{C}=\bigg{(}\bigcap_{i=1}^{m}\mathcal{C}_{g_{i}}\bigg{)}\cap\bigg{(}\bigcap_{j=1}^{k}(\mathcal{C}^{+}_{h_{j}}\cap\mathcal{C}^{-}_{h_{j}})\bigg{)}.

We are now ready to prove Theorem 6.5.

Proof of Theorem 6.5.

By Lemma 6.6, 𝒞\mathcal{C} is input-to-state safe with respect to (39), with gain γ(r)=λmax(Q~)αr\gamma(r)=\frac{\lambda_{\textnormal{max}}(\tilde{Q})}{\alpha}r. Note that, for any solution (x(t),u(t),v(t))(x(t),u(t),v(t)) of (37), the trajectory x(t)x(t) solves (39) with

[e(t)d(t)]=[u(t)v(t)]κ(x(t)).\begin{bmatrix}e(t)\\ d(t)\end{bmatrix}=\begin{bmatrix}u(t)\\ v(t)\end{bmatrix}-\kappa(x(t)).

Next, by Theorem 6.4, for all ϵ\epsilon, there exists τ>0\tau^{*}>0 such that if 0<τ<τ0<\tau<\tau^{*}, then for all t0t\geq 0, (e(t),d(t))β((e(0),d(0)),t)\left\lVert\big{(}e(t),d(t)\big{)}\right\rVert\leq\beta\big{(}\left\lVert\big{(}e(0),d(0)\big{)}\right\rVert,t\big{)} for some class 𝒦\mathcal{K}\mathcal{L} function β\beta. Now, choose δ>0\delta>0 such that α1λmax(Q~)β(δ,0)<ϵ\alpha^{-1}\lambda_{\textnormal{max}}(\tilde{Q})\beta(\delta,0)<\epsilon and let (u(0),v(0))κ(x(0))δ\left\lVert(u(0),v(0))-\kappa(x(0))\right\rVert\leq\delta. Then, for all t0t\geq 0,

γ((e(t),d(t)))\displaystyle\gamma\left(\left\lVert\big{(}e(t),d(t)\big{)}\right\rVert\right) γ(β(δ,t))γ(β(δ,0))<ϵ.\displaystyle\leq\gamma(\beta(\delta,t))\leq\gamma(\beta(\delta,0))<\epsilon.

Hence, for x(0)𝒞𝒞ϵx(0)\in\mathcal{C}\subset\mathcal{C}_{\epsilon}, since 𝒞\mathcal{C} is input-to-state safe with respect to (39), we conclude x(t)𝒞ϵx(t)\in\mathcal{C}_{\epsilon} for all t0t\geq 0. ∎

Remark 6.7 (Implementation of Recursive Safe Monotone Flow on Network Optimization Problem).

One drawback of the projected monotone flow and the safe monotone flow is that neither system admits a distributed implementation for network optimization problems. The reason for this is that both the feedback control and projected implementations of these systems introduce additional couplings between agents that require global knowledge to solve exactly.

We show here that the recursive safe monotone flow can be implemented in a distributed manner on constrained optimization problems with a separable objective function and locally expressible constraints, provided that agents can communicate with their two-hop neighbors. Consider an undirected network (𝒱,)(\mathcal{V},\mathcal{E}) where 𝒱={1,2,,N}\mathcal{V}=\{1,2,\dots,N\} is a set of agents, and \mathcal{E} is the set of edges. For each agent i𝒱i\in\mathcal{V}, let 𝒩i\mathcal{N}_{i} denote the set of neighbors of ii. Consider

minimizexn\displaystyle\underset{x\in^{n}}{\textnormal{minimize}} f(x)=i𝒱fi(xi)\displaystyle f(x)=\sum_{i\in\mathcal{V}}f_{i}(x_{i}) (40)
subject to j𝒩iGijxjcgi,i𝒱\displaystyle\sum_{j\in\mathcal{N}_{i}}G_{ij}x_{j}\leq c_{g_{i}},\;\;\forall i\in\mathcal{V}

where xix_{i} corresponds to the decision variable of the iith agent. For problems with this structure, the constraints corresponding to each node are expressible as a function of information available to that node. This assumption is common in the distributed optimization literature (see e.g. [57]). When implementing (37), the state of the iith agent consists of (xi,ui)(x_{i},u_{i}), and the dynamics are given by

xi˙\displaystyle\dot{x_{i}} =fi(xi)j𝒩iGijuj\displaystyle=-\nabla f_{i}(x_{i})-\sum_{j\in\mathcal{N}_{i}}G_{ij}^{\top}u_{j}
τui˙\displaystyle\tau\dot{u_{i}} =max{βui,j𝒩iGij(fj(xj)k𝒩jGjkuk)+α(j𝒩iGijxjcgi)}\displaystyle=\text{max}\bigg{\{}-\beta u_{i},\sum_{j\in\mathcal{N}_{i}}G_{ij}\bigg{(}-\nabla f_{j}(x_{j})-\sum_{k\in\mathcal{N}_{j}}G_{jk}^{\top}u_{k}\bigg{)}+\alpha\bigg{(}\sum_{j\in\mathcal{N}_{i}}G_{ij}x_{j}-c_{g_{i}}\bigg{)}\bigg{\}}

Observe that from the previous expression, the iith node can implement the dynamics corresponding to its state using only the state information of its two-hop neighbors. \bullet

7 Numerical Examples

Here we illustrate the behavior of the proposed flows on two example problems. The first example is a variational inequality on 2 corresponding to a two-player game with quadratic payoff functions. The second example is a constrained linear-quadratic dynamic game where we implement the safe monotone flow in a receding horizon manner to examine its anytime properties.

