This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

[2]\fnmIlya \surShvartsman

1]\orgdivDepartment of Mathematics, \orgnameMacquarie University, \orgaddress\citySydney, \postcode2109, \stateNSW, \countryAustralia

[2]\orgdivDepartment of Computer Science and Mathematics, \orgnamePenn State Harrisburg, \orgaddress\cityMiddletown, \postcode17057, \statePA, \countryUSA

On Averaging of a Class of “Non-Singularly Perturbed" Control Systems

\fnmVladimir \surGaitsgory [email protected]    [email protected] [ *
Abstract

We study a control system resembling a singularly perturbed system whose variables are decomposed into groups that change their values with rates of different orders of magnitude. We establish that the slow trajectories of this system are dense in the set of solutions of a certain differential inclusion and discuss an implication of this result for optimal control.

keywords:
singular perturbations, averaging method, optimal control
pacs:
[

AMS Subject Classification]34E15, 34C29, 34A60

1 Introduction and the main result

In this paper, we consider the system

ϵdy(t)dt\displaystyle\epsilon\frac{dy(t)}{dt} =\displaystyle= Ay(t)+Bu(t),y(0)=y0,\displaystyle Ay(t)+Bu(t),\ \ \ \ y(0)=y_{0}, (1)
dz(t)dt\displaystyle\frac{dz(t)}{dt} =\displaystyle= g(u(t),y(t),z(t)),z(0)=z0,\displaystyle g(u(t),y(t),z(t)),\ \ \ \ z(0)=z_{0}, (2)

where ϵ>0\epsilon>0 is a parameter, the controls u(t)u(t) are measurable functions taking values in IRk{\rm I\kern-1.99997ptR}^{k}, and AA and BB are m×mm\times m and m×km\times k matrices that satisfy the rank controllability condition:

rank[B,AB,,Am1B]=m.{\rm rank}[B,AB,...,A^{m-1}B]=m. (3)

The function g:IRk×IRm×IRnIRng:{\rm I\kern-1.99997ptR}^{k}\times{\rm I\kern-1.99997ptR}^{m}\times{\rm I\kern-1.99997ptR}^{n}\to{\rm I\kern-1.99997ptR}^{n} is assumed to be bounded

sup(u,y,z)IRk×IRm×IRng(u,y,z):=Mg<\sup_{(u,y,z)\in{\rm I\kern-1.39998ptR}^{k}\times{\rm I\kern-1.39998ptR}^{m}\times{\rm I\kern-1.39998ptR}^{n}}||g(u,y,z)||:=M_{g}<\infty (4)

and satisfying the Lipschitz condition in (y,z)(y,z) uniformly with respect to uIRku\in{\rm I\kern-1.99997ptR}^{k}.

If the parameter ϵ\epsilon was small, the system (1)-(2) would belong to the class of the so-called singularly perturbed (SP) systems. SP systems are characterized by the decomposition of the state variables into groups that change their values with rates of different orders of magnitude (so that some of them can be considered as fast/slow with respect to others), which is due to the presence of the small singular perturbations parameter ϵ\epsilon. Such systems describe processes and interactions in disparate time scales, and they have been extensively studied in the literature (see, e.g., research monographs [5], [11], [12], [13] and surveys [6], [14], [15], [18]).

One of the approaches to SP control systems is the averaging method based on the analysis of asymptotic properties of the sets of time averages of the equation describing the dynamics of the slow state variables over the fast control-state trajectories considered on the intervals [0,S][0,S] for large values of S>0S>0. If the limit of these sets exists as SS tends to infinity, then (under certain additional conditions) it defines the right-hand-side of the differential inclusion, the solutions of which approximate the slow components of the solutions of the SP system when the singular perturbation parameter ϵ\epsilon tends to zero (see [7], [8], [9], [10], [17] and also [1], [2], [3] for related developments).

In this paper, the parameter ϵ\epsilon is not assumed to be small, and to distinguish it from the case of singular perturbations, we call system (1)-(2) non-singularly perturbed. We show that the approximation of the zz-components of the solution of system (1)-(2) by the solution of a certain differential inclusion is still valid, but this result is no longer asymptotic and holds for any ϵ>0\epsilon>0. Such an approximation is possible due to the fact that the yy-components of the state variables of the system (1)-(2) can change their values arbitrarily fast (see Lemma 1 in the next section) while the rates of change of the zz-components are bounded by the constant MgM_{g} (see (4)).

To state our main result, let us introduce the differential inclusion

dz(t)dtV(z(t)),z(0)=z0,\frac{dz(t)}{dt}\in V(z(t)),\ \ \ \ z(0)=z_{0}, (5)

where

V(z):=co¯(g(IRk,IRm,z)),g(IRk,IRm,z):={v|v=g(u,y,z),uIRk,yIRm},V(z):=\bar{\rm co}(g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m},z)),\ \ \ \ g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m},z):=\{v\ |\ v=g(u,y,z),\ u\in{\rm I\kern-1.99997ptR}^{k},\ y\in{\rm I\kern-1.99997ptR}^{m}\}, (6)

with co¯\bar{co} standing for the closure of the convex hull of the corresponding set.

Let TT be an arbitrary positive number. Denote by 𝒵T(ϵ)\mathcal{Z}_{T}(\epsilon) the set of the zz-components of solutions of the system (1)-(2) considered on the interval [0,T][0,T], that is,

𝒵T(ϵ):={z()|(u(),y(),z()) satisfies (1),(2)}.\mathcal{Z}_{T}(\epsilon):=\{z(\cdot)\,|\,(u(\cdot),y(\cdot),z(\cdot))\hbox{ satisfies }(\ref{e:fast}),(\ref{e:slow})\}.

By 𝒵T\mathcal{Z}_{T} denote the set of solutions of the differential inclusion (5) considered on this interval. Note that, as can be readily understood,

𝒵T(ϵ)𝒵Tϵ>0.\mathcal{Z}_{T}(\epsilon)\subset\mathcal{Z}_{T}\ \ \ \ \forall\ \epsilon>0. (7)

The main result of the paper is the following theorem.

Theorem 1.

The equality

cl(𝒵T(ϵ))=𝒵Tϵ>0{\rm cl}(\mathcal{Z}_{T}(\epsilon))=\mathcal{Z}_{T}\ \ \ \ \forall\ \epsilon>0 (8)

is valid, where cl{\rm cl} in the expression above stands for the closure in the uniform convergence metrics. That is, the set of the zz-components of solutions of the system (1)-(2) is dense in the set of solutions of the differential inclusion (5).

Proof.

