2001-3-04 Roubicek Schmidt
2001-3-04 Roubicek Schmidt
1
Mathematical Institute, Charles University, Sokolovskci 83
CZ-186 75 Praha 8, Czech Republic
2
Institute of Information Theory and Automation , Academy of Sciences
Pod vodarenskou vezf 4, CZ-182 08 Praha 8, Czech Republic
3 Institut
fiir Mathematik und Informatik
Ernst-Moritz-Arndt-U niversi tat G reifswald
Friedrich-Ludwig-Jahn-Strai3e 15a, D-17 487 Greifswald, Germany
nonexistence of optimal solutions for a control pro ce~Js with Volterra equation
J;
of the type y(t) = Yo(t) + j(t,r,y(r),u(r))dr , u(t) E S, with the orientor
field f(t, r, r, S) always convex and compact, see (5.2) below, though for some
rather special problems of this kind the existence can be proved, see Schmeling
(1981) or Roubfcek and Schmidt (1997).
Nevertheless, the Filippov- Roxin t heory was extended for the nonlinear in-
tegral equations of the Fredholm (sometimes also called Urysohn) type that are
of Hammerstein type with respect to control by Balder (1993), Bennati (1979),
Cowles (1973) and, for less general equations, also by Zolezzi (1972). For the
Volterra-type equations, even some more references exist; see Remark 2 below.
Besides, the Fredholm case was also addressed by the authors, Roubfcek and
Schmidt (1997), but in completely different situations, via the Bauer's extremal
principle.
In this paper, we want to present the Filippov- Roxin theory for this special
class of Fredholm equations that are of Hammerstein typ e with respect to con-
trol, similarly as already done in Balder (1993), Bcnnati (1979) , but by using
an auxiliary relaxation by Ro ubfcek (1998) similarly as proposed in Mniio~ and
Pedregal (2001) for problems governed by ordinary differential equations and
in Roubfcek (1999) for general situations. This provides a deeper insight and
enables us to refine the existence results to cover also problems with nonconvex
orientor fi eld like it was done in Gabasov and Mordukhovich (1974) , Ioffe and
Tikhomirov (1974), Mordukhovich (1988, 1999), Munoz and Pedregal (2001) for
problems governed by ordinary differential equations.
To pursue this goal, we will treat the following isop crimetrically constrained
optimal control problem:
where n is a subset in a Euclidean space with a finite Lebesgue measure and the
functions K: 0, X 0, XJR" -+ JR" xl, j: 0, XJR" XJRm -+ JR1, tp: 0, X JR" X]Rm-+ JR,
19 : n x JR" x JRm -+ JRk with JRk ordered by a closed convex cone (to give a sense
to "{) :S 0" ), Yo : D -+ JR" and the multivalued mapping S : 0, ::4 lRm will be
subjected to the following basic data qualification:
(1.2)
with£ referring to (1.1h), though other principles which are standard in integral-
equation theory can be used, too. Of course, we need also a feasibility and
coercivity of the whole problem (P). For simplicity, we can assume
for x En andrE .!Rn , see also Balder (1993), Bennati (1979), Zolezzi (1972) .
From the assumptions (1.1a) and (1.3b) it follows that Q(x, r) is closed for a. a. x
and for all r. A solution u to (P) will be called stable if any sequence of controls
converging to some optimal control is minimizing asymptotically feasible, i.e.
for a. a. X E n and all r E R" . Then (P) has an optimal solntion. Moreover,
any solntion to (P) is stable in the sense (1.5).
