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

On universal classes of Lyapunov functions for linear switched systemsthanks: This research was partially supported by the iCODE institute, research project of the Idex Paris-Saclay.

Paolo Mason, Yacine Chitour, and Mario Sigalotti Université Paris-Saclay, CNRS, CentraleSupélec, Laboratoire des signaux et systèmes, 91190, Gif-sur-Yvette, France, paolo.mason@centralesupelec.frUniversité Paris-Saclay, CNRS, CentraleSupélec, Laboratoire des signaux et systèmes, 91190, Gif-sur-Yvette, France, yacine.chitour@centralesupelec.frLaboratoire Jacques-Louis Lions, CNRS, Inria, Sorbonne Université, Université de Paris, France, mario.sigalotti@inria.fr
Abstract

In this paper we discuss the notion of universality for classes of candidate common Lyapunov functions for linear switched systems. On the one hand, we prove that a family of absolutely homogeneous functions is universal as soon as it approximates arbitrarily well every convex absolutely homogeneous function for the C0C^{0} topology of the unit sphere. On the other hand, we prove several obstructions for a class to be universal, showing, in particular, that families of piecewise-polynomial continuous functions whose construction involves at most ll polynomials of degree at most mm (for given positive integers l,ml,m) cannot be universal.

1 Introduction

Common Lyapunov functions constitute the most popular and powerful tool for the stability analysis of switched systems. Roughly speaking, the use of common Lyapunov functions for stability analysis gathers the global behavior of the system and allows to bypass the explicit analysis of single trajectories, which may be extremely complex. Yet, looking for a common Lyapunov function may be a nontrivial task as stability cannot always be checked by means of Lyapunov functions in a simple form, for instance within the class of quadratic forms. Given a family of systems (e.g., the family of all linear switched systems), classes of functions large enough to include a Lyapunov function for each globally asymptotically stable system are called universal [4] and a result establishing the existence of such a class is called a converse Lyapunov theorem. The literature dealing with converse Lyapunov theorems, starting from the works by Massera and Kurzweil in the 1950s (see e.g. [16, 12, 22, 14, 10]) is quite rich. The results concerning the existence of smooth Lyapunov functions for nonlinear systems with global asymptotic stability properties require the development of rather sophisticated techniques. Concerning robust asymptotic stability with respect to a closed invariant set in presence of perturbation terms, converse Lyapunov theorems have been derived in [14]. In the context of switched systems (even in a nonlinear setting) such results establish the equivalence between the global uniform asymptotic stability and the existence of a smooth Lyapunov function. For linear switched systems the construction of a Lyapunov function is much more direct and natural due, essentially, to the homogeneous nature of the system and the equivalence between asymptotic and exponential stability (see e.g. [8]). Furthermore, in the linear case and even for the more general class of uncertain systems, it is well-known that the families of piecewise quadratic functions, polyhedral functions, and homogeneous polynomials are universal [18, 19, 20, 4]. On the other hand, for every positive integer mm, the family of polynomials of degree less or equal than mm is not universal even for the simple class of two-dimensional linear switched systems with two modes [15]. Similarly, it is well accepted in the research community (although, to the authors’ knowledge, no explicit proof is available) that families of piecewise quadratic and polyhedral functions whose construction involves a uniformly bounded number of quadratic or linear functions cannot be universal. For this reason, all numerical methods investigating the existence of Lyapunov functions within these classes are affected by a certain degree of conservativeness.

The contribution of this paper is twofold. First, we provide a general sufficient condition for a class of functions to be universal (Proposition 3.1), which is a formalization of fundamental ideas already present in [18, 19, 20, 4]. As a corollary, we recover the universal classes of functions obtained in these references. We next derive the main results of this paper, which provide some necessary conditions for the universality of classes of functions. The first one, Theorem 4.2, is an abstract result which applies to families of real-valued functions that are analytic outside the origin. The fact that polynomials with a uniform bound mm on their degree do not form a universal class [15] follows as a simple consequence of this result. Finally, Theorem 4.9 states that families of piecewise-polynomial continuous functions whose construction involves at most ll polynomials of degree at most mm (for given positive integers l,ml,m) cannot be universal.

2 Universal classes of common Lyapunov functions

We consider linear switched systems of the form

x˙(t)=A(t)x(t),t0,xn,\dot{x}(t)=A(t)x(t),\qquad t\geq 0,\quad x\in\mathbb{R}^{n}, (Σ)

where the switching law AA is an arbitrary function belonging to the space L(+,)L^{\infty}(\mathbb{R}_{+},\mathscr{M}) of measurable functions from +=[0,+)\mathbb{R}_{+}=[0,+\infty) to a bounded subset \mathscr{M} of the set of n×nn\times n matrices, denoted by Mn()M_{n}(\mathbb{R}). We use ΦA(t,s)\Phi_{A}(t,s) to denote the fundamental matrix from ss to tt for (Σ)(\Sigma_{\mathscr{M}}) associated with the switching law AA so that every solution of (Σ)(\Sigma_{\mathscr{M}}) can be written as x(t)=ΦA(t,0)x(0)x(t)=\Phi_{A}(t,0)x(0). Notice that ΦA(t,s)\Phi_{A}(t,s) exists for every AL(+,)A\in L^{\infty}(\mathbb{R}_{+},\mathscr{M}) and every s,t0s,t\geq 0 (see e.g. [5, Theorem 3.7]). We are interested in the following uniform stability properties.

Definition 2.1.

The switched system (Σ)(\Sigma_{\mathscr{M}}) is said to be

  • uniformly stable if there exists C>0C>0 such that, for every switching law AA and t0t\geq 0, ΦA(t,0)C\|\Phi_{A}(t,0)\|\leq C;

  • uniformly exponentially stable if there exist C,γ>0C,\gamma>0 such that, for every switching law AA and t0t\geq 0, ΦA(t,0)Ceγt\|\Phi_{A}(t,0)\|\leq Ce^{-\gamma t}.

Stability in the previous senses may be assessed through common Lyapunov functions, defined below.

Definition 2.2.

We say that a continuous function V:n+V:\mathbb{R}^{n}\longrightarrow\mathbb{R}_{+} is a nonstrict common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}) if it is positive definite, that is, V(0)=0V(0)=0 and V(x)>0V(x)>0 for every x0x\neq 0, and VV is non-increasing along each trajectory of (Σ)(\Sigma_{\mathscr{M}}). If, moreover, VV is strictly decreasing along each nonzero trajectory of (Σ)(\Sigma_{\mathscr{M}}), we say that VV is a common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}).

Remark 2.3.

If VV is a (possibly nonstrict) common Lyapunov function and φ:++\varphi:\mathbb{R}_{+}\to\mathbb{R}_{+} is continuous, strictly increasing, and such that φ(0)=0\varphi(0)=0, then φV\varphi\circ V is also a (nonstrict) common Lyapunov function. In particular, the positive multiple of a common Lyapunov function is a common Lyapunov function and if there exists an absolutely homogeneous common Lyapunov function111Given α>0\alpha>0, a function V:nV:\mathbb{R}^{n}\to\mathbb{R} is said to be absolutely homogeneous of degree α\alpha if V(λx)=|λ|αV(x)V(\lambda x)=|\lambda|^{\alpha}V(x) for every xnx\in\mathbb{R}^{n} and λ\lambda\in\mathbb{R}. If α=1\alpha=1 we simply say that VV is absolutely homogeneous., then for every α>0\alpha>0 there exists an absolutely homogeneous common Lyapunov function of degree α\alpha.

We state here the classical direct Lyapunov theorem in the linear switched case (see, e.g., [13, Theorem 2.1] or [9, Theorem 4.2] for a formulation involving merely continuous Lyapunov functions).

Theorem 2.4.

A linear switched system (Σ)(\Sigma_{\mathscr{M}}) admitting a nonstrict common Lyapunov function is uniformly stable. If there exists a strict common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}), then the latter is uniformly exponentially stable.