The proof of the theorem is given in the next section. ∎

REMARK I. Theorem 1 resembles results establishing that the Hausdorff distance (induced by the uniform convergence metric) between the set of the slow components of solutions of a SP control system and the set of solutions of the differential inclusion, constructed by averaging of the slow subsystem over the controls and the corresponding solutions of the fast one, tends to zero when the singular perturbations parameter tends to zero (see [7], [8], [9], [10], [17]). As mentioned above, in contrast to these results, Theorem 1 is not of asymptotic nature. The equality (8) is valid for any ϵ>0\epsilon>0, including, e.g., ϵ=1\epsilon=1.

REMARK II. Note that the statement of Theorem 1 would look similar to that of Filippov-Wazewski theorem (see [4]) if y(t)y(t) in (2) was another control (that is, an arbitrary measurable function taking values in IRm{\rm I\kern-1.99997ptR}^{m}) instead of being the solution of (1).

Let us discuss an implication of Theorem 1 for optimal control. Consider the optimal control problem

infu()G(zϵ(T)):=Gϵ,\inf_{u(\cdot)}G(z_{\epsilon}(T)):=G_{\epsilon}^{*}, (9)

where G()G(\cdot) is a continuous function and infinf is taken over the controls u()u(\cdot) and the corresponding solutions (yϵ(),zϵ())(y_{\epsilon}(\cdot),z_{\epsilon}(\cdot)) of system (1)-(2). Consider also the optimal control problem

infu(),y()G(z(T)):=G,\inf_{u(\cdot),y(\cdot)}G(z(T)):=G^{*}, (10)

where infinf is taken over measurable functions (u(),y())IRk×IRm(u(\cdot),y(\cdot))\in{\rm I\kern-1.99997ptR}^{k}\times{\rm I\kern-1.99997ptR}^{m} and the corresponding solutions z()z(\cdot) of the system

dz(t)dt=g(u(t),y(t),z(t)),z(0)=z0\frac{dz(t)}{dt}=g(u(t),y(t),z(t)),\ \ \ \ z(0)=z_{0} (11)

(both u()u(\cdot) and y()y(\cdot) are playing the role of controls in this system).

Corollary 1.

The optimal value in (9) is equal to the optimal value in (10):

Gϵ=Gϵ>0.G_{\epsilon}^{*}=G^{*}\ \ \ \ \forall\ \epsilon>0. (12)
Proof.

Denote by 𝒵T0\mathcal{Z}_{T}^{0} the set of solutions of system (11) considered on the interval [0,T][0,T]. As can be readily seen, the following inclusions are valid:

𝒵T(ϵ)𝒵T0𝒵T.\mathcal{Z}_{T}(\epsilon)\subset\mathcal{Z}_{T}^{0}\subset\mathcal{Z}_{T}.

Therefore, by (8),

cl(𝒵T(ϵ))=cl(𝒵T0)=𝒵T.{\rm cl}(\mathcal{Z}_{T}(\epsilon))={\rm cl}(\mathcal{Z}_{T}^{0})=\mathcal{Z}_{T}.

The latter implies (12). ∎

2 Proof of the Main Result

Consider the system

dy(τ)dτ=Ay(τ)+Bu(τ).\frac{dy(\tau)}{d\tau}=Ay(\tau)+Bu(\tau). (13)

Note that this system looks similar to the fast subsystem (1) but, in contrast to the latter, it evolves in the time scale τ=tϵ\tau=\frac{t}{\epsilon} ((13) will be referred to as the associated system). The following lemma is the key element of the subsequent analysis.

Lemma 1.

For any y,y′′IRmy^{\prime},y^{\prime\prime}\in{\rm I\kern-1.99997ptR}^{m} there exists a control u()u(\cdot) that steers the associated system (13) from yy^{\prime} to y′′y^{\prime\prime} in arbitrarily short period of time.

Proof.

The proof of the lemma follows a standard argument, and we give it at the end of this section. ∎

Let y(τ,u(),y)y(\tau,u(\cdot),y) stand for the solution of the associated system (13) obtained with a control u(τ)u(\tau) and the initial condition y(0)=yy(0)=y. For a fixed zz, denote by V(S,y,z)V(S,y,z) the set of the time averages:

V(S,y,z):=u()1S0Sg(u(τ),y(τ,u(),y),z)𝑑τ,z=const,V(S,y,z):=\bigcup_{u(\cdot)}{\frac{1}{S}\int_{0}^{S}g\bigl{(}u(\tau),y(\tau,u(\cdot),y),z\bigr{)}d\tau},\quad z={\rm const}, (14)

where the union is taken over all controls (that is, over all measurable functions u()u(\cdot) taking values in IRk{\rm I\kern-1.99997ptR}^{k}). The set of the time averages similar to (14) was introduced/used in [7], [8], [9], [10], [17], where it was shown that, as SS tends to infinity, it converges in the Hausdorff metrics to a convex and compact set (provided that the associated system satisfies certain controllability or stability conditions). It was also shown that it is this set that defines the right-hand-side of the differential inclusion, the solutions of which approximate the dynamics of the slow components of the SP control system.

The lemma below establishes that in the case under consideration the set cl(V(S,y0,z)){\rm cl}(V(S,y_{0},z)) (the closure of the set V(S,y0,z)V(S,y_{0},z)) is equal to a convex and compact subset of IRm{\rm I\kern-1.99997ptR}^{m} for any S>0S>0, and it also provides an explicit representation for this set.

Lemma 2.

For any zIRn\ z\in{\rm I\kern-1.99997ptR}^{n}, any yIRmy\in{\rm I\kern-1.99997ptR}^{m}, and any S>0S>0, the following equality is valid:

cl(V(S,y,z))=co¯(g(IRk,IRm,z)).{\rm cl}\,(V(S,y,z))=\bar{\rm co}\,(g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m},z)). (15)
Proof.

Since zz is a constant parameter, we will suppress it in the notation, that is, we will write (14) in the form

V(S,y):=u()1S0Sg(u(τ),y(τ,u(),y))𝑑τ.V(S,y):=\bigcup_{u(\cdot)}{\frac{1}{S}\int_{0}^{S}g\bigl{(}u(\tau),y(\tau,u(\cdot),y)\bigr{)}d\tau}. (16)

The inclusion

V(S,y)co¯(g(IRk,IRm)),V(S,y)\subset\bar{\rm co}\,(g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m})), (17)

follows from the definition of the integral. Indeed, take vV(S,y0)v\in V(S,y_{0}), then there exists u()u(\cdot) such that v=1S0Sg(u(τ),y(τ,u(),y))𝑑τv=\frac{1}{S}\int_{0}^{S}g\bigl{(}u(\tau),y(\tau,u(\cdot),y)\bigr{)}d\tau. If the integrand was a simple function, that is, it was a finite sum of indicator functions of Borel measurable sets, then it would be clear that

1S0Sg(u(τ),y(τ,u(),y))𝑑τco(g(IRk,IRm)),\frac{1}{S}\int_{0}^{S}g\bigl{(}u(\tau),y(\tau,u(\cdot),y)\bigr{)}d\tau\in{\rm co}\,(g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m})),

that is, vcog(IRr,IRm)v\in{\rm co}\,g({\rm I\kern-1.99997ptR}^{r},{\rm I\kern-1.99997ptR}^{m}). For an arbitrary Borel gg, the inclusion vco¯g(IRr,IRm)v\in\bar{\rm co}\,g({\rm I\kern-1.99997ptR}^{r},{\rm I\kern-1.99997ptR}^{m}) follows from the definition of Lebesgue integral as the limit of integrals of simple functions.

