• Aucun résultat trouvé

Local nonsmooth Lyapunov pairs for first-order evolution differential inclusions

N/A
N/A
Protected

Academic year: 2021

Partager "Local nonsmooth Lyapunov pairs for first-order evolution differential inclusions"

Copied!
22
0
0

Texte intégral

(1)

HAL Id: hal-00822994

https://hal.archives-ouvertes.fr/hal-00822994

Submitted on 16 May 2013

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Local nonsmooth Lyapunov pairs for first-order evolution differential inclusions

Samir Adly, Abderrahim Hantoute, Michel Théra

To cite this version:

Samir Adly, Abderrahim Hantoute, Michel Théra. Local nonsmooth Lyapunov pairs for first-order

evolution differential inclusions. 2013. �hal-00822994�

(2)

(will be inserted by the editor)

Local nonsmooth Lyapunov pairs for first-order evolution differential inclusions

Samir Adly · Abderrahim Hantoute · Michel Th´era

Received: date / Accepted: date

Abstract The general theory of Lyapunov’s stability of first-order differential inclusions in Hilbert spaces has been studied by the authors in a previous work [2]. This new contribution focuses on the natural case when the maximally monotone operator governing the given inclusion has a domain with nonempty interior. This setting permits to have nonincreasing Lyapunov functions on the whole trajectory of the solution to the given differential inclusion. It also allows some more explicit criteria for Lyapunov’s pairs. Some consequences to the viability of closed sets are given, as well as some useful cases relying on the continuity or/and convexity of the involved functions. Our analysis makes use of standard tools from convex and variational analysis.

Keywords Evolution differential inclusions, lower semi-continuous functions, invariant sets, proximal subdifferential, Clarke subdifferential, Fr´echet subdifferential, limiting proximal subdifferential, maximally monotone operator, strong solution, weak solution, Lyapunov pair.

Mathematics Subject Classification (2000) 37B25, 47J35, 93B05

1 Introduction and notations

In various applications modeled by ODE’s, one may be forced to work with systems that have non-differentiable solu- tions. Also, Lyapunov’s functions, that is positive definite functions whose decay along the trajectories of the system, which are used to establish a stability property of the system, may be nondifferentiable. The need to extend the classical differentiable Lyapunov’s stability to the nonsmooth case is unavoidable when studying stability properties of discon- tinuous systems. In practice, many systems in physics, engineering, biology etc exhibit generally nonsmooth energy functions, which are usually a typical candidates for Lyapunov functions; thus elements of nonsmooth analysis become essential [3, 16, 18, 25]. A typical example is given by the case of piecewise linear dynamical systems calledLinear Com- plementarity Systems(LCS) for which the analysis of asymptotic and exponential stability uses a piecewise quadratic Lyapunov function [18]. Let us remind that LCS are defined as follows:

LCS(A, B, C, D)

x(t;˙ x0) =Ax(t) +Bu(t), x(t0) =x0, 0≤u(t)⊥Cx+Du≥0,

whereA ∈Rn×n,B ∈ Rn×m,C ∈ Rm×nandD ∈ Rm×mare real matrices,x0is the initial condition,x˙ is the time derivative of the trajectoryx(t)anda⊥bmeans that the two vectorsaandbare orthogonal. Linear and nonlinear complementarity problems belong to the more general mathematical formalism ofDifferential Variational Inequalities (DVI), introduced by J.S. Pang and D. Stewart [21]. It is a combination of an ordinary differential equation (ODE) with

S. Adly

Universit´e de Limoges, Laboratoire XLIM. 123, avenue Albert Thomas, 87060 Limoges CEDEX, France E-mail: samir.adly@unilim.fr

A. Hantoute

University of Chile, CMM Piso 7, Avenida Blanco Encalada 2120, Santiago, Chile . E-mail: ahantoute@dim.uchile.cl

M. Th´era

Universit´e de Limoges, Laboratoire XLIM. 123, avenue Albert Thomas, 87060 Limoges CEDEX, France E-mail: michel.thera@unilim.fr

(3)

a variational inequality or a complementarity constraint. A DVI consists to find trajectoriest7→x(t)andt7→u(t)such that

DV I(f, F, K)

x(t) =˙ f(t, x(t), u(t)), x(t0) =x0,

hF(t, x(t), u(t)), v−u(t)i≥0, ∀v∈K, u(t)∈Kfor a.e.t≥t0,

whereKis a closed convex subset of a Hilbert spaceH,fandFaregiven mappings. WhenKis a closed convex cone, then problemDV I(f, F, K)is equivalent to a Differential Complementarity Problem (DCP):

DCP(f, F, K)

x(t) =˙ f(t, x(t), u(t)), x(t0) =x0,

K∋u(t)⊥F(t, x(t), u(t))∈K, a.e.t≥t0.

Since DVI and DCP formalisms unify several known mathematical problems such that ordinary differentialequations with discontinuous right-hand term, differential algebraic equations, dynamic complementarity problems etc .. (see [9, 10] for more details), it was proved to be powerful for the treatment of many problems in science and engineering such that: unilateral contact problems in mechanics, finance, traffic networks, electrical circuits etc . . . . According also to the fact that LCS formalism has many of applications in various areas including for instance robotics, economics, finance, non smooth mechanics, etc (see Camlibel, Pang and Shen, [18] and the monograph by Facchinei and Pang, [17]), it has received recently a great interest from the mathematical programming and control communities from the theoretical and numerical point of view.

Instead of considering LCSs or DVIs, throughout this contribution we are interested in the general framework of infinite-dimensional dynamical systems, that is systems of the form:

˙

x(t;x0)∈f(x(·;x0))−Ax(·;x0), x0∈cl(DomA) a.e. t≥0. (1.1) Here, and thereafter,cl(DomA)is the closure of the domain of a maximally monotone operatorA:H ⇉Hdefined on a real Hilbert spaceH, possibly nonlinear and multivalued with domainDomAandf is a Lipschitz continuous mapping defined oncl(DomA).

A pair of proper lower semicontinuous (lsc for short) functionsV, W :H→R∪ {+∞}is said to form a Lyapunov pair for (1.1) if for allx0 ∈cl(DomA)the solution of (1.1), in a sense that will be precised in Section 3, denoted by x(·;·, x0)satisfies

V(x(t;x0))−V(x(s;x0)) + Z t

s

W(x(τ;x0))dτ ≤0for allt≥s≥0. (1.2) Observe that whenW ≡0one recovers the classical notion of Lyapunov functions; e.g., [27]. More generally, instead of (1.2), we are going to consider functionsV, Wsatisfying for somea≥0

eatV(x(t;x0))−easV(x(s;x0)) + Z t

s

W(x(τ;x0))dτfor allt≥s≥0.

In this case, the (weighted) pair (V, W)will be refered to as a a-Lyapunov pair. The main motivation in usinga- Lyapunov pairs instead of simply functions is that many stability concepts for the equilibrium sets of (1.1), namely stability, asymptotic or finite-time stability, can be obtained just by choosing appropriate functions W in (1.2). The weighteatis useful for instance when exponential stability is concerned. So, even in autonomous systems like those of (1.1), the functionWor the weighteatmay be of a certain utility since, in some sense, it emphasizes the decreasing of the Lyapunov functionV.

The method of Lyapunov functions is a corner stone of the study of the controllability and stabilizability of control systems. Its history is rich and has been described in several places and various seminal contributions has been made to the subject. We refer to Clarke [14, 15] for an overview of the recent developments of the theory where he pointed out that for nonlinear systems, Lyapunov’s method turns out to be essential to consider nonsmooth Lyapunov functions, even if the underlying control dynamics are themselves smooth.

