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

From local to global asymptotic stabilizability for weakly contractive control systems

Vincent Andrieu Univ. Lyon, Université Claude Bernard Lyon 1, CNRS, LAGEPP UMR 5007, 43 bd du 11 novembre 1918, F-69100 Villeurbanne, France (e-mail: [email protected], [email protected], [email protected]) Lucas Brivadis Univ. Lyon, Université Claude Bernard Lyon 1, CNRS, LAGEPP UMR 5007, 43 bd du 11 novembre 1918, F-69100 Villeurbanne, France (e-mail: [email protected], [email protected], [email protected]) Jean-Paul Gauthier Université de Toulon, Aix Marseille Univ, CNRS, LIS, France (e-mail: [email protected]) Ludovic Sacchelli Department of Mathematics, Lehigh University, Bethlehem, PA, USA (e-mail: [email protected]) Ulysse Serres Univ. Lyon, Université Claude Bernard Lyon 1, CNRS, LAGEPP UMR 5007, 43 bd du 11 novembre 1918, F-69100 Villeurbanne, France (e-mail: [email protected], [email protected], [email protected])
Abstract

A nonlinear control system is said to be weakly contractive in the control if the flow that it generates is non-expanding (in the sense that the distance between two trajectories is a non-increasing function of time) for some fixed Riemannian metric independent of the control. We prove in this paper that for such systems, local asymptotic stabilizability implies global asymptotic stabilizability by means of a dynamic state feedback. We link this result and the so-called Jurdjevic and Quinn approach.

Keywords.

Nonlinear control systems, Feedback stabilization, Asymptotic stability.

Acknowledgments.

This research was funded by the French Grant ANR ODISSE (ANR-19-CE48-0004-01).

1 Main result

1.1 Statement of the result

Consider the following nonlinear continuous-time control system:

x˙=f(x,u)=fu(x),f(0,0)=0,\dot{x}=f(x,u)=f_{u}(x),\quad f(0,0)=0, (1)

where xx lives in n\mathbb{R}^{n} and uu is the control input taking values in an open subset 𝒰\mathcal{U} of m\mathbb{R}^{m} containing zero. We assume that fuC1(n,n)f_{u}\in C^{1}(\mathbb{R}^{n},\mathbb{R}^{n}) for all u𝒰u\in\mathcal{U}, fxC0(n×𝒰,n)\frac{\partial f}{\partial x}\in C^{0}(\mathbb{R}^{n}\times\mathcal{U},\mathbb{R}^{n}) and f(x,)f(x,\cdot) is locally Lipschitz for all xnx\in\mathbb{R}^{n}.

Definition 1 (Static stabilizability).

System (1) is said to be locally (resp. globally) asymptotically stabilizable by a static state feedback if there exists a locally Lipschitz mapping λ:n𝒰\lambda:\mathbb{R}^{n}\rightarrow\mathcal{U} such that

x˙=f(x,λ(x))\dot{x}=f(x,\lambda(x)) (2)

is locally (resp. globally) asymptotically stable at the origin.

Local asymptotic stabilizability is usually obtained by investigating first order or homogeneous approximations of the dynamical system around the origin. Yet obtaining global stabilizability from local stabilizability is not an easy task and may fail in general.

However, there are classes of system for which we know how to bridge the gap between local and global asymptotic stabilizability. This is obviously the case if the feedback law λ\lambda is such that xf(x,λ(x))x\mapsto f(x,\lambda(x)) is a linear vector field. More generally, it still holds for homogeneous systems admitting a homogeneous feedback law (see e.g. [7, 12]). Note also that it is shown in [5] that when the locally stabilizing state feedback fails to share the same homogeneity property than the vector field, global (or semi-global) property can still be achieved by a dynamic state feedback.

Definition 2 (Dynamic stabilizability).