Remark 2.5.

In Definition 2.2 we do not require Lyapunov functions to be proper, i.e., such that the corresponding sublevel sets are compact. This is motivated by the fact that the existence a Lyapunov function in the sense of Definition 2.2 implies local stability of (Σ)(\Sigma_{\mathscr{M}}) and hence, by linearity of the system, its global stability. As a matter of fact, given a possibly non-proper Lyapunov function VV for (Σ)(\Sigma_{\mathscr{M}}), a proper Lyapunov function may always be constructed by considering the Minkowski functional associated with a sublevel set of VV, see e.g. [3].

In case the strict common Lyapunov function V:n+V:\mathbb{R}^{n}\to\mathbb{R}_{+} in the above theorem is of class 𝒞1\mathcal{C}^{1} on n{0}\mathbb{R}^{n}\setminus\{0\}, a standard test for checking the strict decrease of VV along non-trivial trajectories of (Σ)(\Sigma_{\mathscr{M}}) goes as follows:

V(x)Mx<0,M,xn{0}.\nabla V(x)^{\top}Mx<0,\qquad\forall M\in\mathscr{M},\quad\forall x\in\mathbb{R}^{n}\setminus\{0\}. (1)
Definition 2.6.

A set 𝒫\mathcal{P} of functions from n\mathbb{R}^{n} to \mathbb{R} is a universal class of Lyapunov functions if for every bounded set Mn()\mathscr{M}\subset M_{n}(\mathbb{R}) such that (Σ)(\Sigma_{\mathscr{M}}) is uniformly exponentially stable there exists a common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}) in 𝒫\mathcal{P}.

An equivalent formulation of the universality of a class 𝒫\mathcal{P} is that the converse Lyapunov theorem holds true within 𝒫\mathcal{P}.

As mentioned in the introduction, the theoretical construction of a common Lyapunov function for linear switched systems can be easily obtained. For instance, a locally Lipschitz continuous Lyapunov function may be defined as

V(x)=supAL(+,)0+ΦA(t,0)x𝑑t,V(x)=\sup_{A\in L^{\infty}(\mathbb{R}_{+},\mathscr{M})}\int_{0}^{+\infty}\|\Phi_{A}(t,0)x\|dt, (2)

and may be regularized outside the origin by convolution with a smooth function. This classical construction leads to the following result (see e.g. [17, 8, 15]).

Proposition 2.7.

Let α1\alpha\geq 1 and 𝒫\mathcal{P} be the class of convex absolutely homogeneous functions of degree α\alpha on n\mathbb{R}^{n} that are positive and smooth on n{0}\mathbb{R}^{n}\setminus\{0\}. Then 𝒫\mathcal{P} is universal. Moreover, for every bounded set Mn()\mathscr{M}\subset M_{n}(\mathbb{R}) such that (Σ)(\Sigma_{\mathscr{M}}) is uniformly exponentially stable, there exists ε>0\varepsilon>0 and a common Lyapunov function V𝒫V\in\mathcal{P} for (Σ)(\Sigma_{\mathscr{M}}) such that V(x)Mxxα\nabla V(x)^{\top}Mx\leq-\|x\|^{\alpha} for every xn{0}x\in\mathbb{R}^{n}\setminus\{0\} and MM\in\mathscr{M}.

Note that a globally smooth Lyapunov function may be constructed by classical regularization techniques developed in a nonlinear setting (see e.g. [12, 14]), at the price of losing homogeneity.

Similar to Proposition 2.7, the following converse Lyapunov result links the uniform stability of (Σ)(\Sigma_{\mathscr{M}}) with the existence of a nonstrict common Lyapunov function, see e.g. [11, Theorem 2.2].

Proposition 2.8.

Assume that Mn()\mathscr{M}\subset M_{n}(\mathbb{R}) is bounded and (Σ)(\Sigma_{\mathscr{M}}) is uniformly stable. Then the function

V(x)=supt0,AL(+,)ΦA(t,0)xV(x)=\sup_{t\geq 0,A\in L^{\infty}(\mathbb{R}_{+},\mathscr{M})}\|\Phi_{A}(t,0)x\| (3)

is absolutely homogeneous and a nonstrict common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}).

Due to Proposition 2.7, the continuous differentiability outside the origin of the common Lyapunov function is not a restrictive assumption when checking the uniform exponential stability of a linear switched system. On the other hand, the uniform stability of (Σ)(\Sigma_{\mathscr{M}}) does not always imply the existence of a 𝒞1\mathcal{C}^{1} nonstrict common Lyapunov function (see e.g. [7, Example 3]). Furthermore, even in case of uniform exponential stability, it may be useful to provide a criterion to ensure the existence of a common Lyapunov function in a class of non-differentiable functions, such as piecewise linear or piecewise quadratic ones. For these reasons we introduce below a criterion which generalizes Equation (1) and characterizes the family of (possibly nonstrict) common Lyapunov functions in a nonsmooth setting. We refer to [1, Proposition 1] for a similar result in the context of differential inclusions.

We need the following preliminary result, which expresses the variation of a convex function VV along a trajectory in terms of the subdifferential of VV. Recall that the subdifferential V(x)\partial V(x) at a point xnx\in\mathbb{R}^{n} is defined as

V(x)={lnl(yx)V(y)V(x),yn}.\partial V(x)=\{l\in\mathbb{R}^{n}\mid l^{\top}(y-x)\leq V(y)-V(x),\quad\forall y\in\mathbb{R}^{n}\}.

The proof of the lemma is similar to that of [1, Lemma 1] and is provided here for completeness.

Lemma 2.9.

Let V:nV:\mathbb{R}^{n}\to\mathbb{R} be a convex function and φ:In\varphi:I\to\mathbb{R}^{n} be an absolutely continuous function, with II\subseteq\mathbb{R} an open interval. Then VφV\circ\varphi is absolutely continuous and it holds

ddtV(φ(t))=lφ˙(t),lV(φ(t)),for a.e. tI.\frac{d}{dt}V(\varphi(t))=l^{\top}\dot{\varphi}(t),\quad\forall l\in\partial V(\varphi(t)),\quad\hbox{for a.e. }t\in I.
Proof.

As VV is convex, VV is Lipschitz and the composition VφV\circ\varphi is absolutely continuous. Hence for almost every tIt\in I the derivatives of both φ\varphi and VφV\circ\varphi are well-defined. By definition of subdifferential, for every t,sIt,s\in I and lV(φ(t))l\in\partial V(\varphi(t)) we have

l(φ(s)φ(t))V(φ(s))V(φ(t)).l^{\top}(\varphi(s)-\varphi(t))\leq V(\varphi(s))-V(\varphi(t)).

We deduce that

ddtV(φ(t))\displaystyle\frac{d}{dt}V(\varphi(t)) =limst+V(φ(s))V(φ(t))st\displaystyle=\lim_{s\to t^{+}}\frac{V(\varphi(s))-V(\varphi(t))}{s-t}
llimst+φ(s)φ(t)st\displaystyle\geq l^{\top}\lim_{s\to t^{+}}\frac{\varphi(s)-\varphi(t)}{s-t}
=lφ˙(t)\displaystyle=l^{\top}\dot{\varphi}(t)

holds true for almost every tIt\in I and for every lV(φ(t))l\in\partial V(\varphi(t)). Similarly, taking the limit as sts\to t^{-}, we obtain that ddtV(φ(t))lφ˙(t)\frac{d}{dt}V(\varphi(t))\leq l^{\top}\dot{\varphi}(t) for almost every tIt\in I and for every lV(φ(t))l\in\partial V(\varphi(t)). This concludes the proof of the lemma. ∎

Here follows an adaptation to the nonsmooth setting of the characterization of common Lyapunov function.

Proposition 2.10.

