Loading [MathJax]/jax/element/mml/optable/MathOperators.js

Exact and positive controllability of boundary control systems

  • Received: 01 September 2016 Revised: 01 January 2017
  • Primary: 93B05; Secondary: 47N70, 35R02

  • We characterize the space of all exactly reachable states of an abstract boundary control system using a semigroup approach. Moreover, we study the case when the controls of the system are constrained to be positive. The abstract results are then applied to study flows in networks with static as well as dynamic boundary conditions.

    Citation: Klaus-Jochen Engel, Marjeta Kramar FijavŽ. Exact and positive controllability of boundary control systems[J]. Networks and Heterogeneous Media, 2017, 12(2): 319-337. doi: 10.3934/nhm.2017014

    Related Papers:

    [1] Klaus-Jochen Engel, Marjeta Kramar FijavŽ . Exact and positive controllability of boundary control systems. Networks and Heterogeneous Media, 2017, 12(2): 319-337. doi: 10.3934/nhm.2017014
    [2] Klaus-Jochen Engel, Marjeta Kramar Fijavž, Rainer Nagel, Eszter Sikolya . Vertex control of flows in networks. Networks and Heterogeneous Media, 2008, 3(4): 709-722. doi: 10.3934/nhm.2008.3.709
    [3] Didier Georges . Infinite-dimensional nonlinear predictive control design for open-channel hydraulic systems. Networks and Heterogeneous Media, 2009, 4(2): 267-285. doi: 10.3934/nhm.2009.4.267
    [4] Christophe Prieur . Control of systems of conservation laws with boundary errors. Networks and Heterogeneous Media, 2009, 4(2): 393-407. doi: 10.3934/nhm.2009.4.393
    [5] Yassine El Gantouh, Said Hadd . Well-posedness and approximate controllability of neutral network systems. Networks and Heterogeneous Media, 2021, 16(4): 569-589. doi: 10.3934/nhm.2021018
    [6] Ciro D'Apice, Olha P. Kupenko, Rosanna Manzo . On boundary optimal control problem for an arterial system: First-order optimality conditions. Networks and Heterogeneous Media, 2018, 13(4): 585-607. doi: 10.3934/nhm.2018027
    [7] Zhong-Jie Han, Gen-Qi Xu . Spectrum and dynamical behavior of a kind of planar network of non-uniform strings with non-collocated feedbacks. Networks and Heterogeneous Media, 2010, 5(2): 315-334. doi: 10.3934/nhm.2010.5.315
    [8] Luca Di Persio, Giacomo Ziglio . Gaussian estimates on networks with applications to optimal control. Networks and Heterogeneous Media, 2011, 6(2): 279-296. doi: 10.3934/nhm.2011.6.279
    [9] Gildas Besançon, Didier Georges, Zohra Benayache . Towards nonlinear delay-based control for convection-like distributed systems: The example of water flow control in open channel systems. Networks and Heterogeneous Media, 2009, 4(2): 211-221. doi: 10.3934/nhm.2009.4.211
    [10] Xavier Litrico, Vincent Fromion, Gérard Scorletti . Robust feedforward boundary control of hyperbolic conservation laws. Networks and Heterogeneous Media, 2007, 2(4): 717-731. doi: 10.3934/nhm.2007.2.717
  • We characterize the space of all exactly reachable states of an abstract boundary control system using a semigroup approach. Moreover, we study the case when the controls of the system are constrained to be positive. The abstract results are then applied to study flows in networks with static as well as dynamic boundary conditions.



    This paper is a continuation of [16,17] where we introduced a semigroup approach to boundary control problems and applied it to the control of flows in networks. While in these previous works we concentrated on maximal approximate controllability, we now focus on the exact- and positive controllability spaces.

    There is a vast literature on abstract boundary control problems as well as on the application of the abstract theory to various concrete boundary control systems on Euclidean spaces. For an overview and related references on that topic we refer to [17,Sec. 1].

    As a simple motivation for our study, we consider as in [16] a transport process along the edges of a finite network. This system is subject to some transmission conditions in the vertices of the network (imposing, for example, conservation of the mass) which span the "boundary space" for our problem. We then like to control the behavior of this system by acting upon a single vertex only. In this context it is reasonable to ask the following questions.

    ● Can we reach all possible states in finite time? The answer to this question is in general negative since we are limited by the network structure, see, e.g., [16,Sec. 5]. Therefore, we only ask: Can we describe the maximal possible set of reachable states in some finite time?

    ● Does controllability depend on the particular choice of the control node? Here the answer is affirmative, which is again demonstrated by some examples in [16,Sec. 5]. However, to our knowledge there is no simple characterization for nodes yielding maximal exact control.

    ● Which states can be reached if only positive controls are allowed? This question is very important since in many applications only positive controls are meaningful and one expects that the state of the system remains positive for all times.

    In recent decades, the study of different partial differential equations on networks and similar structures gained a lot of interest. Here we restrict ourself to first order equations that model transport problems (or flows) in networks and are motivated by many real-life applications. In fact, such systems can be used to model, to control, and to optimize road traffic [24,23,9,21,19], water supply [27,28], gas flow [5,22] or supply chains [11], to mention just the most frequent applications. Moreover, we refer to [8] for a survey of related results in the context of nonlinear hyperbolic systems.