7.1 Nash Equilibria of Two-Player Game

The first numerical example we discuss is a variational inequality on 2 corresponding to a two-player game, where player i{1,2}i\in\{1,2\} wants to minimize a cost Ji(x1,x2)J_{i}(x_{1},x_{2}) subject to the constraints that xi𝒞ix_{i}\in\mathcal{C}_{i}\subset. We take 𝒞=𝒞1×𝒞22\mathcal{C}=\mathcal{C}_{1}\times\mathcal{C}_{2}\subset^{2}. We have selected a two-dimensional example that allows us to visualize the constraint set and the trajectories of the proposed flows to better illustrate their differences. The problem of finding the Nash equilibria of a game of this form is equivalent to the variational inequality VI(F,𝒞)\textnormal{VI}(F,\mathcal{C}), where FF is the pseudogradient map, given by F(x)=(x1J1(x1,x2),x2J2(x1,x2))F(x)=(\nabla_{x_{1}}J_{1}(x_{1},x_{2}),\nabla_{x_{2}}J_{2}(x_{1},x_{2})). For i{1,2}i\in\{1,2\}, let 𝒞i={x1x1}\mathcal{C}_{i}=\{x\in\mid-1\leq x\leq 1\} and JiJ_{i} be the quadratic function Ji(x1,x2)=12xQix+rixJ_{i}(x_{1},x_{2})=\frac{1}{2}x^{\top}Q_{i}x+r_{i}^{\top}x, with

Q1=[1111]2×2,\displaystyle Q_{1}=\begin{bmatrix}1&-1\\ -1&1\end{bmatrix}\in^{2\times 2},\qquad r1=[00]2,\displaystyle r_{1}=\begin{bmatrix}0\\ 0\end{bmatrix}\in^{2},
Q2=[1111]2×2,\displaystyle Q_{2}=\begin{bmatrix}1&1\\ 1&1\end{bmatrix}\in^{2\times 2},\qquad r2=[0.50.5]2.\displaystyle r_{2}=\begin{bmatrix}0.5\\ 0.5\end{bmatrix}\in^{2}.

The pseudogradient map is given by F(x)=Qx+rF(x)=Qx+r where

Q=[1111]r=[00.5]Q=\begin{bmatrix}1&-1\\ 1&1\end{bmatrix}\qquad r=\begin{bmatrix}0\\ 0.5\end{bmatrix}

Because 12(Q+Q)=I0\frac{1}{2}(Q+Q^{\top})=I\succ 0, it follows that the FF is 11-strongly monotone, and therefore the problem has a unique solution xSOL(F,𝒞)x^{*}\in\textnormal{SOL}(F,\mathcal{C}).

Figure 2 shows the results of implementing each of the proposed flows to find the Nash equilibrium. The projected monotone flow, cf. Figure 2(a), is only well defined in 𝒞\mathcal{C}. However, the constraint set remains forward invariant and all trajectories converge to the solution xx^{*}. The safe monotone flow with α=1.0\alpha=1.0, cf. Figure 2(b), also keeps the constraint set forward invariant and has all trajectories converge to xx^{*}. In addition, the system is well defined outside of 𝒞\mathcal{C}, and trajectories beginning outside the feasible set converge to it.

In Figure 2(c), we consider the recursive safe monotone flow with α=1.0\alpha=1.0, β=1.0\beta=1.0 and τ=0.25\tau=0.25, where u(0)=0u(0)=0. The trajectories converge to xx^{*} and closely approximate the trajectories of the safe monotone flow. Note, however, that the set 𝒞\mathcal{C} is not safe but only practically safe. This is illustrated in the zoomed-in Figure 2(d), where it is apparent that the trajectories do not always remain in 𝒞\mathcal{C} but remain close to it.

Refer to caption
(a) Projected monotone flow
Refer to caption
(b) Safe monotone flow
Refer to caption
(c) Recursive safe monotone flow
Refer to caption
(d) Zoomed-in plot of (c)
Figure 2: Implementation of (a) projected monotone flow, (b) safe monotone flow (α=1.0\alpha=1.0), and (c) recursive safe monotone flow (τ=0.25\tau=0.25) in a two-player game. The shaded region shows the constraint set 𝒞\mathcal{C} and the colored paths represent trajectories of the corresponding flow starting from various initial condition. (d) shows a zoomed-in portion of the boundary of 𝒞\mathcal{C} to illustrate the practical safety of the recursive safe monotone flow.