Let \mathscr{M} be a bounded subset of Mn()M_{n}(\mathbb{R}) and V:n+V:\mathbb{R}^{n}\to\mathbb{R}_{+} be a convex positive definite function. Then VV is a nonstrict common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}) if and only if

lMx0,xn{0},lV(x),M.l^{\top}Mx\leq 0,\qquad\forall x\in\mathbb{R}^{n}\setminus\{0\},\forall l\in\partial V(x),\forall M\in\mathscr{M}. (4)

Moreover, if the inequality in (4) is strict then VV is a common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}).

Proof.

The second part of the proposition and the if implication in the first part directly follow from Lemma 2.9. We are left to show that if VV is a nonstrict common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}), then the inequality (4) holds true. By contradiction, suppose that there exist xn,lV(x)x\in\mathbb{R}^{n},l\in\partial V(x), and MM\in\mathscr{M} such that lMx>0l^{\top}Mx>0. By [21, Theorem 25.6] one may find a differentiability point yy of VV such that the pair (y,V(y))(y,\nabla V(y)) is arbitrarily close to (x,l)(x,l). In particular we may assume V(y)My>0\nabla V(y)^{\top}My>0, that is, VV is increasing at t=0t=0 along the trajectory tetMyt\mapsto e^{tM}y, leading to a contradiction. ∎

3 Sufficient conditions for universality

Given a linear switched system (Σ)(\Sigma_{\mathscr{M}}), the family 𝒫\mathcal{P} identified by Proposition 2.7 is too broad to admit a tractable parameterization, suitable for investigating numerically the existence of a Lyapunov function. With this goal in mind, interesting candidate classes 𝒫\mathcal{P} are those parametric families of functions for which the property of being positive definite and strictly decreasing along all admissible dynamics can be translated into numerically verifiable algebraic relations or inequalities (e.g., linear matrix inequalities). It is well-known that piecewise-quadratic, polynomial, and polyhedral functions represent examples of such families [18, 19, 20, 4].

We next provide a general sufficient condition for a class 𝒫\mathcal{P} to be universal, namely its density in the class of convex absolutely homogeneous functions for the topology of uniform convergence on compact sets. The proof of the sufficient condition exploits the fact, specific to convex functions defined on compact sets, that being close in the uniform norm is equivalent to possessing “close” subdifferentials.

Proposition 3.1.

Let 𝒫\mathcal{P} be a subset of the family of convex absolutely homogeneous functions from n\mathbb{R}^{n} to +\mathbb{R}_{+}. Assume that for every convex absolutely homogeneous function V:n+V:\mathbb{R}^{n}\to\mathbb{R}_{+} and every δ>0\delta>0 there exists a function WW in 𝒫\mathcal{P} such that W(x)V(x)δ\|W(x)-V(x)\|\leq\delta for every xx in the unit sphere Sn1S^{n-1} of n\mathbb{R}^{n}. Then 𝒫\mathcal{P} is a universal class of Lyapunov functions.

Proof.

Let (Σ)(\Sigma_{\mathscr{M}}) be uniformly exponentially stable. Let VV be the absolutely homogeneous common Lyapunov function provided by Proposition 2.7. In particular, VV is smooth on n{0}\mathbb{R}^{n}\setminus\{0\}. In order to prove the proposition, it is enough to show that any convex absolutely homogeneous function close enough to VV on Sn1S^{n-1} in uniform norm is itself a Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}).

We proceed by contradiction: we assume that there exists a sequence of convex absolutely homogeneous functions (Wk)k(W_{k})_{k\in\mathbb{N}} converging uniformly to VV on Sn1S^{n-1} as kk goes to infinity and such that each WkW_{k} is not strictly decreasing along at least one trajectory of the system. In particular the derivative of WkW_{k} along such a trajectory is nonnegative on a set of times of positive measure. By Lemma 2.9 and absolute homogeneity of WkW_{k}, we deduce that there exist xkSn1x_{k}\in S^{n-1} and MkM_{k}\in\mathscr{M} such that, for every fixed lkWk(xk)l_{k}\in\partial W_{k}(x_{k}), one has lkMkxk0l_{k}^{\top}M_{k}x_{k}\geq 0. By compactness, we may assume that xkx_{k} tends to x¯Sn1\bar{x}\in S^{n-1} as kk goes to infinity. Then, by [21, Theorem 24.5], lkl_{k} converges to V(x¯)\nabla V(\bar{x}), so that, by boundedness of \mathscr{M}, lim supkV(x¯)Mkx¯=lim supklkMkxk0\limsup_{k\to\infty}\nabla V(\bar{x})^{\top}M_{k}\bar{x}=\limsup_{k\to\infty}l_{k}^{\top}M_{k}x_{k}\geq 0. However, it follows by the choice of VV and Proposition 2.7 that V(x¯)Mkx¯1\nabla V(\bar{x})^{\top}M_{k}\bar{x}\leq-1, yielding a contradiction. ∎

Remark 3.2.

By the absolute homogeneity property, the statement of Proposition 3.1 could be equivalently reformulated by fixing δ=1\delta=1.

As an application of the previous result, two classical examples of universal classes of Lyapunov functions (cf. [4, 18, 19, 20]) are recalled in the following corollary.

Corollary 3.3.

The family of polyhedral functions {maxk=1,,N|lkx|lkn,N}\{\max_{k=1,\dots,N}|l_{k}^{\top}x|\mid l_{k}\in\mathbb{R}^{n},\,N\in\mathbb{N}\} and that of homogeneous sums of squares {k=1N(lkx)2dlkn,d,N}\{\sum_{k=1}^{N}(l_{k}^{\top}x)^{2d}\mid l_{k}\in\mathbb{R}^{n},\,d,N\in\mathbb{N}\} are universal classes of Lyapunov functions.

Proof.

Let VV be a convex absolutely homogeneous function. Let (xi)i(x_{i})_{i\in\mathbb{N}} be a dense sequence in Sn1S^{n-1}, and liV(xi)l_{i}\in\partial V(x_{i}). We consider the increasing sequence of absolutely homogeneous functions defined by

Wi(x)=maxj=1,,i|ljx|,xn.W_{i}(x)=\max_{j=1,\dots,i}|l_{j}^{\top}x|,\quad\forall x\in\mathbb{R}^{n}.

Observe that each WiW_{i} is convex and Wi(x)V(x)W_{i}(x)\leq V(x) for every xnx\in\mathbb{R}^{n}. Indeed

|ljx|\displaystyle|l_{j}^{\top}x| =max{ljx,lj(x)}\displaystyle=\max\{l_{j}^{\top}x,l_{j}^{\top}(-x)\}
=max{lj(xxj),lj(xxj)}+ljxj\displaystyle=\max\{l_{j}^{\top}(x-x_{j}),l_{j}^{\top}(-x-x_{j})\}+l_{j}^{\top}x_{j}
V(x)V(xj)+ljxj\displaystyle\leq V(x)-V(x_{j})+l_{j}^{\top}x_{j}
=V(x),\displaystyle=V(x),

for every positive integer jj, by definition of subgradient and since V(x)=V(x)V(x)=V(-x) and V(xj)=ljxjV(x_{j})=l_{j}^{\top}x_{j}. We deduce that Wi(xk)=V(xk)W_{i}(x_{k})=V(x_{k}) for every kk\in\mathbb{N} and iki\geq k, hence limiWi(xk)=V(xk)\lim_{i\to\infty}W_{i}(x_{k})=V(x_{k}) and we can apply [21, Theorem 10.8] to conclude that the sequence of functions WiW_{i} converges to VV uniformly on Sn1S^{n-1}. By applying Proposition 3.1 we get that the family of polyhedral functions is a universal class of Lyapunov functions.

Let us now consider the absolutely homogeneous functions

Zi(x)=(j=1i|ljx|2i)12i.Z_{i}(x)=\left(\sum_{j=1}^{i}|l_{j}^{\top}x|^{2i}\right)^{\frac{1}{2i}}.