Over the years, among the various contributions, Kocan & Soravia [19], characterized Lyapunov’s pairs in terms of viscosity solutions of a related partial differential inequality.

Another well-established approach consists of characterizing Lyapunov’s pairs by means of the contingent derivative of the maximally monotone operatorA, see for instance Cˆarj˘a & Motreanu [11], for the case of a maximally linear monotone operator and also whenAis a multivalued m-accretive operator on an arbitrary Banach space [12]. In these approaches the authors used tangency and flow-invariance arguments combined with a priori estimates and approxima- tion.

The starting point of this contribution is the paper by Adly & Goeleven [1] in which smooth Lyapunov functions were used in the framework of the second order differential equations, and non-linear mechanical systems with frictional unilateral constraints.

In this article we provide a different approach that don’t make use of viscosity solutions or contingent derivatives associated to the operatorA. Our objective is to emphasize our previous contribution [2] to the setting where the involved

(4)

maximally monotone operator has a domain with nonempty interior. This case includes the finite dimensional framework since in this case the relative interior of the domain of the operator is always nonempty. Moreover, the criteria for Lyapunov’s pairs are checked only in the interior of the domain (or the relative interior) instead of the closure of the whole domain as in [1]. In contrary to [1], this setting also ensures obtaining global Lyapunov’s pairs and permits in this way to control the whole trajectory of the solution to the given differential inclusion.

The summary of the paper is as follows. In Section 2 we introduce the main tools and basic results used in the paper. In Section 3 we give a new primal and dual criteria for lower semicontinuous Lyapunov pairs. This is achieved in Proposition 2 and Theorem 3.1. In Section 4, we make a review of some old and recent criteria for Lyapunov pairs.

Section 5 is dedicated to complete the proofs of the main results given in Section 3.

2 Notation and main tools

Throughout the paper,H is a (real) Hilbert space endowed with the inner (or scalar) producth·,·iand the associated norm is denoted byk·k. We identifyH(the space of continuous linear functionals defined onH) toH,and we denote the weak limits (w−lim,for short) by the symbol⇀to distinguish it from the usual symbol→used for strong limits.

The zero vector inHis denoted byθ.

We start this section by reviewing some notations used throughout the paper. Given a nonempty setS ⊂ H (or S ⊂ H×R), bycoS,coneS, and affS, we denote theconvex hull, the conic hull, and the affine hull of the set S, respectively. Moreover,IntS is the topologicalinteriorofS, andclS andS are indistinctly used for theclosure ofS(with respect to the norm topology onH). We also useclwS orSw when we deal with the closure ofS with respect to the weak topology. We noteriSthe (topological)relative interiorofS, i.e., the interior ofSin the topology relative toaffSwhatever this set is nonempty (see [23, Chapter 6] for more on this fundamental notion). Forx ∈H (orx∈H×R),ρ≥0, Bρ(x)is the open ball with centerxand radiusρ,andBρ(x)is the closure ofBρ(x), while B:=B1(θ)stands for the unit open ball. Fora, b∈R:=R∪ {+∞,−∞}we denote[a, b)the interval closed ataand open atb([a, b],(a, b), ...are defined similarly); henceR+:= [0,∞). Finally, forα∈R,we noteα+formax{0, α}.

Our notation is the standard one used in convex and variational analysis and in monotone operator theory; see, e.g., [8, 24]. Theindicator functionofSis the function defined as

IS(x) :=

0 ifx∈S +∞otherwise.

Thedistance functiontoSis denoted by

d(x, S) := inf{kx−yk |y∈S}, and theorthogonal projectiononS,πS, is defined as

πS(x) :={y∈S| kx−yk=d(x, S)}. IfSis closed and convex,S⊂H(orH×R) denotes itsrecession cone:

S:={y|x+λy∈Sfor somexand allλ≥0}, while,S ⊂H(orH×R) denotes the polar ofSgiven by

S:={y| hy, vi ≤1for allv∈S}. Given a functionϕ:H→R, its(effective) domainandepigraphare defined by

Domϕ:={x∈H|ϕ(x)<+∞}, epiϕ:={(x, α)∈H×R|ϕ(x)≤α}. Forλ∈R, theopen sublevel setofϕatλis

[ϕ > λ] :={x∈H|ϕ(x)> λ};

[ϕ≥λ],[ϕ≤ λ], and[ϕ < λ]are defined similarly. We say thatϕis proper ifDomϕ6=∅andϕ(x)> −∞for all x∈H.We say thatϕis convex ifepiϕis convex, and (weakly) lower semicontinuous (lsc, for short) ifepiϕis closed with respect to the (weak topology) norm-topology onH. We denote

F(H) :={ϕ:H →R|ϕis proper and lsc}, Fw(H) :={ϕ:H →R|ϕis proper and weakly lsc};

(5)

F(H;R+)andFw(H;R+)stand for the subsets of nonnegative functions ofF(H)andFw(H), respectively.

As maximally monotone set-valued operators play an important role in this work, it is useful to recall some of basic definitions and some of their properties. More generally, they have frequently shown themselves to be a key class of objects in both modern Optimization and Analysis; see, e.g., [4–6, 8, 24, 26].

For an operatorA:H⇉H,thedomainand thegraphofAare given respectively by

DomA:={z∈H|Az6=∅}and gphA:={(x, y)∈H×H|y∈Ax};

for notational simplicity we identify the operatorAto its graph. Theinverse operatorofA, denoted byA−1, is defined as

(y, x)∈A−1⇐⇒(x, y)∈A.

We say that an operatorAismonotoneif

hy1−y2, x1−x2i ≥0for all(x1, y1),(x2, y2)∈A,

andmaximally monotoneifAis monotone and has no proper monotone extension (in the sense of graph inclusion). If Ais maximally monotone, it is well known (e.g., [26]) thatDomAis convex, andAxis convex and closed for every x ∈ DomA. Moreover, ifInt(DomA)6= ∅, thenInt(DomA)is convex,Int(DomA) = Int(DomA), andAis bounded locally onInt(DomA). Note that the domain or the range of a maximally monotone operator may fail to be convex, see, e.g., [24, page 555]. In particular, ifAis the subdifferential∂ϕof some lower semicontinuous (lsc for short) convex and proper functionϕ:H→R, thenAis a classical example of a maximally monotone operator, as is a linear operator with a positive symmetric part. We know that

DomA⊂Domϕ⊂Domϕ= DomA.

Forx∈DomA,we shall use the notation(Ax)to denote theprincipal sectionofA,i.e., the set of points of minimal norm inAx.

Nonsmooth and variational analysis play a central role in this study. Hence, we need to recall briefly some concepts used through the paper. More details can be found for instance in [7, 13, 16, 20, 24]. We assume thatϕ∈ F(H),and take x∈Domϕ.

A vectorξ ∈His called aproximal subgradientofϕatx, writtenξ ∈∂Pϕ(x),if there areρ >0andσ≥0such that

ϕ(y)≥ϕ(x) +hξ, y−xi −σky−xk2 for ally∈Bρ(x);

thedomain ofPϕis then given by

Dom∂Pϕ:={x∈H|∂Pϕ(x)6=∅}. The set∂Pϕ(x)is convex, possibly empty and not necessarily closed.

A vectorξ∈His called aFr´echet subgradientofϕatx, writtenξ∈∂Fϕ(x),if ϕ(y)≥ϕ(x) +hξ, y−xi+o(ky−xk).

