A Maximum Principle for Smooth Infinite Horizon Optimal Control Problems with State Constraints and with Terminal Constraints at Infinity ()
1. Introduction
The aim of this paper is, in a control problem with unilateral state constraints and terminal conditions at infinity, to obtain necessary conditions, with a full set of transversality conditions at infinity, which frequently make it possible to narrow down the set of candidates for optimality to only a few, or sometimes a single one. In infinite horizon problems without unilateral state constraints (pathwise constraints), with or without terminal conditions on the states at the infinite horizon, there exist various types of necessary conditions for optimality, and examples are [1] (without a transversality condition), and a number of results with certain limited types of transversality conditions, for example [2] , slightly generalized in [3] . See the latter paper and [4] for several further references (see also [5] ). The limited types of transversality conditions mentioned are in problems with several states-often insufficient if one wishes to avoid getting an infinite number of candidates. With strong growth conditions there exist necessary conditions, with a full set of transversality conditions at infinity, which in many cases make it possible to narrow down the set of candidates to only a few, or sometimes a single one, see Theorem 16, p. 2441 in [5] . For nonsmooth problems with a full set of transversality conditions in the infinite horizon case, see [6] . For such problems, see also [7] .
The novelty of the results in this paper is hence the establishment of necessary conditions that include a full set of transversality conditions at infinity in an infinite horizon problem with both terminal constraints at the infinite horizon and unilateral state constraints (constraints of the form
for all t). Strong growth conditions are needed for the results to hold.
For Michel-type necessary condition in the case of unilateral state constraints, sees [8] .
The growth conditions used below, ((11), (12), (13)) are more demanding than the conditions applied in [9] for the case of no unilateral state constraints and no terminal constraints (problems with a dominant discount). In later work the authors use even more general conditions, see [10] (see also [11] , and [12] for problems with a special structure).
The results below are of especial interest in the case where not all states are completely constrained at infinity. In the opposite case, generalizations of Halkin’s infinite horizon theorem in [1] to problems with unilateral state constraints where no transversality conditions appear, like Theorem 9, p. 381 in [6] , frequently yield enough information for determining one or a few candidates for optimality. When not all states are completely constrained at infinity, transversality conditions related to the terminal conditions are needed, unless one can accept the possibility of an infinite number of candidates for optimality.
In certain cases there is a danger of degeneracy of multipliers. See the early review in [13] and [14] . We have added conditions that secure nondegeneracy of multipliers in some such cases, in particular in the case where unilateral constraints are satisfied as equalities by the initial state (the state at time zero). See [15] -[17] for a presentation of similar conditions in the finite horizon case, as well as for a number of references for this case (see for example [18] -[22] ).
2. The Control Problem, Necessary Conditions, and Examples
Consider the problem
(1)
where
subject to
(2)
(3)
(4)
where we require that
exists for
and where
Here, 
and n are given natural numbers, and we allow for the case where there are no equality constraints or no inequality constraints in (4) (in which cases
and
so
and/or
are empty sets). Furthermore, 
are fixed entities, u the control. It is possible that
in which case
is replaced by
We want to maximize the objective in (1) over the set
of all measurable functions
taking values in U and being bounded on bounded time intervals, subject
to (2)-(4). When the solution
corresponding to such a
satisfies (2)-(4), we call
admissible. Below
will be a given optimal admissible pair, assumed to exist.
We assume that
is continuous in
that
is measurable in
continuous in
with derivatives
and
, where
is continuous in x and
is continuous in
We also assume, for any bounded sets
and
that ![]()
and that for any x,
These assumptions are called the basic smoothness
assumptions. At various points some strengthening of these assumptions are added.
The following definitions are needed: let
![]()
![]()
let
be the resolvent of the equation
(
I the identity map),
if
![]()
![]()
![]()
![]()
In Theorem 1, in addition to the basic assumptions, assumptions (5)-(15) below are needed. It is assumed that for all ![]()
(5)
We shall make use of some constraint qualifications, (6) and (8) below, related to
Define
,
![]()
![]()
,
![]()
(6)
((6) holds vacuously if
).
(7)
(8)
![]()
![]()
(9)
Either2
![]()
![]()
![]()
(10)
The following growth conditions are also needed: For some ![]()
(11)
and there exist some positive constants
such that
(12)
(13)
where
. In (10) (g), we also need that for some
for all
![]()
Assume finally that, for all j
(14)
(15)
Define
The following necessary conditions for optimality holds.
Theorem 1. (Necessary condition, infinite horizon) Assume (5)-(8), (11)-(15) and the basic smoothness as
sumptions. There exist a number
vectors
,
bounded vector functions
and
nondecreasing and right continuous on
, such that if
3 satisfies, for
, the equation
(16)
then
(the limit does exist), ![]()
is left continuous on
and satisfies (16) for
(the integrals exist), and
(17)
Moreover,
(18)
(19)
![]()
(20)
Finally,
for some
and if (6) fails, then (18) must be replaced by
,
Remark 1. For
the growth condition in (13) can be weakened to : For some
for all
(where still
satisfies
). ,
In the sequel three trivial examples with rather obvious optimal controls will be presented, but to illustrate the use of the necessary conditions, we derive the form of the optimal controls from these conditions.