The function ZiZ_{i} is convex since it is the composition of the 2i2i-norm on i\mathbb{R}^{i}, i.e., y2i=(j=1iyj2i)12i\|y\|_{2i}=\left(\sum_{j=1}^{i}y_{j}^{2i}\right)^{\frac{1}{2i}} for yiy\in\mathbb{R}^{i}, with the linear function from n\mathbb{R}^{n} to i\mathbb{R}^{i} mapping xx to (l1x,,lix)(l_{1}^{\top}x,\dots,l_{i}^{\top}x)^{\top}. Moreover it is immediate to see that Wi(x)Zi(x)i12iWi(x)W_{i}(x)\leq Z_{i}(x)\leq i^{\frac{1}{2i}}W_{i}(x), and in particular

Zi(x)V(x)Zi(x)Wi(x)+Wi(x)V(x)\|Z_{i}(x)-V(x)\|\leq\|Z_{i}(x)-W_{i}(x)\|+\|W_{i}(x)-V(x)\|

tends to zero uniformly on Sn1S^{n-1} as ii goes to infinity. By applying again Proposition 3.1, it follows that the family of homogeneous sums of squares is a universal class of Lyapunov functions. ∎

Remark 3.4.

According to Remark 2.3, the first part of Corollary 3.3 remains valid if one replaces the piecewise linear functions maxk=1,,N|lkx|\max_{k=1,\dots,N}|l_{k}^{\top}x| with the functions (maxk=1,,N|lkx|)q=maxk=1,,N|lkx|q\big{(}\max_{k=1,\dots,N}|l_{k}^{\top}x|\big{)}^{q}=\max_{k=1,\dots,N}|l_{k}^{\top}x|^{q}, for any given q>1q>1. In particular, for q=2q=2, we have that the family of piecewise quadratic functions is a universal class of Lyapunov functions.

Remark 3.5.

The proof of Proposition 3.1 relies on the fact that, whenever a linear switched system is uniformly exponentially stable, there exists a common Lyapunov function which is convex. In the classical construction (2), convexity and homogeneity of the Lyapunov function are direct consequences of the convexity and homogeneity of the map x0ΦA(t,0)x0x_{0}\mapsto\|\Phi_{A}(t,0)x_{0}\| for given t0t\geq 0 and AL(+,)A\in L^{\infty}(\mathbb{R}_{+},\mathscr{M}). In the nonlinear case, the homogeneity property can be recovered from a homogeneity assumption on the vector fields. Similarly, the convexity property can be imposed as an additional requirement. Proposition 3.1, and hence Corollary 3.3, can then be extended to any class of nonlinear switched systems

x˙(t)=fσ(t)(x(t)),σL(+,Σ),\dot{x}(t)=f_{\sigma(t)}(x(t)),\qquad\sigma\in L^{\infty}(\mathbb{R}_{+},\Sigma),

with Σm\Sigma\subset\mathbb{R}^{m} a bounded set of parameters, satisfying the following conditions:

  • fσf_{\sigma} is homogeneous of degree one for every σΣ\sigma\in\Sigma, that is fσ(λx)=λfσ(x)f_{\sigma}(\lambda x)=\lambda f_{\sigma}(x) for every xnx\in\mathbb{R}^{n} and λ\lambda\in\mathbb{R},

  • denoting by x(t,x0,σ())x(t,x_{0},\sigma(\cdot)) the solution at time tt of the system starting at x0x_{0} and corresponding to the switching law σL(+,Σ)\sigma\in L^{\infty}(\mathbb{R}_{+},\Sigma), the function x0x(t,x0,σ())x_{0}\mapsto\|x(t,x_{0},\sigma(\cdot))\| is convex,

  • for every R>0R>0, {fσ|B(0,R)σΣ}\{f_{\sigma}|_{B(0,R)}\mid\sigma\in\Sigma\} is a bounded subset of 𝒞(B(0,R),n)\mathcal{C}(B(0,R),\mathbb{R}^{n}).

The last condition replaces the boundedness of \mathscr{M} which is required in the proofs of both Propositions 2.7 and 3.1.

4 Necessary conditions for universality

Next, we provide restrictions on the classes of functions which may be candidate to be universal. For this purpose, we introduce the following technical result.

Lemma 4.1.

Let 1,2\mathscr{M}_{1},\mathscr{M}_{2} be bounded subsets of Mn()M_{n}(\mathbb{R}) and assume that (Σ1)(\Sigma_{\mathscr{M}_{1}}) is uniformly stable. For ν>0\nu>0, denote by 2ν\mathscr{M}_{2}^{\nu} the set of matrices of the form MνIdnM-\nu\mathrm{Id}_{n} for M2M\in\mathscr{M}_{2}, where Idn\mathrm{Id}_{n} is the n×nn\times n identity matrix. Set =12ν\mathscr{M}=\mathscr{M}_{1}\cup\mathscr{M}_{2}^{\nu}. Then, the switched system (Σ)(\Sigma_{\mathscr{M}}) is uniformly stable for ν>0\nu>0 large enough.

Proof.

Proposition 2.8 guarantees the existence of a convex absolutely homogeneous nonstrict Lyapunov function VV for (Σ1)(\Sigma_{\mathscr{M}_{1}}). Intuitively speaking, the lemma follows from the fact that, for λ\lambda large enough, the vectors MxMx, with M2λM\in\mathscr{M}_{2}^{\lambda} and xnx\in\mathbb{R}^{n}, point towards the interior of the sublevel set V1([0,V(x)])V^{-1}([0,V(x)]). Let us formalize this idea. Since lx=V(x)l^{\top}x=V(x) whenever lV(x)l\in\partial V(x), by boundedness of 2\mathscr{M}_{2} and of xV1(1)V(x)\cup_{x\in V^{-1}(1)}\partial V(x) [21, Theorem 24.7], for λ>0\lambda>0 large enough one has l(MλIdn)x=lMxλlx=lMxλ<0l^{\top}(M-\lambda\mathrm{Id}_{n})x=l^{\top}Mx-\lambda l^{\top}x=l^{\top}Mx-\lambda<0 for every M2M\in\mathscr{M}_{2}, xV1(1)x\in V^{-1}(1), and lV(x)l\in\partial V(x). The result is then an immediate consequence of Proposition 2.10. ∎

Theorem 4.2.

Let n2n\geq 2 and 𝒫\mathcal{P} be a compact subset of the space of continuous functions from n\mathbb{R}^{n} to \mathbb{R} that are analytic on n{0}\mathbb{R}^{n}\setminus\{0\}, endowed with the topology of uniform convergence on bounded sets. Assume that 𝒫\mathcal{P} does not contain the zero function. Then 𝒫\mathcal{P} cannot be a universal class of Lyapunov functions.

Proof.

We start by showing the theorem in the case n=2n=2. We proceed by contradiction, assuming that every uniformly exponentially stable switched system in the case where \mathscr{M} consists of two matrices in M2()M_{2}(\mathbb{R}) admits a Lyapunov function in 𝒫\mathcal{P}. We consider a switched system corresponding to 0={M10,M20}M2()\mathscr{M}^{0}=\{M^{0}_{1},M^{0}_{2}\}\subset M_{2}(\mathbb{R}), where M10,M20M^{0}_{1},M^{0}_{2} are Hurwitz, the corresponding trajectories rotate clockwise around the origin, the system is uniformly stable, but not attractive, and starting from every initial nonzero condition there exists a unique periodic trajectory, with four switches per period. The existence of such a system is obtained in [2, Theorem 1], where it corresponds to the case 𝐒𝟒\mathbf{S4}, =1\mathcal{R}=1. In particular, there exist t1,t2>0t_{1},t_{2}>0 such that et1M10et2M20e^{t_{1}M^{0}_{1}}e^{t_{2}M^{0}_{2}} has an eigenvalue equal to 1-1, corresponding to an eigenvector x0x_{0}. Set T=t1+t2T=t_{1}+t_{2} and consider the switched systems associated with ε={M1ε,M2ε}M2()\mathscr{M}^{\varepsilon}=\{M^{\varepsilon}_{1},M^{\varepsilon}_{2}\}\subset M_{2}(\mathbb{R}), where Miε=Mi0εId2M^{\varepsilon}_{i}=M^{0}_{i}-\varepsilon\mathrm{Id}_{2} for i=1,2i=1,2. For ε0\varepsilon\geq 0 we consider the TT-periodic switching sequence Aε()A^{\varepsilon}(\cdot) which takes values M1εM^{\varepsilon}_{1} for t[0,t1)t\in[0,t_{1}) and M2εM^{\varepsilon}_{2} for t[t1,T)t\in[t_{1},T).