Inclusion (17) implies that

cl(V(S,y))co¯(g(IRk,IRm)).{\rm cl}\,(V(S,y))\subset\bar{\rm co}\,(g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m})).

Let us show that

co(g(IRk,IRm))cl(V(S,y)).{\rm co}\,(g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m}))\subset{\rm cl}\,(V(S,y)). (18)

(This will, obviously, imply that co¯(g(IRk,IRm))cl(V(S,y))\bar{\rm co}\,(g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m}))\subset{\rm cl}\,(V(S,y)).) Take vco(g(IRk,IRm))v\in{\rm co}\,(g({\rm I\kern-1.99997ptR}^{k},{\rm I\kern-1.99997ptR}^{m})). Then, due to the Caratheodori Theorem,

v=i=1lλig(ui,yi),lk+m+1v=\sum_{i=1}^{l}\lambda_{i}g(u_{i},y_{i}),\ \ \ \ l\leq k+m+1

for some (ui,yi)IRk×IRm(u_{i},y_{i})\in{\rm I\kern-1.99997ptR}^{k}\times{\rm I\kern-1.99997ptR}^{m} and some λi>0\lambda_{i}>0 with i=1lλi=1\ \sum_{i=1}^{l}\lambda_{i}=1. Let

S0=0,Sj:=Si=1jλi,j=1,,l,S_{0}=0,\;\ \ \ S_{j}:=S\sum_{i=1}^{j}\lambda_{i},\ \ j=1,\dots,l,\;

which implies that

λj=SjSj1S.\lambda_{j}=\frac{S_{j}-S_{j-1}}{S}. (19)

Let δ>0\delta>0 be arbitrary small and let τ^>0\hat{\tau}>0 be defined by the equation

τ^:=δMg,\hat{\tau}:=\frac{\delta}{M_{g}}, (20)

where MgM_{g} is defined in (4). Note that from (20) it follows that the solution y(τ,u,y)y(\tau,u,y) of the associated system (13) obtained with the constant valued control u(τ)uu(\tau)\equiv u and with the initial condition y(0)=yy(0)=y satisfies the inequality

maxτ[0,τ^]y(τ,u,y)yδ(u,y)IRk×IRm.\max_{\tau\in[0,\hat{\tau}]}\|y(\tau,u,y)-y\|\leq\delta\ \ \ \ \forall\ (u,y)\in{\rm I\kern-1.99997ptR}^{k}\times{\rm I\kern-1.99997ptR}^{m}. (21)

Let us construct a control-state process (u(),y())(u(\cdot),y(\cdot)) of the system (13) on the interval [0,S][0,S] such that, for all j=1,,lj=1,\dots,l, the following inequalities hold true (for brevity, the solution of (13) is denoted below as y(τ)y(\tau) instead of y(τ,u(),y)y(\tau,u(\cdot),y)):

|1SjSj1Sj1Sjg(u(τ),y(τ))𝑑τg(uj,yj)|Kδ\left|\frac{1}{S_{j}-S_{j-1}}\int_{S_{j-1}}^{S_{j}}g(u(\tau),y(\tau))\,d\tau-g(u_{j},y_{j})\right|\leq K\delta

for some constant KK (recall that S0:=0S_{0}:=0 and Sl=SS_{l}=S).

Consider the interval [S0,S1][S_{0},S_{1}]. Assuming (without loss of generality) that δ<S1\delta<S_{1}, define the control u()u(\cdot) on the interval [0,δ][0,S1][0,\delta]\subset[0,S_{1}] in such a way that y(δ)=y1y(\delta)=y_{1}. (This is possible due to Lemma 1.) Set τ11:=δ\tau_{11}:=\delta and τ12:=δ+τ^.\tau_{12}:=\delta+\hat{\tau}. Extend the definition of the control u()u(\cdot) by taking it to be equal to u1u_{1} on the interval (τ11,τ12](\tau_{11},\tau_{12}] if τ12<S1\tau_{12}<S_{1}. In case τ12S1\tau_{12}\geq S_{1}, take u()u(\cdot) to be equal to u1u_{1} on the interval (τ11,S1](\tau_{11},S_{1}]. The control u()u(\cdot) will be, thus, defined on [S0,S1[S_{0},S_{1}]. Note that, by (21),

maxτ[τ11,τ12]y(τ)y1δifτ12<S1\max_{\tau\in[\tau_{11},\tau_{12}]}||y(\tau)-y_{1}||\leq\delta\ \ {\rm if}\ \ \tau_{12}<S_{1}

and

maxτ[τ11,S1]y(τ)y1δifτ12S1.\max_{\tau\in[\tau_{11},S_{1}]}||y(\tau)-y_{1}||\leq\delta\ \ {\rm if}\ \ \tau_{12}\geq S_{1}.

If τ12<S1\tau_{12}<S_{1}, set τ13:=min{τ12+δ2,τ12+S1τ122}\tau_{13}:=\min\{\tau_{12}+\frac{\delta}{2},\tau_{12}+\frac{S_{1}-\tau_{12}}{2}\} and τ14:=τ13+τ^\tau_{14}:=\tau_{13}+\hat{\tau}. Extend the definition of the control u()u(\cdot) to the interval (τ12,τ13](\tau_{12},\tau_{13}] in such a way that the corresponding solution y(τ)y(\tau) of the system (13) satisfies the equation y(τ13)=y1y(\tau_{13})=y_{1}. (Again, this is possible due to Lemma 1.) Also, extend the definition of the control u()u(\cdot) further by taking it to be equal to u1u_{1} on the interval (τ13,τ14](\tau_{13},\tau_{14}] if τ14<S1\tau_{14}<S_{1}. In case τ14S1\tau_{14}\geq S_{1}, take u()u(\cdot) to be equal to u1u_{1} on the interval (τ13,S1](\tau_{13},S_{1}]. The control u()u(\cdot) will be, thus, defined on [S0,S1[S_{0},S_{1}]. By (21),

maxτ[τ11,τ12][τ13,τ14]y(τ)y1δifτ14<S1\max_{\tau\in[\tau_{11},\tau_{12}]\cup[\tau_{13},\tau_{14}]}||y(\tau)-y_{1}||\leq\delta\ \ {\rm if}\ \ \tau_{14}<S_{1}

and

maxτ[τ11,τ12][τ13,S1]y(τ)y1δifτ14S1.\max_{\tau\in[\tau_{11},\tau_{12}]\cup[\tau_{13},S_{1}]}||y(\tau)-y_{1}||\leq\delta\ \ {\rm if}\ \ \tau_{14}\geq S_{1}.