7.2 Receding Horizon Linear-Quadratic Dynamic Game

We now discuss a more complex example, where the input to a plant is specified by the solution to a variational inequality parameterized by the state of the plant. To solve it, we interconnect the plant dynamics with the safe monotone flow, and demonstrate that the anytime property of the latter ensures good performance and satisfaction of the constraints even when terminated terminated early. The plant takes the form of a discrete-time linear time-invariant system with two inputs,

z(s+1)=Az(s)+B1w1(s)+B2w2(s),z(s+1)=Az(s)+B_{1}w_{1}(s)+B_{2}w_{2}(s), (41)

where Anz×nzA\in^{n_{z}\times n_{z}} and Binz×nwB_{i}\in^{n_{z}\times n_{w}} for i{1,2}i\in\{1,2\}.

We consider a linear-quadratic dynamic game (LQDG) between two players, where each player i{1,2}i\in\{1,2\} can influence the system (41) by choosing the corresponding input wiWinww_{i}\in W_{i}\subset^{n_{w}}. We fix a time horizon N>0N>0 and an initial condition z(0)=z0z(0)=z_{0}, and define J:(W1)N×(W2)NJ:(W_{1})^{N}\times(W_{2})^{N}\to as a quadratic payoff,

J(w1(),\displaystyle J(w_{1}(\cdot), w2())=z(N)2Qf+s=0N1z(s)2Q+w1(s)R12w2(s)R22,\displaystyle w_{2}(\cdot))=\left\lVert z(N)\right\rVert^{2}_{Q_{f}}+\sum_{s=0}^{N-1}\left\lVert z(s)\right\rVert^{2}_{Q}+\left\lVert w_{1}(s)\right\rVert_{R_{1}}^{2}-\left\lVert w_{2}(s)\right\rVert_{R_{2}}^{2}, (42)

where Qf,Q0Q_{f},Q\succeq 0 and R1,R20R_{1},R_{2}\succ 0. The goal of player 1 is to minimize the payoff (42), whereas the goal of player 2 is to maximize it. This problem can be solved in closed form when the constraints WiW_{i} are trivial (cf. [58, Chapter 6], [59]), but must be solved numerically for nontrivial constraints.

We first note the LQDG problem can be posed as a variational inequality. Indeed, let z¯=(z(1),,z(N))\bar{z}=(z(1),\dots,z(N)) and, for i{1,2}i\in\{1,2\}, let w¯i=(wi(0),,wi(N1))\bar{w}_{i}=(w_{i}(0),\dots,w_{i}(N-1)). Define

𝒜=[AA2AN],Ci=[Bi00ABiBi0AN1BiAN2BiBi].\mathcal{A}=\begin{bmatrix}A\\ A^{2}\\ \vdots\\ A^{N}\end{bmatrix},\qquad C_{i}=\begin{bmatrix}B_{i}&0&\cdots&0\\ AB_{i}&B_{i}&\cdots&0\\ \vdots&\vdots&\ddots&\vdots\\ A^{N-1}B_{i}&A^{N-2}B_{i}&\cdots&B_{i}\end{bmatrix}.

Next, letting Q¯=diag(Q,,Q,Qf)\bar{Q}=\text{diag}(Q,\dots,Q,Q_{f}) and R¯i=diag(Ri,,Ri)\bar{R}_{i}=\text{diag}(R_{i},\dots,R_{i}), and using the fact that

z¯=𝒜z0+C1w¯1+C2w¯2,\bar{z}=\mathcal{A}z_{0}+C_{1}\bar{w}_{1}+C_{2}\bar{w}_{2},

we see that the payoff function (42) can be written as,

J(w¯1,w¯2)=\displaystyle J(\bar{w}_{1},\bar{w}_{2})= [w¯1w¯2][C1Q¯C1+R¯1C1Q¯C2C2Q¯C1C2Q¯C2R¯2][w¯1w¯2]+2[C1Q¯𝒜z0C2Q¯𝒜z0][w¯1w¯2]+z0𝒜Q¯𝒜z0.\displaystyle\begin{bmatrix}\bar{w}_{1}\\ \bar{w}_{2}\end{bmatrix}^{\top}\begin{bmatrix}C_{1}^{\top}\bar{Q}C_{1}+\bar{R}_{1}&C_{1}^{\top}\bar{Q}C_{2}\\ C_{2}^{\top}\bar{Q}C_{1}&C_{2}^{\top}\bar{Q}C_{2}-\bar{R}_{2}\end{bmatrix}\begin{bmatrix}\bar{w}_{1}\\ \bar{w}_{2}\end{bmatrix}+2\begin{bmatrix}C_{1}^{\top}\bar{Q}\mathcal{A}z_{0}\\ C_{2}^{\top}\bar{Q}\mathcal{A}z_{0}\end{bmatrix}^{\top}\begin{bmatrix}\bar{w}_{1}\\ \bar{w}_{2}\end{bmatrix}+z_{0}^{\top}\mathcal{A}^{\top}\bar{Q}\mathcal{A}z_{0}.