Since every trajectory x()x(\cdot) of (Σε)(\Sigma_{\mathscr{M}^{\varepsilon}}) can be written as teεty(t)t\mapsto e^{-\varepsilon t}y(t) where y()y(\cdot) is a trajectory of (Σ)(\Sigma_{\mathscr{M}}), then (Σε)(\Sigma_{\mathscr{M}^{\varepsilon}}) is uniformly exponentially stable for ε>0\varepsilon>0. Hence, by assumption, it admits a Lyapunov function Vε()V_{\varepsilon}(\cdot) in 𝒫\mathcal{P}. Since the latter is compact, there exists a sequence (εk)k(\varepsilon_{k})_{k\in\mathbb{N}} converging to zero such that (Vεk)k(V_{\varepsilon_{k}})_{k\in\mathbb{N}} converges to some V¯𝒫\bar{V}\in\mathcal{P}. Moreover, for every t0t\geq 0,

V¯(ΦA0(t,0)x0)\displaystyle\bar{V}(\Phi_{A^{0}}(t,0)x_{0}) =limkVεk(ΦAεk(t,0)x0)\displaystyle=\lim_{k\to\infty}V_{\varepsilon_{k}}(\Phi_{A^{\varepsilon_{k}}}(t,0)x_{0})
limkVεk(x0)=V¯(x0).\displaystyle\leq\lim_{k\to\infty}V_{\varepsilon_{k}}(x_{0})=\bar{V}(x_{0}).

Since V¯(ΦA0(2T,0)x0)=V¯(x0)\bar{V}(\Phi_{A^{0}}(2T,0)x_{0})=\bar{V}(x_{0}) we deduce that V¯\bar{V} is constant along the trajectory ΦA0(,0)x0\Phi_{A^{0}}(\cdot,0)x_{0}. The function tV¯(etM10x0)t\mapsto\bar{V}(e^{tM^{0}_{1}}x_{0}) is analytic for x00x_{0}\neq 0, being the composition of analytic functions, and it is constantly equal to V¯(x0)\bar{V}(x_{0}) for t[0,t1]t\in[0,t_{1}]. By analyticity, tV¯(etM10x0)t\mapsto\bar{V}(e^{tM^{0}_{1}}x_{0}) is constant for all t>0t>0 and therefore it must be identically equal to 0 since V¯(limtetM10x0)=V¯(0)=0\bar{V}(\lim_{t\to\infty}e^{tM^{0}_{1}}x_{0})=\bar{V}(0)=0. Since every nonzero point of 2\mathbb{R}^{2} may be written as μetM10x0\mu e^{tM^{0}_{1}}x_{0} for some positive μ\mu and tt, we deduce that V¯\bar{V} must be identically zero, contradicting the assumptions on 𝒫\mathcal{P}.

We are left to prove the result for n>2n>2. For this purpose we consider 1={M¯10,M¯20}\mathscr{M}_{1}=\{\bar{M}^{0}_{1},\bar{M}^{0}_{2}\} with

M¯i0=(Mi000Idn2),\bar{M}^{0}_{i}=\left(\begin{array}[]{cc}M^{0}_{i}&0\\ 0&-\mathrm{Id}_{n-2}\end{array}\right),

where the matrices Mi0M^{0}_{i} are defined as above. Let λ>0\lambda>0 and 2λ\mathscr{M}_{2}^{\lambda} be given by Lemma 4.1 with 2={Mso(n)M1}\mathscr{M}_{2}=\{M\in{\rm so}(n)\mid\|M\|\leq 1\}, where so(n){\rm so}(n) denotes the space of skew-symmetric n×nn\times n matrices. Define ¯0=12λ\bar{\mathscr{M}}^{0}=\mathscr{M}_{1}\cup\mathscr{M}_{2}^{\lambda} and, for ε>0\varepsilon>0, consider the switched system corresponding to ¯ε=¯0εIdn\bar{\mathscr{M}}^{\varepsilon}=\bar{\mathscr{M}}^{0}-\varepsilon\mathrm{Id}_{n}. It is clear that (Σ¯ε)(\Sigma_{\bar{\mathscr{M}}^{\varepsilon}}) is uniformly exponentially stable for every ε>0\varepsilon>0.

Letting Π1,2\Pi_{1,2} be the (x1,x2)(x_{1},x_{2}) plane, i.e., Π1,2={xnx3==xn=0}\Pi_{1,2}=\{x\in\mathbb{R}^{n}\mid x_{3}=\dots=x_{n}=0\}, we notice that, starting from every x¯Π1,2\bar{x}\in\Pi_{1,2}, there exists a periodic trajectory of (Σ¯0)(\Sigma_{\bar{\mathscr{M}}^{0}}) lying on Π1,2\Pi_{1,2}. The restrictions of functions in 𝒫\mathcal{P} to Π1,2\Pi_{1,2} form a compact set of functions on Π1,2\Pi_{1,2} that are analytic outside the origin. As in the case n=2n=2, we prove by contradiction that 𝒫\mathcal{P} is not universal. Assume that there exists a sequence (Vεk())k(V_{\varepsilon_{k}}(\cdot))_{k\in\mathbb{N}} of Lyapunov functions in 𝒫\mathcal{P} for (Σ¯εk)(\Sigma_{\bar{\mathscr{M}}^{\varepsilon_{k}}}) converging to V¯𝒫\bar{V}\in\mathcal{P}. We can show as before that V¯\bar{V} is equal to 0 on Π1,2\Pi_{1,2}. Because of the choice of 2\mathscr{M}_{2} and by construction of ¯0\bar{\mathscr{M}}^{0}, every 11-dimensional linear subspace of n\mathbb{R}^{n} may be reached in finite time from Π1,2\Pi_{1,2} via a trajectory of (Σ¯0)(\Sigma_{\bar{\mathscr{M}}^{0}}). Since V¯\bar{V} is non-increasing along such a trajectory, we deduce that V¯0\bar{V}\equiv 0 on n\mathbb{R}^{n}, obtaining a contradiction. ∎

Remark 4.3.

The assumption that the zero function is not in 𝒫\mathcal{P} cannot be removed from the hypotheses of Theorem 4.2. Indeed, consider the subset of polynomial functions made of the zero polynomial and, for every N0N\geq 0, the polynomials of degree NN with absolute value of the coefficients upper bounded by a positive constant cNc_{N}, chosen in such a way that the supremum on the ball of radius NN of the polynomial is less than or equal to 1/(N+1)1/(N+1). Since the class 𝒫\mathcal{P} contains a multiple of any polynomial, it is universal by Corollary 3.3. It is also compact since every sequence in 𝒫\mathcal{P} admits a subsequence with either degree going to infinity or constant degree. In the former case, the subsequence converges to zero for the topology of uniform convergence on bounded sets, while in the latter one the coefficients are uniformly bounded and hence the sequence admits a further converging subsequence.

Example 4.4.

Consider the class of absolutely homogeneous functions of degree two

𝒫={xexQ1xx2xQ2xQ11,Q2=1}.\mathcal{P}=\left\{x\mapsto e^{\frac{x^{\top}Q_{1}x}{\|x\|^{2}}}x^{\top}Q_{2}x\mid\|Q_{1}\|\leq 1,\,\|Q_{2}\|=1\right\}.