Associated to proximal and Fr´echet subdifferentials, limiting objects have been introduced. A vectorξ ∈Hbelongs to thelimiting proximal subdifferentialofϕatx, written∂Lϕ(x), if there exist sequences(xk)kNand(ξk)kNsuch that xk

ϕ x(that is,xk⇀ xandϕ(xk)→ϕ(x)),ξk∈∂Pϕ(xk)andξk⇀ ξ.

A vectorξ ∈ H is called a horisontal subgradient of ϕat x, writtenξ ∈ ∂ϕ(x), if there exist sequences (αk)kN⊂R+,(xk)kNand(ξk)kNsuch thatαk→0+, xkϕ x,ξk∈∂Pϕ(xk)andαkξk⇀ ξ.

TheClarke subdifferentialofϕatxis defined by the following so-calledrepresentation formula; see, e.g., Mor- dukhovich [20] and Rockafellar [24],

Cϕ(x) = cow{∂Lϕ(x) +∂ϕ(x)}.

From a geometrical point of view, ifS⊂His closed andx∈S,theproximal normal cone toSatxis NPS(x) :=∂PIS(x).

We also denote byNePS(x)the subset ofNPS(x)given by

NePS(x) :={ξ∈H| hξ, y−xi ≤ ky−xk2 for ally∈Sclosed tox}. It can be proved; e.g., [13], that

NPS(x) =

cone(πS−1(x)−x), ifπ−1S (x)6=∅, {θ} ifπ−1S (x) =∅,

(6)

whereπS−1(x) :={y∈H\S|x∈πS(y)}.

Similarly,NLS(x) :=∂LIS(x)(=∂IS(x)) is thelimiting normal cone toSatx,andNCS(x) := cow{NLS(x)}is theClarke normal cone toSatx.

In that way, the above subdifferentials ofϕ∈ F(H)can be geometrically described as

Pϕ(x) ={ξ∈H |(ξ,−1)∈NPepiϕ(x, ϕ(x))},

ϕ(x) ={ξ∈H |(ξ,0)∈NPepiϕ(x, ϕ(x))}.

We callcontingent cone toSatx∈S(or theBouligand tangent cone), writtenTS(x),the cone given by TS(x) :={ξ∈H|x+τkξk∈Sfor someξk→ξandτk→0+}.

The Dini directional derivative of the functionϕ(∈ F(H))atx∈Domϕin the directionv∈His given by ϕ(x, v) = lim inf

t→0+,wv

ϕ(x+tw)−ϕ(x)

t .

Hence,epiϕ(x,·) = Tepiϕ(x, ϕ(x)).TheGˆateaux derivativeofϕatxis a linear continuous form onH,written ϕG(x),satisfying

tlim→0+

ϕ(x+tv)−ϕ(x)

t =hϕG(x), vi for allv∈H.

We close this section by giving some properties of the subdifferential sets defined above that will be used later on. First, it follows easily from the definitions that

Pϕ(x)⊂∂Fϕ(x)⊂∂Lϕ(x)⊂∂Cϕ(x).

Ifϕis convex, then

Pϕ(x) =∂Cϕ(x) =∂ϕ(x), where∂ϕ(x)is the usualMoreau-Rockafellarsubdifferential ofϕatx:

∂ϕ(x) :={ξ∈H|ϕ(y)−ϕ(x)≥ hξ, y−xifor ally∈H}. Ifϕ∈ F(H)is Gˆateaux-differentiable atx∈Domϕ,we have

Pϕ(x)⊂ {ϕG(x)} ⊂∂Cϕ(x).

IfϕisC1then

Pϕ(x)⊂ {ϕ(x)}=∂Cϕ(x)and∂ϕ(x) ={θ}. IfϕisC2then

Pϕ(x) =∂Cϕ(x) ={ϕ(x)}. In particular, ifϕ:=d(·, S)withS⊂Hclosed, forx∈Swe have that

Cϕ(x) = NCS(x)∩B, while, forx6∈Ssuch that∂Pϕ(x)6=∅, πS(x)is a singleton and (e.g., [16])

Pϕ(x) = x−πS(x) ϕ(x) ; hence

Lϕ(x) =

w−lim

k

xk−πS(xk)

ϕ(x) ;xk⇀ x

. More generally, we have that

NPS(x) =R+PdS(x)andNCS(x) =R+CdS(x)w (with the convention that0.∅={θ}).

Finally, we recall thatϕ∈ F(R)is nonincreasing if and only ifξ≤0for everyξ∈∂Pϕ(x)andx∈R,(e.g., [16]).

We shall use the following version of the Gronwall Lemma (e.g., [1, Lemma 1]).

Lemma 1 Givent2 > t1≥0,a6= 0, andb≥0,we assume that an absolutely continuous functionψ: [t1, t2]→R+ satisfies

ψ(t)≤aψ(t) +b a.e.t∈[t1, t2].

Then, for allt∈[t1, t2],

ψ(t)≤(ψ(t1) + b

a)ea(tt1)− b a.

(7)

3 Local characterization of Lyapunov pairs on the interior of the domain ofA

In this section we provide the desired explicit criterion for lower semicontinuous (weighted-) Lyapunov pairs associated to the differential inclusion (1.1):

˙

x(t;x0)∈f(x(·;x0))−Ax(·;x0), x0∈cl (DomA),

whereA : H ⇉ H is a maximally monotone operator andf : cl (DomA) ⊂ H → H is a Lipschitz continuous mapping. Recall that for fixedT >0andx0 ∈cl (DomA),a strong solution of (1.1),x(·;x0) : [0, T] →H,is a uniquely defined absolute continuous function which satisfiesx(0;x0) =x0together with (see, e.g., [8])

˙

x(t;x0)∈Lloc((0, T], H), (3.3)

x(t;x0)∈DomA, for allt >0, (3.4)

˙

x(t;x0)∈f(x(t;x0))−Ax(t;x0), a.e.t≥0. (3.5) Existence of strong solutions is known to occur if for instance:

– x0∈DomA,Int (co (DomA))6=∅; – dimH <∞;

– or ifA≡∂ϕwhereϕ:H→R∪ {+∞}is a lsc extended-real-valued convex proper function.

Moreover, we have thatx(˙ ·;x0)∈L([0, T], H)if and only ifx0∈DomA. In this later case,x(·;x0)is derivable from right at eachs∈[0, T)and

d+x(·;x0)

t (s) =f(x(s;x0))−πAx(s;x0)(f(x(s;x0))).

The strong solution also satisfies the so-called semi-group property,

x(s;x(t;x0)) =x(s+t;x0)for alls, t≥0, (3.6) together with the relationship

kx(t;x0)−x(t;y0)k ≤eLftkx0−y0k

whenevert≥0andx0, y0∈cl(DomA);hereafter,Lfdenotes the Lipschitz constant of the mappingfoncl(DomA).

In the general case, it is well established that (1.1) admits a unique weak solutionx(·;x0) ∈ C(0, T;H)which satisfiesx(t;x0)∈cl(DomA)for allt≥0. More precisely, there exists a sequence(xn)n∈N ⊂DomAconverging tox0such that the strong solutionxk(·;zk)of the equation

˙

xk(t;zk)∈f(x(t;zk))−Axk(t;zk), xk(0, zk) =zk, (3.7) converges uniformly tox(·;x0)on[0, T].Moreover, we have that

x(s;x(t;x0)) =x(s+t;x0)for alls, t≥0 (3.8) (called the semigroup property). IfLf denotes the Lipschitz constant off oncl(DomA),then for everyt ≥ 0and x0, y0∈cl(DomA)we have that

kx(t;x0)−x(t;y0)k ≤eLftkx0−y0k.