Finally, letting x=(w¯1,w¯2)x=(\bar{w}_{1},\bar{w}_{2}), we see that the LQDG problem corresponds to the variational inequality VI(F(,z0),𝒞)\textnormal{VI}(F(\cdot,z_{0}),\mathcal{C}), where

F(x,z0)=[C1Q¯C1+R¯1C1Q¯C2C2Q¯C1R¯2C2Q¯C2][w¯1w¯2]+[C1Q¯𝒜z0C2Q¯𝒜z0]\displaystyle F(x,z_{0})\!=\!\begin{bmatrix}C_{1}^{\top}\bar{Q}C_{1}+\bar{R}_{1}&C_{1}^{\top}\bar{Q}C_{2}\\ -C_{2}^{\top}\bar{Q}C_{1}&\bar{R}_{2}-C_{2}^{\top}\bar{Q}C_{2}\end{bmatrix}\!\begin{bmatrix}\bar{w}_{1}\\ \bar{w}_{2}\end{bmatrix}\!+\!\begin{bmatrix}C_{1}^{\top}\bar{Q}\mathcal{A}z_{0}\\ -C_{2}^{\top}\bar{Q}\mathcal{A}z_{0}\end{bmatrix}

and the constraint set is

𝒞=W1×W1×W1N times×W2×W2××W2N times.\mathcal{C}=\underset{N\text{ times}}{\underbrace{W_{1}\times W_{1}\times\cdots W_{1}}}\times\underset{N\text{ times}}{\underbrace{W_{2}\times W_{2}\times\cdots\times W_{2}}}.

We assume that the problem data satisfies

[C1Q¯C1+R¯100R¯2C2Q¯C2]0,\begin{bmatrix}C_{1}^{\top}\bar{Q}C_{1}+\bar{R}_{1}&0\\ 0&\bar{R}_{2}-C_{2}^{\top}\bar{Q}C_{2}\end{bmatrix}\succ 0,

in which case FF is strongly monotone.

For simulation purposes, we take nz=5n_{z}=5, nw=2n_{w}=2, Bi=IB_{i}=I, Wi=02W_{i}=^{2}_{\geq 0}, and AA a marginally stable matrix selected randomly. We use the safe monotone flow to solve the variational inequality and implement the solution in a receding horizon manner: given the initial state z0z_{0}, we solve for the optimal input sequence (w1(),w2())(w_{1}(\cdot),w_{2}(\cdot)) over the entire time horizon, apply the input (w1(0),w2(0))(w_{1}(0),w_{2}(0)) to (41) to obtain z(1)z(1), update the initial condition z0z(1)z_{0}\leftarrow z(1) and repeat. When FF is strongly monotone, on each iteration the flow converges to the exact solution as tt\to\infty. However, we also consider here the effect of terminating the solver early at some t=tf<t=t_{f}<\infty.

Figure 3 shows the results of the simulation. In Figure 3(a), we plot z(s)\left\lVert z(s)\right\rVert for various values of termination times. We denote the exact solution with tf=t_{f}=\infty. The closed-loop dynamics with the exact solution to the receding horizon LQDG is stabilizing, and as tft_{f} grows larger, the early terminated solution drives the state of the system closer to the origin. In Figure 3(b), we plot the first component of w1(s)w_{1}(s) in blue and the first component of w2(s)w_{2}(s) in red. Regardless of when terminated, the inputs satisfy the input constraints on each iteration due to the safety properties of the safe monotone flow.

Refer to caption
(a) z(s)\left\lVert z(s)\right\rVert on each iteration
Refer to caption
(b) First component of {wi(s)}i{1,2}\{w_{i}(s)\}_{i\in\{1,2\}} on each iteration
Figure 3: Receding horizon implementation of the safe monotone flow solving a linear quadratic dynamic game for different choices of termination time tft_{f}. The closed-loop implementation of the exact solution corresponds to tf=t_{f}=\infty (dashed lines). (a) We plot the evolution of z(s)\left\lVert z(s)\right\rVert in green. (b) We plot the evolution of the first component of w1(s)w_{1}(s) in blue-green (scale in left yy-axis) and the first component of w2(x)w_{2}(x) in red-orange (scale in right yy-axis).

8 Conclusions

We have tackled the design of anytime algorithms to solve variational inequalities as a feedback control problem. Using techniques from safety-critical control, we have synthesized three continuous-time dynamics which find solutions to monotone variational inequalities: the projected monotone flow, already well known in the literature, and the novel safe monotone and recursive safe monotone flows. The equilibria of these flows correspond to solutions of the variational inequality, and so we have embarked in the precise characterization of their asymptotic stability properties. We have established asymptotic stability of equilibria in the case of strong monotonicity, and contractivity and exponential stability in the case of polyhedral constraints. We have also shown that the safe monotone flow renders the constraint forward invariant and asymptotically stable. The recursive safe monotone flow offers an alternative implementation that does not necessitate the solution of a quadratic program along the trajectories. This flow results from coupling two systems evolving on different timescales, and we have established local exponential stability and global attractivity of equilibria, as well as practical safety guarantees. We have illustrated in two game scenarios the properties of the proposed flows and, in particular, their amenability for interconnection and regulation of physical processes. Future work will develop methods for distributed network problems, study time-varying and parametric problems, consider applications to feedback optimization, explore connections between the recursive safe monotone flow and primal-dual flows, and analyze discretizations and the implementation on computational platforms of the proposed flows.

Appendix A Appendix: Proof of Lemma 5.8

Proof.