    In the present work we will, however, only consider linear models and make heavy use of the theory of semigroups of linear operators in the spirit of [18]. The application of semigroup theory to flows in networks was initiated in [26,30], see also the survey paper [13] and the detailed accounts in the monographs [6,Ch. 18] and [29]. For an example of an application of this approach to population models in biology we refer to [4]. The stability and control problems of linear flows in networks using semigroup approach were investigated in [16,17,25,7]. Our aim here is to further generalize and refine these latter results.

    This paper is organized as follows. In Section 2 we first recall our abstract framework from [17] as well as some basic results concerning boundary control systems. In Section 3 we then characterize boundary admissible control operators and describe the corresponding exact reachability space. In Section 4 we turn our attention to positive boundary control systems on Banach lattices. Finally, in Section 5 we apply our abstract results and explicitly compute the exact (positive) reachability spaces in three different examples of transport equations controlled at the boundary: in Cm, in a network, and in a network with dynamic boundary conditions. Moreover, we return to the motivating questions stated above.

    We start by recalling our setting from [17].

    Abstract Framework 2.1. We consider

    (ⅰ) three Banach spaces X, X and U, called state, boundary and control space, respectively;

    (ⅱ) a closed, densely defined system operator Am:D(Am)XX;

    (ⅲ) a boundary operator QL([D(Am)],X);

    (ⅳ) a control operator BL(U,X).

    For these operators and spaces and a control function uL1loc(R+,U) we then consider the abstract Cauchy problem with boundary control1

    1We denote by ˙x(t) the derivative of x with respect to the "time" variable t.

    {˙x(t)=Amx(t),t0,Qx(t)=Bu(t),t0,x(0)=x0X. (1)

    A function x()=x(,x0,u)C1(R+,X) with x(t)D(Am) for all t0 satisfying (1) is called a classical solution. Moreover, we denote the abstract boundary control system associated to (1) by BC(Am,B,Q).

    In order to investigate (1) we make the following standing assumptions which in particular ensure that the uncontrolled abstract Cauchy problem, i.e., (1) with B=0, is well-posed.

    Main Assumptions 2.2. (i) The restriction AAm with domain D(A):=kerQ generates a strongly continuous semigroup (T(t))t0 on X;

    (ii) the boundary operator Q:D(Am)X is surjective.

    Under these assumptions the following has been shown in [20,Lem. 1.2].

    Lemma 2.3. Let Assumptions 2.2 be satisfied. Then the following assertions are true for all λ,μρ(A).

    (i) D(Am)=D(A)ker(λAm);

    (ii) Q|ker(λAm) is invertible and the operator

    Qλ:=(Q|ker(λAm))1:Xker(λAm)X (2)

    is bounded;

    (iii) R(μ,A)Qλ=R(λ,A)Qμ.

    The following operators are essential to obtain explicit representations of the solutions of the boundary control problem (1).

    Definition 2.4. For λρ(A) we call the operator Qλ introduced in (2) abstract Dirichlet operator and define

    Bλ:=QλBL(U,ker(λAm))L(U,X).

    By [17,Prop. 2.7] the solutions of (1) can be represented by the following extrapolated version of the variation of parameters formula. Here we use the standard notation for the extrapolated spaces and operators: X1 denotes the completion of X with respect to the norm

    x1:=R(λ0,A)x,xX

    for some fixed λ0ρ(A), T1(t)L(X1) is the unique bounded extension of the operator T(t) to X1, and A1 is the generator of the extrapolated semigroup (T1(t))t0 with domain D(A1)=X, cf. [18,Sect. Ⅱ.5.a].

    Proposition 2.5. Let x0X, uL1loc(R+,U) and λρ(A). If x()=x(,x0,u) is a classical solution of (1), then it is given by the variation of parameters formula

    x(t)=T(t)x0+(λA1)t0T(ts)Bλu(s)ds,t0. (3)

    Our aim in the sequel is to investigate which states in X can be exactly reached from x0=0 by solutions of (1). To this end we have to impose an additional assumption which, by (3), ensures that solutions for Lp-controls remain in X.

    Definition 2.6. Let 1p+. Then the control operator BL(U,X) is called p-boundary admissible if there exist t>0 and λρ(A) such that

    t0T(ts)Bλu(s)dsD(A)for all uLP([0,t],U). (4)

    Remark 2.7. From Lemma 2.3.(ⅲ) it follows that (λA1)QλL(X,X1), hence also

    BA:=(λA1)BλL(U,X1),

    are independent of λρ(A). Then BL(U,X) is p-boundary admissible if and only if BA is p-admissible in the usual sense, cf. [31,Def. 4.1]. This implies that if (4) is satisfied for some t>0 then it is satisfied for every t>0. Moreover, we note that by [17,Lem. A.3], B is 1-boundary admissible if ker(λAm)FA1, the Favard class of A (see [17,Def. A.1] and references there). Finally, since LP([0,t],U)L1([0,t],U) it follows that 1-boundary admissibility implies p-boundary admissibility for all p>1.

    Now assume that BL(U,X) is p-boundary admissible. Then for fixed λρ(A) and t>0 the operators BBCt:LP([0,t],U)X given by

    BBCtu:=(λA)t0T(ts)Bλu(s)ds=t0T1(ts)BAu(s)ds (5)

    are called the controllability maps of the system BC(Am,B,Q), where the second integral initially is taken in the extrapolation space X1. Note that by the closed graph theorem BBCtL(LP([0,t],U),X). Hence, this definition is independent of the particular choice of λρ(A) and gives the (unique) classical solution of (1) for given uW2,1([0,t],U) and x0=0. This motivates the following definition.

