Discontinuous Feedback and Nonlinear Systems
29 pages
English

Découvre YouScribe en t'inscrivant gratuitement

Je m'inscris

Discontinuous Feedback and Nonlinear Systems

-

Découvre YouScribe en t'inscrivant gratuitement

Je m'inscris
Obtenez un accès à la bibliothèque pour le consulter en ligne
En savoir plus
29 pages
English
Obtenez un accès à la bibliothèque pour le consulter en ligne
En savoir plus

Description

Niveau: Supérieur, Doctorat, Bac+8
Discontinuous Feedback and Nonlinear Systems ? Francis Clarke ? ?Universite de Lyon, Institut Camille Jordan 69622 Villeurbanne, France (e-mail: ) Abstract: This tutorial paper is devoted to the controllability and stability of control systems that are nonlinear, and for which, for whatever reason, linearization fails. We begin by motivating the need for two seemingly exotic tools: nonsmooth control-Lyapunov functions, and discontinuous feedbacks. Then, after a (very) short course on nonsmooth analysis, we build a theory around these tools. We proceed to apply it in various contexts, focusing principally on the design of discontinuous stabilizing feedbacks. Keywords: controllability, discontinuous control, feedback, nonlinear theory, stabilization 1. INTRODUCTION Our interest centers throughout on the standard control system x?(t) = f ( x(t), u(t) ) a.e., u(t) ? U a.e., (?) where the dynamics function f : Rn ? Rm ? Rn and the control set U ? Rm are given, and ‘a.e.' is the abbreviation of ‘almost everywhere'. A control on some interval [a, b] of interest refers to a measurable function u(·) defined on [a, b] and having values in U .