In the general case (for an arbitrary small δ\delta), one can proceed in a similar way to:

(i) Define the sequence of moments of time τ11,τ12,,τ1s¯,τ1s¯+1\tau_{11},\tau_{12},\cdots,\tau_{1\bar{s}},\tau_{1\bar{s}+1}, where s¯\bar{s} is odd, τs¯+1+τ¯S1\tau_{\bar{s}+1}+\bar{\tau}\geq S_{1}, and, for any odd ss such that 3s<s¯3\leq s<\bar{s},

τ1s:=min{τ1s1+δ2s2,τ1s1+S1τ1s12},τ1s+1:=τ1s+τ^<S1;\tau_{1s}:=\min\{\tau_{1s-1}+\frac{\delta}{2^{s-2}},\tau_{1s-1}+\frac{S_{1}-\tau_{1s-1}}{2}\},\ \ \ \ \tau_{1s+1}:=\tau_{1s}+\hat{\tau}<S_{1}; (22)

and

(ii) Construct the control u()u(\cdot) that along with the corresponding solution y()y(\cdot) of (13) satisfy the relations

u(τ)u1,y(τ)y1δτD:=(τ11,τ12](τ13,τ14](τ1s¯2,τ1s¯1](τ1s¯,S1],u(\tau)\equiv u_{1},\;||y(\tau)-y_{1}||\leq\delta\;\forall\tau\in D:=(\tau_{11},\tau_{12}]\cup(\tau_{13},\tau_{14}]\cup\ldots\cup(\tau_{1\bar{s}-2},\tau_{1\bar{s}-1}]\cup(\tau_{1\bar{s}},S_{1}], (23)

with y(τ1,s)=y1y(\tau_{1,s})=y_{1} for all s=1,3,,s¯s=1,3,\ldots,\bar{s}.

By construction (see (22) and (23)), the Lebesgue measure of the set [0,S1]D[0,S_{1}]\setminus D (which we denote as meas([0,S1]D)meas([0,S_{1}]\setminus D) satisfies the inequality

meas([0,S1]D)=τ11+(τ13τ12)++(τ1s¯τ1s¯1)δ+δ2++δ2s¯22δ.meas([0,S_{1}]\setminus D)=\tau_{11}+(\tau_{13}-\tau_{12})+\cdots+(\tau_{1\bar{s}}-\tau_{1\bar{s}-1})\leq\delta+\frac{\delta}{2}+\cdots+\frac{\delta}{2^{\bar{s}-2}}\leq 2\delta.

Therefore,

|1S1[0,S1]D(g(u(τ),y(τ))g(u1,y1))𝑑τ|2MgS1meas([0,S1]D)4MgS1δ,\left|\frac{1}{S_{1}}\int_{[0,S_{1}]\setminus D}(g(u(\tau),y(\tau))-g(u_{1},y_{1}))\,d\tau\right|\leq\frac{2M_{g}}{S_{1}}meas([0,S_{1}]\setminus D)\leq\frac{4M_{g}}{S_{1}}\delta,

where MgM_{g} is as in (4). Also, due to (23),

|1S1D(g(u(τ),y(τ))g(u1,y1))𝑑τ|=|1S1Dg(u1,y(τ))𝑑τg(u1,y1)|Lδ,\left|\frac{1}{S_{1}}\int_{D}\big{(}g(u(\tau),y(\tau))\,-g(u_{1},y_{1})\big{)}d\tau\right|=\left|\frac{1}{S_{1}}\int_{D}g(u_{1},y(\tau))\,d\tau-g(u_{1},y_{1})\right|\leq L\delta,

where LL is a Lipschitz constant of gg. Consequently,

|1S10S1g(u(τ),y(τ))𝑑τg(u1,y1)|=|1S10S1(g(u(τ),y(τ))g(u1,y1))𝑑τ|\displaystyle\left|\frac{1}{S_{1}}\int_{0}^{S_{1}}g(u(\tau),y(\tau))\,d\tau-g(u_{1},y_{1})\right|=\left|\frac{1}{S_{1}}\int_{0}^{S_{1}}(g(u(\tau),y(\tau))-g(u_{1},y_{1}))\,d\tau\right|\leq
|1S1[0,S1]D(g(u(τ),y(τ))g(u1,y1))𝑑τ|+|1S1D(g(u(τ),y(τ))g(u1,y1))𝑑τ|\displaystyle\left|\frac{1}{S_{1}}\int_{[0,S_{1}]\setminus D}(g(u(\tau),y(\tau))-g(u_{1},y_{1}))\,d\tau\right|+\left|\frac{1}{S_{1}}\int_{D}(g(u(\tau),y(\tau))-g(u_{1},y_{1}))\,d\tau\right|\leq
(4MgS1+L)δ.\displaystyle\left(\frac{4M_{g}}{S_{1}}+L\right)\delta.

Continuing similarly on each subinterval [Sj1,Sj][S_{j-1},S_{j}], we construct the control-state process (u(),y())(u(\cdot),y(\cdot)) on [0,S][0,S] such that for all j=1,,lj=1,\dots,l

|1SjSj1Sj1Sjg(u(τ),y(τ))𝑑τg(uj,yj)|(4MgSjSj1+L)δ.\left|\frac{1}{S_{j}-S_{j-1}}\int_{S_{j-1}}^{S_{j}}g(u(\tau),y(\tau))\,d\tau-g(u_{j},y_{j})\right|\leq\left(\frac{4M_{g}}{S_{j}-S_{j-1}}+L\right)\delta. (24)

For this process, taking into account (19) and (24), we have

|1S0Sg(u(τ),y(τ))𝑑τv|\displaystyle\left|\frac{1}{S}\int_{0}^{S}g(u(\tau),y(\tau))\,d\tau-v\right|
=\displaystyle= |i=1lSiSi1S1SiSi1Si1Sig(u(τ),y(τ))𝑑τi=1lλig(ui,yi)|\displaystyle\left|\sum_{i=1}^{l}\frac{S_{i}-S_{i-1}}{S}\frac{1}{S_{i}-S_{i-1}}\int_{S_{i-1}}^{S_{i}}g(u(\tau),y(\tau))\,d\tau-\sum_{i=1}^{l}\lambda_{i}g(u_{i},y_{i})\right|
=\displaystyle= i=1lλi|1SiSi1Si1Sig(u(τ),y(τ))𝑑τg(ui,yi)|\displaystyle\sum_{i=1}^{l}\lambda_{i}\left|\frac{1}{S_{i}-S_{i-1}}\int_{S_{i-1}}^{S_{i}}g(u(\tau),y(\tau))\,d\tau-g(u_{i},y_{i})\right|
\displaystyle\leq δi=1lλi(4MgSiSi1+L)δ(4Mgmini{SiSi1}+L).\displaystyle\,\delta\sum_{i=1}^{l}\lambda_{i}\left(\frac{4M_{g}}{S_{i}-S_{i-1}}+L\right)\leq\delta\left(\frac{4M_{g}}{\min_{i}\{S_{i}-S_{i-1}\}}+L\right).