Note that

D𝒢α+V~(x)=(xx)F(x)i=1mui(xx)gi(x)j=1kvj(xx)hj(x).\displaystyle D^{+}_{\mathcal{G}_{\alpha}}\tilde{V}(x)=-(x-x^{*})^{\top}F(x)-\sum_{i=1}^{m}u_{i}(x-x^{*})^{\top}\nabla g_{i}(x)-\sum_{j=1}^{k}v_{j}(x-x^{*})^{\top}\nabla h_{j}(x).

When FF is μ\mu-strongly monotone,

(xx)F(x)μxx2(xx)F(x),-(x-x^{*})^{\top}F(x)\leq-\mu\left\lVert x-x^{*}\right\rVert^{2}-(x-x^{*})^{\top}F(x^{*}),

and when FF is monotone, the previous inequality holds with μ=0\mu=0. Next, we rearrange (3a) and use that gig_{i} is convex for all i=1,,mi=1,\dots,m and hjh_{j} is affine for all j=1,,mj=1,\dots,m to obtain

(xx)F(x)\displaystyle-(x-x^{*})^{\top}F(x^{*}) =i=1mui(xx)gi(x)+j=1kvj(xx)hj(x)\displaystyle=\sum_{i=1}^{m}u_{i}^{*}(x-x^{*})^{\top}\nabla g_{i}(x^{*})+\sum_{j=1}^{k}v_{j}^{*}(x-x^{*})^{\top}\nabla h_{j}(x^{*})
i=1mui(gi(x)gi(x))+j=1kvj(hj(x)hj(x))\displaystyle\leq\sum_{i=1}^{m}u_{i}^{*}(g_{i}(x)-g_{i}(x^{*}))+\sum_{j=1}^{k}v_{j}^{*}(h_{j}(x)-h_{j}(x^{*}))
=(u)(g(x)g(x))+(v)h(x).\displaystyle=(u^{*})^{\top}(g(x)-g(x^{*}))+(v^{*})^{\top}h(x).

where the last equality follows from the fact that h(x)=0h(x^{*})=0. By a similar line of reasoning, we have

i=1m\displaystyle-\sum_{i=1}^{m} ui(xx)gi(x)j=1kvj(xx)hj(x)\displaystyle u_{i}(x-x^{*})^{\top}\nabla g_{i}(x)-\sum_{j=1}^{k}v_{j}(x-x^{*})^{\top}\nabla h_{j}(x)
i=1mui(gi(x)gi(x))j=1kvj(hj(x)hj(x))\displaystyle\leq-\sum_{i=1}^{m}u_{i}(g_{i}(x)-g_{i}(x^{*}))-\sum_{j=1}^{k}v_{j}(h_{j}(x)-h_{j}(x^{*}))
=u(g(x)g(x))vh(x).\displaystyle=-u^{\top}(g(x)-g(x^{*}))-v^{\top}h(x).

The result follows by summing the previous two expressions. ∎

Appendix B Appendix: Proof of Lemma 5.9

Proof.

We first show that the solution to the optimization problem in the definition of WW is ξ=𝒢α(x)\xi=\mathcal{G}_{\alpha}(x). Note that the constraints in the definition of WW in (17) and (15) are identical. Let J(x,ξ)J(x,\xi) denote the objective function in the definition of WW. Then J(x,ξ)12ξ+F(x)2=12F(x)2J(x,\xi)-\frac{1}{2}\left\lVert\xi+F(x)\right\rVert^{2}=-\frac{1}{2}\left\lVert F(x)\right\rVert^{2}. Because the difference between the objective functions of (15) and the definition of WW is independent of ξ\xi, the two optimization problems have the same solution. The claim now follows because 𝒢α(x)\mathcal{G}_{\alpha}(x) is the minimizer solving (15).

Next we show that WW can be expressed in closed form as (18). Because the optimizer is ξ=𝒢α(x)\xi=\mathcal{G}_{\alpha}(x), we have

W(x)=𝒢α(x)F(x)+12𝒢α(x)2.W(x)=\mathcal{G}_{\alpha}(x)^{\top}F(x)+\frac{1}{2}\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}. (43)

Note that (𝒢α(x),u,v)(\mathcal{G}_{\alpha}(x),u,v) satisfies the optimality conditions (16) for all (u,v)Λα(x)(u,v)\in\Lambda_{\alpha}(x). Therefore we can rearrange (16a) to obtain F(x)=𝒢α(x)g(x)xuh(x)xvF(x)=-\mathcal{G}_{\alpha}(x)-\frac{\partial g(x)}{\partial x}^{\top}u-\frac{\partial h(x)}{\partial x}^{\top}v. Next

𝒢α(x)F(x)\displaystyle\mathcal{G}_{\alpha}(x)^{\top}F(x) =𝒢α(x)2ug(x)x𝒢α(x)vh(x)x𝒢α(x)\displaystyle=-\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}-u^{\top}\frac{\partial g(x)}{\partial x}\mathcal{G}_{\alpha}(x)-v^{\top}\frac{\partial h(x)}{\partial x}\mathcal{G}_{\alpha}(x)
=𝒢α(x)2+αug(x)+αvh(x),\displaystyle=-\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}+\alpha u^{\top}g(x)+\alpha v^{\top}h(x),

where the second equality follows by rearranging (16c) and (16e). Then, (18) follows by substituting the previous expression into (43).