REMARK 3 (Volterra integral equations) The special case (of th e Volt er'Ta type)
of our integml equations has been already investigated by Angel (1976), Balder
(1993), Gadson (1990} and Yeh (1978}; they used n := [0, T], l := 2, K(t, T, y)
Optimal control of Fredholm integra l equations 307
which is a norm, see Roubfcek (1997), Example 3.4.13. Since 1 > 1 and q < +oo,
both p/(1'-l)(fl.) and U(fl.) are separable, and t hus also H is separable if
1
equipped with the norm (2.3) . For a ranging C(fl.; ffi. ), this separability was
shown in Roubfcek (1999), Lemma 1. Here however, instead of the obvious
inequality \\a· h\\H ~ 1\a\\L"' (fl) \\h\\H , we must use \\a· h\\H ~ \\a\!l()i~!1l(fl) +
1\\h\'\\H, which allows us to estimate
1h- l )
\\a·(! o y)- a· (f o y) \\H -< b1\ llf o vi' IIH + c6, \ a- all'£1/(>-l)(fl;u-..) TlJ)I
with b1, b2 > 0 arbitrarily small and C 0 ,, C6 2 depending on b's . Then, for a given
c > 0, a E L'/(1-ll(fl.; ffi. 1) andy E U(fl.; ffi.n) , one can take b1 small enough so
a
that It ~ c I 4, then take from a (chosen fixed) dense countable subset close
enough to a so that h ~ c I 4, further take b2 sma ll enough so that h ~ c I 4,
and finally y from a (chosen fixed) dense countable s ubset close enough toy so
that 14 ~ c I 4 by using (1.1e). This proves the separa bility of the set {a· (f o y );
a E p/(-r-l)(fl.;ffi.1), y E U(fl.;ffi.")}. From this, the separability of the whole
space H follows easily by using also (1.1c,j).
Furthermore, we define a (norm,weak*)-continuous (possibly not injective)
embedding i: LP(fl.; ffi.m) --t H* by
Minimize (i(u),cpoy) }
subject to y(x) = y0 (x) + (i(u), (Kx f) o y) for a .a. x E 0., (P')
(i(u), {) o y) ~ 0, (i(u), hs) = 0,
y E Lq(fl.; ffi.")) , u E LP(fl.; ffi.m).
Note that, as S(x) is closed (see (1.1 k)), we have hs(x, s) > 0 for s E ffi.m \ S(x)
while hs(x, s) = 0 for s E S(x), and therefore (i(u), hs) = 0 is indeed equivalent
to u(x) E S(x) for a .a. x E 0. .
Furthermore, we define the set of the so-called generalized Young functionals
by Y);(0.; ffi.m) := w* -cl-i(LP(fl.; ffi.m)). It is known from Roubfcek (1997) that , as
" rrm<:PnllPnrP nf (? 1) wit.h (1 .::\l . Y~ (fl.: ~rn) is a convex locallv (seauentiallv)
Optimal co nt rol of Fredholm integral equations 309
Minimize (17, cp o y) }
subject to y.(x) = y 0 (x) + (17, (.K "' f) o y) for a.a. x E rl,
(RP)
(17, 19 o y) ::; 0, (ry, hs) = 0,
y E Lq(rl; JRn)), 17 E Y};(H; lRm) .
PROPOSITION 1 Let {1.1} - {1.3} hold. Then the relaxed problem (RP ) always
has a solution 17 E Y};(rl ;lRm). Moreove1·, min( RP )::; inf(P).
Proof. It just follows from Roubfcek (1998), Proposition 4.1 , together with Re-
mark 3.2 . Note t hat it can be explicit ly seen from Roubfcek (1998), Remark 3.3
that Ro ubfcek (1998), Condition (10) is ind eed fulfilled. •
Let L';'(rl ; rca(JR"')) denote the space of weakly measurable essentially
bounded fun ctions on rl with values in t he space of Borel measures on JRm ,
and rcai(JRm) the set of all probability measures on JR'" . Further, let us define
the set of t he so-called LP- Young measures as
r /'
.JI! ./'Rm
19(x, y(x), s) l!x(cls) dx ::; 0,
supp(vx) C S(x) for a.e. x E rl ,
y E Lq(rl; JR"), v E Y~'(rl; JR"') .
3. Proof of Theorem 1
Now we are ready to give a quite simple proof of Theorem 1, following Ro ubfcek
(1999), Lemma 2. Let us remark that the optimal Young-measure solution
and the measurable-selection technique has already been used in Munoz and
Optimal contro l of Fredholm integral equations 311
Indeed, (3.1) implies (1.6) because, takingq 1 ,q 2 E Q(x,r), one has s 1 ,s 2 E S(x)
such that qt 2: cp(x,r,si), q~ = f(1:,r,si), and q1 2: cp(x,T,si) for 'i = 1, 2,
and then (3.1) guarantees existence of s 3 E S(x) such that :Z:i=l,Z ~(cp(x, r, s t
f(x, r, si), cp(x, r, si)) E Q, which eventually results in :Z:i=l, 2 ~qi E Q(x, r). Con-
versely, (1.6) implies (3.1) because always co[cp x f x 19](x, r, S(x)) C coQ(x, r).
