APM3344
APM3344
APM3344
net/publication/278675542
CITATIONS READS
34 339
2 authors:
All content following this page was uploaded by Ke Tran Dinh on 18 June 2015.
1. Introduction
We consider the following differential variational inequality (DVI)
x0 (t) = Ax(t) + h(x(t)) + B(x(t), xt )u(t), t ∈ J = [0, T ], (1.1)
hv − u(t), F (x(t)) + G(u(t))i ≥ 0, ∀v ∈ K, for a.e t ∈ J, (1.2)
x(s) = ϕ(s), s ∈ [−τ, 0], (1.3)
n m
where x(t) ∈ R , u(t) ∈ K with K being a closed convex subset in R , xt denotes
the history of the state function up to time t; A, B, F, G and h are given maps
which will be specified in the next section.
The notion of differential variational inequality can be traced back to Aubin
and Cellina [1] in 1984. In a later work of Avgerinous and Papageorgiou [3], this
concept was revisited and expanded. However, it is seen that DVIs were first
systematically studied by Pang and Stewart [14]. As mentioned in this paper,
DVIs are useful for representing models involving both dynamics and constraints
in the form of inequalities which arise in many problems in reality, for example,
mechanical impact problems, electrical circuits with ideal diodes, Coulomb friction
problems for contacting bodies, economical dynamics and related problems such as
dynamic traffic networks. In case K = Rm , system (1.1)-(1.3) becomes a differential
algebraic equation with the unknown y = (x, u), while if K is a cone, it is a
differential complementarity problem. Some existence results for DVIs can be found
in [8, 9, 12] and the references therein.
In the theory of differential equations, problem (1.1)-(1.3) is called a differential
system with unilateral constraints. As a matter of fact, it can be seen as a control
problem subject to constraints. In this work, we will look for admissible conditions
to gain the existence and behavior of solutions to this problem. Specifically, after
giving a short proof for the solvability of (1.1)-(1.3) on bounded intervals, we show
that this system has a compact set of exponentially decay solutions. In addition, the
multi-valued semiflow generated by (1.1)-(1.3) admits a compact global attractor
in C([−τ, 0]; Rn ). Up to our knowledge, no attempt has been made to study the
1
2 N.T.V. ANH, T.D. KE
2. Preliminaries
2.1. Measure of noncompactness. Let E be a Banach space. Denote
P(E) = {B ⊂ E : B 6= ∅},
B(E) = {B ∈ P(E) : B is bounded}.
We recall the definition of measure of noncompactness introduced in [2].
Definition 2.1. A function β : B(E) → R+ is called a measure of noncompactness
(MNC) on E if
β(co Ω) = β(Ω) for every Ω ∈ B(E),
where co Ω is the closure of the convex hull of Ω. An MNC β is called
(i) monotone if Ω0 , Ω1 ∈ B(E), Ω0 ⊂ Ω1 implies β(Ω0 ) ⊂ β(Ω1 );
(ii) nonsingular if β({a} ∪ Ω) = β(Ω) for any a ∈ E, Ω ∈ B(E);
(iii) invariant with respect to union with compact set if β(K ∪ Ω) = β(Ω) for
every relatively compact set K ⊂ E and Ω ∈ B(E);
(iv) algebraically semi-additive if β(Ω0 + Ω1 ) ≤ β(Ω0 ) + β(Ω1 ) for any Ω0 , Ω1 ∈
B(E);
(v) regular if β(Ω) = 0 is equivalent to the relative compactness of Ω.
An important example of MNC is the Hausdorff MNC χ(·), which is defined as
follows: for Ω ∈ B(E),
χ(Ω) = inf{ > 0 : Ω has a finite − net}.
In particular, it is known that the Hausdorff MNC on C([0, T ]; Rn ), the space of
continuous functions on [0, T ] taking values in Rn , is given by (see [2])
1
χT (D) = lim sup max kx(t) − x(s)k. (2.1)
2 δ→0 x∈D t,s∈[0,T ],|t−s|<δ
The measure χT (D) of a subset D can be seen as the modulus of equicontinuity of
a subset in C([0, T ]; Rn ).