Finally we show (19). By [60, Theorem 4.2], it follows that

D𝒢α+W(x)\displaystyle D^{+}_{\mathcal{G}_{\alpha}}W(x) =sup(u,v)Λα(x){xL(x;𝒢α(x),u,v)𝒢α(x)}\displaystyle=\sup_{(u^{\prime},v^{\prime})\in\Lambda_{\alpha}(x)}\left\{\nabla_{x}L(x;\mathcal{G}_{\alpha}(x),u^{\prime},v^{\prime})^{\top}\mathcal{G}_{\alpha}(x)\right\}
xL(x;𝒢α(x),u,v)𝒢α(x),\displaystyle\geq\nabla_{x}L(x;\mathcal{G}_{\alpha}(x),u,v)^{\top}\mathcal{G}_{\alpha}(x),

where LL is defined in (30). By direct computation, we can verify that

xL(x;ξ,u,v)=Q(x,u)ξ+αg(x)xu+αh(x)xv\displaystyle\nabla_{x}L(x;\xi,u,v)=Q(x,u)\xi+\alpha\frac{\partial g(x)}{\partial x}^{\top}u+\alpha\frac{\partial h(x)}{\partial x}^{\top}v

Therefore

xL(x;𝒢α(x),u,v)𝒢α(x)\displaystyle\nabla_{x}L(x;\mathcal{G}_{\alpha}(x),u,v)^{\top}\mathcal{G}_{\alpha}(x) =𝒢α(x)Q(x,u)2+αug(x)x𝒢α(x)+αvh(x)x𝒢α(x)\displaystyle=\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}_{Q(x,u)}+\alpha u^{\top}\frac{\partial g(x)}{\partial x}\mathcal{G}_{\alpha}(x)+\alpha v^{\top}\frac{\partial h(x)}{\partial x}\mathcal{G}_{\alpha}(x)
=𝒢α(x)Q(x,u)2α2ug(x)α2vh(x),\displaystyle=\left\lVert\mathcal{G}_{\alpha}(x)\right\rVert^{2}_{Q(x,u)}-\alpha^{2}u^{\top}g(x)-\alpha^{2}v^{\top}h(x),

where the last equality above follows by rearranging (16c) and (16e). ∎

Appendix C Appendix: Proof of Lemma 5.11

Proof.

Note that δϵ\delta_{\epsilon} corresponds to the 1\ell^{1}-penalty function for the set 𝒞\mathcal{C}. By [61, Proposition 3], the directional derivative of δϵ\delta_{\epsilon} is

δϵ(x;ξ)=\displaystyle\delta^{\prime}_{\epsilon}(x;\xi)= 1ϵiI+(x)gi(x)ξ+1ϵiI0(x)[gi(x)ξ]+\displaystyle\frac{1}{\epsilon}\sum_{i\in I_{+}(x)}\nabla g_{i}(x)^{\top}\xi+\frac{1}{\epsilon}\sum_{i\in I_{0}(x)}[\nabla g_{i}(x)^{\top}\xi]_{+}
+1ϵjIh(x)sgn(hj(x))hj(x)ξ+1ϵjIh(x)|hj(x)ξ|.\displaystyle+\frac{1}{\epsilon}\sum_{j\in I_{h}(x)}\textnormal{sgn}(h_{j}(x))\nabla h_{j}(x)^{\top}\xi+\frac{1}{\epsilon}\sum_{j\not\in I_{h}(x)}\left|\nabla h_{j}(x)^{\top}\xi\right|.

Note D𝒢α+δϵ(x)=δϵ(x;𝒢α(x))D^{+}_{\mathcal{G}_{\alpha}}\delta_{\epsilon}(x)=\delta^{\prime}_{\epsilon}(x;\mathcal{G}_{\alpha}(x)). Then (21) follows by noting that gi(x)𝒢α(x)αgi(x)\nabla g_{i}(x)^{\top}\mathcal{G}_{\alpha}(x)\leq-\alpha g_{i}(x) for all i=1,,mi=1,\dots,m and hj(x)𝒢α(x)=αhj(x)\nabla h_{j}(x)^{\top}\mathcal{G}_{\alpha}(x)=-\alpha h_{j}(x) for all j=1,kj=1,\dots k. ∎