System (1) is said to be locally (resp. globally) asymptotically stabilizable by a dynamic state feedback if there exist f^:n×n×𝒰n\hat{f}:\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathcal{U}\to\mathbb{R}^{n} such that f^(,,u)C1(n×n,n)\hat{f}(\cdot,\cdot,u)\in C^{1}(\mathbb{R}^{n}\times\mathbb{R}^{n},\mathbb{R}^{n}) for all u𝒰u\in\mathcal{U},

f^(x,x^)C0(n×n×𝒰,n)\frac{\partial\hat{f}}{\partial(x,\hat{x})}\in C^{0}(\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathcal{U},\mathbb{R}^{n})

and f^(x,x^,)\hat{f}(x,\hat{x},\cdot) is locally Lipschitz for all (x,x^)n×n(x,\hat{x})\in\mathbb{R}^{n}\times\mathbb{R}^{n} and a locally Lipschitz mapping λ:n𝒰\lambda:~{}\mathbb{R}^{n}\rightarrow\mathcal{U} such that

x˙=f(x,λ(x^)),x^˙=f^(x,x^,λ(x^))\dot{x}=f(x,\lambda(\hat{x})),\quad\dot{\hat{x}}=\hat{f}(x,\hat{x},\lambda(\hat{x})) (3)

is locally (resp. globally) asymptotically stable at the origin.

In this paper, we give another class of dynamical systems which share the same property that static local asymptotic stabilizability implies dynamic global asymptotic stabilizability: namely, weakly contractive control systems.

Definition 3 (Weakly contractive).

Let gg be a C1C^{1} Riemannian metric on n\mathbb{R}^{n}. System (1) is said to be weakly contractive with respect to gg if

u𝒰,Lfug0,\forall u\in\mathcal{U},\quad L_{f_{u}}g\leqslant 0, (4)

where LfugL_{f_{u}}g denotes the Lie derivative of the metric gg with respect to the vector field fuf_{u}.

A vector field FF over n\mathbb{R}^{n} is usually said to be contractive with respect to a metric gg if LFgL_{F}g is negative. Here we insist on the fact that the vector fields fuf_{u} are only weakly contractive with respect to the metric gg, in the sense that LfugL_{f_{u}}g is only non-positive.

For all pair of vectors (φ,ψ)n×n(\varphi,\psi)\in\mathbb{R}^{n}\times\mathbb{R}^{n}, we denote by φ,ψ\langle\varphi,\psi\rangle and |φ||\varphi| the canonical Euclidean inner product and induced norm over n\mathbb{R}^{n}. For all point xnx\in\mathbb{R}^{n}, let φ,ψg(x)=g(x)(φ,ψ)\langle\varphi,\psi\rangle_{g(x)}=g(x)(\varphi,\psi) denote the inner product between the two vectors φ\varphi and ψ\psi at the point xx for the metric gg, and set |φ|g(x)=φ,φg(x)|\varphi|_{g(x)}=\langle\varphi,\varphi\rangle_{g(x)}.

Recall that associated to the metric gg we can define a distance dgd_{g} between a pair of points of n\mathbb{R}^{n} in the following way. The length of any piecewise C1C^{1} path γ:[s1,s2]n\gamma:[s_{1},s_{2}]\to\mathbb{R}^{n} between two arbitrary points x1=γ(s1)x_{1}=\gamma(s_{1}) and x2=γ(s2)x_{2}=\gamma(s_{2}) in n\mathbb{R}^{n} is defined as:

(γ)=s1s2|γ(s)|g(γ(s))ds\ell(\gamma)=\int_{s_{1}}^{s_{2}}|\gamma^{\prime}(s)|_{g(\gamma(s))}\mathrm{d}s (5)

The distance dg(x1,x2)d_{g}(x_{1},x_{2}) is defined as the infimum of the length over all such paths. We denote dg2d_{g}^{2} the square of the distance function.

For all point (x,x^)n×n(x,\hat{x})\in\mathbb{R}^{n}\times\mathbb{R}^{n}, we denote (if it exists) g(x^)dg2(x,x^)\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x}) the gradient of the function x^dg2(x,x^)\hat{x}\mapsto d_{g}^{2}(x,\hat{x}) at the point x^\hat{x} for the metric gg.