Consider the space BC(0, ∞; Rn ) of bounded continuous functions on [0, ∞) tak-
ing value on Rn . Denote by πT the restriction operator on the space BC(0, ∞; Rn ),
that is πT (x) is the restriction of x on [0, T ]. Then
χ∞ (D) = sup χT (πT (D)), D ⊂ BC(0, ∞; Rn ), (2.2)
T >0
is an MNC. One can check that χ∞ satisfies all properties given in Definition
2.1, but regularity. Indeed, we will testify this claim by choosing the sequence
ASYMPTOTIC BEHAVIOR FOR A CLASS OF DVIS 3
Then it is obvious that {πT (fk )} is compact (converging to 0 in C([0, T ]; R)) for
any T > 0. However, one sees that
sup |fk (t) − fl (t)| = 1 for k 6= l,
t≥0
and then {fk } is not a Cauchy sequence in BC(0, ∞; R). This fact tells us that
χT (πT ({fk }))) = 0 for any T > 0 and then χ∞ ({fk }) = 0, but {fk } is not relatively
compact.
We now construct a regular MNC on BC(0, ∞; Rn ). Recall the following MNCs
on BC(0, ∞; Rn ) (see [4]):
dT (D) = sup sup kx(t)k,
x∈D t≥T
d∞ (D) = lim dT (D).
T →∞
Define
χ∗ (D) = χ∞ (D) + d∞ (D). (2.3)
∗ n
By a simple check, we get that χ is an MNC on BC(0, ∞; R ).
Lemma 2.1. The MNC χ∗ defined by (2.3) is regular.
Proof. Let D ⊂ BC(0, ∞; Rn ) be a bounded set such that χ∗ (D) = 0. We show that
D is relatively compact. Let P BC(0, ∞; Rn ) be the space of piecewise continuous
and bounded functions on R+ , taking values in Rn . This is a Banach space with
the norm
kxkP BC = sup kx(t)k,
t≥0
n
and contains BC(0, ∞; R ) as a closed subspace.
For > 0, since d∞ (D) = 0 one can choose T > 0 such that supt≥T kx(t)k <
, ∀x ∈ D. This means that
2
kx − πT (x)kP BC < , ∀x ∈ D,
2
here πT (x) agrees with a function in P BC(0, ∞; Rn ) by the following manner
(
x(t), t ∈ [0, T ],
πT (x) =
0, t > T.
where xi ∈ C([0, T ]; Rn ), i = 1, ..., N , the notation BT (x, r) stands for the ball in
C([0, T ]; Rn ) centered at x with radius r. Put
(
xi (t), t ∈ [0, T ],
x̂i (t) =
0, t > T,
then {x̂i }N n
i=1 belong to P BC(0, ∞; R ). We assert that
N
[
D⊂ B∞ (x̂i , ),
i=1
here B∞ (x, r) is the ball in P BC(0, ∞; Rn ) with center x and radius r. Indeed, let
x ∈ D then by (2.4), there is a number k ∈ {1, ..., N } such that
kπT (x) − xk kC < ,
2
here k · kC is the norm in C([0, T ]; Rn ). This implies
kπT (x) − x̂k kP BC < .
2
Then
kx − x̂k kP BC ≤ kx − πT (x)kP BC + kπT (x) − x̂k kP BC
≤ + = .
2 2
N
S
Thus x ∈ B∞ (x̂k , ). We have D ⊂ B∞ (x̂i , ), and hence D is relatively
i=1
compact in P BC(0, ∞; Rn ). In order to show that D is also relatively compact
in BC(0, ∞; Rn ), we observe that if {xn } ⊂ D then one can find a function
x ∈ P BC(0, ∞; Rn ) such that
lim kxn − xkP BC = lim sup kxn (t) − x(t)k = 0,
n→∞ n→∞ t≥0
We also make use of some notions and facts of set-valued analysis. Let Y be a
metric space.
Definition 2.2. A multivalued map (multimap) F : Y → P(E) is said to be
(i) upper semicontinuous (u.s.c) if F −1 (V ) = {y ∈ V : F(y) ∩ V 6= ∅} is a
closed subset of Y for every closed set V ⊂ E;
(ii) weakly upper semicontinuous (weakly u.s.c) if F −1 (V ) is closed subset of Y
for all weakly closed set V ⊂ E;
(iii) closed if its graph ΓF = {(y, z) : z ∈ F(y)} is a closed subset of Y × E;
(iv) compact if F(B) is relatively compact in E for any bounded set B ⊂ Y ;
(v) quasicompact if its restriction to any compact subset A ⊂ Y is compact.