References

  • [1] A. Allibhoy and J. Cortés. Safety-critical control as a design paradigm of anytime solvers of variational inequalities. In IEEE Conf. on Decision and Control, pages 6211–6216, Cancun, Mexico, December 2022.
  • [2] K. Arrow, L Hurwitz, and H. Uzawa. Studies in Linear and Non-Linear Programming. Stanford University Press, Stanford, CA, 1958.
  • [3] R. W. Brockett. Dynamical systems that sort lists, diagonalize matrices, and solve linear programming problems. Linear Algebra and Its Applications, 146:79–91, 1991.
  • [4] U. Helmke and J. B. Moore. Optimization and Dynamical Systems. Springer, 1994.
  • [5] C. Henry. An existence theorem for a class of differential equations with multivalued right-hand side. Journal of Mathematical Analysis and Applications, 41:179–186, 1973.
  • [6] H. Brezis. Operateurs Maximaux Monotones et Semi-groupes de Contractions dans Les Espaces de Hilbert, volume 5 of North-Holland Mathematics Studies. Elsevier, Oxford, UK, 1973.
  • [7] A. Nagurney and D. Zhang. Projected Dynamical Systems and Variational Inequalities with Applications, volume 2 of International Series in Operations Research and Management Science. Kluwer Academic Publishers, Dordrecht, The Netherlands, 1996.
  • [8] B. Brogliato, A. Daniilidis, C. Lemaréchal, and V. Acary. On the equivalence between complementarity systems, projected systems, and differential inclusions. Systems & Control Letters, 55(1):45–51, 2006.
  • [9] W. P. M. H. Heemels, J. M. Schumacher, and S. Weiland. Projected dynamical systems in a complementarity formalism. Operations Research Letters, 27:83–91, 2000.
  • [10] W. P. M. H. Heemels, M. K. Camlibel, and M. F. Heertjes. Oblique projected dynamical systems and incremental stability under state constraints. IEEE Control Systems Letters, 4(4):1060–1065, 2020.
  • [11] A. Hauswirth, S. Bolognani, and F. Dörfler. Projected dynamical systems on irregular, non-Euclidean domains for nonlinear optimization. SIAM Journal on Control and Optimization, 59(1):635–668, 2021.
  • [12] A. Tanwani, B. Brogliato, and C. Prieur. Well-posedness and output regulation for implicit time-varying evolution variational inequalities. SIAM Journal on Control and Optimization, 56(2):751–781, 2018.
  • [13] B. Brogliato and A. Tanwani. Dynamical systems coupled with monotone set-valued operators: Formalisms, applications, well-posedness, and stability. SIAM Review, 62(1):3–129, 2020.
  • [14] M. Colombino, E. Dall’Anese, and A. Bernstein. Online optimization as a feedback controller: Stability and tracking. IEEE Transactions on Control of Network Systems, 7(1):422–432, 2020.
  • [15] A. Hauswirth, Z. He, S. Bolognani, G. Hug, and F. Dörfler. Optimization algorithms as robust feedback controllers. Annual Reviews in Control, 57:100941, 2024.
  • [16] E. Dall’Anese and A. Simonetto. Optimal power flow pursuit. IEEE Transactions on Smart Grid, 9(2):942–952, 2016.
  • [17] L. S. P. Lawrence, J. W. Simpson-Porco, and E. Mallada. Linear-convex optimal steady-state control. IEEE Transactions on Automatic Control, 66(11):5377–5384, 2021.
  • [18] S. H. Low, F. Paganini, and J. C. Doyle. Internet congestion control. IEEE Control Systems, 22(1):28–43, 2002.
  • [19] G. Bianchin, J. Cortés, J. I. Poveda, and E. Dall’Anese. Time-varying optimization of LTI systems via projected primal-dual gradient flows. IEEE Transactions on Control of Network Systems, 9(1):474–486, 2022.
  • [20] V. Häberle, A. Hauswirth, L. Ortmann, S. Bolognani, and F. Dörfler. Non-convex feedback optimization with input and output constraints. IEEE Control Systems Letters, 5(1):343–348, 2021.
  • [21] A. Agarwal, J. W. Simpson-Porco, and L. Pavel. Game-theoretic feedback-based optimization. IFAC-PapersOnLine, 55(13):174–179, 2022.
  • [22] G. Belgioioso, D. Liao-McPherson, M. H. de Badyn, S. Bolognani, R. S. Smith, J. Lygeros, and F. Dörfler. Online feedback equilibrium seeking. IEEE Transactions on Automatic Control, 2024. To appear.
  • [23] K. B. Ariyur and M. Krstić. Real-Time Optimization by Extremum-Seeking Control. Wiley, New York, 2003.
  • [24] P. Frihauf, M. Krstić, and T. Başar. Nash equilibrium seeking in noncooperative games. IEEE Transactions on Automatic Control, 57(5):1192–1207, 2012.
  • [25] A. Williams, M. Krstic, and A. Scheinker. Semi-global practical extremum seeking with practical safety. In IEEE Conf. on Decision and Control, pages 6774–6779. IEEE, 2023.
  • [26] A. Allibhoy and J. Cortés. Control barrier function-based design of gradient flows for constrained nonlinear programming. IEEE Transactions on Automatic Control, 69(6):3499–3514, 2024.
  • [27] A. D. Ames, S. Coogan, M. Egerstedt, G. Notomista, K. Sreenath, and P. Tabuada. Control barrier functions: theory and applications. In European Control Conference, pages 3420–3431, Naples, Italy, 2019.
  • [28] M. Cohen and C. Belta. Adaptive and Learning-Based Control of Safety-Critical Systems. Springer, New York, 2023.
  • [29] F. Blanchini. Set invariance in control. Automatica, 35(11):1747–1767, 1999.