In the remaining part of the paper,x(·;x0)denotes the weak solution of Equation (1.1) (which is also, a strong one whenever a strong solution exists.)

From now on, we suppose throughout this section that

Int (co (DomA))6=∅.

Hence,Int (DomA)is convex,Int (DomA) = Int (co (DomA)) = Int (cl (DomA)),andAis locally bounded on Int (DomA). Therefore, a (unique) strong solution of (1.1) always exists [8]. We have the following technical lemma, adding more information about the qualitative behavior of this solution.

(8)

Lemma 2 Lety¯∈DomAandρ >0be such thatBρ(¯y)⊂Int (co (DomA)).Then,M := supz∈Bρy)k(f(z)−Az)k<

and for ally∈Bρ(¯y)andt≤1we have that

d+x(·;y) dt (t)

≤eLfM.

Proof By virtue of the semi-group property (3.6), the following inequality holds for ally∈cl(DomA)and0≤t < s (e.g., [8, Lemma 1.1])

kx(t+s;y)−x(t;y)k=kx(t;x(s;y))−x(t;y)k ≤eLftkx(s;y)−yk. (3.9) In particular, fory∈Bρ(¯y)andt≤1we get that

d+x(·;y) dt (t)

= lim

s↓0s−1kx(t+s;y)−x(t;y)k ≤eLftlim

s↓0s−1kx(s;y)−yk

=eLft

d+x(·;y) dt (0)

=eLft(f(y)−Ay)≤eLfM.

The fact thatMis finite follows from the maximal monotonicity ofAtogether with the Lipschitz continuity off.△ Definition 1 Let be given functionsV ∈ F(H), W ∈ F(H;R+)and a numbera∈R+.We say that(V, W)forms a a-Lyapunov pair for(1.1)with respect to a setD⊂cl(DomA)if for ally∈Dwe have that

eatV(x(t;y)) + Zt

0

W(x(τ;y))dτ≤V(y)for allt≥0. (3.10) a-Lyapunov pairs with respect tocl(DomA)are simply calleda-Lyapunov pairs (see [2]); in addition, ifa = 0and W = 0, we recover the classical concept of Lyapunov functions. The caseD = Int(DomA)(when nonmepty), or D = ri(DomA)in the finite-dimensional setting, is useful too since it allows recovering the behaviour ofV on the whole setcl(DomA)when, as in Proposition 1 below, some continuity conditions onV are known. More precisely, our characterization theorem, Theorem 3.1 below, provides criteria for Lyapunov pairs with respect to small sets, for instance balls, rather than the whole setInt(DomA).The lack of regularity properties ofa-Lyapunov pairs(V, W)in Definition 1 is mainly due to the non-smoothness of the functionV.Let us remind that inequality (3.10) also holds if instead of W one considers its Moreau-Yosida regularization, which is Lipschitz continuous on every bounded subset ofH. This follows from the next Lemma 3 (e.g [2]).

Lemma 3 For everyW ∈ F(H;R+),there exists a sequence of functions(Wk)k∈N ⊂ F(H,R+)converging toW (for instance,Wk↑W) such that eachWkis Lipschitz continuous on every bounded subset ofH,and satisfiesV(y)>0 if and only ifVk(y)>0.

Consequently, ifV, D ⊂ cl(DomA),anda ∈ R+ are as in Definition 1 then, with respect toD,(V, W)forms an a-Lyapunov pair for(1.1)if and only if each pair(V, Wk)forms ana-Lyapunov pair for(1.1).

Proposition 1 Let be given functionsV ∈ F(H), W ∈ F(H;R+)and a numbera∈R+.IfV verifies lim inf

DomAzyV(z) =V(y) for ally∈cl(DomA)∩DomV, (3.11) then it is equivalent to saying that(V, W)forms ana-Lyapunov pair with respect to eitherDomAorcl(DomA).

Property (3.11) has been already used in [19], and implicitely in [22], among other works. It holds, if for instance,V (∈F(H)) is convex and its effective domain has a nonempty interior such thatInt(DomV)⊂DomA.

Our starting point is the next result which characterizesa-Lyapunov pairs locally inInt(DomA).The general form corresponding toa-Lyapunov pairs incl(DomA)was recently established in [2]. For the reader convenience we include here a sketch of the proof.

Proposition 2 Assume thatInt (co{DomA})6=∅.LetV ∈Fw(H)satisfyDomV ⊂cl(DomA),W ∈F(H;R+), anda∈R+.Lety¯∈H,λ¯∈[−∞, V(¯y)),andρ¯∈(0,+∞]be such that

DomV ∩Bρ¯(¯y)∩[V >λ]¯ ⊂Int(DomA).

(9)

Then, the following statements are equivalent:

(i)∀y∈DomV ∩Bρ¯(¯y)∩[V >¯λ]

sup

ξPV(y)

υ∈Aymin hξ, f(y)−υi+aV(y) +W(y)≤0;

(ii)∀y∈DomV ∩Bρ¯(¯y)∩[V >¯λ]

sup

ξPV(y)hξ, f(y)−πAy(f(y))i+aV(y) +W(y)≤0;

(iii)∀y∈Bρ¯(¯y)∩[V >λ]¯ we have that eatV(x(t;y)) +

Z t 0

W(x(τ;y))dτ≤V(y) ∀t∈[0, ρ(y)], where

ρ(y) := sup



ν >0

∃ρ >0s.t. Bρ(y)⊂Bρ¯(¯y)∩[V >λ],¯ and for allt∈[0, ν]

2kx(t;y)−yk< ρ2 and

(eat−1)V(y)−R0tW(x(τ;y))dτ< ρ2



. (3.12) Remark 1 (Before the proof) the constantρ(y)defined in (3.12) is positive whenevery∈cl(DomA)∩Bρ¯(¯y)∩[V >¯λ].

Hence, whenρ¯=−¯λ=∞one can easily show that (iii) is equivalent to (see [2, Proposition 3.2]) eatV(x(t;y)) +

Z t 0

W(x(τ;y))dτ ≤V(y) for allt≥0;

that is,(V, W)forms a Lyapunov pair with respect toInt(DomA).

Proof For simplicity, we suppose thatW ≡0.

(iii) =⇒ (ii) Let us fixy∈Bρ¯(¯y)∩[V >λ]¯ andξ∈∂PV(y)so thaty∈Bρ¯(¯y)∩[V >λ]¯ ∩DomV ⊂DomA and there existα >0andT ∈(0, ρ(y))such that

hξ, x(t;y)−yi ≤V(x(t;y))−V(y) +αkx(t;y)−yk2≤αkx(t;y)−yk2 for allt∈[0, T).

Buty∈DomAand so there exists a constantl≥0such that

hξ, t−1(x(t;y)−y)i ≤lkx(t;y)−yk for allt∈[0, T);

hence, taking the limit ast→0+we obtain that

hξ, f(y)−πAy(f(y))i ≤0;

that is, (ii) follows.