Fix xnx\in\mathbb{R}^{n}. Then g(x^)dg2(x,x^)\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x}) is well-defined if and only if, for all x^n\hat{x}\in\mathbb{R}^{n}, there exists a unique length-minimizing curve γ\gamma joining xx to x^\hat{x}, i.e. such that (γ)=dg(x,x^)\ell(\gamma)=d_{g}(x,\hat{x}). Equivalently, the Riemannian exponential map at the point x^\hat{x} (denoted by expx^\exp_{\hat{x}}) is invertible111see e.g. [3, Chap. 7, Theorem 3.1] for sufficient geometric conditions. and we have

g(x^)dg2(x,x^)=2expx^1(x)\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x})=-2\exp_{\hat{x}}^{-1}(x)

for all x^n\hat{x}\in\mathbb{R}^{n}, which yields

g(x^)dg2(x,x^)=0if and only ifx=x^.\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x})=0\quad\textrm{if and only if}\quad x=\hat{x}. (6)

Also, by definition of the Riemannian gradient, for all vectors φn\varphi\in\mathbb{R}^{n},

g(x^)dg2(x,x^),φg(x^)=dg2x^(x,x^),φ.\left\langle\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x}),\varphi\right\rangle_{g(\hat{x})}=\left\langle\frac{\partial d_{g}^{2}}{\partial\hat{x}}(x,\hat{x}),\varphi\right\rangle. (7)

Assume that ff is C1C^{1}. If (1) is a weakly contractive vector field, then for all C1C^{1} control u:+𝒰u:\mathbb{R}_{+}\to\mathcal{U} the time-varying vector field fuf_{u} generates a non-expanding flow in the sense that, if x1x_{1} and x2x_{2} satisfy x˙i=fu(xi)\dot{x}_{i}=f_{u}(x_{i}) for i{1,2}i\in\{1,2\}, then the distance dg(x1,x2)d_{g}(x_{1},x_{2}) between the two trajectories is a non-increasing function of time. We give in appendix a short proof of this well-known statement to be self-contained.

The following theorem is the main result of the paper.

Theorem 4.

Let gg be a C2C^{2} complete Riemannian metric on n\mathbb{R}^{n} such that dg2d_{g}^{2} is a C2C^{2} function. Assume that (1) is weakly contractive with respect to gg, and fC1(n×𝒰,n)f\in C^{1}(\mathbb{R}^{n}\times\mathcal{U},\mathbb{R}^{n}). If (1) is locally asymptotically stabilizable by a static state feedback λC1(n,𝒰)\lambda\in C^{1}(\mathbb{R}^{n},\mathcal{U}), then it is also globally asymptotically stabilizable by a dynamic state feedback given by

x˙=f(x,λ(x^)),x^˙=f(x^,λ(x^))+k(x,x^)\dot{x}=f(x,\lambda(\hat{x})),\quad\dot{\hat{x}}=f(\hat{x},\lambda(\hat{x}))+k(x,\hat{x}) (8)

where

k(x,x^)=α(x,x^)g(x^)dg2(x,x^)k(x,\hat{x})=-\alpha(x,\hat{x})\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x})

in which the function α\alpha has to be selected sufficiently small.

1.2 Discussion on the result

The idea of the proof is somehow counter-intuitive. Indeed, the feedback depends only on x^\hat{x}. By selecting α\alpha sufficiently small, we make sure that x^\hat{x} remains in the basin of attraction of the origin for the vector field associated to the state feedback. On the other hand, the correction terms kk acting on x^˙\dot{\hat{x}} forces xx to converge to x^\hat{x}, which implies that xx goes to zero.

An interesting aspect of our approach is that no structural constraints is imposed on the local asymptotic stabilizer. This one can be designed for qualitative purposes and can be for instance bounded or optimal as long as this one ensures a local asymptotic stability property. This technique offers another approach to solve the global asymptotic stabilization with local optimal behavior as for instance studied in [2] or [4]. The main difference with these studies being that the local optimal behavior is reproduced asymptotically in time (as xx converges to x^\hat{x}).

To construct the feedback law one needs to compute g(x^)dg2(x,x^)\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x}) which may be difficult to obtain analytically in general (except in some simple cases, e.g., if the metric is constant). Some ways of constructing similar correction terms may be obtained following observer designs based on Riemannian approaches as in [1, 13]. In particular in [13, Lemma 3.6], the authors introduced a “distance-like” function δ\delta, that is of crucial importance in the construction of the correction term.