Example 1.![]()
, ![]()
free, ![]()
![]()
![]()
![]()
free.
Solution:
Evidently
and, because
by necessity,
for all t. For
from (16) we get
Then
The maxi- mum condition is that
(21)
Consider first the case that we might have
Let
satisfy
Now, for
means that
for t close to T, ![]()
see the expression for
so
for t close to T. But this surely continues back to
(see
the expression for
again). So even
for
(for such t, in fact, ![]()
when
). Let
be the smallest
such that
for
Consider first the subcase
Then
and by the expression for
,
and
By the maximum condition
for
for
We must have
in order to obtain
we get
![]()
![]()
for
for
so
For
(
an arbitrary constant). Using
we get
hence
Thus,
,
with k satisfying
i.e.
so
(By the way, note here that as
satisfies the equation for
we would know that
for some constant C, but
would not determine the constant. This shows the usefulness of the formula
) The subcase
is impossible, then
and
The case
is impossible, then
for t close to T, so
for such t and then for all t (see the expression for
). In fact,
when
so
for all t, implying
and
a contradiction. Consider finally
Then, by (18),
so
and then
for t close to T, so
for such t, in fact for all t (see the expression for
), and
Hence,
for all t, which gives
and
contradicting
,
Remark 2. (Further non-triviality properties)
a) Replace (6) by the assumption that either
is empty, or (if not), for some
some
for any
there exists a
such that
for all
where
is assumed to exist. Assume also that
and
are bounded4. Then ![]()
b) Assume in addition that, for any
either
is empty, or (if not), there exists a
such that
for all
and, in case
that for each u,
is continuous, that
is left continuous at each
and has a limit when
Then
,
For finite horizon normality conditions, see [23] and [24] .
The main reason for including the next theorem is that it forms a basis for obtaining Theorem 1, but it has some interest of its own.
It contains necessary conditions for the case where (14) and/or (15) fail, in particular where
also depends on
We then need three conditions, see (25)-(27) below, that automatically hold if (14) and (15) are satisfied.
Theorem 2. In the situation of Theorem 1, with (5), (6), (8), (14), and (15) deleted, assume that the three conditions (25), (26), (27) below are satisfied. Then the following necessary conditions hold: for some
for some vector
and some bounded nonnegative finitely additive set functions
vanishing on sets of Lebesgue measure zero, for a.e. s, for all ![]()
(22)
where
(
and the integrals exist). Moreover, ![]()
satisfying (20). Finally, defining
we have
(23)
(24)
If (7) and (8) hold,
Moreover,
if, for some
and some positive
when
for all
Finally, if both the last condition and (7) and (8) hold, then
,
As before when
is replaced by ![]()
Assume, for some arbitrarily large
that the conditions (25)-(27) hold
(25)
Let
be a solution on
of
for
given. For some positive second order term
(i.e.
), if
then, for all j, for all ![]()
(26)
Moreover, for any given number
and any given positive second order term
a positive second order term
exists such that the following property holds. Let
be a solution on
of
for
given,
. Then, if ![]()
(27)
As an example in which (25)-(27) hold, consider a case where
(11) and (12) hold for
where f
is concave in x, and where, for some positive
is
and convex,
and
all j,
(For
in a short
hand notation
![]()
where
which means that, a.e.,
so
If
then
and then ![]()
Remark 3. For Theorem 2 to hold, we can weaken (7) and the basic assumptions on
and
as follows: the derivatives
and
exist at
for all t and the three conditions on
below are satisfied: For all ![]()
(28)
(29)
(30)
Moreover in the growth conditions (12) and (13), roughly speaking, the inequalities need not hold for states x that cannot possibly occur, more precisely, the conditions can be modified as follows. Define for each t,
![]()
![]()
the solution of (2) corresponding to
. Then (12), (13) need only hold for some
for
such that
In (10)(g) we must add the assumptions that, for some
is differentiable in
at
uniformly in
with a derivative at this point bounded uniformly in ![]()
Finally, U can be replaced by a time dependent subset
where
,
continuous, we then require all control functions to satisfy
We assume
and in case
in (9), we require
Then the maximum conditions (17) and (22) hold only for
(The set U can even be replaced by
with (17) and (22) holding for
We must then still require
and, in (9),
In the proof below, the perturbations
of the optimal control have to belong to
. ,
Example 2.