(i) =⇒ (iii) To simplify the proof of this part, we assume thatf ≡ 0, W ≡ 0anda = 0.For this aim we fix y∈DomV ∩Bρ¯(¯y)∩[V >¯λ]and letρ >0andv >0be such that

Bρ(y)⊂Bρ¯(¯y)∩[V >¯λ] and (3.13) sup

t∈[0,ν]

2kx(t;y)−yk< ρ; (3.14)

the existence of such scalarsρandvis a consequence of the lower semicontinuity ofV and the Lipschitz continuity of x(·;·)(see Lemma 2). LetT < νbe fixed and define the functionsz(·) : [0, T]⊂R+ →H×Randη(·) : [0, T]⊂ R+→R+as

z(t) := (x(t;y), V(y)), η(t) := 1

2d2(z(t),epiV); (3.15)

observe thatz(·)andη(·)are Lipschitz continuous on[0, T).Now, using a standard chain rule (e.g. [13]), for fixed t∈(0, T)it holds that

Cη(t) =d(z(t),epiV)∂Cd(z(·),epiV)(t).

So, from one hand we get∂Cη(t) ={θ}wheneverz(t)∈epiV.On the other hand, whenz(t)6∈epiV we obtain that

Cη(t)⊂co

 [

(u,µ)∈ΠepiV(z(t)), uBρ(y)

hx(t;y)−u,−Ax(t;y)i

; (3.16)

(10)

the fact thatu∈Bρ(y)is a consequence of the following inequalities:

ku−yk ≤ kx(t;y)−uk+kx(t;y)−yk

≤ k(x(t;y), V(y))−(u, µ)k+kx(t;y)−yk

≤ k(x(t;y), V(y))−(y, V(y))k+kx(t;y)−yk

≤2kx(t;y)−yk< ρ.

Take nowξ∈Ax(t;y)and(u, µ)∈ΠepiV(z(t))withu∈Bρ(y)so thatV(y)−µ≤0andu∈DomV∩Bρ¯(¯y)∩ [V >λ]¯ (recall (3.13)).

IfV(y)−µ < 0,we write(µ−V(y))−1(x(t;y)−u) ∈∂VP(u).Then, by the current assumption (i), select

υ∈Ausuch that D

(µ−V(y))−1(x(t;y)−u),−υE≤0.

Therefore, invoking the monotonicty ofAwe get

hx(t;y)−u,−ϑi=hx(t;y)−u,−υi+hx(t;y)−u, υ−ϑi ≤ hx(t;y)−u,−υi ≤0.

Sinceξ∈Ax(t;y)is arbitrary and according to (3.16), we deduce that∂Cη(t)⊂R.

IfV(y)−µ= 0so that(x(t;y)−u,0)∈ NepiV(u, V(u))andx(t;y)−u 6= θ.Then, for a fixedε > 0such thatBε(u)⊂Bρ(y)∩Int(Dom (A))(recall thatu∈Bρ(y)∩DomV ∩Bρ¯(¯y)∩[V >λ]¯ ⊂Int(Dom (A))), take uε ∈Bε(u)∩DomV with|V(u)−V(uε)| ≤ε, α ∈(0, ε)andξ ∈Bε(x(t;y)−u)such thatα−1ξ ∈∂VP(uε) (see, e.g., [13, Theorem 2.4]). Hence, using the current assumption, selectξε∈Auεsuch thathξ,−ξεi ≤αε. Hence,

hx(t;y)−u,−ξεi ≤εkξεk+ξ,−uε

≤εkξεk+αε≤εkξεk+ε2. By the monotonicity ofAthis yields

hx(t;y)−u,−ϑi ≤ hx(t;y)−uε,−ϑi+εkϑk

≤ hx(t;y)−uε,−ξεi+εkϑk

≤ hx(t;y)−u,−ξεi+kuε−uk kξεk+εkϑk

≤2εkξεk+εkϑk+ε2.

Moreover, as(uε)ε≤1is bounded inInt(Dom (A)),the net(ξε)εis also bounded and passing to the limit asεgoes to 0we get

hx(t;y)−u,−ϑi ≤0.

This gives the desired inclusion∂Cη(t)⊂R(recall (3.16)) and so establishes the proof of (iii). △

We are now ready to give the main result of this section, which provides a precise improvement of Proposition 2.

Theorem 3.1 Assume thatInt (co{DomA})6=∅.LetV ∈Fw(H)withinfV >−∞, W ∈F(H;R+),anda∈R+

be given. Fixy¯∈DomV,¯λ∈(−∞, V(¯y))and letρ >¯ 0be such that

DomV ∩[V >¯λ]∩Bρ¯(¯y)⊂Int(DomA).

Then, the following statements are equivalent:

(i)∀y∈DomV ∩Bρ¯(¯y)∩[V >¯λ]

sup

ξPV(y)

υ∈Aymin hξ, f(y)−υi+aV(y) +W(y)≤0;

(ii) (IfV is weakly continuous when restricted toBρ(¯y))∀y∈DomV ∩Bρ¯(¯y)∩[V >¯λ]

eatV(x(t;y)) + Z t

0

W(x(τ;y))dτ ≤V(y) for allt≥0.

Consequently, if (i)-(ii) holds onInt(DomA),the pair(V, W)is ana-Lyapunov pair for (1.1) with respect tocl(DomA).

(11)

Proof The consequence is immediate once we prove the main conclusion.

First, invoking Lemma 3 we may assume w.l.o.g. thatW is Lipschitz continuous on every bounded subset ofH.In the rest of the proof, we takeyˆinDomV ∩Bρ¯(¯y)∩[V >¯λ](⊂Int(DomA)) and, taking into account the lsc ofV, chooseρ >0such thatB(ˆy)⊂Bρ¯(¯y)∩[V >¯λ]∩Int(DomA)and

V(z)≥V(ˆy)−1 ∀z∈B(˜y). (3.17)

Also, by virtue of Lemma 2, we consider a positive constantM such that, for all0≤t≤1and allz∈B(ˆy),

d+x(·;z) dt (t)

≤M; (3.18)

hence,kx(t;z)−zk ≤M tand so, by (3.17),

V(x(t;z))≥V(ˆy)−1≥λ¯−1 ∀z∈Bρ(ˆy)and∀t∈h0, ρ M

i. (3.19)

Let us fixγ≥1and define the set

G(ˆy) := [|V| ≤ |V(ˆy)|+γ]. (3.20) Claim: there existsT >0such that

eatV(x(t;y)) + Z t

0

W(x(τ;y))dτ ≤V(y) ∀y∈Bρ(ˆy)∩G(ˆy), ∀t∈[0, T]. (3.21) Using the (LW-)Lipschitz continuity ofW on the (bounded) set{x(t;y)|0 ≤t≤ 1, y ∈B(ˆy)},we write, for all y∈B(ˆy)∩G(ˆy)∩DomV and0≤t≤1,

2kx(t;y)−yk+

eat−1V(y)− Zt

0

W(x(τ;y))dτ

≤2M t+1−eat(|V(ˆy)|+γ) + (W(ˆy) +LW(M+ 2ρ))t.

Therefore, we can chooseT >0so that for ally∈B(ˆy)∩G(ˆy)we have that sup

t∈[0,T]

2kx(t;y)−yk+

e−at−1V(y)− Z t

0

W(x(τ;y))dτ < ρ

2.

We also observe that for any giveny∈Bρ(ˆy)∩G(ˆy)we have thatBρ(y)⊂B(ˆy)∩[V >λ].¯ Therefore, since Bρ(ˆy)∩G(ˆy)⊂B(ˆy)∩[V >λ]¯ ⊂Bρ¯(¯y)∩[V >¯λ]∩DomV,

the claim follows from Theorem 2.