1.3 Proof

Let λ\lambda be a C1C^{1} locally asymptotically stabilizing feedback law. Let 𝒟\mathcal{D} be the basin of attraction of the origin for the vector field xf(x,λ(x))x\mapsto f(x,\lambda(x)), which is a non-empty open subset of n\mathbb{R}^{n}.

According to the converse Lyapunov theorem [14] (based on the previous works of [8, 9, 10]) , there exists a proper function VC(𝒟,+)V\in C^{\infty}(\mathcal{D},\mathbb{R}_{+}) such that V(0)=0V(0)=0 and

Vx(x)f(x,λ(x))V(x),x𝒟.\frac{\partial V}{\partial x}(x)f(x,\lambda(x))\leqslant-V(x),\quad\forall x\in\mathcal{D}\ . (9)

For all r>0r>0, set D(r)={xnV(x)r}D(r)=\{x\in\mathbb{R}^{n}\mid V(x)\leqslant r\} which is a compact subset of 𝒟\mathcal{D}. Let α:n×𝒟+\alpha:\mathbb{R}^{n}\times\mathcal{D}\to\mathbb{R}_{+} be the positive and locally Lipschitz function given by

α(x,x^)=max{V(x^),1}2(1+|Vx(x^)|)(1+|g(x^)dg2(x,x^)|).\alpha(x,\hat{x})=\frac{-\max\{V(\hat{x}),1\}}{2\left(1+\left|\frac{\partial V}{\partial x}(\hat{x})\right|\right)\left(1+\left|\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x})\right|\right)}. (10)

It yields

|k(x,x^)|max{V(x^),1}2(1+|Vx(x^)|),(x,x^)n×𝒟.|k(x,\hat{x})|\leqslant\frac{\max\{V(\hat{x}),1\}}{2\left(1+\left|\frac{\partial V}{\partial x}(\hat{x})\right|\right)},\quad\forall(x,\hat{x})\in\mathbb{R}^{n}\times\mathcal{D}. (11)

We prove Theorem 4 in three steps.

Step 1 : the x^\hat{x}-component of semi-trajectories of (8) remain in a compact subset of 𝒟\mathcal{D}. For all (x,x^)n×𝒟(x,\hat{x})\in\mathbb{R}^{n}\times\mathcal{D}, it follows from (9) and (11) that

Vx(x^)[f(x^,λ(x^))+k(x,x^)]\displaystyle\frac{\partial V}{\partial x}(\hat{x})[f(\hat{x},\lambda(\hat{x}))+k(x,\hat{x})] V(x^)+|Vx(x^)|max{V(x^),1}2(1+|Vx(x^)|)\displaystyle\leqslant-V(\hat{x})+\left|\frac{\partial V}{\partial x}(\hat{x})\right|\frac{\max\{V(\hat{x}),1\}}{2\left(1+\left|\frac{\partial V}{\partial x}(\hat{x})\right|\right)}
V(x^)+12max{V(x^),1}.\displaystyle\leqslant-V(\hat{x})+\frac{1}{2}\max\{V(\hat{x}),1\}.

Hence, if x^𝒟D(1)\hat{x}\in\mathcal{D}\setminus D(1),

Vx(x^)(f(x^,λ(x^))+k(x,x^))12V(x^).\frac{\partial V}{\partial x}(\hat{x})(f(\hat{x},\lambda(\hat{x}))+k(x,\hat{x}))\leqslant-\frac{1}{2}V(\hat{x})\ . (12)

For all initial conditions (x0,x^0)n×𝒟(x_{0},\hat{x}_{0})\in\mathbb{R}^{n}\times\mathcal{D}, the solution (x,x^)(x,\hat{x}) of the closed-loop system (8) satisfies

V(x^(t))max{V(x^0),1},V(\hat{x}(t))\leqslant\max\{V(\hat{x}_{0}),1\},

for all t0t\geqslant 0, in the time domain of existence of the solution. In other words, x^(t)D(1)D(V(x^0))\hat{x}(t)\in D(1)\cup D(V(\hat{x}_{0})) which is a compact subset of 𝒟\mathcal{D}.