    Definition 2.8.(a) The exact reachability space in time t0 of BC(Am,B,Q) is defined by2

    2By rg(T) we denote the range TXY of an operator T:XY.

    eRBCt:=rg(BBCt).

    Moreover, we define the exact reachability space (in arbitrary time) by

    eRBC:=t0rg(BBCt)

    and call BC(Am,B,Q) exactly controllable (in arbitrary time) if eRBC=X.

    (b) The approximate reachability space in time t0 of BC(Am,B,Q) is defined by

    aRBCt:=¯eRBCt.

    Moreover, we define the approximate reachability space (in arbitrary time) by

    aRBC:=¯t0aRBCt

    and call BC(Am,B,Q) approximately controllable if aRBC=X.

    From [17,Thm. 2.12 & Cor. 2.13] we obtain the following properties and representations of the approximate reachability space.

    Proposition 2.9. Assume that BL(U,X) is p-boundary admissible. Then the following holds.

    (i) aRBC is a closed linear subspace of X which is invariant under (T(t))t0 and R(λ,A) for λ>ω0(A).

    (ii) aRBC=¯spanλ>ωrg(Bλ) for every ω>ω0(A).

    (iii) aRBC¯spanλ>ω0(A)ker(λAm).

    Part (ⅲ) shows that there is an upper bound for the reachability space depending on the eigenvectors of Am only, independent of the control operator B. This justifies the following notion.

    Definition 2.10. The maximal reachability space of BC(Am,B,Q) is defined by

    RBCmax:=¯spanλ>ω0(A)ker(λAm).

    The system BC(Am,B,Q) is called maximally controllable if eRBC=RBCmax.

    We stress again that RBCmaxX may happen (cf. [16,Sec. 5] for an example), hence the relevant question about exact or approximate controllability for boundary systems is indeed to compare eRBC or aRBC to the space RBCmax and not to the whole space X, as it is usually done in the classical situation in systems theory (see [10]).

    After this short summary on boundary control systems BC(Am,B,Q) taken mainly from [17] in the context of approximate controllability, we now turn our attention to the case of exact controllability.

    We start this section by giving two characterizations of p-boundary admissibility for a control operator B which frequently also simplifies the explicit computation of the associated controllability map BBCt. Here for λC we introduce the function ελ:RC by ελ(s):=eλs. Moreover, for fLp[0,t] and uU we define

    fuLP([0,t],U)by(fu)(s):=f(s)u.

    Finally, we denote by 1[α,β] the characteristic function of the interval [α,β][0,t].

    Proposition 3.1. For a control operator BL(U,X) the following are equivalent.

    (a) B is p-boundary admissible.

    (b) There exist λρ(A), t>0 and ML(LP([0,t],U),X) such that for all 0αβt and vU

    (eλβT(tβ)eλαT(tα))Bλv=M(ελ1[α,β]v). (6)

    (c) There exist t>0, λ0>ω0(A) and ML(LP([0,t],U),X) such that for all λλ0 and vU

    (eλtT(t))Bλv=M(ελv). (7)

    Moreover, in this case the controllability map is given by BBCt=M.

    Proof. Let u=ελ1[α,β]v for some λρ(A), 0αβt and vU. Then

    t0T(ts)Bλu(s)ds=eλtβαeλ(ts)T(ts)Bλvds=eλttαtβeλsT(s)Bλvds=R(λ,A)(eλβT(tβ)eλαT(tα))Bλv. (8)

    (a)(b). Since by assumption B is p-boundary admissible we have BBCtL(LP([0,t],U),X). Hence, (5) and (8) imply (6) for M=BBCt.

    (b)(a). We start by proving (4). The idea is to show this first for functions of the type u=ελ1[α,β]v. Then by linearity it also holds for linear combinations of such functions and a density argument implies (4) for arbitrary uLP([0,t],U). To this end let u=ελ1[α,β]v for [α,β][0,t] and vU. Then (6) and (8) imply

    t0T(ts)Bλu(s)ds=R(λ,A)M(ελ1[α,β]v)=R(λ,A)Mu. (9)

    Note that the multiplication operator MλL(LP([0,t],U)) defined by Mλu:=ελu is an isomorphism (with bounded inverse Mλ). Hence, it maps dense sets of LP([0,t],U) into dense sets. Since the step functions are dense in LP([0,t],U) (see [3,p.14]), the linear combinations of functions of the type ελ1[α,β]v for [α,β][0,t] and vU form a dense subspace of LP([0,t],U). Thus, we conclude that (9) holds for all uLP([0,t],U). Clearly this implies that B is p-boundary admissible and BBCt=M.

    Recall that BA=(λA1)Bλ is independent of λρ(A). Hence, (a)(c) follows as before by replacing the total set {ελ1[α,β]v:0α<βt,vU} by the set {ελv:λλ0,vU} which by the Stone-Weierstraẞ theorem is total as well in LP([0,t],U) for all λ0>ω0(A).

    We note that by linearity it would suffice that Part (b) of Proposition 3.1 is satisfied for α=0 and all 0βt (or for all 0αt and β=t).

    Corollary 3.2. Let3 nN1 and assume that B is p-boundary admissible. Then for all uLP([0,nt],U)

    3We use the notation Nl:={l,l+1,l+2,} for the set of natural numbers starting at lN.

    BBCntu=n1k=0T(t)kMuk (10)

    where ukLP([0,t],U) is defined by

    uk(s)=u((nk1)t+s) (11)

    and ML(LP([0,t],U),X) is the operator introduced in Proposition 3.1.