Since δ\delta is arbitrarily small, this implies that vclV(S,y)v\in{\rm cl}\,V(S,y), which implies (18). The lemma is proved. ∎

Lemma 3.

The multivalued function V(z)\ V(z) defined in accordance with (6) is Lipschitz continuous. That is,

dH(V(z),V(z′′))Lzz′′z,z′′,d_{H}(V(z^{\prime}),V(z^{\prime\prime}))\leq L||z^{\prime}-z^{\prime\prime}||\ \ \ \ \forall z^{\prime},z^{\prime\prime}, (25)

where dHd_{H} stands for the Hausdorff distance between sets and LL is the Lipschitz constant of g(u,y,z)g(u,y,z) in zz.

Proof.

Let z,z′′z^{\prime},z^{\prime\prime} be arbitrary elements of IRn{\rm I\kern-1.99997ptR}^{n}. Take some S>0S>0 and choose an arbitrary element v\ v^{\prime} from V(z,S,y0)\ V(z^{\prime},S,y_{0}). From the definition of V(z,S,y0)\ V(z^{\prime},S,y_{0})\ it follows that there exists a control u()\ u(\cdot) such that v=1S0Sg(z,u(τ),y(τ))𝑑τ.\ v^{\prime}=\frac{1}{S}\int_{0}^{S}g(z^{\prime},u(\tau),y(\tau))d\tau. Define v′′\ v^{\prime\prime} by

v′′:=1S0Sg(z′′,u(τ),y(τ))𝑑τV(z′′,S,y0).v^{\prime\prime}:=\frac{1}{S}\int_{0}^{S}g(z^{\prime\prime},u(\tau),y(\tau))d\tau\in V(z^{\prime\prime},S,y_{0}).

We have

vv′′1S0Sg(z,u(τ),y(τ))g(z′′,u(τ),y(τ))𝑑τ||v^{\prime}-v^{\prime\prime}||\leq\frac{1}{S}\int_{0}^{S}||g(z^{\prime},u(\tau),y(\tau))-g(z^{\prime\prime},u(\tau),y(\tau))||d\tau
Lzz′′d(v,V(z′′))Lzz′′,\leq L||z^{\prime}-z^{\prime\prime}||\ \ \ \ \Rightarrow\ \ \ \ \ d(v^{\prime},V(z^{\prime\prime}))\leq L||z^{\prime}-z^{\prime\prime}||,

where d(v,V)d(v,V) stands for the distance from a vector vv to a set VV (d(v,V):=infwVvwd(v,V):=\inf_{w\in V}||v-w||). Since v\ v^{\prime} is an arbitrary element of V(z,S,y0)\ V(z^{\prime},S,y_{0}), it implies that

supvV(z,S,y0)d(v,V(z′′,S,y0))Lzz′′.\ \sup_{v\in V(z^{\prime},S,y_{0})}d(v,V(z^{\prime\prime},S,y_{0}))\leq L||z^{\prime}-z^{\prime\prime}||.

Since V(z)=clV(S,y0,z)V(z)={\rm cl}\,V(S,y_{0},z), we conclude that

supvV(z)d(v,V(z′′))Lzz′′.\sup_{v\in V(z^{\prime})}d(v,V(z^{\prime\prime}))\leq L||z^{\prime}-z^{\prime\prime}||.

Similarly, it is established that   supvV(z′′)d(v,V(z))Lzz′′.\ \sup_{v\in V(z^{\prime\prime})}d(v,V(z^{\prime}))\leq L||z^{\prime}-z^{\prime\prime}||.   ∎

Proof of Theorem 1. Due to (7), to prove the theorem it is sufficient to show that

𝒵Tcl(𝒵T(ϵ))ϵ>0.\mathcal{Z}_{T}\subset{\rm cl}(\mathcal{Z}_{T}(\epsilon))\ \ \ \ \forall\ \epsilon>0.

This, in turn, will be shown, if we establish that, for any δ>0\delta>0, corresponding to any solution z(t)z(t) of the differential inclusion (5), there exists a control uϵ(t)u_{\epsilon}(t), which, being used in the system (1)-(2), generates the solution (yϵ(t),zϵ(t))(y_{\epsilon}(t),z_{\epsilon}(t)) that satisfies the inequality

maxt[0,T]zϵ(t)z(t)δ.\max_{t\in[0,T]}||z_{\epsilon}(t)-z(t)||\leq\delta. (26)

To prove the latter statement, take an arbitrary solution z(t)z(t) of (5), choose an arbitrary δ>0\delta>0 and construct the control uϵ(t)u_{\epsilon}(t) that insures the validity of (26).

Take S(0,Tϵ]S\in(0,\frac{T}{\epsilon}] and partition the interval [0,T]\ [0,T]\ by the points

tl:=l(ϵS),l=0,1,,NϵS:=TϵS,t_{l}:=l\left(\epsilon S\right)\ ,\ \ \ \ l=0,1,...,N_{\epsilon S}:=\left\lfloor\frac{T}{\epsilon S}\right\rfloor, (27)

where \lfloor\cdot\rfloor is the floor function (that is, for any real xx, x\lfloor x\rfloor is the maximal integer that is less or equal than xx).

Let v0v_{0} be the projection of the vector (ϵS)10t1dz(t)dt𝑑t\ \left(\epsilon S\right)^{-1}\int_{0}^{t_{1}}\frac{dz(t)}{dt}dt\ onto the set V(z0)V(z_{0}). That is,

v0:=argminvV(z0){(ϵS)10t1dz(t)dt𝑑tv}.v_{0}:={\rm argmin}_{v\in V\left(z_{0}\right)}\left\{\left\|\left(\epsilon S\right)^{-1}\int_{0}^{t_{1}}\frac{dz(t)}{dt}dt-v\right\|\right\}.

By Lemma 2, clV(S,y0,z0)=V(z0){\rm cl}\,V(S,y_{0},z_{0})=V(z_{0}). Therefore, there exists a control u0(τ)u_{0}(\tau) such that, being used in the associated system (13) on the interval [0,t1ϵ],\ [0,\frac{t_{1}}{\epsilon}],\ ensures that

v01S0t1ϵg(u0(τ),y0(τ),z0)𝑑τϵS,\left\|v_{0}-\frac{1}{S}\int_{0}^{\frac{t_{1}}{\epsilon}}g(u_{0}(\tau),y_{0}(\tau),z_{0})d\tau\right\|\leq\epsilon S,

where y0(τ):=y(τ,u0(),y0)y_{0}(\tau):=y(\tau,u_{0}(\cdot),y_{0}).