  • existence theory

  • optimal control

  • time function

  • nonlinear systems

  • lyapunov function

  • system

  • ‘max' has

  • hamilton–jacobi equation


Sujets

Informations

Publié par
Nombre de lectures 29
Langue English
Poids de l'ouvrage 1 Mo

Extrait

Discontinuous Feedback and Nonlinear Systems?
Francis Clarke
deLyit´enstion,ImaliutCtdrnaeloJinUsrev 69622 Villeurbanne, France (e-mail: clarke@math.univ-lyon1.fr)
Abstract:This tutorial paper is devoted to the controllability and stability of control systems that are nonlinear, and for which, for whatever reason, linearization fails. We begin by motivating the need for two seemingly exotic tools: nonsmooth control-Lyapunov functions, and discontinuous feedbacks. Then, after a (very) short course on nonsmooth analysis, we build a theory around these tools. We proceed to apply it in various contexts, focusing principally on the design of discontinuous stabilizing feedbacks.
Keywords: controllability, discontinuous control, feedback, nonlinear theory, stabilization
1. INTRODUCTION strategy that we know.) The positive features of such an approach, as well as certain inherent difficulties which arise Our interest centers throughout on the standard control when we employ it, are well illustrated by what is called system thedynamic programmingtechnique in optimal control. It x0(t) =fx(t), u(t)a.e., u(t)Ua.e.,(uobaehttdiugecnaalicthmaatem)iwllufswhuisrnuaalhvitgisnielbuootnithbilirstaonprzati,mnaboelivedpdor where the dynamics functionf:Rn×RmRnand the tools needed. control setURmare given, and ‘a.e.’ is the abbreviation of ‘almost everywhere’. Acontrolon some interval [a, b] of interest refers to a measurable functionu() definedDynamic programming and minimal time on [a, b] and having values inU. By atrajectoryof the Themini al-time problemrefers to finding a trajectory system ( ( of) we mean (as usual) an absolutely continuous) thatmreaches the origin as quickly as possible from state functionx: [a, b]Rncorresponding to some choice of controlu(nioplaittalfninivegidaimttniagoct.nor)αunioctsnuhT.ue(wse0ekthn[)o,eT]ivahleastTnght0e Standing hypotheses.It is assumed throughout thatf that the resulting trajectoryis propertyxwithx(0) =α continuous,Uis compact, andfis locally Lipschitz with satisfiesx(T) = 0. The dynamic programming approach respect to the state variable in the following sense: for centers upon theminimal-time functionT(), defined on every bounded subsetSRn, there existsK=Kf(S)Rnas follows:T(α) is the least timeTdefined above. such that Theprinciple of optimalitymakes two observations about f(x, u)f(y, u)Kfxyx, yS, uU.(1)T(). The first of these is that, for any trajectoryx() beginning atα, for any two timess, twith 0s < t, We remark that this Lipschitz behavior is automatically we have present iffis continuously differentiable, but differentia-bility offis irrelevant to our discussion. Much more to theTx(s)Tx(t)+ts.(2) point are the assumptions that arenotbeing made:fis This reflects the fact that, starting at the pointx(s), not linear,U strateis not simply ‘large enough’ to be effectively ignored.twreajemctaoyrychxtnuitliwtoosethemetylmilaogtpdenirocehenpandt,yg-otspegtinhefoofowll Thecentralissueunderdiscussionwillbetheconvergencetfrwoom-sttehpepstorianttex(gytmitehT.nigiroehtto)reqeiuerfdrohtsi);(2eithquneitalhtsigireisthfoed of state trajectoriesx(tbearemaerstbettsdebhyloteehacsuebygetargninnigewhcihw,ekatnequ)toariumilib to be the origin: stability, controllability, and feedback y stabilization. Stabilization to the origin is a simple rep- fromx(s). resentative of various other objectives that can be treated The second observation is th ho bythetechniquesthatweshalldescribe.(Stabilizationtoatequalitnyinlmdisniinm(2)ifx othertargetsetswillalsobeinvolvedlater.)itshaattirsa,jiefctxor(Tr=0fo)ojnisyhttaαTo=tT(ethαtcgirio;emitla).ehaftctserehTsi Onewaytosteertrajectoriestozeroistoinventacosttsthraattewghyenthxtehtedenopetdaberisc.veboiimasmanitil-trmeecajotyrt,ehersionebtteran whose minimization will have that effect. (Indeed, in a wo-s certain sense, this is rather close to being the only effective Combining these two observations, we find that, for any ?ˆrlooctneh´emematueduatiqahCeeri´htniroetrajectoryx(), the functiont7→Tx(t)+tis nondecreas-Institut universitaire de France.ing; it is constant whenxis a minimal-time trajectory.