    Proof. Let uLP([0,nt],U). Then by (5)

    BBCntu=(λA)nt0T(nts)Bλu(s)ds=(λA)nk=1T((nk)t)kt(k1)tT(kts)Bλu(s)ds=nk=1T((nk)t)(λA)t0T(ts)Bλunk(s)ds=n1k=0T(t)kBBCtuk.

    In Section 5 we will see that (6), (7), and (10) allow us to easily compute the controllability map in the situations studied in [16,Sect. 4] and [17,Sect. 3] dealing with the control of flows in networks.

    Corollary 3.3. If B is p-boundary admissible, then the exact reachability space in time nt for nN1 is given by

    eRBCnt={n1k=0T(t)kMuk:ukLP([0,t],U),1kn1},

    where ML(LP([0,t],U),X) is the operator from Proposition 3.1.

    In this section we are interested in positive control functions yielding positive states. To this end we will make the following

    Additional Assumption 4.1. The spaces X and U are Banach lattices.

    Moreover, by Y+:={yY:Y0} we denote the positive cone in a Banach lattice Y. For a detailed account of the theory of semigroups of positive linear operators we refer to [2,6].

    Note that in the sequel we do not make any positivity assumptions on (T(t))t0, B or Qλ if not stated otherwise.

    Definition 4.2.(a) The exact positive reachability space in time t0 of system BC(Am,B,Q) is defined by

    e+RBCt:={BBCtu:uLP([0,t],U+)}.

    Moreover, we define the exact positive reachability space (in arbitrary time) by

    e+RBC:=t0e+RBCt

    and call BC(Am,B,Q) exactly positive controllable (in arbitrary time) if e+RBC=X+.

    (b) The approximate positive reachability space in time t0 of BC(Am,B,Q) is defined by

    a+RBCt:=¯e+RBCt.

    Moreover, we define the approximate positive reachability space (in arbitrary time) by

    a+RBC:=¯t0a+RBCt

    and call BC(Am,B,Q) approximately positive controllable if a+RBC=X+.

    First we give necessary and sufficient conditions implying that starting from the initial state x0=0 positive controls result in positive states.

    Proposition 4.3. Assume that BL(U,X) is p-boundary admissible. Then

    e+RBCtX+ (12)

    if and only if

    a+RBCtX+ (13)

    if and only if there exists λRρ(A) such that

    (eλβT(tβ)eλαT(tα))Bλ0forall0αβt. (14)

    Moreover, if (T(t))t0 is positive, then the above assertions are satisfied if and only if

    e+RBCX+ (15)

    if and only if

    a+RBCX+ (16)

    if and only if there exists λ>ω0(A) and t>0 such that

    (eλsT(s))Bλ0forall0st (17)

    if and only if there exists λ0>ω0(A) such that

    Bλ0forallλλ0. (18)

    Proof. The equivalence of (12) and (13) follows from the closedness of X+. To show the equivalence of (12) and (14) recall that by [3,p.14] the step functions are dense in LP([0,t],U). Since the map uu+ on LP([0,t],U) is continuous, we conclude that the positive step functions are dense in LP([0,t],U+). The claim then follows from (the proof of) Proposition 3.1 using the boundedness of the controllability map BBCt.

    Now assume that (T(t))t0 is positive. Then the equivalences of (12), (13) with (15), (16) follow from Corollary 3.3 using the fact that the reachability spaces are growing in time. In particular, this implies that if (14) holds for some t>0 it holds for arbitrary t>0 and choosing β=t and α=0 we obtain (17) for arbitrary t>0.

    To show the remaining assertions we fix some λ>ω0(A) and define on X:=X×X the operator matrix

    A:=(Aλ000),D(A):={(xy)D(Am)×X:Qx=By}.

    Then by [14,Cor. 3.4] the matrix A generates a C0-semigroup (T(t))t0 given by

    T(s)=(eλsT(s)(IeλsT(s))Bλ0I),s0.

    Moreover, by [14,Lem. 3.1] we have (0,+)ρ(A) and

    R(μ,A)=(R(μ+λ,A)1μBμ+λ01μ)for μ>0. (19)

    Now, if (17) holds then T(s)0 for all 0st which implies that (T(t))t0 is positive which is equivalent to the fact that A is resolvent positive. However, by (19) the latter is the case if and only if (18) is satisfied which shows the equivalence of (17) and (18). Finally, if (17) holds, then

    (eλβT(tβ)eλαT(tα))Bλ=eλβT(tβ)(Ieλ(βα)T(βα))Bλ0

    for all 0αβt. This proves (14) and completes the proof.

    In the sequel we use the notation coM and ¯coM to indicate the convex hull and the closed convex hull of a set MX, respectively.

    Proposition 4.4. Assume that BL(U,X) is p-boundary admissible and that e+RBCtX+. Then the following holds.

    (i) a+RBC is a closed convex cone, invariant under (T(t))t0 and R(λ,A) for λ>ω0(A).

    (ii) a+RBC=¯co{(eλβT(tβ)eλαT(tα))Bλv:0αβt,vU+} for all λ>ω0(A).

    (iii) a+RBC=¯co{T(t)Bλv:t0,λ>w,vU+} for all w>ω0(A).

    (iv) a+RBC=¯co{R(λ,A)nBλv:nN0,λ>w,vU+} for some (and hence for all) w>ω0(A).

    Proof. (ⅰ). Clearly, a+RBC is a closed convex cone. Its invariance under (T(t))t0 and R(λ,A) for λ>ω0(A) follows from the representations in (ⅲ) and (ⅳ).