![]()
![]()
for all t, ![]()
![]()
free. It is convenient to replace
by
(Then (25)-(27) will be satisfied). Choose
as the largest possible
such that
for
If
and even, for some
for all t, then
so
by the maximum condition
(31)
and
, contradicting the last inequality. Consider now the case where ![]()
Let
be the set of time points s for which (31) holds. Now,
implies the existence of some
for which
5, (
see the state equation). If
can be chosen arbitrarily large, if
can be chosen arbitrarily close to
. Now,
so
If
then
(use (31) for
), implying
contradicting
Hence,
(Note that this
argument would not work if we had replaced
by, say,
). For
by (31),
and because
is strictly decreasing on
(
is con-
stant on
), in fact
on
in fact on all
by (31). Hence, on ![]()
so
is finite and
and
It is easily seen by a similar argument that on
![]()
: to see this, having
on an interval
, assuming that
is as large as
possible, is impossible: Let
and define
Now,
leads to
on
which is impossible, and both for
and
and for
and
certain time points in
close to
(i.e. arbitrarily large if
) exist at which
but then
in
by (31), as
is strictly decreasing in
. But
in
contradicts
in
. Hence, on
,
by (31), so
or
on
(
can be represented by an integrable function
here). If we put
, we have a valid proposal for the multipliers (It can be seen that
is even necessary, compare (89(ii)) on p. 333 in [6] ).
Example 3.
![]()
![]()
![]()
, ![]()
free. It is convenient to replace
by
The maximum condition is
(32)
Again, assuming, for some
that
for all t, and (in the opposite case)
both yield contradictions. So
Now,
in (32) is impossible, so
all the time. But, due to the constraint
then
all the time (see the state equation). Let
The maximum condition (32) yields
, for
Now,
. The general solution is
To have the initial condition and
satisfied we need
and hence ![]()
Remark 4. Assume in the problem (1)-(7), (11)-(15), that
is convex and that there are given additional constraints in the problem of the form
(Perhaps
In this case, here and below, replace
by
.) Assume that
is optimal in this problem. We assume for
that
is continuous and depends only on
, that
and
exist, that
that
is continuous in
uniformly in t and that
is
in
and measurable in t. We assume, for some positive constants
for all
that for
and
for
and
for
and, for
that
and
for
and
for
Define
{
: For all
such that
}. Assume that
and that (8) holds, for
for
in (9). Write
Then, in addition to
and
satisfying (20), (23), and (24),
there exist bounded non-
negative finitely additive set functions
also vanishing on Lebesgue null sets, such that (22) holds
for
, summing now over
Moreover, for
for all ![]()
(33)
(34)
(
and the integrals exist). Furthermore, for
we have
on
and on
for all
Finally,
where
(If (6) fails, then the last property must be replaced by
).
When for some vectors
, ![]()
the following properties hold: the maxi-
mum condition (17) holds for
, together with
(35)
for all
where now ![]()
(36)
and where, for some nonnegative
,
,
for
and
for
if
or if
for
Moreover,
and in (36), for
can be
represented by a bounded nondecreasing right-continuous function
and, finally,
(If (6) fails, then this property must be replaced by
).
When, in addition, for some
some
for all
the inequality
holds for all t, then
for
and, for
in both (36) and (35),
can be represented by a nonnegative function
in
(replace
by
) and, moreover,
(If (6) fails, then this property must be replaced by
). Finally, in this case, for a.e. t, for all ![]()
,
3. Proofs of the Results
Proof of Theorem 2. To simplify the notation, instead of the criterion (1), we can and shall assume that
is the criterion to be maximized,
that
is free, hence is not required to be equal to
The proof will be carried out under the assumptions of Theorem 2, allowing for the weakening of these assumptions in Remark 3.
Overview of the proof. A rough outline of the proof is as follows. We are going to make a number of strong (needleshaped) perturbations of
This gives rise to first order variations of the optimal trajectory (the
- functions below). We introduce a convex subset of these variations (
below) consisting of variations satisfying a first order version of the unilateral constraint. We then introduce the convex set of endpoints (at infinity) of these variations as is standard in traditional proofs of the maximum principle, and show that it has to be separated from the set of “better, first order admissible” points, the set {
for
if
} (The endpoints we consider consist actually only of the first
components of the state). The separation argument (carried out in
) consists of a standard use of the Brouwer fixed point theorem combined with the fact the endpoints are “good” first order approximations of the endpoints of the exact solutions following from the perturbations. We need the fact that these exact solutions satisfy the unilateral state constraint, and this is shown first. The separating functional (
-vector) is denoted
. Another separation argument, carried out in
-space’ gives the multipliers
related to the unilateral state constraints.
Detailed proof. To avoid certain problems connected with coinciding perturbation time points, the following construction is helpful (we then avoid coinciding perturbation time points). Let
be a countable dense set in
and let
be the set of right Lebesgue points of
and all
in
Then choose some set
of full measure (i.e. meas
),
such that for each
a subset
of
exists with the property that if
and
then
and with the property that for each
each
there exists a sequence
such
that
and
For any given
let
be the collection of
-tuples of the type
![]()
![]()
![]()
and for all
such that
belongs
to
and
belongs to
(This means that for any
if
then
which implies
). The separate treatment of the case
where we can have several perturbations at the same time, is useful for obtaining nondegeneracy results (i.e. informative necessary conditions). Below
is varying ![]()
Let
and for
define
(37)
and
![]()
![]()
![]()
where a sum over an empty set is put equal to zero. For
![]()
Let
and note that
is convex6. Define
to be the convex subset of
consisting of functions
that satisfy: For some
![]()
(38)
where ![]()
The linear variations
are the ones that will appear in the necessary conditions that will be obtained (see (53), (54) below). These variations are jumping at each perturbation time points, so, near these points, they do not approximate (to the first order) the corresponding (continuous) exact solutions. Yet, we are able to show that the latter solutions satisfy the unilateral constraints when
belongs to
. To show this, the “better”, continuous, approximations
are used.