  • [30] P. Wieland and F. Allgöwer. Constructive safety using control barrier functions. IFAC Proceedings Volumes, 40(12):462–467, 2007.
  • [31] 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, 62(8):3861–3876, 2017.
  • [32] W. Xiao, C. G. Cassandras, and C. Belta. Safe Autonomy with Control Barrier Functions: Theory and Applications. Synthesis Lectures on Computer Science. Springer, New York, 2023.
  • [33] G. Delimpaltadakis and W. P. M. H. Heemels. On the relationship between control barrier functions and projected dynamical systems. In IEEE Conf. on Decision and Control, page 770–775, Singapore, 2023.
  • [34] G. Delimpaltadakis, J. Cortés, and W. P. M. H. Heemels. Continuous approximations of projected dynamical systems via control barrier functions. IEEE Transactions on Automatic Control, 2024. Submitted.
  • [35] F. Facchinei and J.-S. Pang. Finite-Dimensional Variational Inequalities and Complementarity Problems. Springer, 2003.
  • [36] R. T. Rockafellar and R. J. B. Wets. Variational Analysis, volume 317 of Comprehensive Studies in Mathematics. Springer, New York, 1998.
  • [37] J. P. Aubin and A. Cellina. Differential Inclusions, volume 264 of Grundlehren der mathematischen Wissenschaften. Springer, New York, 1984.
  • [38] V. Acary and B. Brogliato. Numerical Methods for Nonsmooth Dynamical Systems: Applications in Mechanics and Electronics, volume 35 of Lecture Notes in Applied and Computational Mechanics. Springer, New York, 2008.
  • [39] R. T. Rockafellar. Convex Analysis. Princeton University Press, 1970.
  • [40] D. W. Peterson. A review of constraint qualifications in finite-dimensional spaces. SIAM Review, 15(3):639–654, 1973.
  • [41] J. Cortés. Discontinuous dynamical systems - a tutorial on solutions, nonsmooth analysis, and stability. IEEE Control Systems, 28(3):36–73, 2008.
  • [42] D. Goeleven and B. Brogliato. Stability and instability matrices for linear evolution variational inequalities. IEEE Transactions on Automatic Control, 49(4):521–534, 2004.
  • [43] S. Boyd and L. Vandenberghe. Convex Optimization. Cambridge University Press, Cambridge, UK, 2009.
  • [44] S. M. Robinson. Stability theory for systems of inequalities. Part I: Linear systems. SIAM Journal on Numerical Analysis, 12(5):754–769, 1975.
  • [45] J. Liu. Sensitivity analysis in nonlinear programs and variational inequalities via continuous selections. SIAM Journal on Control and Optimization, 33(4):1040–1060, 1995.
  • [46] A. Davydov, S. Jafarpour, and F. Bullo. Non-Euclidean contraction theory for robust nonlinear stability. IEEE Transactions on Automatic Control, 67(12):6667–6681, 2022.
  • [47] F. Bullo. Contraction Theory for Dynamical Systems. Kindle Direct Publishing, 1.1 edition, 2023.
  • [48] N. D. Yen. Lipschitz continuity of solutions of variational inequalities with a parametric polyhedral constraint. Mathematics of Operations Research, 20(3):695–708, 1995.
  • [49] H. L. Royden and P. Fitzpatrick. Real Analysis. Prentice Hall, 2010.
  • [50] T. Kose. Solutions of saddle value problems by differential equations. Econometrica, 24(1):59–70, 1956.
  • [51] D. Feijer and F. Paganini. Stability of primal-dual gradient dynamics and applications to network optimization. Automatica, 46:1974–1981, 2010.
  • [52] A. Cherukuri, B. Gharesifard, and J. Cortés. Saddle-point dynamics: conditions for asymptotic stability of saddle points. SIAM Journal on Control and Optimization, 55(1):486–511, 2017.
  • [53] A. Cherukuri, E. Mallada, S. H. Low, and J. Cortés. The role of convexity in saddle-point dynamics: Lyapunov function and robustness. IEEE Transactions on Automatic Control, 63(8):2449–2464, 2018.
  • [54] J. Cortés and S. K. Niederländer. Distributed coordination for nonsmooth convex optimization via saddle-point dynamics. Journal of Nonlinear Science, 29(4):1247–1272, 2019.
  • [55] L. Cothren, F. Bullo, and E. Dall’Anese. Singular perturbation via contraction theory. arXiv preprint arXiv:2310.07966, 2023.
  • [56] S. Kolathaya and A. D. Ames. Input-to-state safety with control barrier functions. IEEE Control Systems Letters, 3(1):108–113, 2018.
  • [57] S. Boyd, N. Parikh, E. Chu, B. Peleato, and J. Eckstein. Distributed optimization and statistical learning via the alternating direction method of multipliers. Foundations and Trends in Machine Learning, 3(1):1–122, 2011.
  • [58] T. Başar and G. J. Olsder. Dynamic Noncooperative Game Theory. SIAM, 2 edition, 1999.
  • [59] M. Pachter and K. D. Pham. Discrete-time linear-quadratic dynamic games. Journal of Optimization Theory & Applications, 146:151–179, 2010.
  • [60] V. Bondarevsky, A. Leschov, and L. Minchenko. Value functions and their directional derivatives in parametric nonlinear programming. Journal of Optimization Theory & Applications, 171(2):440–464, 2016.
  • [61] G. Di Pillo and L. Grippo. Exact penalty functions in constrained optimization. SIAM Journal on Control and Optimization, 27(6):1333–1360, 1989.