    To show (ⅱ) we note that by (5) and (8) the inclusion "" holds. Now recall that the positive step functions are dense in LP([0,t],U+) and invariant under positive convex combinations. Hence, the boundedness of the controllability maps implies equality of the spaces in (ⅱ).

    To obtain (ⅲ) we note that by (5) and (8) we have

    (eλβT(tβ)eλαT(tα))Bλve+RBC

    for all 0αβt and vU+. Multiplying this inclusion by eλβ>0 and putting s:=tβ and r:=tα implies

    (T(s)eλ(sr)T(r))Bλve+RBC

    for all 0sr and vU+. Since λ>ω0(A) we obtain

    limr+eλ(sr)T(r)=0

    and hence

    T(s)Bλva+RBC

    for all s0 and vU+. This shows the inclusion "" in (ⅲ). For the converse inclusion in (ⅲ) it suffices to prove that

    e+RBCt¯co{T(s)Bμy:s0,μ>w,yU+} (20)

    for all t>0 and w>ω0(A). Since BL(U,X) is p-boundary admissible the controllability map BBCt is continuous. Moreover, the positive step functions are dense in LP([0,t],U+) and ¯co{T(s)Bμy:s0,μ>w,yU+} is a convex cone. Combining these facts and (8) it follows that (20) holds if

    (eλβT(tβ)eλαT(tα))Bλv¯co{T(s)Bμy:s0,μ>w,yU+} (21)

    for all 0αβt, kN0 and vX+. Since (T(t))t0 is strongly continuous the following integral is the limit of Riemann sums, hence for ν>max{0,w} we obtain using Lemma 2.3.(ⅲ)

    ¯co{T(s)Bμy:s0,μ>w,yU+}νtαtβeλ(tr)T(r)Bνvdr=(eλβT(tβ)eλαT(tα))νR(λ,A)Bνv=(eλβT(tβ)eλαT(tα))νR(ν,A)Bλv(eλβT(tβ)eλαT(tα))Bλv,

    as ν+. This proves (21) and completes the proof of (ⅲ).

    That the right-hand-sides of the equalities in (ⅲ) and (ⅳ) coincide follows from the integral representation of the resolvent (see [18,Cor. Ⅱ.1.11]) and the Post-Widder inversion formula (see [18,Cor. Ⅲ.5.5]). For the details we refer to the proof of [7,Prop. 3.3].

    Corollary 4.5. Assume that BL(U,X) is p-boundary admissible and that a+RBCX+. Then the following are equivalent.

    (a) The system BC(Am,B,Q) is approximately positive controllable.

    (b) There exists w>ω0(A) such that the following implication holds for all ϕX

    T(s)Bλv,ϕ0forallvU+,s0andλ>wϕ0.

    (c) There exists w>ω0(A) such that the following implication holds for all ϕX

    R(λ,A)nBλv,ϕ0forallvU+,nNandλ>wϕ0.

    Proof. This follows from the proof of [7,Thm. 3.4] by replacing [7,Prop. 3.3] with our Proposition 4.4.

    Remark 4.6. The previous two results generalize [7,Prop. 3.3 and Thm. 3.4], respectively, where it is assumed that (T(t))t0, B and Qλ for all λ>λ0 are all positive and, in particular, where the additional hypothesis

    (H) There exists γ>0 and λ0R such that Qxγλx for all λ>λ0 and xker(λAm)

    is made. We note that Hypothesis (H) is quite strong, e.g., in reflexive state spaces X it implies that A=Am, cf. [1,Lem. A.1]. Hence, the results of [7] are not applicable to state spaces like X=Lp([a,b],Y) for p(1,+) and reflexive Y.

    Combining Corollary 3.2 and Proposition 4.3 we finally obtain the following characterization of an exact positive reachability space.

    Corollary 4.7. Assume that B is p-boundary admissible, t>0 and nN1. Then the exact positive reachability space in time nt is given by

    e+RBCnt={n1k=0T(t)kMuk:ukLP([0,t],U+),1kn1},

    where ML(LP([0,t],U+),X) is the operator from Proposition 3.1. Moreover, the operator M is positive if and only if a+RBCtX+.

    In this section we will show how our abstract results can be applied to a linear transport equation with boundary control and to vertex control of linear flows in networks subject to static and dynamic boundary conditions.

    In this subsection we study the controlled transport equation in Cm given by4

    4We denote by x(t,s) the derivative of x(t,s) with respect to the "space" variable s.

    {˙x(t,s)=x(t,s),s[0,1], t0,x(t,1)=Bx(t,0)+u(t)b,t0,x(0,s)=0,s[0,1]. (22)

    Here x:R+×[0,1]Cm (i.e., x(t)=(xj(t,s))mj=1), BMm(C) implements the boundary conditions for the functions xj(t,s), u:R+C is a control function, and bCm is the vector that assigns the control. Roughly spoken, B determines how the material (flowing on the s-interval [0,1] from right to left) leaving the system at the left end point s=0 is again fed into the system at the right end point s=1.

    We note that by our approach one can also deal with other boundary conditions and controls. However, the above choice is useful for studying the network example in Section 5.2.