To go further in the proof, we fix two parametersε, δ >0and we introduce the setEε,δ⊂R+defined as Eε,δ:=



λ∈R+

∃ρ1, ρ2∈(ρ2, ρ), ρ1< ρ2,∃ρλ∈(ρ2, ρ2), ∀y∈Bρλ(ˆy)∩G(ˆy), ∀t≤λ: eatVδ(x(t;y)) +R0tW(x(τ;y))dτ ≤V(y) +ε(ρ1ρ2)(ρ−ρ2)



, whereVδ:H→Ris the function given by

Vδ(y) := inf

zH{V(z) + 1

δky−zk2}.

Vδis dominated byV and is Lipschitz continuous on the bounded sets ofH.Then, we have that[0, T]⊂Eε,δ,that is, Eε,δ6=∅.Next, we shall show thatEε,δ=R+or, equivalentely, thatEε,δis closed and open with respect to the usual topolgy onR+.

Claim:Eε,δis closed.

Let a sequence(λn)nN⊂Eε,δbe such thatλn→λ˜and, by the definition ofEε,δ,take(ρ1,n)nN,(ρ2,n)nN, (ρ3,n)nN⊂(ρ2, ρ)be such that

ρ1,n< ρ2,n, ρ3,n∈(ρ 2, ρ2,n), together with the relation

eatVδ(x(t;y)) + Z t

0

W(x(τ;y))dτ≤V(y) +ε(ρ1,n−ρ

2)(ρ−ρ2,n) ∀y∈Bρ3,n(ˆy)∩G(ˆy),

(12)

valid for allt ≤ λn. Because all the sequences(ρ1,n)nN,(ρ2,n)nN,and(ρ3,n)nN are bounded, on relabeling if necessary, we may suppose thatρ1,n → ρ1 ∈[ρ2, ρ],ρ2,n → ρ2 ∈ [ρ2, ρ2],andρ3,n →ρˆ ∈[ρ2, ρ2].As well, it is enough to suppose thatλ > T˜ andλ > λ˜ nfor allnbecause, otherwise, eitherλ˜≤λnfor somenorλ˜≤T;hence in both cases we haveλ˜∈Eε,δ.

Ify∈Bρˆ(ˆy)∩G(ˆy)andt <λ,˜ for allnlarge enough we get thaty∈Bρn(ˆy)∩G(ˆy)andt < λnand, so, eatVδ(x(t;y)) +

Zt 0

W(x(τ;y))dτ ≤V(y) +ε(ρ1,n− ρ

2)(ρ−ρ2,n).

Asngoes to∞we obtain that

eatVδ(x(t;y)) + Z t

0

W(x(τ;y))dτ≤V(y) +ε(ρ1− ρ

2)(ρ−ρ2);

this inequality also holds for t = ˜λin view of the continuity ofVδ.It is also useful to notice here that for ally ∈ Bρ

2(ˆy)∩G(ˆy)andt≤λ˜

eatVδ(x(t;y)) + Z t

0

W(x(τ;y))dτ≤V(y). (3.22)

Now, checking the possible values ofρ1, ρ2,andρˆwe observe that only two cases may occur: the first corresponds to (ρ1ρ2)(ρ−ρ2) = 0and happens whenρ1= ρ2, ρ2=ρ,orρ2=ρ;this last equality implies thatρ2 ≤ρ1≤ρ2ρ2

and, so,(ρ1ρ2)(ρ−ρ2) = 0.While the second case corresponds to(ρ1ρ2)(ρ−ρ2) > 0 and happens when ρ1, ρ2∈(ρ2, ρ).

To begin with, we analyze the case(ρ1ρ2)(ρ−ρ2)> 0.This necessarily implies thatρ2 > ρ2 in view of the inequalityρ2 ≤ ρ1 ≤ρ2. We may suppose thatρˆ= ρ2 because otherwiseρˆ∈(ρ2, ρ2)trivially yieldsλ˜∈Eε,δ.So, in order to prove thatλ˜∈Eε,δ,we only need to find someβ >0such thatρ2+β∈(ρ2, ρ2)and for ally∈Bρ

2(ˆy)∩G(ˆy) andt≤˜λ,

eatVδ(x(t;y)) + Z t

0

W(x(τ;y))dτ≤V(y) +ε(ρ1− ρ

2)(ρ−ρ2). (3.23)

Proceeding by contradiction, we assume that for eachk≥1verifying ρ2 +1k ∈(ρ2, ρ2),there existyk∈Bρ

2+1k(ˆy)∩ G(ˆy)and0< tk≤λ˜such that

eatkVδ(x(tk;yk)) + Z tk

0

W(x(τ;yk))dτ > V(yk) +ε(ρ1−ρ

2)(ρ−ρ2). (3.24)

Because of (3.22) we must have(yk)k⊂Bρ

2+1k(ˆy)\Bρ

2(ˆy).W.l.o.g. we may suppose thattk →˜t≤λ.˜ For eachk, we denote byy˜k∈Bρ

2(ˆy)the orthogonal projection ofykontoBρ

2(ˆy).Thus, from one hand, we may also suppose that (˜yk)kweakly converges to somey˜∈Bρ

2(ˆy).Furthermore, from the inequalitykyk−y˜kk ≤ k1 we infer thatykalso weakly converges toy˜and, so, by the weak continuity ofV onBρ(ˆy),

V(˜y) = lim

k V(˜yk) = lim

k V(yk). (3.25)

Hence,

|V(˜y)|= lim

k |V(˜yk)|= lim

k |V(yk)| ≤ |V(ˆy)|+ 1.

In particular, (w.l.o.g.) this implies that (˜yk)k∪ {y˜} ⊂Bρ

2(ˆy)∩[|V| ≤ |V(ˆy)|+ 1] =Bρ

2(ˆy)∩G(ˆy).

On the other hand, the absolute continuity ofx(·; ˜yk)yields x(tk; ˜yk)−y˜k=

Z tk 0

˙

x(τ; ˜yk)dτ and, since thatx(˙ ·; ˜yk)∈L([0,λ];˜ H),the following holds:

kx(tk; ˜yk)−y˜kk ≤tk sup

τ∈[0,tk]kx(τ˙ ; ˜yk)k ≤tk sup

τ∈[0,tk]

eLfτ(f(˜yk)−A˜yk)

≤λe˜ Lfλ˜ sup

z∈Bρ 2y)

eLfτ(f(z)−Az)≤M˜λeLf˜λ.

(13)

Hence, w.l.o.g. we may suppose that the bounded sequence(x(tk; ˜yk))kN weakly converges inH.Furthermore, the inequality

kx(tk;yk)−x(tk; ˜yk)k ≤eLftkkyk−y˜kk ≤ eLft˜ k ,

infers that the both sequences(x(tk;yk))kNand(x(tk; ˜yk))kNweakly converge to the same point inH.