The level sets of each element of 𝒫\mathcal{P} are obtained by deforming those of the quadratic functions xxQ2xx\mapsto x^{\top}Q_{2}x by a positive 0-homogeneous term.

As each function in 𝒫\mathcal{P} is analytic outside the origin, by Theorem 4.2 the class 𝒫\mathcal{P} is not universal, despite being richer than that of quadratic functions.

As a consequence of Theorem 4.2 we obtain the following corollary which provides, in particular, a partial counterpart to Corollary 3.3 for homogeneous polynomial functions. Namely, we recover that, if we impose a uniform bound on the degree, such functions do not form a universal class of Lyapunov functions, as already established in [15].

Corollary 4.5.

Let n2n\geq 2 and 𝒫\mathcal{P} be a finite-dimensional vector subspace of the space of continuous functions from n\mathbb{R}^{n} to \mathbb{R} that are analytic on n{0}\mathbb{R}^{n}\setminus\{0\}. Then 𝒫\mathcal{P} is not universal. In particular, for every positive integer mm, the set of polynomial functions of degree at most mm from n\mathbb{R}^{n} to \mathbb{R} is not a universal class of Lyapunov functions.

Proof.

Let {f1,,fN}\{f_{1},\dots,f_{N}\} be a basis of 𝒫\mathcal{P}. The linear map φ:N𝒫\varphi:\mathbb{R}^{N}\to\mathcal{P} defined as φ(α)=i=1Nαifi\varphi(\alpha)=\sum_{i=1}^{N}\alpha_{i}f_{i} is continuous when 𝒫\mathcal{P} is endowed with the topology of uniform convergence on compact sets. Indeed, on every compact set KnK\subset\mathbb{R}^{n} and for every α1,α2N\alpha^{1},\alpha^{2}\in\mathbb{R}^{N}, one has

φ(α1)|Kφ(α2)|Kα1α2maxi=1,,Nfi|K.\|\varphi(\alpha^{1})|_{K}-\varphi(\alpha^{2})|_{K}\|_{\infty}\leq\|\alpha^{1}-\alpha^{2}\|\max_{i=1,\dots,N}\|f_{i}|_{K}\|_{\infty}.

Hence the image of the unit sphere via the map φ\varphi is a compact set 𝒫¯𝒫\bar{\mathcal{P}}\subset\mathcal{P}. Furthermore, 𝒫¯\bar{\mathcal{P}} does not contain the zero function because {f1,,fN}\{f_{1},\dots,f_{N}\} is a linearly independent subset of 𝒫\mathcal{P}. Applying Theorem 4.2 we obtain that 𝒫¯\bar{\mathcal{P}} is not universal. As each element of 𝒫\mathcal{P} is a scalar multiple of an element of 𝒫¯\bar{\mathcal{P}} we deduce that 𝒫\mathcal{P} is not universal either, concluding the proof of the first part of the corollary. Concerning the second part, it is enough to observe that the set of polynomial functions of degree at most mm from n\mathbb{R}^{n} to \mathbb{R} is a finite-dimensional vector space of analytic functions. ∎

Remark 4.6.

In order to avoid confusion, let us stress that a finite-dimensional space as the set 𝒫\mathcal{P} appearing in the statement of Corollary 4.5 obviously contains {0}\{0\}. This is not in contradiction with Theorem 4.2, since the latter is applied to 𝒫¯\bar{\mathcal{P}} and not to 𝒫\mathcal{P} in our proof of the corollary.

Example 4.7.

Let n=2n=2, fix a positive integer mm, and consider the vector space 𝒫\mathcal{P} of functions from 2\mathbb{R}^{2} to \mathbb{R} defined, in polar coordinates, by

(r,θ)ra0+ri=1M(aicos(2iθ)+bisin(2iθ)),(r,\theta)\mapsto ra_{0}+r\sum_{i=1}^{M}(a_{i}\cos(2i\theta)+b_{i}\sin(2i\theta)),

where a0,,am,b1,,bma_{0},\dots,a_{m},b_{1},\dots,b_{m} are in \mathbb{R}. Each function in 𝒫\mathcal{P} is analytic outside the origin of 2\mathbb{R}^{2} and absolutely homogeneous.

By Corollary 4.5 the class 𝒫\mathcal{P} is not universal.

Example 4.8.

Let n=2n=2 and consider the matrices

M1=(0.1110.1),M2=(0.1α1α0.1),M_{1}=\begin{pmatrix}-0.1&-1\\ 1&-0.1\end{pmatrix},\qquad M_{2}=\begin{pmatrix}-0.1&-\alpha\\ \frac{1}{\alpha}&-0.1\end{pmatrix},

where α>0\alpha>0. Set ={M1,M2}\mathscr{M}=\{M_{1},M_{2}\}. By using the results in [2], we can deduce that (Σ)(\Sigma_{\mathscr{M}}) is uniformly exponentially stable if and only if α1<α<α2\alpha_{1}<\alpha<\alpha_{2}, where α10.819\alpha_{1}\approx 0.819 and α21.367\alpha_{2}\approx 1.367.

Taken α(α1,α2)\alpha\in(\alpha_{1},\alpha_{2}), let us focus on the minimal degree of a polynomial homogeneous common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}). The existence of a polynomial Lyapunov function homogeneous of a given degree nn can be tested using LMIs, as detailed in [6, Theorems 3.4 and 3.6]. For n32n\leq 32 we established numerically the existence of homogeneous polynomial Lyapunov functions for α(α1,α2(n))\alpha\in(\alpha_{1},\alpha_{2}^{(n)}), where α2(n)\alpha_{2}^{(n)} is given in Table 1.

Table 1:
nn 2 4 6 8 12
α2(n)\alpha_{2}^{(n)} 1.22 1.325 1.325 1.348 1.356
nn 16 20 24 28 32
α2(n)\alpha_{2}^{(n)} 1.36 1.3621 1.3634 1.3642 1.3647

In accordance with Corollary 4.5, the minimal degree appears to diverge as α\alpha tends to α2\alpha_{2}.

The conclusion of Corollary 4.5 can be proved to hold true for functions involving maxima and minima within a finite family of polynomials such as the class of polyhedral functions VV of the form

V(x)\displaystyle V(x) =max{|l1x|,,|lNx|}\displaystyle=\max\{|l_{1}^{\top}x|,\dots,|l_{N}^{\top}x|\}
=max{l1x,,lNx,l1x,,lNx},\displaystyle=\max\{l_{1}^{\top}x,\dots,l_{N}^{\top}x,-l_{1}^{\top}x,\dots,-l_{N}^{\top}x\},

with l1,,lNnl_{1},\dots,l_{N}\in\mathbb{R}^{n}, where NN is fixed. This partial counterpart to Corollary 3.3 is a consequence of the following more general result.

Theorem 4.9.

Let n2n\geq 2 and 𝒫dn\mathcal{P}^{n}_{d} be the family of polynomial functions in n\mathbb{R}^{n} of degree at most dd and ll be a positive integer. Consider the family

𝒫d,ln\displaystyle\mathcal{P}^{n}_{d,l} ={V𝒞(n,)V1,,Vl𝒫dn\displaystyle=\{V\in\mathcal{C}(\mathbb{R}^{n},\mathbb{R})\mid\exists V_{1},\dots,V_{l}\in\mathcal{P}^{n}_{d}
s.t. V(x){V1(x),,Vl(x)},xn}.\displaystyle\mbox{ s.t. }V(x)\in\{V_{1}(x),\dots,V_{l}(x)\},\ \forall x\in\mathbb{R}^{n}\}.

Then, 𝒫d,ln\mathcal{P}^{n}_{d,l} is not universal.

Proof.