Step 2 : the distance between x^\hat{x} and xx is non-increasing and has limit zero. System (8) can be rewritten as

[x˙x^˙]=F(x,x^)+K(x,x^)\begin{bmatrix}\dot{x}\\ \dot{\hat{x}}\end{bmatrix}=F(x,\hat{x})+K(x,\hat{x}) (13)

by setting F(x,x^)=[f(x,λ(x^))f(x^,λ(x^))]F(x,\hat{x})=\begin{bmatrix}f(x,\lambda(\hat{x}))\\ f(\hat{x},\lambda(\hat{x}))\end{bmatrix} and K(x,x^)=[0α(x,x^)g(x^)dg2(x,x^)]K(x,\hat{x})=\begin{bmatrix}0\\ -\alpha(x,\hat{x})\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x})\end{bmatrix}.

Since (1) is weakly contractive with respect to gg, the result proved in appendix applied to the control u=λ(x^)u=\lambda(\hat{x}) shows that

LFdg2(x,x^)0.L_{F}d_{g}^{2}(x,\hat{x})\leqslant 0.

Thus, by (7),

LF+Kdg2(x,x^)α(x,x^)|g(x^)dg2(x,x^)|g(x^)2.L_{F+K}d_{g}^{2}(x,\hat{x})\leqslant-\alpha(x,\hat{x})\left|\nabla_{g(\hat{x})}d_{g}^{2}(x,\hat{x})\right|_{g(\hat{x})}^{2}. (14)

Hence, for all (x0,x^0)n×𝒟(x_{0},\hat{x}_{0})\in\mathbb{R}^{n}\times\mathcal{D}, tdg(x(t),x^(t))t\mapsto d_{g}(x(t),\hat{x}(t)) is non-increasing and for all t0t\geqslant 0 on the time domain of existence of the solution we have

(x(t),x^(t))Γ(x0,x^0),(x(t),\hat{x}(t))\in\Gamma(x_{0},\hat{x}_{0}),

where

Γ(x0,x^0)={(ξ,ξ^)n×𝒟ξ^D(1)D(V(x^0)),dg(ξ,ξ^)dg(x0,x^0)}.\Gamma(x_{0},\hat{x}_{0})=\Big{\{}(\xi,\hat{\xi})\in\mathbb{R}^{n}\times\mathcal{D}\mid\hat{\xi}\in D(1)\cup D(V(\hat{x}_{0})),d_{g}(\xi,\hat{\xi})\leqslant d_{g}(x_{0},\hat{x}_{0})\Big{\}}.

Moreover, gg is a complete metric. Then, according to the Hopf-Rinow theorem, Γ(x0,x^0)\Gamma(x_{0},\hat{x}_{0}) is compact. Hence, solutions of (8) are complete in positive time.

Given (x0,x^0)n×𝒟(x_{0},\hat{x}_{0})\in\mathbb{R}^{n}\times\mathcal{D}, let κ:++\kappa:\mathbb{R}_{+}\to\mathbb{R}_{+} be the function defined by

κ(s)=min(ξ,ξ^)Γ(x0,x^0)dg(ξ,ξ^)=sα(ξ,ξ^)|g(ξ^)dg2(ξ,ξ^)|g(ξ^)2.\kappa(s)=\min_{(\xi,\hat{\xi})\in\Gamma(x_{0},\hat{x}_{0})\mid d_{g}(\xi,\hat{\xi})=s}\alpha(\xi,\hat{\xi})\left|\nabla_{g(\hat{\xi})}d_{g}^{2}(\xi,\hat{\xi})\right|_{g(\hat{\xi})}^{2}.

Note that if x0x^0x_{0}\neq\hat{x}_{0}, then, for all s>0s>0, κ(s)>0\kappa(s)>0 since α\alpha takes positive values and (6) holds. Hence, (14) leads to

ddtdg2(x(t),x^(t))κ(dg2(x(t),x^(t))),t0.\frac{\mathrm{d}}{\mathrm{d}t}d_{g}^{2}(x(t),\hat{x}(t))\leqslant-\kappa(d_{g}^{2}(x(t),\hat{x}(t)))\ ,\ \forall t\geqslant 0. (15)

Thus limt+dg(x(t),x^(t))=0\lim_{t\to+\infty}d_{g}(x(t),\hat{x}(t))=0.