Lemma 2.2. [11, Theorem 1.1.12] Let G : Y → P(E) be a closed quasicompact
multimap with compact values. Then G is u.s.c.
ASYMPTOTIC BEHAVIOR FOR A CLASS OF DVIS 5
Lemma 3.1. Let (H4) hold. Then for each z ∈ Rm , the solution set SOL(K, z +
G(·)) is non-empty, convex and closed. Moreover, there exists ηG > 0 such that
kvk ≤ ηG (1 + kzk), ∀v ∈ SOL(K, z + G(·)). (3.2)
In order to solve (1.1) − (1.3), we convert it to a differential inclusion. Let
U (z) = SOL(K, z + G(·)), z ∈ Rm .
Then we get that U : Rm → P(Rm ) has compact convex values, thanks to Lemma
3.1. Moreover, it is easy to verify that U is a closed map. By (3.2), we see that U
is locally bounded, then it is u.s.c.
Now we define Φ : Rn × Cτ → P(Rn ) as follows:
Φ(v, w) = {B(v, w)y + h(v) : y ∈ U (F (v))}. (3.3)
n
Since B(v, w) is a linear operator for each v ∈ R , w ∈ Cτ and U has compact
convex values, Φ also has compact convex values. Furthermore, thanks to the
continuity of B, F , h and the fact that U is u.s.c, the composition multimap Φ is
u.s.c as well.
Due to the above setting, DVI (1.1)-(1.3) is converted to the following differential
inclusion
x0 (t) ∈ Ax(t) + Φ(x(t), xt ), t ∈ J, (3.4)
x(t) = ϕ(t), t ∈ [−τ, 0]. (3.5)
Denote
PΦ (x) = {f ∈ L1 (J; Rn ) : f (t) ∈ Φ(x(t), xt )}, for x ∈ C. (3.6)
Then we deduce that a solution x ∈ C of DVI (1.1)-(1.3) is given by the following
formula
Zt
x(t) = etA ϕ(0) + e(t−s)A f (s)ds, f ∈ PΦ (x), t ∈ J, (3.7)
0
x(t) = ϕ(t), t ∈ [−τ, 0]. (3.8)
For y ∈ CT and ϕ ∈ Cτ , we define the function y[ϕ] ∈ C as follows
(
y(t), if t ∈ [0, T ],
y[ϕ](t) =
ϕ(t), if t ∈ [−τ, 0].
By defining the following operator
W : L1 (J; Rn ) → CT
Zt
W(f )(t) = e(t−s)A f (s)ds, (3.9)
0
Proof. Based on the hypotheses and the result of Lemma 3.1, we get
kΦ(v, w)k := sup{kzk : z ∈ Φ(v, w)}
≤ kB(v, w)kηG (1 + kF (v)k) + kh(v)k
≤ ηG (1 + ηF )[ηB (kvk + kwkCτ ) + ζB ] + ηh kvk + ζh . (3.10)
Since Φ is u.s.c with compact convex values, the multimap Λ(t) = Φ(x(t), xt ) is
strongly measurable due to [11, Proposition 1.3.1]. Therefore it has a Castaing
representation (see [11, Definition 1.3.3]) and hence PΦ (x) 6= ∅ for x ∈ C.