PROPOSITION 3 Assume {1.1) - {1.3) hold. Let v be the solution to (RP') and
let also {1.6) hold for a. a. x E 0 and aliT E lRn. Then there is u E LP(O; lRm)
such that
u(x) E U(x) := { s E S(x); cp(:r, y(.1:), s) :S: .L"' cp(x, y(x), a) v,c(da),
f(x, y(x), s) = { f(x, y(x), a) Vc(da),
}Rm
19(.7:, y(x), s) :S: { 19(x, y(x), a) llx(da)}, X E f2, (3.2)
J'R"'
and any such u is an optimal control for (P).
r [cp
km X f X 19](x,y(:r),s)l/x(ds) =lim /
J-oo km [cp X f X 79](x , y(x),s)l/~(ds)
E clco[cp X f x 19](x,y(x),supp(vx))
C clco[cp x f x 19](.7:, y(.1:), S(x)) C Q(x, y(x)); (3.3)
this limit passage is indeed correct at each 1: E n for which JR"' !s iP 1/x(ds) is
finite, sec Roubicck (1999), Lemma. 2 for details.
This enables us to show that the set U(:r) defined by (3.2) is nonempty.
Indeed, because of definition (1.4), for any (qo , q1 , qz) E Q(x, y(x)) there is
s E S(x) such that qo 2: cp(x, y(:r), .s), q1 = f(x, y(x), s) and q2 2: 19(:r, y(x), s).
Hence, for the particular choice
r -
312 T. ROUBICEK, W.H. SCHMIDT
the inclusion (3.3) implies that qo 2: 'P(J;, y(x), s), q1 = f(x, y(x), s) and q2 2:
?'J(x, y(x), s) for somes E S(x).
Moreover, by Aubin and Frankowska (1990) the multivalued mapping U :
n =l !Rm defined by (3.2) is measurable because s is measurable, 1/ is weakly
measurable, and 'P, f and ?') are Caratheodory mappings, see , again, Roubicek
(1999), Lemma 2 for details.
0 bviously, U (X) is closed for a.a. X E f2 because S (X) is closed and 'P( X, r, ·),
f(x, r, ·)and ?'J(x, r, ·)are continuous, see (1.1k) and (1.1a), respectively. Then,
by Aubin and Frankowska (1990), Theorem 8.1.4, the rnultivalued mapping U
possesses a measurable selection u( x) E U (x).
In particular, u(x) E S(x). Moreover, in view of (3.2) with (3.4),
and also
fo K(x,~,y(0)(1m f(~,y(0,s)vt:(ds)) d~
y(x) = Yo(x) +
fo 'P(x,y(x),u(x))dx :S fo qo(x)dx
= r r ¢(x, y(x), s) Vx(ds) dx = min(RP') = min(RP)::; inf(P).
Jn }Rm
(3.8)
In particular, (3.8) and the coercivity (1.3b) together with (1.3a) and (1.1b)
imply that
REMARK 5 (Zero relaxation gap) Under the assumptions (1.1) - (1..'3) and (1.6),
Proposition 3 implies that the first term in (.'1.8) equals to inf(P) , so that (3.8)
gives min(RP) = inf(P) = min(P) , i.e. th ere is no relaxation gap. Let us em-
phasize that this is a nontrivial fact for problems involving state constraints.
with £1 E L'~(D.; Lqf (q- 2 l(D. )) ; here <p;. denotes the differential of J( x, ·, s), K~ is
the differential of K(x, ~' ·) , etc., and the remaining notation is as in (1.1) .
Moreover, (l.lf) is to be strengthened a bit:
PROPOSITION 4 Let the assumvtions (1 .1 )-(1.3 ) and (.4..1 J hold. o.nd {p_f. f.h.P.
314 T. ROUBICEK, W.H. SCHMIDT
for a.a. X E n with (-)T denoting the transposition . Moreover, the following
transversality condition holds:
'THF.nRRM?. T.Pf th P n.RRnm.nfinnR (1.1 )- (1 . .'1) and (Ll) hold. and let the cone
Optimal control of Fredholm integral eq uations 315
with c01-responding >. 's from Proposition 4, and let the following condition hold
for a.a. X E [2;
co [<p x f x 19](.r-,y(x),M(x)) C Q(x,y(:r;)), (4.6)
where M(x) C S(x) is an estimate of the set of maximizers of the Hamiltonian
H ,\ 0 ,.x 1 ,.X,y, i.e.