Step 3 : attractivity and local asymptotic stability of the origin. Given (x0,x^0)(x_{0},\hat{x}_{0}) in n×𝒟\mathbb{R}^{n}\times\mathcal{D}, let μ:++\mu:\mathbb{R}_{+}\to\mathbb{R}_{+} be the function defined by

μ(s)=max(ξ,ξ^)Γ(x0,x^0)dg(ξ,ξ^)s|Vx(ξ^)k(ξ,ξ^)|.\mu(s)=\max_{(\xi,\hat{\xi})\in\Gamma(x_{0},\hat{x}_{0})\mid d_{g}(\xi,\hat{\xi})\leqslant s}\left|\frac{\partial V}{\partial x}(\hat{\xi})k(\xi,\hat{\xi})\right|.

Then μ\mu is non-decreasing, continuous and μ(0)=0\mu(0)=0. Moreover, the solution (x,x^)(x,\hat{x}) of (8) initialized at (x0,x^0)n×𝒟(x_{0},\hat{x}_{0})\in\mathbb{R}^{n}\times\mathcal{D} satisfies

ddtV(x^(t))V(x^(t))+μ(dg(x(t),x^(t)).\frac{\mathrm{d}}{\mathrm{d}t}V(\hat{x}(t))\leqslant-V(\hat{x}(t))+\mu(d_{g}(x(t),\hat{x}(t)). (16)

From this inequality and Step 2 we conclude that limt+(x(t),x^(t))=(0,0)\lim_{t\rightarrow+\infty}(x(t),\hat{x}(t))=(0,0).

Inequalities (14) and (16) being true for all solutions starting in Γ(x0,x^0)\Gamma(x_{0},\hat{x}_{0}), this implies also stability of (0,0)(0,0).

2 Link with Jurdjevic and Quinn approach

2.1 Jurdjevic and Quinn result

The next result follows from the work of Jurdjevic and Quinn in [6]. The version that we state here is a direct corollary of [11, Theorem II.1]

Theorem 5 (Jurdjevic and Quinn approach).

Consider the control system

x˙=a(x)+b(x,u)u,\dot{x}=a(x)+b(x,u)u, (17)

with aa and bb two C1C^{1} functions. Assume that there exists a C1C^{1} positive definite proper function V:n+V:\mathbb{R}^{n}\mapsto\mathbb{R}_{+} such that

LaV0.L_{a}V\leqslant 0.

If the only solution of the system

x˙=a(x),Lb(,0)V(x)=0,LaV(x)=0\dot{x}=a(x),\quad L_{b(\cdot,0)}V(x)=0,\quad L_{a}V(x)=0 (18)

is x0x\equiv 0, then (17) is globally asymptotically stabilizable by a static state feedback.

In the context of weakly contractive control systems, the Jurdjevic and Quinn approach leads to the following corollary.

Corollary 6.

Let gg be a complete Riemannian metric on n\mathbb{R}^{n}. Assume that (1) is weakly contractive with respect to gg and that fC2(n×𝒰,n)f\in C^{2}(\mathbb{R}^{n}\times\mathcal{U},\mathbb{R}^{n}). If the only solution of the system

x˙=f(x,0),(Lb(,0)dg2(,0))(x)=0,Lf0dg2(x,0)=0\dot{x}=f(x,0),\ \left(L_{b(\cdot,0)}d_{g}^{2}(\cdot,0)\right)(x)=0,\ L_{f_{0}}d_{g}^{2}(x,0)=0

where b(,0)=fu(,0)b(\cdot,0)=\frac{\partial f}{\partial u}(\cdot,0) is x0x\equiv 0, then (1) is globally asymptotically stabilizable by a static state feedback.

To prove this corollary, it is sufficient to apply Theorem 5 with V:xdg(x,0)2V:x\mapsto d_{g}(x,0)^{2}, a:xf(x,0)a:x\mapsto f(x,0) and b:(x,u)01fu(x,su)𝑑sb:(x,u)\mapsto\int_{0}^{1}\frac{\partial f}{\partial u}(x,su)ds.