    In order to fit the system (22) in our general framework we choose

    ● the state space X:=Lp([0,1],Cm) for some 1p<+,

    ● the boundary space X:=Cm,

    ● the control space U:=C,

    ● the control operator B:=bCmL(U,X)=L(C,Cm),

    ● the system operator

    Am:=diag(dds)m×mwith domainD(Am):=W1,p([0,1],Cm),

    ● the boundary operator Q:W1,p([0,1],Cm)Cm, Qf:=f(1)Bf(0),

    ● the operator AAm with domain D(A)=kerQ,

    ● the state trajectory x:R+Lp([0,1],Cm), x(t):=x(t,).

    For these choices the controlled transport equation (22) can be reformulated as an abstract Cauchy problem with boundary control of the form (1). Clearly, the above boundary operator Q is surjective.

    Observe that the operator A is a difference operator as considered in [12,25,15,6]. By [6,Cor. 18.4] we know that for λC and A=Am|ker(Q) as above we have

    λρ(A)eλρ(B).

    Moreover, by [6,Prop. 18.7] the operator A generates a strongly continuous semigroup given by

    (T(t)f)(s)=Bkf(t+sk)if t+s[k,k+1) for kN0, (23)

    where B0:=Id. This shows that the Assumptions 2.2 are satisfied. To proceed we have to compute the associated Dirichlet operator.

    Lemma 5.1. For λρ(A) the Dirichlet operator QλL(Cm,Lp([0,1],Cm)) is given by

    Qλ=ελR(eλ,B). (24)

    Proof. By Lemma 2.3.(ⅱ) we know that Q:ker(λAm)X is invertible. Moreover, for dCm=X we have

    Q(ελR(eλ,B)d)=eλR(eλ,B)dBR(eλ,B)d=d

    which proves (24).

    In order to apply Proposition 3.1 to the present situation we need the following.

    Lemma 5.2. Let λρ(A). Then for all 0α1

    (eλαT(1α)Bλ)(s)={ελ(1+s)R(eλ,B)bif0s<α,ελ(1+s)R(eλ,B)bελ(s)bifαs1.

    Hence, (6) is satisfied for

    M=bL(Lp[0,1],Lp([0,1],Cm)), (Mu)(s)=u(s)b.

    Proof. The claim follows from (24) and (23) by the following simple computation.

    (eλαT(1α)Bλ)(s)=eλα(T(1α)(ελR(eλ,B)b))(s)=eλα{ελ(1α+s)R(eλ,B)bif 0s<α,ελ(sα)BR(eλ,B)bif αs1,={ελ(1+s)R(eλ,B)bif 0s<α,ελ(1+s)R(eλ,B)bελ(s)bif αs1.

    Thus, by Proposition 3.1 the operator B is p-boundary admissible. Next we compute the appropriate reachability space.

    Corollary 5.3. If tm then the exact reachability space of the controlled transport equation (22) is given by

    eRBCt=eRBC=Lp[0,1]span{b,Bb,,Bm1b}.

    Proof. Note that by (23) we have T(1)f=Bf. Hence, for t=m the assertion follows immediately from Corollary 3.3 and Lemma 5.2. Clearly, eRBCt increases in time t0. However, by the Cayley-Hamilton theorem span{b,Bb,,Blb}=span{b,Bb,,Bm1b} for all lm1 and the claim follows.

    Remark 5.4. Let lm be the degree of the minimal polynomial of B. Then the previous proof shows that for all tl we even have

    eRBCt=eRBC=Lp[0,1]span{b,Bb,,Bl1b}.

    Corollary 5.5. The following assertions are equivalent.

    (a) Equation (22) is exactly boundary controllable in time tm, i.e., eRBCt=X.

    (b) Equation (22) is maximally controllable in time tm, i.e., eRBCt=RBCmax.

    (c) span{b,Bb,,Bm1b}=Cm.

    Proof. Note that ker(λAm)=ελCm. Since by the Stone-Weierstraẞ theorem we have

    ¯spanλ>ω0(A){ελ}=Lp[0,1],

    the maximal reachability space equals

    RBCmax=Lp[0,1]Cm=X

    and the assertions follow immediately from Corollary 5.3.

    Remark 5.6. The previous result characterizes the exact maximal boundary controllability by a one-dimensional control in terms of a Kalman-type condition which is well-known in control theory.

    Combining Remark 5.4 and Corollary 5.5 we furthermore obtain the following

    Corollary 5.7. Let lN be the degree of the minimal polynomial of B. If l<m, the transport equation (22) is not maximally controllable, i.e., eRBC.

    Finally, we investigate positive controllability and consider

    ● the positive cone X^+:={\rm L}^p([0,1],\mathbb{R}_+^m) in the state space X,

    ● the positive cone U^+:=\mathbb{R}_+ in the control space U,

    ● a positive matrix \mathbb{B}\in{\rm M}_m(\mathbb{R}_+),

    ● a positive control operator B:=b\in\mathbb{R}_+^m.

    Then by (23)-(24) the operators T(t)\in\mathcal{L}(X) for t\ge0 and B_\lambda \in\mathcal{L}(U,X) for \lambda >\omega_0(A) are positive. Thus arguing as above using Proposition 4.3 and Corollary 4.7 we obtain the following.

    Corollary 5.8. The exact positive reachability space of the controlled transport equation (22) is given by

    e^+\mathcal{R}^{BC}={\rm L}^p([0,1],\mathbb{R}_+)\otimes{\rm{co}}\bigl\{\mathbb{B}^{k}b \;:\; k\in\mathbb{N}_0\bigr\}.

    Hence, the problem is exactly positive controllable if and only if

    {\rm{co}}\bigl\{\mathbb{B}^{k}b \;:\; k\in\mathbb{N}_0\bigr\} = \mathbb{R}_+^m.