On another hand, since the sequences(x(tk;yk))kNand(x(tk; ˜yk))kNare bounded, there exits somel≥0such that for allt≤t˜

|W(x(t;yk))−W(x(t; ˜yk))|+|Vδ(x(t;yk))−Vδ(x(t; ˜yk))| ≤lkx(t;yk)−x(t; ˜yk)k

≤ leLf˜t k and, so, we deduce that (w.l.o.g.)

limk Vδ(x(tk;yk)) = lim

k Vδ(x(tk; ˜yk))and lim

k W(x(tk;yk)) = lim

k W(x(tk; ˜yk). (3.26) Using Lebesgue’s Theorem, this infers

limk

"

etVδ(x(tk; ˜yk)) + Z ˜t

0

W(x(τ; ˜yk))dτ

#

=etlim

k Vδ(x(tk; ˜yk)) + Z ˜t

0

limk W(x(τ; ˜yk))dτ

=etlim

k Vδ(x(tk;yk)) + Z ˜t

0

limk W(x(τ;yk))dτ

= lim

k

eatkVδ(x(tk;yk)) + Z tk

0

W(x(τ;yk))dτ

. Consequently, taking limits in (3.24), and using (3.25) we obtain

limk

"

eat˜Vδ(x(tk; ˜yk)) + Z ˜t

0

W(x(τ; ˜yk))dτ

#

(3.27)

≥lim

k V(yk) +ε(ρ1−ρ

2)(ρ−ρ2)

=V(˜y) +ε(ρ1−ρ

2)(ρ−ρ2)

=V(lim

kk) +ε(ρ1−ρ

2)(ρ−ρ2)

= lim

k V(˜yk) +ε(ρ1−ρ

2)(ρ−ρ2)).

In other words, forklarge enough we have etVδ(x(tk; ˜yk)) +

Z t˜ 0

W(x(τ; ˜yk))dτ ≥V(˜yk) +ε 2(ρ1−ρ

2)(ρ−ρ2)> V(˜yk), and a contradiction to (3.22) asy˜k∈Bρ

2(ˆy)∩G(ˆy),andt˜≤˜λ.Hence, we conclude that someρ˜λ∈(ρ2, ρ2)exists so that (3.23) holds for ally∈Bρ

2(ˆy)∩G(ˆy)andt≤˜λ.This fact shows that˜λ∈Eε,δ.

It remains to analyse the other case corresponding to(ρ1ρ2)(ρ−ρ2) = 0.If this happens, we chooseρ˜1,ρ˜2∈(ρ2, ρ) such thatρ˜1<ρ˜2and(ρ1ρ2)(ρ−ρ2)>0.Thus, following the same argument as in the first case, taking into account (3.22) we can find someβ >0,withρˆ+β ∈(ρ2, ρ2),so that (3.23) holds for ally∈Bρ

2(ˆy)∩G(ˆy).This shows that˜λ∈Eε,δand, hence, establishes the proof of the closedness ofEε,δ.

Claim:Eε,δis open.

Fixλ∈Eε,δ(it is sufficient to takeλ≥ν >0), and letρ1, ρ2∈(ρ2, ρ)andρˆ∈(ρ2, ρ2)be such thatρ1< ρ2and, for ally∈Bρˆ(ˆy)∩G(ˆy)andt≤λ,

eatVδ(x(t;y)) + Z t

0

W(x(τ;y))dτ≤V(y) +ε(ρ1− ρ

2)(ρ−ρ2). (3.28)

We letν >ˆ 0verify

ˆ

ν≤min{ν, λ}, ρ

2 <ρˆ−Mˆν < ρ2, and ρ

2 < eνρ1< ρ2.

(14)

So, from one hand, for all0≤α≤νˆandy∈Bρˆν(ˆy)∩G(ˆy)it holds, by Lemma 2,

kx(α;y)−yˆk ≤M α+ky−yˆk< M α+ ˆρ−Mνˆ≤ρ,ˆ (3.29) whereM ≥0is defined in (3.18). Hence, by the choice ofν(ν≤λ),from (3.28) we infer that

V(x(α;y))≤V(y)≤ |V(ˆy)|+ 1.

Thus, taking into account (3.19) we obtain that

x(α;y)∈Bρˆ(ˆy)∩G(ˆy).

Now fixy∈Bρˆν(ˆy)∩G(ˆy)andt∈[0, λ].From above we have that

x(ˆν, y)∈Bρˆ(ˆy)∩G(ˆy) (3.30) and, so, applying (3.28) we get that

eatVδ(x(t;x(ˆν, y))) + Z t

0

W(x(τ;x(ˆν, y)))dτ ≤V(x(ˆν, y)) +ε(ρ1−ρ

2)(ρ−ρ2).

Thus, using the semi-group property together with (3.28) and (3.30), we infer that ea(ˆν+t)Vδ(x(ˆν+t, y)) +

Z ν+tˆ

0

W(x(τ;y))dτ

=eνeatVδ(x(t;x(ˆν, y))) + Z t

0

W(x(τ;x(ˆν, y)))dτ+ Zνˆ

0

W(x(τ;y))dτ

≤eν

eatVδ(x(t;x(ˆν, y))) + Z t

0

W(x(τ;x(ˆν, y)))dτ

+ Zνˆ

0

W(x(τ;y))dτ

≤eνV(x(ˆν, y)) + Z νˆ

0

W(x(τ, y))dτ+εeν1−ρ

2)(ρ−ρ2).

At this step, for the choice that we made onνˆ(ˆν ≤ ν),the last inequality above reads, for ally ∈ Bρ−Mˆˆ ν(ˆy)and t∈[0,νˆ+λ],

ea(ˆν+t)Vδ(x(ˆν+t, y)) + Z ν+tˆ

0

W(x(τ;y))dτ ≤V(y) +εeν1−ρ

2)(ρ−ρ2)

≤V(y) +ε(eνρ1−ρ

2)(ρ−ρ2).

Consequently, since thatρˆ−Mˆν∈(ρ2, ρ2)andeνρ1∈(ρ2, ρ2)it follows that[0, λ+ ˆν]⊂Eε,δand, so, the openness ofEε,δfollows.

In order to conclude the proof, lety∈Bρ

2(ˆy)∩G(ˆy)be given. Then, for everyt≥0we have thatt∈ ∩ε>0Eε,δ; that is for allε >0it holds

eatVδ(x(t;y)) + Z t

0

W(x(τ;y))dτ≤V(y) +ε(ρ−ρ 2)(ρ−ρ

2) =V(y) +ερ2 4. Hence, lettingε→0it follows that

eatVδ(x(t;y)) + Z t

0

W(x(τ;y))dτ≤V(y), which asδ→0yields (using the fact thatlimδ→0Vδ(x(t;y)) =V(x(t;y))

eatV(x(t;y)) + Z t

0

W(x(τ;y))dτ ≤V(y).

Now, ifz¯∈ Bρ(ˆy)∩DomV,then similarly as above, we can findρz¯ > 0such that for everyz ∈ Bρ¯z

2 (¯z)∩G(¯z) (whereG(¯z)is defined as in (3.20)) we have that

eatV(x(t;z)) + Z t

0

W(x(τ;z))dτ≤V(z) for allt≥0.

Thus, the main conclusion of the current theorem follows since that the last inequality obviously holds whenz /¯∈DomV.

(15)

Remark 2 The conclusion of Theorem 3.1 also holds if, instead of V being weak continuous onBρ(¯y),we assume that eitherHis finite-dimensional orV is convex.

Proof The only difference with the proof of Theorem 3.1 arises in showing (3.23).

(a) Assume thatH is finite-dimensional. Let us show that (3.23) holds. Assuming the contrary, we find bounded sequencseyk∈Bρ

2+1k(¯y)∩G(¯y)and0< tk≤ ˜λsuch that (3.24) holds. W.l.o.g. we may suppose thattk→t˜≤˜λ andyk⇀y˜∈Bρ

2(¯y).Furthermore, we have that V(˜y)≤lim inf

k V(yk)≤ |V(¯y)|+ 1,

while (3.19) guarantees thatV(˜y)≥V(¯y)−1.Hence, we also have thaty˜∈[|V| ≤ |V(¯y)|+ 1].Now, recalling that x(tk;yk)converges tox(˜t,y)˜ in this case, it follows that

ea˜tVδ(x(˜t,y)) +˜ Z ˜t

0

W(x(τ; ˜y))dτ

=etVδ(lim

k x(tk, yk)) + Z ˜t

0

W(lim

k x(τ;yk))dτ

= lim

k

"

etVδ(x(tk;yk)) + Z ˜t

0

W(x(τ;yk))dτ

#

≥lim inf

k V(yk) +ε(ρ1−ρ

2)(ρ−ρ2)

≥V(˜y) +ε(ρ1−ρ

2)(ρ−ρ2)> V(˜y), which contradicts (3.22).