2.2 Link with our result

Note that the Jurdjevic-Quinn approach guarantees the existence of a static state feedback, contrarily to our main Theorem 4 which build a dynamic state feedback. However, the feedback obtained by their approach is implicit, while our dynamic state feedback is explicitly given by (8).

Moreover, our feedback law differs strongly with the one given in Jurdjevic-Quinn approach. Indeed, in their approach the feedback is designed small enough to make sure that it acts in a good direction related to the Lyapunov function. In our framework, this is no more a small feedback approach but more a small correction term for an observer approach.

Let us consider the particular case in which

f(x,u)=Ax+b(x)u.f(x,u)=Ax+b(x)u. (19)

where An×nA\in\mathbb{R}^{n\times n} and bC1(n,n)b\in C^{1}(\mathbb{R}^{n},\mathbb{R}^{n}). Then (19) is weakly contractive with respect to some constant metric gg if and only if LAg0L_{A}g\leqslant 0 and Lbg=0L_{b}g=0222It is easy to check that it is the case if and only if b(x)=b(0)+Jxb(x)=b(0)+Jx with LJg=0L_{J}g=0.. Moreover, the pair (A,b(0))(A,b(0)) is controllable if and only if (19) is locally asymptotically stabilizable by a static feedback. Then, if all these hypotheses hold, a dynamic globally stabilizing state feedback is given by Theorem 4.

We can also show under the same hypotheses that the Jurdjevic and Quinn approach can be applied. Indeed, the system in Corollary 6 is equivalent to

x˙=Ax,(Lbdg2(,0))(x)=0,LAdg2(x,0)=0\dot{x}=Ax,\quad\left(L_{b}d_{g}^{2}(\cdot,0)\right)(x)=0,\quad L_{A}d_{g}^{2}(x,0)=0 (20)

which implies that x0x\equiv 0 when the pair (A,b(0))(A,b(0)) is controllable. Then, according to Corollary 6, (19) is globally asymptotically stabilizable by a static state feedback. However, it is not clear in general that both contexts are equivalent, and finding an example fitting in the framework of Theorem 4 but in which the Jurdjevic and Quinn approach of Corollary 6 remains an open question.

3 Appendix on weakly contractive vector fields

For all u:+𝒰u:\mathbb{R}_{+}\to\mathcal{U} and all xnx\in\mathbb{R}^{n}, denote by tXu(x,t)t\mapsto X_{u}(x,t) the solution of (1) with initial condition xx. Let u:+𝒰u:\mathbb{R}_{+}\to\mathcal{U} be such that XuX_{u} is well-defined and C2C^{2} on n×+\mathbb{R}^{n}\times\mathbb{R}_{+}.

Let (x1,x2)n×n(x_{1},x_{2})\in\mathbb{R}^{n}\times\mathbb{R}^{n} and γ:[s1,s2]n\gamma:[s_{1},s_{2}]\to\mathbb{R}^{n} be a C2C^{2} path between the points x1=γ(s1)x_{1}=\gamma(s_{1}) and x2=γ(s2)x_{2}=\gamma(s_{2}). For all (s,t)[s1,s2]×+(s,t)\in[s_{1},s_{2}]\times\mathbb{R}_{+}, set Γ(s,t)=Xu(γ(s),t)\Gamma(s,t)=X_{u}(\gamma(s),t) and ρ(s,t)=|Γs(s,t)|g(Γ(s,t))2\rho(s,t)=\left|\frac{\partial\Gamma}{\partial s}(s,t)\right|^{2}_{g(\Gamma(s,t))}. Then ρ\rho is C1C^{1} and

ρt(s,t)=Lfug(Γ(s,t))(Γs(s,t),Γs(s,t))0,\displaystyle\frac{\partial\rho}{\partial t}(s,t)=L_{f_{u}}g(\Gamma(s,t))\left(\frac{\partial\Gamma}{\partial s}(s,t),\frac{\partial\Gamma}{\partial s}(s,t)\right)\leqslant 0,

which yields

d(Γ(,t))dt\displaystyle\frac{\mathrm{d}\ell(\Gamma(\cdot,t))}{\mathrm{d}t} =ddts1s2ρ(s,t)ds\displaystyle=\frac{\mathrm{d}}{\mathrm{d}t}\int_{s_{1}}^{s_{2}}\sqrt{\rho(s,t)}\mathrm{d}s
=s1s212ρ(s,t)ρt(s,t)ds\displaystyle=\int_{s_{1}}^{s_{2}}\frac{1}{2\sqrt{\rho(s,t)}}\frac{\partial\rho}{\partial t}(s,t)\mathrm{d}s
0.\displaystyle\leqslant 0.