Define the control uϵ(t)u_{\epsilon}(t) on the interval [0,t1)[0,t_{1}) by the equation

uϵ(t):=u0(tϵ)t[0,t1)u_{\epsilon}(t):=u_{0}\left(\frac{t}{\epsilon}\right)\ \ \forall\ t\in[0,t_{1})

and show how this definition can be extended to the interval [0,T][0,T].

Assume that the control uϵ(t)u_{\epsilon}(t) has been defined on the interval [0,tl)[0,t_{l}) and denote by (yϵ(t),zϵ(t))(y_{\epsilon}(t),z_{\epsilon}(t)) the corresponding solution of the system (1)-(2) on this interval. Extend the definition of uϵ(t)u_{\epsilon}(t) to the interval [0,tl+1)[0,t_{l+1}) by following the steps:

(a) Define vlv_{l}\ as the projection of the vector (ϵS)1tltl+1dz(t)dt𝑑t\ \left(\epsilon S\right)^{-1}\int_{t_{l}}^{t_{l+1}}\frac{dz(t)}{dt}dt\ onto the set V(zϵ(tl))V(z_{\epsilon}(t_{l})). That is,

vl:=argminvV(zϵ(tl)){(ϵS)1tltl+1dz(t)dt𝑑tv}.v_{l}:={\rm argmin}_{v\in V(z_{\epsilon}(t_{l}))}\left\{\left\|\left(\epsilon S\right)^{-1}\int_{t_{l}}^{t_{l+1}}\frac{dz(t)}{dt}dt-v\right\|\right\}. (28)

(b) Define a control ul(τ)u_{l}(\tau) such that, being used in the associated system (13) on the interval [tlϵ,tl+1ϵ],\ [\frac{t_{l}}{\epsilon},\frac{t_{l+1}}{\epsilon}],\ ensures that

vl1Stlϵtl+1ϵg(ul(τ),yl(τ),zϵ(tl))𝑑τϵS,\left\|v_{l}-\frac{1}{S}\int_{\frac{t_{l}}{\epsilon}}^{\frac{t_{l+1}}{\epsilon}}g(u_{l}(\tau),y_{l}(\tau),z_{\epsilon}(t_{l}))d\tau\right\|\leq\epsilon S, (29)

where yl(τ)y_{l}(\tau) is the solution of the associated system (13) obtained with the control ul(τ)u_{l}(\tau) and with the initial condition yl(tlϵ)=yϵ(tl)y_{l}(\frac{t_{l}}{\epsilon})=y_{\epsilon}(t_{l}). The existence of such control follows the fact that, by Lemma 2, clV(S,yϵ(tl),zϵ(tl))=V(zϵ(tl)){\rm cl}\,V(S,y_{\epsilon}(t_{l}),z_{\epsilon}(t_{l}))=V(z_{\epsilon}(t_{l})).

(c) Define the control uϵ(t)u_{\epsilon}(t) on the interval [tl,tl+1)[t_{l},t_{l+1}) as equal to ul(tϵ)u_{l}(\frac{t}{\epsilon}). That is,

uϵ(t):=ul(tϵ)t[tl,tl+1).\ u_{\epsilon}(t):=u_{l}\left(\frac{t}{\epsilon}\right)\ \ \forall\ t\in[t_{l},t_{l+1}). (30)

Proceeding in a similar way, one can extend the definition of the control uϵ(t)u_{\epsilon}(t) to the interval [0,tNϵS)[0,t_{N_{\epsilon S}}). On the interval [tNϵS,T][t_{N_{\epsilon S}},T], the control uϵ(t)u_{\epsilon}(t) can be taken to be equal to an arbitrary element uu. Denote by (yϵ(t),zϵ(t))(y_{\epsilon}(t),z_{\epsilon}(t)) the corresponding solution of the system (1)-(2) on the interval [0,T][0,T].

Denote

(y¯ϵ(τ),z¯ϵ(τ)):=(yϵ(ϵτ),zϵ(ϵτ))τ[0,Tϵ].(\bar{y}_{\epsilon}(\tau),\bar{z}_{\epsilon}(\tau)):=(y_{\epsilon}(\epsilon\tau),z_{\epsilon}(\epsilon\tau))\ \ \ \ \forall\ \tau\in[0,\frac{T}{\epsilon}].

Then

(y¯ϵ(τl),z¯ϵ(τl))=(yϵ(tl),zϵ(tl))l=0,1,,Nϵ,(\bar{y}_{\epsilon}(\tau_{l}),\bar{z}_{\epsilon}(\tau_{l}))=(y_{\epsilon}(t_{l}),z_{\epsilon}(t_{l}))\ \ \ \ \forall\ l=0,1,...,N_{\epsilon},

where τl=tlϵ=lS\ \tau_{l}=\frac{t_{l}}{\epsilon}=lS (see (27)), and

maxτ[τl,τl+1]z¯ϵ(τ)z¯ϵ(τl)=maxt[tl,tl+1]zϵ(t)zϵ(tl)MgϵS,\max_{\tau\in[\tau_{l},\tau_{l+1}]}||\bar{z}_{\epsilon}(\tau)-\bar{z}_{\epsilon}(\tau_{l})||=\max_{t\in[t_{l},t_{l+1}]}||z_{\epsilon}(t)-z_{\epsilon}(t_{l})||\leq M_{g}\epsilon S, (31)

where MgM_{g} is defined in (4). Also notice that

maxt[tl,tl+1]z(t)z(tl)=maxτ[τl,τl+1]z(ϵτ)z(tl)MgϵS,\max_{t\in[t_{l},t_{l+1}]}||z(t)-z(t_{l})||=\max_{\tau\in[\tau_{l},\tau_{l+1}]}||z(\epsilon\tau)-z(t_{l})||\leq M_{g}\epsilon S, (32)

since

max{v|vV(z),zZ}Mg.\max\{||v||\ |\ v\in V(z),\ z\in Z\}\leq M_{g}.