3.1. Satisfaction of the Unilateral Constraints by Perturbed Solutions
Fix a pair
such that
satisfies (38) for certain numbers ![]()
![]()
. Let
be some number for which (25), (26) and (27) hold. There exists a
so small that
(39)
To see this, choose
such that (39) is satisfied in this manner both for
and (by using (25)) for
Then, for
satisfies
(40)
For some positive
for
for all
, when
then
and
are disjoint for all
and, moreover,
if ![]()
(
is closed). Let
be the set of points
in
for which
(41)
Now,
(42)
by (40) because
in
Let
and define
inductively by the for
mula
(
). Let
and, for
let
![]()
(see (37) for
). Then, assuming right continuity at
, (see (9)7), it is easily seen that
(43)
is small when
is small,
uniformly in
See the arguments connected with (74) below. For
let
for
let
for
(recall that these intervals are disjoint when the left ends differ and that
), and
elsewhere. Let
be the corresponding solution. Define
![]()
Now, for
small,
for
hence by (43) and the Lebesgue point property of
and
at
is of the second order in
uniformly in
and
Let
We want to prove that for some ![]()
(44)
Because
is bounded by assumption (25), and
is of the second order,
is of the second order, uniformly in
Moreover,
is of the second order, uniformly in
by the boundedness of
Hence, by (41), for some positive
for ![]()
(45)
Next, it is well-known that
(46)
uniformly in
(see Lemma D in Appendix). Moreover, by the differentiability assumption on
at
((28)), for some positive second order term
, we have
(47)
for all
(because
is of the first order in
). Hence, for some positive second order term
for ![]()
(48)
both for all
by boundedness of
on
(combine (46) and (47)) and for all
(com-
bine (27) and (26)). Then, by (45), for some positive ![]()
for
Moreover,
for some positive
for
for all t, by (48),
and (by (25)),
uniformly in
so uniformly for
![]()
(49)
Hence, (44) has been shown, and in particular, (see (42))
(50)
So far, we have only used the basic assumptions and the first of the three conditions on
in Remark 3, namely (28). The other two properties, (30) and (29), will be used in what follows.
We want to show that
when
Now,
implies
for
close to
(recall
). Thus, by (40),
for
and
For
small (![]()
),
when
(by continuity,
is small, when
is small). Note that if
for some
,
is a convex combination of
and
(the former one has weight
). Hence, for any
belongs to the segment between
and
hence
for
![]()
So far, we have proved (44) for
recall that if
then for
,
(in particular this holds for
so for
(44) holds for all
).
Finally, let us prove (44) for
We can assume that
is so small that
if
So consider the case where
, ![]()
Then the right derivative
First
consider the subcase where
and
(only one
). By (39),
![]()
Combining the two last weak inequalities we get
Then
so
for
for
small (![]()
).
Consider next the subcase where
contains several pairs
,
Using (43), it is easily seen that
When
and
we have that
Now,
by (39), so
Hence we again get
for
for
small (![]()
).
Thus, when
satisfies (38), then
(51)
3.2. Local Controllability at Infinity
Observation 1. Define
and
{
for
for all
with
if
}. Note that by Lemma B in Appendix, for all
and
,
and
exist. Let
be the convex set
(for
see (38)). If
, then for some positive
, some
where each
equals
linearly independent. Let
Then ![]()
for some vector
In fact, for each
there is a unique
such that
evidently depends linearly on z, note that
extends linearly to all
. Let
and let
consist of all
for
the moment we allow the pairs in
to be doubly indexed, (and the time points not to be ordered). Then
where
(as we have double
indices on
, we have double indices on the components of
). Note that ![]()
Of course, we can re-index the pairs in
(and so also the entities
) by using a
single index, with the time points in the pairs in increasing order. Let
be the number of pairs
We use
also as the name of the vector of reindexed pairs, and
for the vector consisting of all entities
reindexed in the same manner as the pairs in
. Then for ![]()
![]()
![]()
linear.
The following result should surprise nobody, a proof however is given in Appendix.
Lemma 1. Assume that for
linear,![]()
. Then there exist some first order term
(i.e.