We prove the second assertion by using Lemma 2.3. Let {xk } ⊂ C such that
xk → x∗ , fk ∈ PΦ (xk ). Then {fk (t)} ⊂ C(t) := Φ({xk (t), (xk )t }), and C(t) is a
compact set for each t ∈ J. Furthermore, by (3.10), {fk } is integrably bounded
(bounded by an integrable function). Thus {fk } is weakly relatively compact in
L1 (J; Rn ) (see [7, Corollary 2.6]). Let fk * f ∗ in L1 (J; Rn ), then by Mazur’s
lemma (see, e.g [5]) there are f¯k ∈ co{fi : i ≥ k} such that f¯k → f ∗ in L1 (J; Rn )
and then f¯k (t) → f ∗ (t) for a.e. t ∈ J, up to a subsequence. Observe that in our
case, the upper semicontinuity of Φ implies that for a given > 0
Φ(xk (t), (xk )t ) ⊂ Φ(x∗ (t), x∗t ) + B for all large k,
here B is the ball in Rn centered at origin with radius . So
fk (t) ∈ Φ(x∗ (t), x∗t ) + B for a.e. t ∈ J,
and
f¯k (t) ∈ Φ(x∗ (t), x∗t ) + B for a.e. t ∈ J,
thanks to the convexity of Φ(x∗ (t), x∗t ) + B . The last inclusion deduces f ∗ (t) ∈
Φ(x∗ (t), x∗t ) + B for a.e. t ∈ J. Since is arbitrary, one obtains f ∗ ∈ PΦ (x∗ ). The
lemma is proved.
Lemma 3.3. The operator W defined by (3.9) is compact.
Proof. We show that W(Ω) is relatively compact in CT for any bounded set Ω ⊂
L1 (J; Rn ). Obviously, we get that W(Ω)(t) is bounded in Rn . In addition, W(Ω) is
equicontinuous thanks to the fact that S(t) = etA is a norm-continuous semigroup.
So we get the conclusion by using Arzelà-Ascoli theorem.
Lemma 3.4. Let (H1)-(H5) hold. Then the solution operator F is compact and
has a closed graph.
Proof. Since W is compact, it is easy to check that F(B) is relatively compact for
any bounded set B ⊂ CT . So F is a compact multimap.
Now let {xk } ⊂ CT , xk → x∗ , yk ∈ F(xk [ϕ]) and yk → y ∗ . We will verify that
y ∈ F(x∗ ). By the formulation of F, one can take fk ∈ PΦ (xk [ϕ]) such that
∗
Theorem 3.5. Assume (H1)-(H5). Then problem (3.4)-(3.5) has at least one so-
lution on [−τ, T ]. Moreover, the solution set is compact.
Proof. We first prove using Theorem 2.4 that Fix(F) 6= ∅. According to Lemma
3.4, it suffices to show that there exists a closed convex set M0 ⊂ CT satisfying that
F(M0 ) ⊂ M0 . Let y ∈ F(x). Then it follows from the definition of the solution
operator and the estimate (3.10) that there exists f ∈ PΦ (x[ϕ]) verifying
Zt
tA
ky(t)k = ke ϕ(0) + e(t−s)A f (s)dsk
0
Zt
≤ M kϕ(0)k + ke(t−s)A kkf (s)kds
0
Zt
≤ M kϕ(0)k + M [(η + ηh )kx(s)k + ηkx[ϕ]s kCτ + ηG (1 + ηF )ζB + ζh ] ds, ∀t ∈ J,
0
one has
Z t
ky(t)k ≤ M1 + M ((η + ηh )kx(s)k + η sup kx(ρ)k)ds
0 ρ∈[0,s]
Z t
≤ M1 + M (2η + ηh ) sup kx(ρ)kds,
0 ρ∈[0,s]
Zt
sup ky(ρ)k ≤ M1 + M (2η + ηh ) sup kx(ρ)k)ds. (3.12)
ρ∈[0,t] ρ∈[0,s]
0
Denote
M0 = {x ∈ CT : sup kx(s)k ≤ ψ(t), t ∈ [0, T ]},
s∈[0,t]
4. Decay solutions
In this section, we consider the solution operator F on BC(0, ∞; Rn ). For a
positive number γ and ϕ ∈ Cτ , denote
Bϕγ (R) = {x ∈ C([0, ∞); Rn ) : x(0) = ϕ(0), eγt kx(t)k ≤ R for all t ≥ 0}.
Lemma 4.1. Under hypotheses (H1*), (H2*), (H3)-(H4) and (H5*), F(Bϕγ (R)) ⊂
Bϕγ (R) for some R > 0, provided that
Zt
tA
yn (t) = e ϕ(0) + e(t−s)A fn (s)ds, ∀t ≥ 0,
0
Now one observes that eγt kxn (t)k ≤ n for all t ≥ 0. Then for all t ≥ τ , we get
On the other hand, for t ∈ [0, τ ] one has eγt kxn [ϕ]t kCτ ≤ eγτ kϕkCτ . Hence
This implies dT (D) ≤ Ce−γT , so d∞ (D) = lim dT (D) = 0. Combining this with
T →∞
(4.5) yields
χ∗ (F(D)) = 0.