We first claim that if 𝒫d,ln\mathcal{P}^{n}_{d,l} is universal, the same is true for 𝒫d,l2\mathcal{P}^{2}_{d,l}. Indeed, for every M2()\mathscr{M}\subset M_{2}(\mathbb{R}) such that (Σ)(\Sigma_{\mathscr{M}}) is uniformly exponentially stable, consider ^Mn()\hat{\mathscr{M}}\subset M_{n}(\mathbb{R}) given by

^={(M00Idn2)M}.\hat{\mathscr{M}}=\left\{\begin{pmatrix}M&0\\ 0&-\mathrm{Id}_{n-2}\end{pmatrix}\mid M\in\mathscr{M}\right\}.

If V^𝒫d,ln\hat{V}\in\mathcal{P}^{n}_{d,l} is a common Lyapunov function for (Σ^)(\Sigma_{\hat{\mathscr{M}}}), then V:2xV^(x,0)V:\mathbb{R}^{2}\ni x\mapsto\hat{V}(x,0) is a common Lyapunov function for (Σ)(\Sigma_{\mathscr{M}}) and V𝒫d,l2V\in\mathcal{P}^{2}_{d,l}.

We are left to prove that 𝒫d,l2\mathcal{P}^{2}_{d,l} is not universal. Consider the switched systems (Σε)(\Sigma_{\mathscr{M}^{\varepsilon}}) introduced in the proof of Theorem 4.2, which are uniformly exponentially stable for ε>0\varepsilon>0, and only uniformly stable for ε=0\varepsilon=0. Assume by contradiction that 𝒫d,l2\mathcal{P}^{2}_{d,l} is universal and, in particular, that for every ε>0\varepsilon>0 there exists a Lyapunov function Vε𝒫d,l2V^{\varepsilon}\in\mathcal{P}^{2}_{d,l} for (Σε)(\Sigma_{\mathscr{M}^{\varepsilon}}). By definition of 𝒫d,l2\mathcal{P}^{2}_{d,l}, for every ε>0\varepsilon>0 there exist ll polynomials P1ε,,PlεP_{1}^{\varepsilon},\dots,P_{l}^{\varepsilon} of degree at most dd such that Vε(x){P1ε(x),,Plε(x)}V^{\varepsilon}(x)\in\{P_{1}^{\varepsilon}(x),\dots,P_{l}^{\varepsilon}(x)\}. Given j,k{1,,l}j,k\in\{1,\dots,l\}, we investigate the set of zeroes of the polynomial QjkεQ_{jk}^{\varepsilon} defined as the homogeneous polynomial corresponding to the terms of maximal degree of PjεPkεP_{j}^{\varepsilon}-P_{k}^{\varepsilon}. For this purpose, recall that, by the fundamental theorem of algebra, every homogeneous polynomial QQ of positive degree mm may be factorized as Q=k=1m(αkx1+βkx2)Q=\prod_{k=1}^{m}(\alpha_{k}x_{1}+\beta_{k}x_{2}), where αk,βk\alpha_{k},\beta_{k}\in\mathbb{C} for k=1,,mk=1,\dots,m, so that its zeroes correspond to the intersection of the unit circle S1S^{1} with at most mm lines through the origin. Hence, it follows that either Qjkε0Q_{jk}^{\varepsilon}\equiv 0 (i.e., PjεPkεP_{j}^{\varepsilon}\equiv P_{k}^{\varepsilon}) or QjkεQ_{jk}^{\varepsilon} vanishes at most 2d2d times on the unit circle. Moreover, for every ε>0\varepsilon>0, the integer N=2d(l12)+1N=2d\binom{l-1}{2}+1 is a strict upper bound for the total number of zeroes of QjkεQ_{jk}^{\varepsilon} for j,k{1,,l}j,k\in\{1,\dots,l\}. Partitioning the circle into NN arcs 𝒞1,,𝒞N\mathcal{C}_{1},\dots,\mathcal{C}_{N} of equal length, for every ε>0\varepsilon>0 there exists an arc 𝒞nε\mathcal{C}_{n_{\varepsilon}} which contains no zero of the nontrivial polynomials QjkεQ_{jk}^{\varepsilon} in its interior. Denote by 𝒜nε\mathcal{A}_{n_{\varepsilon}} the closed middle third of 𝒞nε\mathcal{C}_{n_{\varepsilon}} (see Figure 1).

Refer to caption
Figure 1: Construction of the arc 𝒜nε\mathcal{A}_{n_{\varepsilon}}

We next claim that for every ε>0\varepsilon>0 there exists νε>0\nu_{\varepsilon}>0 large enough such that the restriction of VεV^{\varepsilon} to the dilated arc νε𝒜nε\nu_{\varepsilon}\mathcal{A}_{n_{\varepsilon}} coincides with the restriction to the same arc of one of the polynomials P1ε,,PlεP_{1}^{\varepsilon},\dots,P_{l}^{\varepsilon}. By definition of the function VεV^{\varepsilon} and taking into account its continuity, it is enough to prove that, for every ε>0\varepsilon>0 there exists νε>0\nu_{\varepsilon}>0 large enough such that, in the interior of the arc νε𝒜nε\nu_{\varepsilon}\mathcal{A}_{n_{\varepsilon}}, one has for every j,k{1,,l}j,k\in\{1,\dots,l\} that either PjεPkεP_{j}^{\varepsilon}\equiv P_{k}^{\varepsilon} or PjεPkεP_{j}^{\varepsilon}-P_{k}^{\varepsilon} is never vanishing. To see that, it is enough to prove that if QjkεQ_{jk}^{\varepsilon} does not vanish on 𝒜nε\mathcal{A}_{n_{\varepsilon}} then PjεPkεP_{j}^{\varepsilon}-P_{k}^{\varepsilon} does not vanish on νε𝒜nε\nu_{\varepsilon}\mathcal{A}_{n_{\varepsilon}} for νε>0\nu_{\varepsilon}>0 large enough independent of j,kj,k. In that case, one has, for ν>0\nu>0 large enough, xS1x\in S^{1} and j,k{1,,l}j,k\in\{1,\dots,l\},

(PjεPkε)(νx)=νd(Qjkε(x)+o(1)),(P_{j}^{\varepsilon}-P_{k}^{\varepsilon})(\nu x)=\nu^{d^{\prime}}\left(Q_{jk}^{\varepsilon}(x)+o(1)\right),

where dd^{\prime} is the positive degree of QjkεQ_{jk}^{\varepsilon} and o(1)o(1) is a function of xx and ν\nu tending to 0 as ν\nu tends to infinity uniformly with respect to xS1x\in S^{1} and j,k{1,,l}j,k\in\{1,\dots,l\}. This concludes the proof of the claim.

Since the arcs 𝒜1,,𝒜N\mathcal{A}_{1},\dots,\mathcal{A}_{N} do not depend on ε\varepsilon, there exist one of them, denoted by 𝒜{\mathcal{A}}, and sequences (εm)m,(νm)m(\varepsilon_{m})_{m\in\mathbb{N}},(\nu_{m})_{m\in\mathbb{N}} in (0,+)(0,+\infty), and (Vm)m(V_{m})_{m\in\mathbb{N}} in 𝒫d2\mathcal{P}^{2}_{d} such that limmεm=0\lim_{m\to\infty}\varepsilon_{m}=0 and Vεm=VmV^{\varepsilon_{m}}=V_{m} on ν𝒜\nu{\mathcal{A}} for every mm\in\mathbb{N} and every ννm\nu\geq\nu_{m}. Let V^m𝒫d2\hat{V}_{m}\in\mathcal{P}^{2}_{d} be the homogeneous term of maximal degree of VmV_{m}. Notice that

V^m(x)=limν+νdmVm(νx),x2,\hat{V}_{m}(x)=\lim_{\nu\to+\infty}\nu^{-d_{m}}V_{m}(\nu x),\qquad\forall x\in\mathbb{R}^{2},

where dmd_{m} denotes the degree of VmV_{m}. Up to normalizing VεmV^{\varepsilon_{m}}, we may assume that the maximum of the moduli of the coefficients of the polynomial V^j\hat{V}_{j} is equal to 11. Thus, up to extracting a subsequence, V^m\hat{V}_{m} converges uniformly on compact sets to some nonzero V¯𝒫d2\bar{V}\in\mathcal{P}^{2}_{d}.