) and some
such that for each
for some
some
, ![]()
,
3.3. Separation Arguments That Yield the Multipliers
By optimality, for all
To see this, assume, by contradiction the opposite,
that for some
Then by Observation 1 and Lemma 1, for
and small,
for some
and
satisfies the unilateral time constraint (3), see (51) (because
and
). The last equality gives that
satisfies the terminal constraint (4) for
small, and that
This contradicts the optimality of
. Thus the sets
and
are disjoint (this is trivial if
). Thus these sets can be separated8: there exists a nonzero vector
such that
As
this inequality gives that
(52)
Define
{
: for all j, for some positive
,
,
when
}. Recall that
for all
see (25), and write
. Note that
(for
, see (38)) has to be disjoint from the convex body
in
otherwise the inequality
is contradicted. By separation, for some continuous linear functional
on
and some number
![]()
(53)
for all
all
Evidently, by this inequality,
and
are nonnegative on
, with
for all
Each
can be represented by a bounded
finitely additive nonnegative set function
vanishing on sets of Lebesgue measure zero. Evidently,
va-
nishes on
for all
in particular
vanishes on
The inequality (53) gives that, for
), for all pairs
(
still fixed in
) and for all pairs
and ![]()
(54)
where
(To obtain (54), in (53) let
). Moreover, (54) also holds for
, for
for any given ![]()
Let us now choose a sequence
converging to zero when
such that (54) holds in the manner
described for
for certain multipliers
In particular (54) holds for
for any given
Let us fix such a sequence
assuming it to be bounded. We can assume that
(we extend
to
by letting
). Using the weak* topology on
there exists a cluster point
of the sequence
satisfying
(55)
(for some subsequence
). (by the cluster point property, so the last
equality holds). The cluster point
is a bounded nonnegative finitely additive set function that vanishes on Lebesgue null sets. It is furthermore easily seen that (54) holds for
and
, for
,
and for
, both for
equal to the cluster point
9 as well as for
provided this limit exists, for any
for any given
![]()
Now
is nondecreasing and bounded. Let
be the continuity points in
of
For any
and for any
a sequence
exists, such that
and
(see the very beginning of this proof), and because (54) holds for
it is easily seen by taking limits that (54) holds for
and
10.
Finally, let us extract an additional property. If
for all t in some interval
and
then, for some
for
when
which by (55) implies
and hence
. Thus
(56)
for (say)
11.
3.4. Further Information on the Multipliers in Special Cases
Let us prove the results concerning the multipliers in the three last sentences in Theorem 2 in the case where
is maximized.
Define
(57)
Now, assume (9) and (10) (a), with
We may assume of the sequence
used above that
for the single
(there exists a sequence
such that
). Here (29) was used. Let
satisfy the inequality in (9), define
and assume that
Then, by (30), for some
for
close to 0,
which combined with the previous inequality gives
So
any given cluster point of
. We can assume that (54)
holds for this cluster point
, for ![]()
Evidently,
for all t close to 0. Now, if
and
for all
then for any
From
we get, for all
close to zero, that
which gives
for
close to zero, and so
a contradiction. Hence, ![]()
When
and 0 is a right Lebesgue point of
(i.e. (10) (b) holds), we can choose the sequence
Evidently, by (29),
when
so again
and we get the same conclusion regarding
(with
again) and in this case, (as well as in the case that
is differentiable,
i.e. (10) (g), see Appendix), we don’t need the assumption that
contains a single element12. In fact,
when (8) holds, in Theorem 2, we can assume ![]()
Define
Note that, by (23),
for
large enough if
Assume that a
and positive numbers
and
exist such that
(58)
(see the end of Theorem 2). For ![]()
![]()
Assume for the moment that
Choose a sequence
such that
Letting
in the preceding inequality, and using
(see Appendix, Lemma A), we get
so ![]()
Can
? No, we have shown that then
, and then
so
a contradiction. So when (58) holds
(59)
Finally, assume that both (58) and (8) are satisfied. By contradiction assume now that
Then
(Þ
and
, so
a contradiction. So (58) and (8) imply
(60)
Proof of Theorem 1.
We still keep the assumption that
is maximized. Using lemmas in Appendix, note that (25)-(28) are implied by the basic smoothness assumptions, the growth conditions (11)-(13), (7), (14) (i.e.
depends only
on
,
), and (15) implying that
is differentiable at 0, uniformly in t ((27) follows from Lemma E in Appendix). Moreover, also (29), (30) evidently follow. So all the above results also hold in the situation of Theorem 1. Using the definition (57) the maximum condition (54) can be written
(61)
Now,
Using
(5) and (14),
when
Then, by (57), also
exists, and
Hence,
when ![]()
Let
Let
be the continuity points of all
Write for the moment
![]()
and
Then13, for
and then
With
(
occurring in the definition of
) it is well-known that
satisfies (16)14. Evidently, (61) yields
(62)
In Theorem 1, we have written
and
instead of
and
.
Note that (6) implies that (58) holds, as
,
. Thus, (60) holds, which means that
in the situation of Theorem 1, because then
exists and equals
.
Proof of Remark 2. We give a proof for the case where
is maximized. Note that
for
for
where
is a set of Lebesgue points of
of full measure. The inequality holds for all
in case of left continuity of
.
Proof of a)
Let
and recall
Assume
Assume by contradiction that
Then, for all
large enough
![]()
![]()
(the last term is
all square brackets
when
). Then
which is a contradiction because
for
So
cannot be
when
. Hence
implies ![]()
Proof of b) Assume by contradiction that ![]()
If
then for some arbitrary large
we must have
For
for all j, for large
the left hand side is
i.e. it does not change much if
is replaced by
. Hence, for
(whether
or =1), for large
, for
as in a),
![]()
Using the vectors
in a), for
large we then have
Because
when
s large, we finally get
when
is large. When s is large
means
Using
and (22) (which even holds for
), for
large, we get
Hence, both for
and for
for some (large) s.