Since the MNC χ∗ is regular, we conclude that F(D) is relatively compact.
To prove that F is u.s.c, it suffices to show that F has closed graph. But
this is done by the same arguments as in the proof of Lemma 3.4. The proof is
complete.
Proof. Since G(t, ·) is a compact multimap with compact values, it suffices to prove
that G(t, ·) is closed for each t ≥ 0, thanks to Lemma 2.2. Let ϕn → ϕ∗ in Cτ and
zn ∈ G(t, ϕn ) such that zn → z ∗ . We show that z ∗ ∈ G(t, ϕ∗ ), i.e. z ∗ = x∗ [ϕ∗ ]t
for an x∗ ∈ Σ(ϕ∗ ). Taking xn ∈ Σ(ϕn ) such that zn = xn [ϕn ]t , one can find
fn ∈ PΦ (xn [ϕn ]) verifying
xn = e(·)A ϕn (0) + W(fn ). (5.2)
n
Since {ϕn } is bounded in Cτ , {xn } is a bounded sequence in C([0, T ]; R ) for
any T > 0. Thus {fn } is integrably bounded in L1 (0, T ; Rn ). The compactness
of W implies that {W(fn )} is relatively compact in C([0, T ]; Rn ). In addition,
{e(·)A ϕn (0)} is a convergent sequence in C([0, T ]; Rn ), then taking into account
(5.2) we see that {xn } has a convergent subsequence (still denoted by {xn }). Let
x∗ = lim xn in C([0, T ]; Rn ). Then xn [ϕn ] → x∗ [ϕ∗ ] in C([−τ, T ]; Rn ). Since PΦ
n→∞
is weakly u.s.c, we have fn * f ∗ ∈ PΦ (x∗ [ϕ∗ ]) up to a subsequence, thanks to
Lemma 2.3. Therefore one can pass (5.2) into limits to get
x∗ = e(·)A ϕ∗ (0) + W(f ∗ ),
for f ∗ ∈ PΦ (x∗ [ϕ∗ ]). That is, x∗ [ϕ∗ ] is a solution of (1.1)-(1.3) and then x∗ [ϕ∗ ]t ∈
G(t, ϕ∗ ). Obviously, we have zn = xn [ϕn ]t → z ∗ = x∗ [ϕ∗ ]t and z ∗ ∈ G(t, ϕ∗ ). The
proof is complete.
In order to apply Theorem 2.5, it remains to show that G has an absorbing set
in Cτ . We make use of the following result (see [10]).
Proposition 5.3 (Halanay’s inequality). Let the function f : [t0 − τ, T ) → R+ , 0 ≤
t0 < T < +∞ satisfy the functional differential inequality
f 0 (t) ≤ −γf (t) + ν sup f (s),
s∈[t−τ,t]
Let
t
e−at kϕ(0)k + R e−a(t−s) [(η + η )kx(s)k + dkx k ]ds, t ≥ 0,
h s Cτ
y(t) = 0
kx(t)k, t ∈ [−τ, 0].
This inequality indicates that kxt kCτ tends to zero as t → ∞, so one can find t1 > 0
such that kxt1 kCτ < R. We get a contradiction.
We have just proved that if kϕkCτ ≤ C, then there exists t0 > 0 such that
kxt0 kCτ ≤ R. In the sequel, we assert that kut kCτ ≤ R, ∀t ≥ t0 . Assume to the
contrary that there exists t1 ≥ t0 satisfying
kxt1 kCτ ≤ R but kxt kCτ > R for all t ∈ (t1 , t1 + θ),
where θ > 0. Regarding the solution x[ϕ] on [t1 , t1 + θ), we have
Zt
(t−t1 )A
x(t) = e x(t1 ) + e(t−s)A f (s)ds, f ∈ PΦ (x[ϕ]).
t1
Then
Zt
−a(t−t1 )
kx(t)k ≤ e kϕ(0)k + e−a(t−s) [(η + ηh )kx(s)k + dkxs kCτ ]ds, t ∈ [t1 , t1 + θ).
t1
Using the same arguments as above, we see that for all t ∈ [t1 , t1 + θ),
kx(t)k ≤ kxt1 kCτ e−`(t−t1 ) ≤ kxt1 kCτ ≤ R.