    The previous example can be easily adapted to cover a transport problem on a network controlled in a single vertex. More precisely, consider a network consisting of n vertices \{v_1,\dots ,v_n\} and m edges \{e_1,\dots ,e_m\}. As shown in [6,Sec. 18.1], its structure can be described by either the transposed weighted adjacency matrix \mathbb{A}\in{\rm M}_n(\mathbb{C}) given by

    \mathbb{A}_{ij}:=\begin{cases} w _{jk}&\mbox{if }v_j\stackrel{e_k}{\longrightarrow}v_i, \\ 0&\text{otherwise,} \end{cases}

    or by the transposed weighted adjacency matrix of the line graph \mathbb{B}\in{\rm M}_m(\mathbb{C}) where

    \mathbb{B}_{ij}:=\begin{cases} w_{ki}&\mbox{if }\stackrel{e_j}{\longrightarrow}v_k \stackrel{e_i}{\longrightarrow}\ , \\ 0&\text{otherwise.} \end{cases}

    We also need the transposed weighted outgoing incidence matrix (\Phi_w^-)^\top=:\Psi\in{\rm M}_{m \times n}(\mathbb{C}) defined by

    \Psi_{ij} := \begin{cases} w_{ij}&\mbox{if }v_j\stackrel{e_i}{\longrightarrow}\ , \\ 0&\text{otherwise} \end{cases}

    and the corresponding unweighted outgoing incidence matrix denoted by \Phi^-\in{\rm M}_{n \times m}(\mathbb{C}). For the weights we assume 0\le w_{ij} \le 1, thus all these matrices are positive. Moreover, we assume that \Psi is column stochastic, i.e., the weights on all outgoing edges from a given vertex sum up to 1. This implies that \mathbb{B} is column stochastic as well. For a detailed account of the various graph matrices we refer to [6,Sec. 18.1]. Here we only mention the following relations

    \Psi\mathbb{A} = \mathbb{B}\Psi, \Psi R(\lambda ,\mathbb{A}) = R(\lambda ,\mathbb{B})\Psi, \;\;\;\;\text{and}\;\;\;\; \Phi^-\Psi = Id_{\mathbb{C}^n} (25)

    which we will need in the sequel.

    We then consider a transport equation on the m edges, which are all parametrized on the interval [0,1], imposing n boundary conditions in the vertices, controlled in a single vertex v_i, i.e.,

    \begin{cases} \dot x(t,s)=x'(t,s),& s\in[0,1],\ t\ge0,\\ x(t,1)=\mathbb{B} x(t,0)+u(t)\cdot \Psi v,&t\ge0,\\ x(0,s)=0,& s\in[0,1]. \end{cases} (26)

    Here x:\mathbb{R}_+\times[0,1]\to\mathbb{C}^m, that is, x(t)=(x_j(t,\cdot))_{j=1}^m consists of functions on the parametrized edges, and u:\mathbb{R}_+\to\mathbb{C} is a control function acting on the vertex v=v_i, which is represented by the i-th canonical basis vector in \mathbb{C}^n. The matrix \Psi applied to v then takes the control with the appropriate weights to the outgoing edges. Moreover, the boundary conditions are encoded into the matrix \mathbb{B}. Note that by applying this adjacency matrix we "glue" together the relevant values of the functions x_j(t) at the endpoints s=0 and s=1 of the edges that share a common vertex. Since \mathbb{B} is column stochastic, this also implements conservation of mass in every vertex (the so-called Kirchhoff's condition).

    To rewrite equation (26) in our abstract form we take as in Section 5.1 the state space X:={\rm L}^p([0,1],\mathbb{C}^m), the control space U:=\mathbb{C} and the boundary space \partial X:=\mathbb{C}^m. Adapting the domain of A_m as

    D(A_m):=\left\{f\in{\rm W}^{1,p}([0,1],\mathbb{C}^m): f(1) \in {\rm{rg}}\Psi \right\}

    and choosing the control operator B=b:=\Psi v\in\mathbb{C}^m we are in the situation considered in [16] and [7], see also [6,Sec. 18.4].

    Then the approximate controllability space for the network flow problem computed in [16,Cor. 4.3] by Corollary 5.3 above indeed coincides with the exact controllability space.

    Corollary 5.9. If t\ge \min\{m,n\}=:l then the exact reachability space of the controlled transport in network problem (26) equals

    \begin{align*} e\mathcal{R}_t^{BC}=e\mathcal{R}^{BC}&={\rm L}^p[0,1]\otimes{\rm{span}}\left\{\Psi v,\mathbb{B} \Psi v,\ldots,\mathbb{B}^{l-1}\Psi v\right\}\\ & = {\rm L}^p[0,1]\otimes \Psi {\rm{span}}\left\{v,\mathbb{A} v,\ldots,\mathbb{A}^{l-1} v\right\} . \end{align*}

    Note that in big connected networks one usually has n\le m, hence the latter space is more relevant for applications.

    Positive control for this problem was already studied in [7] and the approximate positive reachability space was computed. However, our approach even yields in this context the exact reachability space.