Next, let
By contradiction, assume
Let
where
has the property that
for
see b). By continuity, for any
and close to ![]()
(63)
There exist
and arbitrarily close to
such that for
. If
for all j, the left hand side is
so for
close to
for all
whether
or
Combining this inequality with (63), we get for
s close to
that
for
15. From this, we finally get, by Lipschitz continuity of
on
, uniformly in s, that there exists a
and close to
such that
when
and
can be chosen so close to
that
when
The last inequality and (22) then yields
which gives
both for
and for
and so
for all j. Evidently we cannot have
so
Thus
contradicting
Hence ![]()
Proof of Remark 4. We construct an auxiliary problem: assume for given functions
that we want to maximize
subject to
(64)
![]()
![]()
measurable. Here
are auxiliary states, governed by
where
are auxiliary controls. Write
For
define
{
if
},
, and let
Below,
is so small that
Given any measurable control functions
let
be the solutions of (64) and
corresponding to
. For any
there exists a
such that if
and
then
for all t, by Lemmas B and C in Appendix, hence, by continuity of
at
uniformly in t, for some
for
when
and
(
and
perhaps dependent on
). In the auxiliary problem the constraints are the terminal constraints (4),
for all
for
for all
for
Hence, if
is admissible in the auxiliary problem, we have seen that
is admissible in the original problem when
We assume that
is
(for
see the beginning of Remark 4, then
and then, for
the property related to
in Remark 3 is satisfied in the auxiliary problem for
). So, in the auxiliary problem, ![]()
are optimal in the set of controls {
}. The arguments in the proof of Theorem 2 apply also in the present situation, with one modification: For
the inequality
in
for
automatically holds for
Hence the arguments in the section between (50) and (51) are not needed16 (and do not work) for ![]()
The necessary conditions in Theorem 2 are now applied to this auxiliary system (they apply even when admissible controls are restricted as above, see the inequalities involving
and even for
replaced by
see the end of Remark 317. In the auxiliary system, the linearized system is
where
is the transposed of
The resolvent of the linearized system becomes
(65)
where
From Lemma A in Appendix, we get that
(66)
for some constant Q, independent of t and
, where
is the i’th row of B and
(to apply Lemma A, note that for
in an obvious notation, ![]()
where
). Note that (22), or actually (54), applied to the auxiliary system, holds for
for the limit point
given,
any given element in
(see remarks subsequent to (55)). From this we get, for
and
and
that
(67)
From now on assume
in (67). Moreover, for
from (54) applied to the auxiliary system, we get that (54) holds as written. Finally, ![]()
defined below.
Let
consist of all pairs
such that
satisfies (52) and
satisfies
for all
and (56) for
with
Let
consist of all pairs
such that (67) holds for the given
and (54) are satisfied for all
for a.e. s, in particular for
and
any given
cluster point of any given sequence
, each
corresponding to some collection
from
.
Let
consist of all finite set
We have just proved that for each
,
is nonempty, so by compactness
is nonempty (the weak* topology is applied on the m’s ). Let
be any given element in the latter intersection. Then, for
, both
(67) holds for all
and (54) is satisfied for
for a.e. s. (To obtain this last property, preferably the set
of point s for which (54) holds should be independent of the
’s in each
, one can use that (54) now holds for
, for
for
,
and hence by earlier limit arguments (54) holds for
, for all
, for a.e. s). We also have that (54) is satisfied by
for
for any given cluster point
for any given sequence
, each
corresponding to some
for some
.
The proof of
is the same as the proof for the analogous condition in the case
noting that
for some
means for some
for t near 0 (so
again leads to a contradiction in the same way as before)18. Similarly,
has essentially the same proof as before. To show
in case
holds, we now assume
and we replace
by
in the definition of
where
Then from (66), we get
(68)
Using the inequality
(69)
(i.e. (67)), we get
(70)
Note that
So, from (68), (70) and
,
it follows that
Then
![]()
(note the
-norm on
). But then
,
can be
represented by nonnegative functions
in
in fact in
because
is bounded.
Let us finally show that
when
exists. By (66), for
independent of
and
So, for
,
By this inequality, there exists a
such that, for
for
Next, for some
, for
for
. To see this, for k chosen such that
for
note that
when
By (67) and the two inequalities involving
for
, and for
for any ![]()
![]()
But then
because
was arbitrary. A contradiction of
has arisen, so ![]()
4. Conclusion
The paper establishes necessary conditions for optimality in a smooth infinite horizon optimal control problem with unilateral state constraints and terminal constraints at the infinite horizon. The necessary conditions include a complete set of transversality conditions at infinity. The specific growth conditions placed upon the system in this paper can easily be modified, but strong growth conditions are in any case needed for the full set of necessary conditions to hold.
Acknowledgements
I was deeply grateful to the referees. Their detailed comments made it possible to correct omissions and improve the exposition.