ASYMPTOTIC BEHAVIOR FOR A CLASS OF DVIS 15
≤ sup kx(r)k
r∈[t1 −τ,t]
This is a contradiction. In summary, one can take a ball centered at origin with
radius R as an absorbing set for the m-semiflow G, where R is chosen such that
ηG (1 + ηF )ζB + ζh
R> .
a − (2η + ηh )
Theorem 5.5. Let (H1*), (H2)-(H5) hold. Then the m-semiflow G generated by
(1.1)-(1.3) admits a compact global attractor provided that
2ηB ηG (1 + ηF ) + ηh < a.
Proof. The conclusion follows from Corollary 5.1, Lemma 5.2 and 5.4.
Acknowledgment. The authors are grateful to anonymous referee for the careful
reading and constructive comments and suggestions, that help to improve the man-
uscript. The work was supported by Vietnam Ministry of Education and Training,
under project B2014-17-70.
References
[1] J.-P. Aubin, A. Cellina, Differential inclusions. Set-valued maps and viability theory,
Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical
Sciences], 264. Springer-Verlag, Berlin, 1984.
[2] R.R. Akhmerov, M.I. Kamenskii, A.S. Potapov, A.E. Rodkina, B.N. Sadovskii, Measures of
noncompactness and condensing operators, Birkhäuser, Boston-Basel-Berlin, 1992.
[3] E.P. Avgerinos, N.S. Papageorgiou, Differential variational inequalities in RN , Indian J.
Pure Appl. Math. 28 (1997), no.9, 1267-1287.
[4] J. Banas and L. Olszowy, On a class of measures of noncompactness in Banach algebras and
their application to nonlinear integral equations, J. Anal. Appl. 28 (2009), 475-498.
[5] D. Bothe, Multivalued perturbations of m-accretive differential inclusions, Israel J. Math. 108
(1998), 109-138.
[6] T. Caraballo, P.E. Kloeden, Non-autonomous attractors for integro-differential evolution
equations, Discrete Contin. Dyn. Syst. Ser. S 2 (2009), 17-36.
[7] J. Diestel, W. M. Ruess, W Schachermayer, Weak compactness in Ll (µ, X), Proc. Amer.
Math. Soc. 118 (1993), 447-453.
[8] J. Gwinner, On differential variational inequalities and projected dynamical systems - equiv-
alence and a stability result, Discrete Contin. Dynam. Syst. 2007, Dynamical Systems and
Differential Equations. Proceedings of the 6th AIMS International Conference, suppl., 467-
476.
[9] J. Gwinner, A note on linear differential variational inequalities in Hilbert Spaces, in: IFIP
Advances in Information and Communication Technology, Syst. Model. Optim. Vol. 391,
2013, 85-91.
[10] A. Hanalay, Differential equations, stability, oscillations, time lags, Academic Press, New
York, London 1996.
[11] M. Kamenskii, V. Obukhovskii, P. Zecca, Condensing multivalued maps and semilinear dif-
ferential inclusions in Banach spaces, in: de Gruyter Series in Nonlinear Analysis and Ap-
plications, vol. 7, Walter de Gruyter, Berlin, New York, 2001.
16 N.T.V. ANH, T.D. KE
[12] Z. Liu, N.V. Loi, V. Obukhovskii, Existence and global bifurcation of periodic solutions to a
class of differential variational inequalities, Int. J. Bifur. Chaos 23 (2013) No.7, 1350125.
[13] V.S. Melnik, J. Valero, On attractors of multivalued semi-flows and differential inclusions,
Set-Valued Anal. 6 (1998), 83-111.
[14] J.-S. Pang, D.E.Stewart, Differential variational inequalities, Math. Program. Ser. A 113
(2008), 345-424.
[15] I. Ekeland, R. Temam, Convex analysis and variational problems, Society for Industrial and
Applied Mathematics (SIAM), Philadenphia, PA, 1999.