Let us verify that thus defined control ensures the validity of (26). Subtracting the equation

z(tl+1)=z(tl)+tltl+1dz(t)dt𝑑tz(t_{l+1})=z(t_{l})+\int_{t_{l}}^{t_{l+1}}\frac{dz(t)}{dt}dt

from the equation

zϵ(tl+1)=zϵ(tl)+tltl+1g(uϵ(t),yϵ(t),zϵ(t))𝑑t,z_{\epsilon}(t_{l+1})=z_{\epsilon}(t_{l})+\int_{t_{l}}^{t_{l+1}}g(u_{\epsilon}(t),y_{\epsilon}(t),z_{\epsilon}(t))dt,

one obtains, taking into account (30), that

zϵ(tl+1)z(tl+1)zϵ(tl)z(tl)\displaystyle||z_{\epsilon}(t_{l+1})-z(t_{l+1})||\leq||z_{\epsilon}(t_{l})-z(t_{l})|| (33)
+ϵτlτl+1g(ul(τ),y¯ϵ(τ),z¯ϵ(τ))g(ul(τ),y¯ϵ(τ),z¯ϵ(τl))𝑑τ\displaystyle+\epsilon\int_{\tau_{l}}^{\tau_{l+1}}||g(u_{l}(\tau),\bar{y}_{\epsilon}(\tau),\bar{z}_{\epsilon}(\tau))-g(u_{l}(\tau),\bar{y}_{\epsilon}(\tau),\bar{z}_{\epsilon}(\tau_{l}))||d\tau
+ϵS1Sτlτl+1g(ul(τ),y¯ϵ(τ),z¯ϵ(τl))𝑑τvl+ϵSvl1ϵStltl+1dz(t)dt𝑑t.\displaystyle+\epsilon S\left\|\frac{1}{S}\int_{\tau_{l}}^{\tau_{l+1}}g(u_{l}(\tau),\bar{y}_{\epsilon}(\tau),\bar{z}_{\epsilon}(\tau_{l}))d\tau-v_{l}\right\|+\epsilon S\left\|v_{l}-\frac{1}{\epsilon S}\int_{t_{l}}^{t_{l+1}}\frac{dz(t)}{dt}dt\right\|.

From Lipschitz continuity of g(u,y,z)g(u,y,z) in zz and (31) we obtain

ϵτlτl+1g(ul(τ),y¯ϵ(τ),z¯ϵ(τ))g(ul(τ),y¯ϵ(τ),z¯ϵ(τl))𝑑τϵSL(MgϵS).\epsilon\int_{\tau_{l}}^{\tau_{l+1}}||g(u_{l}(\tau),\bar{y}_{\epsilon}(\tau),\bar{z}_{\epsilon}(\tau))-g(u_{l}(\tau),\bar{y}_{\epsilon}(\tau),\bar{z}_{\epsilon}(\tau_{l}))||d\tau\leq\epsilon SL\left(M_{g}\epsilon S\right). (34)

Due to (25) and (32),

dz(t)dtV(z(t))V(z(tl))+Lz(t)z(tl)B¯n\displaystyle\frac{dz(t)}{dt}\in V(z(t))\subset V(z(t_{l}))+L||z(t)-z(t_{l})||\bar{B}^{n}
V(zϵ(tl))+L(z(t)z(tl)+z(tl)zϵ(tl))B¯n\displaystyle\subset V(z_{\epsilon}(t_{l}))+L(||z(t)-z(t_{l})||+||z(t_{l})-z_{\epsilon}(t_{l})||)\bar{B}^{n}
V(zϵ(tl))+(LMgϵS+Lz(tl)zϵ(tl))B¯nt[tl,tl+1],\displaystyle\subset V(z_{\epsilon}(t_{l}))+(LM_{g}\epsilon S+L||z(t_{l})-z_{\epsilon}(t_{l})||)\bar{B}^{n}\ \ \ \forall\ t\in[t_{l},t_{l+1}],

where B¯n\bar{B}^{n} is the closed unit ball in RnR^{n}. Consequently,

1ϵStltl+1dz(t)dt𝑑tV(zϵ(tl))+(LMϵS+Lz(tl)zϵ(tl))B¯n,\frac{1}{\epsilon S}\int_{t_{l}}^{t_{l+1}}\frac{dz(t)}{dt}dt\in V(z_{\epsilon}(t_{l}))+(LM\epsilon S+L||z(t_{l})-z_{\epsilon}(t_{l})||)\bar{B}^{n},

and, by (28),

1ϵStltl+1dz(t)dt𝑑tvl=d(1ϵStltl+1dz(t)dt𝑑t,V(zϵ(tl)))\displaystyle\left\|\frac{1}{\epsilon S}\int_{t_{l}}^{t_{l+1}}\frac{dz(t)}{dt}dt-v_{l}\right\|=d\left(\frac{1}{\epsilon S}\int_{t_{l}}^{t_{l+1}}\frac{dz(t)}{dt}dt,V(z_{\epsilon}(t_{l}))\right)
LMgϵS+Lz(tl)zϵ(tl).\displaystyle\leq LM_{g}\epsilon S+L||z(t_{l})-z_{\epsilon}(t_{l})||.

Using the latter and (34), (29), one can obtain from (33)

zϵ(tl+1)z(tl+1)zϵ(tl)z(tl)+ϵSL(MgϵS)+(ϵS)2\displaystyle||z_{\epsilon}(t_{l+1})-z(t_{l+1})||\leq||z_{\epsilon}(t_{l})-z(t_{l})||+\epsilon SL\left(M_{g}\epsilon S\right)+(\epsilon S)^{2}
+LMg(ϵS)2+LϵSz(tl)zϵ(tl)=(1+LϵS)z(tl)zϵ(tl)+(ϵS)2(2LMg+1)\displaystyle+LM_{g}(\epsilon S)^{2}+L\epsilon S||z(t_{l})-z_{\epsilon}(t_{l})||=(1+L\epsilon S)||z(t_{l})-z_{\epsilon}(t_{l})||+(\epsilon S)^{2}(2LM_{g}+1)
(1+LTNϵS)z(tl)zϵ(tl)+(TNϵS)(ϵS)(2LMg+1),\displaystyle\leq\left(1+\frac{LT}{N_{\epsilon S}}\right)||z(t_{l})-z_{\epsilon}(t_{l})||+\left(\frac{T}{N_{\epsilon S}}\right)(\epsilon S)(2LM_{g}+1),

where the last inequality follows from the definition of NϵSN_{\epsilon S} (see (27)). Using an argument similar to that of Gronwall’s lemma (see, e.g., Proposition 5.1 in [9]), one can now derive that

zϵ(tl)z(tl)L1eLT(2LMg+1)ϵS,l=0,1,,NϵS.||z_{\epsilon}(t_{l})-z(t_{l})||\leq L^{-1}e^{LT}(2LM_{g}+1)\epsilon S,\ \ \ l=0,1,...,N_{\epsilon S}. (35)

Estimate (35), along with (31) and (32), imply that

zϵ(t)z(t)L1eLT(2LMg+1)ϵS+2MgϵSt[0,T].||z_{\epsilon}(t)-z(t)||\leq L^{-1}e^{LT}(2LM_{g}+1)\epsilon S+2M_{g}\epsilon S\ \ \;\forall t\in[0,T].

The right-hand-side in the expression above tends to zero with SS tending to zero. Therefore, it can be made less or equal than δ\delta if SS is chosen small enough. This proves (26).