Appendix
Below, for any matrix
.
Lemma A. Let
let
![]()
where, in
, the matrices in the first row are respectively
and
and in the second row
and
, and, in
,
is
and
is
all entries being measurable functions of
and
locally integrable. Assume that for some positive numbers
Write
for the solution
on
,
, of
for
given. Define
Then for some positive number Q, only dependent on
![]()
Hence ![]()
Lemma B. Let
and
be measurable in t, and with
, ![]()
and
Lipschitz continuous with Lipschitz constants
,
,
and
, respectively,
Write ![]()
There exists a positive number
such that the following properties hold: let
and
be two solution on
of
for
, respectively
given, assumed to exist. Then for all t,
and
, so
which implies that
exists if
exists. ,
Note that
and
do exist whenever
and
are integrable.
The proofs of the lemmas A and B are of a standard type and omitted in order to save space.
Let
be given and let
Let ![]()
Lemma C. Let
be a family of functions
such that all
are Lipschitz
continuous in
with a common Lipschitz rank
integrable, and with
measurable. Let
be a solution of
and
given (assumed to exist). Assume that all
![]()
are bounded by a common constant
. Then a constant
exists such
that for any
for any given
, a solution of
exists, and
for all
.
Proof of Lemma C. Note that
so by Gronwall’s inequality, ![]()
Lemma D. Let
be a family of functions
all Lipschitz continuous in x with a
common Lipschitz rank
,
integrable, and with
measurable. Let
be a given function in
and let
be a solution of
,
given (assumed to exist). We assume that all
![]()
are bounded by a common constant
. We also assume that
is differentiable at
for a.e. t. For
let
be a solution of
given. Then, for some
for all t,
when
Moreover, for some
and some second order term
for all t, all
all
all f such that
where
is the solution of
given (it does exist).
Proof of Lemma D. The proof of
follows from Lemma C. Let
and let
on
We have that
and, by differentiability of
at
we have that
for some second order increasing term
, so for
![]()
![]()
for a.e. t. Dividing by d, we get
![]()
By Lebesgue’s dominated convergence theorem,
when
. As
![]()
then
![]()
and then by Gronwall’s inequality,
![]()
where
a second order term in d, equals ![]()
Lemma E. In the situation of Lemma B, let
be differentiable at
where
is a given solution on
of
given, assumed to exist. For each
let
be a solution on
of
given (it does exist), and let
be a solution of
given. Assume that, for some K,
for all
a second order term in d. Then for some second order term
, ![]()
Proof of Lemma E. By Lemma D, for some term
being of the second order in d,
when
is of the second order. For some
for
Hence, by Lemma B, for some constants
we have that
so
![]()
By Lebesgue’s dominated convergence theorem the conclusion in the lemma follows if we can prove that for
each t,
when
To obtain the latter fact, let
and note that
![]()
where the second order term ![]()
Proof of Lemma 1.
Consider the map
for
d any number in
. Let
and note that, by Lemma C,
and
are continuous in
Then, by Lemma B,
and
are continuous in
Let
be a Lipschitz rank of
and
. For
let the second order term
satisfy
when
![]()
for the existence of
see Lemmas D and E. Recall that
and that
when
for some second order term
, see an argument preceding (44) and Lemma A. Hence, for
,
when
![]()
Note that for
we have
as
, and for d small (
for some
),
Fix such a
. Now,
is continuous in
and has a fixed point
here, by Brouwer’s fixed point theorem. As
and
Then let
to obtain Lemma 1.
Observation A. On the space of continuous real-valued functions on
with compact support, furnished with the sup-norm,
can be represented by a nondecreasing bounded function
such that
for all bounded continuous
with compact support. In fact, we can let
, (
right continuous for
). Let
be the continuity points in
of
Then
for any
hence for
and
so
is also left continuous at s. For piecewise constant functions
with bounded support, jumping only at points in
evidently
By approximating continuous function (or even piecewise continuous functions jumping only at points in
) by such piecewise constant functions, one see that the same equality holds for continuous functions (or such piecewise continuous functions)
with bounded support.
Note that if
uniformly (
continuous, with a common bounded support), and
then
: Assuming for simplicity (say)
, this follows from
![]()
Let
arbitrary. For
large,
hence, for any
Thus for
Hence,
when
If
is bounded and continuous, but with unbounded support, evidently, by the last inequality,
so
exists. As also
(![]()
), exactly the same argument works for
the latter limit written ![]()
Proof of (10) (g) Þ ![]()
Assume again that
is maximized and postulate the conditions in Theorem 2 (allow even for the conditions in Remark 3), in particular postulate (10) (g). For simplicity, assume
for
for
We want to replace each condition
by two conditions,
for
for
It can be done by requiring that
holds for
and, by adding new constraints
required to hold on
, where
for
. We now first assume both that
,
is independent of t, and that
is independent of t for
. Assume that there exist
, ![]()
![]()
(see (9)) such that
(71)
Then, for some
both
(72)
and, for ![]()
(73)
Let
have the property that
and
for
Let
be a first order term (i.e.
when
), such that
when
for all
, i. As we did in connection with (43), let
and define inductively
. Let
. Choose a partition
of
such that
. Define
It is easily seen that for ![]()
(74)
for any
To see this, note that for
the left hand side vanishes, while for
the left hand side is smaller that
Let
Define
Let
be any given number, ![]()
such that
and
when
.