M(x) :J {s E S(x); 7i.x 0 ,.x 1 ,.x,y(x,s) = ~~(:) 7i.x 0 ,.x 1 ,.x,y(x,s)}.
8 (4.7)
Then ( P) has at least one solution and, moreover, every solution to ( P) is stable
in the sense (1.5).
Proof. Let I/ denote some Young-measure representation of the optimal relaxed
control 11 in question. By Proposition 4, for a.a. X E n, 1/x must be supported
on M(x) . T hen, Proposition 3 gives a solution ·u to (P) if used in a refined
way, namely (3 .3) can simply exploit M(x) and (4.6) instead of S(x) a nd (3.1),
respectively. The stability (1.5) is again the consequence of zero relaxation gap
(by arguments as in Remark 5) and of the correctness of the relaxed problem. •
REMARK 6 There are two extreme situations. First, the maximum principle
does not give any specific information, see Section 5.1 below, or the particular
problem often is so complicated that one is unable to extract such informa-
tion; then we can say that M(x) = S(x) only, and (4.6) coincides with (3.1),
i. e. with the Pilippov- Roxin-type condition (1.6}. Second, sometimes it may
happen that one can a pr-iori guarantee, by a specific analysis, that the Hamil-
tonian 7i.x 0 ,.x 1 ,,\,y (x, ·) is maximized only at a single point for a.a. x E n, see
Remark 7 below, i.e. the set on the right-hand side of (4. 7) is a singleton for
a.a. x En, and by choosing M(.r,) equal to this set we make the condition (4JJ)
trivially satisfied.
5. Examples
In this last section we present three concrete, rather simple, illustrative ex-
amples .
l
To present the example by Schmeling (1979, 1981) in our context, we must
modify it slightly. For example, we consider the following problem:
As usual for Volterra equations, we put n := (0, T) and write t and T instead
of x and ~, respectively.
For reader's convenience, let us remind the slightly modified arguments from
Schmeling (1981) to show t he nonexistence of solutions of (PI): Taking a fast os-
cillating sequence {uk: (0, T) ___, {1, -1} hEN converging to 0 weakly in LP(O, T),
we get a sequence of corresponding states {YdkEN converging to y(t ) = ~t 2
in C(O, T), which shows inf (Pr) = 0. Therefore, if there exists a solution (11, y)
to (Pl), then inevitably y(t ) = ~t 2 would hold. By differentiating the Volterra
equation involved in (Pl), one would get
for a.a. t E [O,T]. Note that, since u is bounded, t f-+ J;[(T- T)u(T ) + (t-
T)u( T) 2 ] dT is Lipschitz continuous and thus a.e. differentiable. If possibly (5.1)
does not hold just fort:= T, we can pass to the limit with t ___,Tin (5.1), which
J
gives T = 0T u(7) 2 dT. T his would be possible only if lui = 1 a.e. on [0, T].
Coming back to (5.1), we would get t = (T- t)u(t) + t, which would give u = 0,
a contradiction showing that a solution (u,y) to (PI) cannot exist.
Of course, ( P 1) must somehow violate the assumptions of the above presented
theory. We consider n = m = 1, p, q arbitrary, cp(t, 7', s) := (2r- t 2 ) 2 (we
can add the term like c-max(ls!P, 1) t o satisfy formally (1.3b)), {) := 0, and
S(t) := [-1, 1]. Then, we can think of choosing l = 1 and thus have to p ut
f( t, T, r, s ) ·-
.- (T - T)s + (t - T) s,2 , K ( t, T, 1·) -_ { 1 'f t
if 2 T, (5.2)
0 1 t < T.