Hence dg(Xu(x1,t),Xu(x2,t))(Γ(,t))(γ).d_{g}(X_{u}(x_{1},t),X_{u}(x_{2},t))\leqslant\ell(\Gamma(\cdot,t))\leqslant\ell(\gamma). Choosing a sequence of paths (γn)n(\gamma_{n})_{n\in\mathbb{N}} such that (γn)dg(x1,x2)\ell(\gamma_{n})\to d_{g}(x_{1},x_{2}) and passing to the limit we get

dg(Xu(x1,t),Xu(x2,t))dg(x1,x2).\displaystyle d_{g}(X_{u}(x_{1},t),X_{u}(x_{2},t))\leqslant d_{g}(x_{1},x_{2}).

Since this inequality is true for any control input uu, tdg(Xu(x1,t),Xu(x2,t))t\mapsto d_{g}(X_{u}(x_{1},t),X_{u}(x_{2},t)) is non-increasing for all control uu and all points x1x_{1}, x2x_{2}.

References

  • [1] N. Aghannan and P. Rouchon. An intrinsic observer for a class of Lagrangian systems. IEEE Trans. Automat. Control, 48(6):936–945, 2003.
  • [2] S. Benachour, H. Stein Shiromoto, and V. Andrieu. Locally optimal controllers and globally inverse optimal controllers. Automatica J. IFAC, 50(11):2918–2923, 2014.
  • [3] M. P. do Carmo. Riemannian geometry. Mathematics: Theory & Applications. Birkhäuser Boston, Inc., Boston, MA, 1992. Translated from the second Portuguese edition by Francis Flaherty.
  • [4] K. Ezal, Z. Pan, and P. V. Kokotović. Locally optimal and robust backstepping design. IEEE Trans. Automat. Control, 45(2):260–271, 2000.
  • [5] H. Hammouri and J.-C. Marquès. Two global stabilizability results for homogeneous systems from a local stabilizability assumption. IEEE Trans. Automat. Control, 54(9):2239–2244, 2009.
  • [6] V. Jurdjevic and J. P. Quinn. Controllability and stability. J. Differential Equations, 28(3):381–389, 1978.
  • [7] M. Kawski. Homogeneous stabilizing feedback laws. Control Theory Adv. Tech., 6(4):497–516, 1990.
  • [8] J. Kurzweil. On the inversion of Lyapunov’s second theorem on stability of motion. Czechoslovak Math. J., 06(2):217–219, 1956.
  • [9] J. Kurzweil. On the inversion of Lyapunov’s second theorem on stability of motion. Amer. Math. Soc. Transl., 2(24):19–77, 1963.
  • [10] J. L. Massera. Contributions to stability theory. Ann. of Math. (2), 64:182–206, 1956.
  • [11] F. Mazenc and L. Praly. Adding integrations, saturated controls, and stabilization for feedforward systems. IEEE Trans. Automat. Control, 41(11):1559–1578, 1996.
  • [12] L. Rosier. Homogeneous Lyapunov function for homogeneous continuous vector fields. Systems Control Lett., 19(6):467–473, 1992.
  • [13] R. G. Sanfelice and L. Praly. Convergence of nonlinear observers on n\mathbb{R}^{n} with a Riemannian metric (Part I). IEEE Trans. Automat. Control, 57(7):1709–1722, 2012.
  • [14] A. R. Teel and L. Praly. A smooth Lyapunov function from a class-𝒦\mathcal{K}\mathcal{L} estimate involving two positive semidefinite functions. ESAIM Control Optim. Calc. Var., 5:313–367, 2000.