Let
Moreover, let
Then
when
as
. So
when ![]()
Consider the following auxiliary control problem on
. Define
Let
and introduce the two state equations
, ![]()
free, and let
and
be the controls. We require
on
for
and
on
for
and
described below. The end conditions on
are as before. Then
and
(on
) are optimal in this problem, see below. Applying Theorem 2 to this problem, with
and
as costates corresponding to
and
, gives ![]()
![]()
for all
for
and the maximum condition (22), i.e. (a.e.)
![]()
![]()
![]()
,
where
![]()
,
In particular, because
when
see (72), then
(75)
Assume by contradiction both that
and that
where
which by necessity means that
Then, fo r
![]()
,
for
for ![]()
for
which gives
and
for
and
for
. Also,
by (73). As
, using (74), we have, for
,
![]()
contradicting
for
(as
is optimal).
The optimality of
follows from the following argument: Let
be an arbitrary admissible quadruple in the auxiliary problem. Let
If
let
for
let
for
and let
,
for
Next, let
Then
![]()
and hence
Moreover, ![]()
So
is a solution in the original system evidently satisfying the end restrictions, and
for
Because
for
and then
for
Finally, on
as we shall see in a moment, so
for
hence
for all
For
let
in which case
is automatically an original solution.
From (74) and
for
we get
(76)
Using
when
for ![]()
![]()
Then, by (72), (73), (76), for all
all
all ![]()
![]()
![]()
19In case we have constraints
,
, (in which case
,
is required), then for
and
small, for
,
, as
.
As
then
for
Moreover, for
for some positive
small enough,
when
(For
there is nothing to prove.)
Now,
for
so
belongs to the set of admissible pairs in the original problem. We have
hence
are optimal in the auxiliary problem19.
Let now
where
and let
and
be corresponding multipliers, satisfying the normalization
We put
Now,
has a cluster point
satisfying (22), (55), and (56). Assume now that
. Then there exist some
such that
and hence
Hence,
which leads to a contradiction, as was shown above. Thus,
.
We can extend this result to problems that are nonautonomous on
, by using t as a new state variable, say z, governed by
with x governed by
provided that
is jointly differentiable in
at
for all t, and that
is differentiable in
at
uniformly in
with a derivative at this point bounded uniformly in ![]()
NOTES
1In that theorem, correct the inequality
by replacing it by
.
![]()
2A right Lebesgue point s of any integrable function
means a point s such that
. If
in (9) even belongs to
and (10) (b) holds, then the right continuity in (9) can be weakened to 0 being a right Lebesgue point of
for all
.
![]()
3
always exists, and is left continuous on
.
![]()
4For
it suffices that
when
,
, note that for
,
for some C.
![]()
5Or even
, an observation needed in the next example.
![]()
6One may consult Observation 1 below at this point.
![]()
7Only when there are more than one perturbation
at
right continuity is needed to obtain (43). If there is a single perturbation at
, (i.e. for all
,
= some
) it suffices that
is a right Lebesgue point of
. Finally, if
, then we shall throughout assume a single perturbation
at
.
![]()
8If
and
, then K and
can be separated. If
, then for some nonzero
,
, and
can be chosen such that
.
![]()
9For simplicity, assume that
. Then, for any
, for all n large enough,
for all for
, so, by (25); for
,
, for any
; for all n large enough,
for all
. As
,
, so
, which converges to zero at least for a subsequence of
.
![]()
10If U is replaced by
as in Remark 3, we asume that
. The perturbations
will belong to
, because in the proof we now require the
’s to belong to
and then
for t in
, for d small. Then (54) holds for a.e. s for
(Þ
for n) and also for
, for any given cluster point
of any given bounded sequence
where
,
,
In addition, (54) holds for
for all limits
of all such sequences
that are convergent. In particular, (54) holds for a given cluster point
of a given sequence
, where
,
given points in
,
.
11In case
, it can be assumed that
, for all j, and, as said before, we then assume that for all P,
,
some arbitrarily given element in
(in
in case U is replaced by
,
for n large).
![]()
12If (8) holds only for a subset
of
then it is easily seen that the collection
,
,
,
is nonzero (at least one entity is nonzero).
![]()
13Observation A in Appendix may be consulted at this point.
14Compare e.g. Section 9.5 in [16] .
![]()
15As an alternative to the left continuity assumption on
in Remark 2 (for
), we may assume that, if
in the necessary conditions, then these conditions imply left continuity of
on
.
![]()
16Thus we don’t need (and often don’t have!) differentiability of
for
.
17The growth conditions related to
are of the same type as those related to
in problem (1)-(4). Note that the perturbations
now belong to
.
![]()
18We can again let
,
also in the present case, (67) does not change for this change of
.