Sincet7→Tx(t)+tis nondecreasing, we expect to have rTx(t), x0(t)+ 10, with equality whenx() is an optimal trajectory. The possible values ofx0f(t) fotr),aUtctjeraew,ytheiegnrobysilerpce elements of the setx at( arrive miUnrT(x), f(x, u)+ 1 = 0.(3) u We define the (lower)Hamiltonian functionhas follows: h(x, p) :=umiUnhp, f(x, u)i.(4) In terms ofh, the partial differential equation (3) above reads hx,rT(x)+ 1 = 0,(5) a special case of theHamilton–Jacobi equation. We have now reached the first stage in the dynamic pro-gramming approach: solve the Hamilton–Jacobi equation (5), together with the boundary conditionT(0) = 0, to find T(). How will this help us find minimal-time trajectories? To answer this question, we recall that an optimal trajec-tory is such that equality holds in (3). This suggests the following procedure: For eachx, letk(x) be a point inU satisfying umiUnrT(x), f(x, u)=rT(x), fx, k(x)=1.(6) Then, if we constructx() via the initial-value problem x0(t) =fx(t), kx(t), x(0) =α,(7)
we obtain a minimum-time trajectory (fromα). Let us see why this so: Ifx() satisfies (7), then, in light of (6), we have (d/dt)Tx(t)=rTx(t), x0(t)=rTx(t), fx(t), kx(t)=1.
Integrating, we find Tx(t)=T(α)t, which implies that atτ=T(α), we haT whencex(τ) = 0 (sinceTvetaylnooirezstxeh(τo)r0,).=inig Thereforex() is a minimal-time trajectory. This second stage of the dynamic programming approach has provided a feedbackk() which, fromanyinitial valueα, generates via (7) a minimal-time trajectory;k constitutes what can be considered the ultimate solution to our problem: anoptimal feedback synthesis. We remark that the Hamilton-Jacobi equation (5) has an-other use, when we know that it has a unique solutionT() satisfyingT(0) = 0 (namely, the minimal-time function). We refer to theverification methodin optimal control (see for example Clarke (1989)). It would work here as follows: Suppose we have formulated a conjecture that, for eachα, a certain trajectoryxαis a minimal-time one from the initial conditionα. We proceed to calculateT(α) (provisionally) based on this conjecture; that is, by setting T(αequal to the time required for) xαto joinαto 0. Then, if the resulting functionTsatisfies (5), our conjecture is
verified (since, by uniqueness,Tmust then coincide with the minimal-time function). IfTfails to satisfy (5), then our conjecture is certainly false (and the way in which (5) fails may help us amend it). We now rain on this parade by pointing out that there are serious obstacles to rigorously justifying the route that we have just outlined. There is, to begin with, the issue ofcontrollability: Is it always possible to steerαto 0 in finite time? And if this holds, do minimal-time trajectories exist? Even if this is true, how do we know thatT() is differentiable? If this fails to be the case, then we shall need to replace the gradientrTused above by some suitably generalized derivative. Next, we would have to examine anew the argument that led to the Hamilton– Jacobi equation (5), which itself will require reformulation in some way that allows for nonsmooth solutions. Will the Hamilton–Jacobi equation generalized in such a way admit Tas the unique solution? Assuming that all this can be done, the second stage above offers fresh difficulties of its own. Even ifTis smooth, there is in general nocontinuousfunctionk() satisfying (6) for eachx. Whenkis discontinuous, the classical concept of ‘solution’ to (7) is inappropriate; what solution concept should we use instead? Would optimal trajectories still result? That these difficulties are real, and indeed that they arise in the simplest problems, can be illustrated by the following example, familiar from any introductory text in optimal control. The double integrator.This refers to the systemx00=u, or, in terms of the standard formulation (): x0(t) =y(t), y0(t) =u(t), u(t)[1,+1].(8) Thusn= 2, m= 1, and the dynamics are linear. It is not difficult to show that all initial points (x(0), y(0)) = (α, β) are controllable to the origin in finite time; existence theory tells us that minimal-time trajectories exist. The Maximum Principle (see Section 2) helps us to identify them: they turn out to be bang-bang with at most one switch between +1 and1. We can then calculate the minimal-time functionT(): T(α, β) =+ββ++pp22ββ22+44ααwh(ewnhen(β,α,βαgiir)sflostfheotf)iSS where theswitching curveSin thex-yplane is given by y2= 2|x|; see Fig. 1. The resulting functionT() is seen to be continuous, but it fails to be differentiable or even locally Lipschitz along the switching curve. The optimal feedback synthesis consists of takingk=1 to the right or on the upper branch) ofS, andk= +1 otherwise. We see therefore that our doubts correspond to real diffi-culties, and they explain why the dynamic programming approach to optimal control, very prominent in the 1950s and 60s, is now frequently ignored in engineering texts, or else relegated to a heuristic role, perhaps in exercises. In fact, however, the difficulties have now been successfully and rigorously resolved, through the use of nonsmooth analysis, viscosity solutions, and discontinuous feedbacks.
  • Univers Univers
  • Ebooks Ebooks
  • Livres audio Livres audio
  • Presse Presse
  • Podcasts Podcasts
  • BD BD
  • Documents Documents