(b) Assume thatV is convex. We consider again the sequences of the proof of Theorem 3.1,(yk)k∈N⊂Bρ

2+1k(¯y)∩ G(¯y)\Bρ

2(¯y)and(˜yk)k∈N⊂Bρ

2(¯y),which both converge toy˜∈Bρ

2(¯y)∩G(¯z).Since that eachy˜k∈[yk,y],¯ with k≥1,we findβk∈[0,1]such thaty˜k:=βkyk+ (1−βk)¯y, this yields

V(˜yk)≤βkV(yk) + (1−βk)V(¯y).

We notice that1≥βkkρ+2 since by construction,y˜kis on the boundary ofBρ

2(¯y)andyk ∈Bρ

2+1k(¯y).Thus, we may suppose thatβk→1.Consequently, taking limits in the inequality above,

lim inf

k V(˜yk)≤lim

k βklim inf

k V(yk) = lim inf

k V(yk).

Hence, as in (3.27), using (3.26) we obtain that limk

"

etVδ(x(tk; ˜yk)) + Z ˜t

0

W(x(τ; ˜yk))dτ

#

=etlim

k Vδ(x(tk; ˜yk)) + Zt˜

0

limk W(x(τ; ˜yk))dτ

=etlim

k Vδ(x(tk;yk)) + Zt˜

0

limk W(x(τ;yk))dτ

= lim

k

eatkVδ(x(tk;yk)) + Z tk

0

W(x(τ;yk))dτ

≥lim inf

k V(yk) +ε(ρ1−ρ

2)(ρ−ρ2)

≥lim inf

k V(˜yk) +ε(ρ1−ρ

2)(ρ−ρ2),

which contradicts (3.22). △

Corollary 1 Assume thatInt (co{DomA})6=∅.LetV ∈ F(H)be convex, and letW ∈ F(H;R+)anda∈R+be given. Fixy¯∈Int(DomA)∩DomV,and letρ >0be such thatB(¯y)⊂Int(DomA).For ally∈B(¯y)∩DomV we assume that

sup

ξ∈∂PV(y)

υinfAyhξ, f(y)−υi+aV(y) +W(y)≤0.

Then, for ally∈Bρ(¯y)we have that

eatV(x(t;y)) + Z t

0

W(x(τ;y))dτ ≤V(y) for allt≥0.

(16)

Proof According to Theorem 3.1 and Remark 2, it suffices to show that the current assumptionimplies that, for every giveny∈B(¯y)∩DomV andξ∈∂V(y) = NDomV(y)(if any), there existsυ∈Aysuch that

hξ, f(y)−υi ≤0. (3.31)

To prove this fact, by the lsc ofV we letε >0be such that

Bε(y)⊂Int(cl(DomA)), V(Bε(y))≥V(y)−1.

Pickyε ∈ ∂εV(y);this last set is not empty since thatV ∈ F(H)is a convex function. Then, from the relationship NDomV(y) = (∂εV(y))(e.g. ), for everyk∈Nwe have that

yε+kξ∈∂εV(y).

According to the Brøndsted-Rockafellar Theorem, there areyk∈Bε(y)anduk∈Bε(θ)such that yε+kξ∈∂εV(yk) +uk;

that is, in particular,yk∈DomV.Consequently, by the current assumption we get that

khξ, f(y)−πAyk(f(yk))i ≤ huk−yε, f(yk)−πAyk(f(yk))i −aV(yk)−W(yk)

≤ huk−yε, f(yk)−πAyk(f(yk))i −aV(y) +a +khξ, f(y)−f(yk)i

≤ huk−yε, f(yk)−πAyk(f(yk))i −aV(y) +a+kLf√ εkξk. Since thathuk−yε, f(yk)−πAyk(f(yk))iis bounded independently ofk, forkε≥1big enough we get that

hξ, f(y)−πAy(f(ykε))i ≤√

ε+Lf√ εkξk.

Moreover, asykε ∈ Bε(y)andζε := πAy(f(ykε))(∈ Aykε) is bounded independently ofkε,we may suppose asε → 0that(ζε)weakly converges to someυ ∈ Ay.Thus, taking limits in the last inequality above we get that

hξ, f(y)−υi ≤0;that is (3.31) follows. △

Corollary 2 Assume thatdimH <∞.LetV ∈ F(H),W ∈ F(H;R+),anda∈R+be given. Fixy¯∈Int(DomA), and letρ >0be such thatB(¯y)⊂Int(DomA).For ally∈B(¯y)∩DomV we assume that

sup

ξPV(y)

υ∈Ayinf hξ, f(y)−υi+aV(y) +W(y)≤0.

Then, for ally∈Bρ(¯y)we have that

eatV(x(t;y)) + Z t

0

W(x(τ;y))dτ ≤V(y) for allt≥0.

Proof As in the proof of Corollary 1, giveny ∈ B(¯y)∩DomV andξ ∈ ∂V(y)(if any),we only to find some υ∈Aysuch that

hξ, f(y)−υi ≤0.

Fixε >0.By definition, we letξk∈∂PV(yk)andαk↓0such thatyk→y,V(yk)→V(y),andαkξk⇀ ξ.Then, by the current assumption, for eachkthere existsyk∈Ayksuch that

ξk, f(yk)−yk

+aV(yk) +W(yk)≤ε.

BecausedimH <∞andyk, ykare bounded, we may suppose thatykconverges to someυ∈Ay.Thus, multiplying the equation above byαkand next passing to the limit asε → 0and finally invoking the lsc ofV and the Lipschitz continuity off, we obtain that

hξ, f(y)−υi ≤lim

k

αkξk, f(yk)−yk

+alim inf

k αkV(yk)≤lim

k αkε= 0.

The conclusion follows. △

Références

Documents relatifs

We give different conditions for the invariance of closed sets with respect to differential inclusions governed by a maximal monotone operator defined on Hilbert

Nonsmooth Lyapunov pairs for differential inclusions governed by operators with nonempty interior domain.. Samir Adly, Abderrahim Hantoute,

The aim of this paper is to provide explicit primal and dual criteria for weakly lower semi- continuous (lsc, for short) Lyapunov functions, or, more generally, for Lyapunov

Namely, we prove existence of viable solutions of problem (1.1) in the case where the set- valued map F (· , ·) is upper semicontinuous compact valued and contained in the

1. Preliminaries and Definitions. The sets of admissible velocities. Discretization and convergence of approximate solutions. Uniform estimates on the computed velocities. Extraction

2.2 The various generalized subdifferentials of a discontinuous function We recall here various notions of generalized subdifferentials of a discontinuous (in fact

Each of these enlargements is nonempty for the Rainwater function in Example 1.1 and satisfies the fundamen- tal properties of the Moreau–Rockafellar subdifferential: convexity, weak

Invariant sets and Lyapunov pairs or functions associated to general differential inclusions/equations of the form of (1.1) have been the subject of extensive research during the