Similarly to the proof of Theorem 4.2, we can construct a periodic trajectory tΦA0(t,0)x¯t\mapsto\Phi_{A^{0}}(t,0)\bar{x} starting at x¯\bar{x} in the interior of the arc 𝒜{\mathcal{A}}, with A0()A^{0}(\cdot) piecewise constant taking values in 0\mathscr{M}^{0}. Consider the switching laws Aε()=A0()εId2A^{\varepsilon}(\cdot)=A^{0}(\cdot)-\varepsilon\mathrm{Id}_{2} taking values in ε\mathscr{M}^{\varepsilon}. For every t0t\geq 0 such that ΦA0(t,0)x¯𝒜\Phi_{A^{0}}(t,0)\bar{x}\in{\mathcal{A}} and for every mm\in\mathbb{N}, we have

V^m(ΦAεm(t,0)x¯)\displaystyle\hat{V}_{m}(\Phi_{A^{\varepsilon_{m}}}(t,0)\bar{x}) =limν+νdmVm(νΦAεm(t,0)x¯)\displaystyle=\lim_{\nu\to+\infty}\nu^{-d_{m}}V_{m}(\nu\Phi_{A^{\varepsilon_{m}}}(t,0)\bar{x})
=limν+νdmVεm(ΦAεm(t,0)νx¯)\displaystyle=\lim_{\nu\to+\infty}\nu^{-d_{m}}V^{\varepsilon_{m}}(\Phi_{A^{\varepsilon_{m}}}(t,0)\nu\bar{x})
limν+νdmVεm(νx¯)\displaystyle\leq\lim_{\nu\to+\infty}\nu^{-d_{m}}V^{\varepsilon_{m}}(\nu\bar{x})
=limν+νdmVm(νx¯)=V^m(x¯),\displaystyle=\lim_{\nu\to+\infty}\nu^{-d_{m}}V_{m}(\nu\bar{x})=\hat{V}_{m}(\bar{x}),

and therefore

V¯(ΦA0(t,0)x¯)\displaystyle\bar{V}(\Phi_{A^{0}}(t,0)\bar{x}) =limmV^m(ΦAεm(t,0)x¯)\displaystyle=\lim_{m\to\infty}\hat{V}_{m}(\Phi_{A^{\varepsilon_{m}}}(t,0)\bar{x})
limmV^m(x¯)=V¯(x¯).\displaystyle\leq\lim_{m\to\infty}\hat{V}_{m}(\bar{x})=\bar{V}(\bar{x}).

We then deduce that tV¯(ΦA0(t,0)x¯0)t\mapsto\bar{V}(\Phi_{A^{0}}(t,0)\bar{x}_{0}) is constant on {t0ΦA0(t,0)x¯0𝒜}\{t\geq 0\mid\Phi_{A^{0}}(t,0)\bar{x}_{0}\in{\mathcal{A}}\}. Moreover ΦA0(t,0)=etMi0\Phi_{A^{0}}(t,0)=e^{tM^{0}_{i}} for some i=1,2i=1,2, for tt small enough, and by repeating the argument in the proof of Theorem 4.2 we obtain that V¯0\bar{V}\equiv 0, yielding a contradiction. ∎

Remark 4.10.

At the light of the previous results, one may wonder if it is possible to identify universal classes of Lyapunov functions defined with a finite number of parameters. The results proved in this section show that this is not the case for linear spaces of analytic functions and families of piecewise polynomials. In [15] an explicit construction of a universal class depending only on six parameters has been provided in the special case of two-dimensional linear switched systems with two modes. Unfortunately, since such a construction is based on an explicit characterization of the stability properties of the switched system in terms of the matrix coefficients, it seems unlikely that it can be adjusted to higher dimensional switched systems.

References

  • [1] A. Bacciotti and F. Ceragioli. Stability and stabilization of discontinuous systems and nonsmooth Lyapunov functions. ESAIM Control Optim. Calc. Var., 4:361–376, 1999.
  • [2] M. Balde, U. Boscain, and P. Mason. A note on stability conditions for planar switched systems. Internat. J. Control, 82(10):1882–1888, 2009.
  • [3] F. Blanchini. Set invariance in control. Automatica, 35(11):1747–1767, 1999.
  • [4] F. Blanchini and S. Miani. A new class of universal Lyapunov functions for the control of uncertain linear systems. IEEE Trans. Automat. Control, 44(3):641–647, 1999.
  • [5] A. Bressan and B. Piccoli. Introduction to the mathematical theory of control, volume 2 of AIMS Series on Applied Mathematics. American Institute of Mathematical Sciences (AIMS), Springfield, MO, 2007.
  • [6] G. Chesi, A. Garulli, A. Tesi, and A. Vicino. Homogeneous polynomial forms for robustness analysis of uncertain systems, volume 390 of Lecture Notes in Control and Information Sciences. Springer-Verlag, Berlin, 2009.
  • [7] Y. Chitour, M. Gaye, and P. Mason. Geometric and asymptotic properties associated with linear switched systems. Journal of Differential Equations, 259(11):5582–5616, 2015.
  • [8] W. P. Dayawansa and C. F. Martin. A converse Lyapunov theorem for a class of dynamical systems which undergo switching. IEEE Trans. Automat. Control, 44(4):751–760, 1999.
  • [9] S. F. Hafstein. An algorithm for constructing Lyapunov functions, volume 8 of Electronic Journal of Differential Equations. Monograph. Texas State University–San Marcos, Department of Mathematics, San Marcos, TX, 2007.
  • [10] F. M. Hante and M. Sigalotti. Converse Lyapunov theorems for switched systems in Banach and Hilbert spaces. SIAM J. Control Optim., 49(2):752–770, 2011.
  • [11] R. Jungers. The joint spectral radius. Theory and applications, volume 385 of Lecture Notes in Control and Information Sciences. Springer-Verlag, Berlin, 2009.
  • [12] J. Kurzweil. On the inversion of Lyapunov’s second theorem on stability of motion. American Mathematical Society Translations, 24(2):19–77, 1956.
  • [13] D. Liberzon. Switching in systems and control. Systems & Control: Foundations & Applications. Birkhäuser Boston Inc., Boston, MA, 2003.
  • [14] Y. Lin, E. D. Sontag, and Y. Wang. A smooth converse Lyapunov theorem for robust stability. SIAM Journal on Control and Optimization, 34(1):124–160, 1996.
  • [15] P. Mason, U. Boscain, and Y. Chitour. Common polynomial Lyapunov functions for linear switched systems. SIAM J. Control Optimization, 45(1):226–245, 2006.
  • [16] J. L. Massera. Contributions to stability theory. Annals of Mathematics, pages 182–206, 1956.
  • [17] A. Meilakhs. Design of stable control systems subject to parametric perturbation. Automation and Remote Control, 39(10):1409–1418, 1979.
  • [18] A. P. Molchanov and E. S. Pyatnitskiĭ. Lyapunov functions that define necessary and sufficient conditions for absolute stability of nonlinear nonstationary control systems. I. Automat. Remote Control, 47(3):343–354, 1986.
  • [19] A. P. Molchanov and E. S. Pyatnitskiĭ. Lyapunov functions that define necessary and sufficient conditions for absolute stability of nonlinear nonstationary control systems. II. Automat. Remote Control, 47(4):443–451, 1986.
  • [20] A. P. Molchanov and Y. S. Pyatnitskiy. Criteria of asymptotic stability of differential and difference inclusions encountered in control theory. Systems Control Lett., 13(1):59–64, 1989.
  • [21] R. Rockafellar. Convex Analysis. Princeton Landmarks in Mathematics and Physics. Princeton University Press, 1970.
  • [22] F. W. Wilson. Smoothing derivatives of functions and applications. Transactions of the American Mathematical Society, 139:413–428, 1969.