This makes the orientor field f(t, T, r, S) always convex compact, but evidently
such f does not have the form required in Theorems 1 and 2 because it depends
also on t. Alternatively, we can choose l = 2 and then put
It has already the above considered form, but the orientor field
Minimize k y(x) 2 dx
s ubj ect to y( x) = Yo(x)
for x fixed (but a rbitra ry). This ensures that, for a.a . x En, t he set My(x) :=
{8 E S(.1:) ; 7-ly(x,s) = max 7-ly (x,S( :r ))} satisfi es
(r; 7\
Optimal control of Fredholm integral equations 319
d·)
0
r
d.~;
d (
= d::2 a(1- x) + bx + ./o (1- x)~w(O d~ + 1
x
1
x(1- Ow(O d~
)
d ( b-a-)
=d.?.: r ~w(Od~+ 11"'(1-~)w(Ocl~ )
0
= -:z:w(x)- (1- x)w(x) = -w(x).
Let us mention an illustrative interpretation of (5.8) as a deflection of stretched,
homogeneous, elastically supported (c determines the linear response of the
support) string with fixed end points, loaded in a perpendicular direction by the
force sin1t. This also shows that the solution y E L 2 (0, 1) is, in fact, smooth,
namely y E W 2 •00 (0, 1) if one uses the standard notation for Sobolev spaces. It
is an interesting observation that, although f(x, r, ·) is nonlinear, the condition
(l.G) is valid, i.e. Q(x, r), here independent of (x, T), is always convex. Thus
Theorem 1 yields existence on an (unspecified) optimal control.
Let us now modify (P3) by restricting the admissible control values to the set
S(x) = [-37!', -21r] u [-1r, OJ u [21r, 37r]. (5.9)
Then Q(x, T), again independent of (x, r ), is no longer convex so that Theorem 1
does not apply. Let us analyze the optimality conditions. The adjoint equation
(4.4) now looks as
.\(x) = -c 1 1
K(.1:, 0.\(~) d~- 3.\ 1 y(0 2 (5.10)
for K(x, 0 = K(x, ~, T) defined by (5.6) and for some .\o, .\ 1 non-negative, .\o+.\1
> 0. We claim that
·1
A(x) := K(~, x).\(0 d~::; 0 for a.a ..1: E [0, 1]. (5.11)
./ 0
For c::; 0, we can see directly from (5.10) that even A ::; 0 as a consequence of
non-negativity of]{ and /\ 1. Using again K;::: 0, we get (5.11). For c > 0, we
must usc finer argument: A defined by (5.10)- (5.11) solves, in fact, the adjoint
two-point boundary-value problem
320 T. ROUBICEK, W.H. SCHMIDT
and then (5.11) follows from the maximum principle (in the sense used in theory
of 2nd-order differential equation) by use of a cont radiction argument, saying
that, if maxxE[D,l] A(x) > 0, then at some x E [0, 1] there must be A(x ) > 0 and
~A(.T):::; 0 simultaneously, which, however, contradicts (5.12).
Now, the Hamiltonian (4.3) takes the form 'H>.o)q ,>.,y(x, s) = A(x) sin s -,\ 0 s 2
(up to a function constant ins-variable) and, having (5.11) at our disposal, we
can see that its maximum cannot be attained for s positive. Thus, from the
maximum principle (4.2), we can see that any solu tion v to (RP3) satisfies
supp(vx) C [-1r, OJ U [-31!', -211'] =: M(x) for a.a. x E [0, 1]. (5.13)
References
ANGELL, T.S. (1976) On the optimal control of systems governed by nonlinear
Volterra equations. J. Optim. Theory Appl., 19, 29- 45.
AUBIN, J.P. and FRANKOWSKA, H. (1990) Set-val-ued Analysis. Birkhanser.
BALDER, E.J. (1984) On existence problems for the optimal control of certain
nonlinear integral equations of Urysohn type . .J. Optimization Th . Appl.,
42, 447- 465 .
BALDER, E. (1993) Exist ence of optimal solutions for control and variational
problems with recursive objectives. J. Math. Anal. Appl., 178, 418- 437.
BENNATI, M.L. (1979) Un theorem di esistenza per controlli ottimi di sistemi
definiti da equazioni integrali di Urysohn. Ann. Mat . Pura Appl. (IV) ,
121, 187- 197.
BITTNER, L. (1994) Existence of optimal control for integral processes . In: XI.
Herbstschule "Variationsrech., opt. Prozesse und Anwendungen", Schmidt ,
W .H .. r.d .. Prenrint- Reihe Mathematik No. 1. Universitiit Greifswald.
Optimal control of Fredholm integral equations 321