Proof of Lemma 1. By Cauchy formula, the lemma will be proved if we show that, for any τ>0\tau>0, there exists a control u()u(\cdot) such that

y′′=eAτy+0τeA(τs)Bu(s)𝑑seAτy′′y=0τeAsBu(s)𝑑s.y^{\prime\prime}=e^{A\tau}y^{\prime}+\int_{0}^{\tau}e^{A(\tau-s)}Bu(s)\,ds\ \ \ \ \Leftrightarrow\ \ \ \ e^{-A\tau}y^{\prime\prime}-y^{\prime}=\int_{0}^{\tau}e^{-As}Bu(s)\,ds. (36)

Take u(s)=BTeATsξu(s)=B^{T}e^{-A^{T}s}\xi and show that there exists ξIRm\xi\in{\rm I\kern-1.99997ptR}^{m} such that these equalities are satisfied. To this end, substitute u(s)=BTeATsξu(s)=B^{T}e^{-A^{T}s}\xi into the second equality in (36) to obtain

eAτy′′y=(0τeAsBBTeATs𝑑s)ξ,e^{-A\tau}y^{\prime\prime}-y^{\prime}=\left(\int_{0}^{\tau}e^{-As}BB^{T}e^{-A^{T}s}\,ds\right)\xi,

therefore,

(0τeAsBBTeATs𝑑s)1(eAτy′′y)=ξ,\left(\int_{0}^{\tau}e^{-As}BB^{T}e^{-A^{T}s}\,ds\right)^{-1}\Big{(}e^{-A\tau}y^{\prime\prime}-y^{\prime}\Big{)}=\xi,

where the latter implication is valid if the matrix (0τeAsBBTeATs𝑑s)\left(\int_{0}^{\tau}e^{-As}BB^{T}e^{-A^{T}s}\,ds\right) is non-singular.

Thus, a sufficient condition for (36) to be valid is non-singularity of the matrix (0τeAsBBTeATs𝑑s)\left(\int_{0}^{\tau}e^{-As}BB^{T}e^{-A^{T}s}\,ds\right). Since this matrix (0τeAsBBTeATs𝑑s)\left(\int_{0}^{\tau}e^{-As}BB^{T}e^{-A^{T}s}\,ds\right) is non-negative definite, to show that it is non-singular, one needs to show that it is positive definite. Assume the contrary, that is, there exists a nonzero vector vv and time τ\tau such that

vT(0τeAsBBTeATs𝑑s)v=0.v^{T}\left(\int_{0}^{\tau}e^{-As}BB^{T}e^{-A^{T}s}\,ds\right)v=0.

This is possible only if

vTeAsBBTeATsv0for all s[0,τ].v^{T}e^{-As}BB^{T}e^{-A^{T}s}v\equiv 0\quad\hbox{for all }s\in[0,\tau].

The latter identity implies that

vTeAsB0for all s[0,τ].v^{T}e^{-As}B\equiv 0\quad\hbox{for all }s\in[0,\tau].

Differentiating this identity with respect to ss jj times, j=0,,m1j=0,\dots,m-1 and plugging in s=0s=0 we get

vTB=0,vTAB=0,,vTAm1B=0,v^{T}B=0,\,v^{T}AB=0,\,\dots,v^{T}A^{m-1}B=0,

or, in a matrix form,

vT[B,AB,,Am1B]=0.v^{T}[B,AB,\dots,A^{m-1}B]=0.

This implies that the rank of the matrix [B,AB,,Am1B][B,AB,\dots,A^{m-1}B] is less than mm, which is a contradiction to our assumption (3).

References

  • [1] Alvarez, O., Bardi, M.: Viscosity Solutions Methods for Singular Perturbations in Deterministic and Stochastic Control. SIAM J. Control and Optimization, 40(4), 1159–1188 (2001)
  • [2] Artstein, Z.: Invariant Measures of Differential Inclusions Applied to Singular Perturbations. J. Differential Equations, 152, 289-307 (1999)
  • [3] Artstein, Z., Gaitsgory, V.: The Value Function of Singularly Perturbed Control Systems. Applied Mathematics and Optimization, 41(3), 425-445 (2000)
  • [4] Aubin, J.-P., Frankowska, H.: Set-Valued Analysis. Birkhauser, Boston (1990)
  • [5] Bensoussan, A.: Perturbation Methods in Optimal Control. John Wiley, New York (1989)
  • [6] Dmitriev, M.G., Kurina, G.A.: Singular Perturbations in Control Problems (Review). Automation and Remote Control, 67(1), 1-43 (2006)
  • [7] Donchev, T.D., Dontchev, A.L.: Singular Perturbations in Infinite-Dimensional Control Systems. SIAM J. Control and Optimization, 42, 1795-1812 (2003)
  • [8] Gaitsgory, V.G.: Use of the averaging method in control problems. Differential Equations (translated from Russian), 22, 1290-1299 (1986)
  • [9] Gaitsgory, V.: Suboptimization of Singularly Perturbed Control Problems. SIAM J. Control and Optimization, 30(5),1228 - 1240 (1992)
  • [10] Grammel, G.: Averaging of Singularly Perturbed Systems. Nonlinear Analysis, 28, 1851-1865 (1997)
  • [11] Kabanov, Y., Pergamenshchikov, S.: Two-Scale Stochastic Systems. Springer Verlag, Berlin-Heidelberg (2003)
  • [12] Kokotovic, P.V., Khalil, H.K., O’Reilly, J.: Singular Perturbation Methods in Control: Analysis and Design, 2nd Edition, SIAM Classics in Applied Mathematics, 25 (1999)
  • [13] Kushner, H.J.: Weak Convergence Methods and Singularly Perturbed Stochastic Control and Filtering Problems. Birkhauser, Boston (1990)
  • [14] Naidu, S.D.: Singular Perturbations and Time Scales in Control Theory and Applications: An Overview. Dynamics of Continuous Discrete and Impulsive Systems, Series B: Applications and Algorithms, 9, 233-278 (2002)
  • [15] O’Malley, Jr. R.E.: Singular Perturbations and Optimal Control. In: Mathematical Control Theory, Copel, W.A. (ed.) Lecture Notes in Mathematics, 680, Springer-Verlag, Berlin (1978)
  • [16] Plotnikov, V.A.: Differential Equations with Multivalued Right-Hand Sides: Asymptotic Methods. AstroPrint, Odessa (1999) (in Russian).
  • [17] Quincampoix, M, Watbled, F.: Averaging Method for Discontinuous Mayer’s problem of Singularly Perturbed Control Systems. Nonlinear Analysis: Theory, Methods & Applications, 54, 819-837 (2003)
  • [18] Zhang, Y., Naidu, D.S., Cai, C., Zou, Y.: Singular Perturbations and Time Scales in Control Theory and Applications: an Overview 2002 \textendash- 2012. Int. J. Inf. Syst. Sci. 9(1), 1-36 (2014)