    Corollary 5.10. The exact positive reachability space of the controlled transport in network problem (26) is given by

    \begin{align*} e^+\mathcal{R}^{BC}&={\rm L}^p([0,1],\mathbb{R}_+)\otimes{\rm{co}}\left\{\mathbb{B}^{k} \Psi v : k\in\mathbb{N}_0\right\}\\ & = {\rm L}^p([0,1],\mathbb{R}_+)\otimes \Psi {\rm{co}}\left\{\mathbb{A}^k v : k\in\mathbb{N}_0\right\}. \end{align*}

    Remark 5.11. Let us revisit the motivating questions from the introduction. We have answered the first and the third one by giving explicit descriptions of the appropriate reachability spaces. From these descriptions one can also see that the choice of the vertex v where the control acts is important. This answers also the second question. For concrete examples we refer to [16,Sec. 5]. A characterization of a vertex yielding maximal control remains open.

    Finally, we note that by adding more vertices where the control takes place the reachability space increases since the appropriate span or convex hull in Corollaries 5.9 and 5.10, respectively, obtain additional terms in the added vertex. Hence, in this way it is easier to achieve maximal controllability.

    In this subsection we investigate exact and positive controllability in the situation of [17,Sect. 3]. Without going much into details we only introduce the necessary facts to state the problem and to compute the corresponding reachability spaces.

    We start from the transport problem on the network introduced in the previous example, but now change the transmission process in the vertices allowing for dynamical boundary conditions. To encode the structure of the underlying network and the imposed boundary conditions we use the incidence matrices introduced above as well as the weighted incoming incidence matrix \Phi_w^+ given by

    \bigl(\Phi_w^+\bigr)_{ij}:= \begin{cases} w^+_{ij}&\mbox{if }\stackrel{e_j}{\longrightarrow}v_i\ , \\ 0&\text{otherwise,} \end{cases}

    for some 0\le w^+_{ij} \le 1. Defining

    \mathbb{A} := \Phi_w^+\Psi \;\;\;\;\text{and}\;\;\;\; \mathbb{B}:=\Psi\Phi_w^+ (27)

    we obtain the adjacency matrices as above (with different nonzero weights). We mention that the relations (25) remain valid also in this case.

    We are then interested in the network transport problem with dynamical boundary conditions in s=1 considered already in [30] and [17,Sect. 3], i.e.,

    \begin{cases} \dot x(t,s)=x'(t,s),& s\in[0,1],\ t\ge0,\\ \dot x(t,1)=\mathbb{B} x(t,0)+u(t)\cdot \Psi v,&t\ge0,\\ x(0,s)=0,& s\in[0,1],\\ \Phi^- x(1,0) = 0. \end{cases} (28)

    To embed this example in our setting we introduce

    ● the state space X:={\rm L}^p([0,1],\mathbb{C}^m)\times\mathbb{C}^n where 1\le p<+\infty,

    ● the boundary space \partial X:=\mathbb{C}^m,

    ● the control space U:=\mathbb{C},

    ● the control operator B:=\Psi v\in\mathbb{C}^m\simeq\mathcal{L}(U,\partial X)=\mathcal{L}(\mathbb{C},\mathbb{C}^m) where v=v_i denotes the i-th canonical basis vector of \mathbb{C}^n meaning that the control acts in the i-th vertex of the network,

    ● the system operator5

    5By \delta_s we denote the point evaluation in s\in[0,1], i.e., \delta_s(f)=f(s).

    \begin{align*} A_m:&=\begin{pmatrix} {\rm {diag}}\bigl(\frac{d}{ds}\bigr)_{m\times m}&0\\\Phi_w^+\delta_0&0 \end{pmatrix} \text{with domain}\\ D(A_m):&=\left\{\tbinom fd\in{\rm W}^{1,p}([0,1],\mathbb{C}^m)\times\mathbb{C}^n:f(1)\in{\rm{rg}}\Psi\right\}, \end{align*}

    ● the boundary operator Q:D(A_m)\times\mathbb{C}^n\to\mathbb{C}^m, Q\binom fd:=\Phi^- f(1)-d,

    ● the operator A\subset A_m with domain D(A)=\ker Q.

    As is shown in [17,Prop. 3.4] these spaces and operators satisfy all assumptions of Section 2. To proceed we first need to compute the associated Dirichlet operator Q_\lambda and an explicit representation of the semigroup operators T(t) for t\in[0,1].

    Lemma 5.12. (i) For each 0\neq \lambda \in\rho(A), the Dirichlet operator Q_{\lambda }\in \mathcal{L}(\mathbb{C}^n, {X}) is given by

    Q_{\lambda } = \binom{ \lambda \varepsilon_\lambda \otimes\Psi R(\lambda e^\lambda ,\mathbb{A})} {\mathbb{A} R(\lambda e^\lambda ,\mathbb{A})}.

    (ii) The semigroup (T(t))_{t \ge 0} generated by A is given by6

    6We use the notations \bigl[\binom fd\bigr]_1:=f and \bigl[\binom fd\bigr]_2:=d for the canonical projections of \binom fd\in X.

    \begin{align} \left[T(t)\tbinom fd\right]_1(s)&= \begin{cases} f(t+s)&if\;\;\kern22pt 0\le t < 1-s,\\ \mathbb{B}\, V_{t+s-1} f +\Psi d&if \;\;\,1-s\le t\le1, \end{cases}\end{align} (29)
    \begin{align} \left[T(t)\tbinom fd\right]_2&=\Phi_w^+\, V_t f +d\kern45pt\;\;\;\;for \;\;\;\;0\le t\le1,\end{align} (30)

    where

    V_s f:=\int_0^s f(r)dr\;\;for \;\;f\in{\rm L}^p([0,1],\mathbb{C}^m). (31)

    Proof. Assertion (ⅰ) is proved in [17,Prop. 3.8]. Equation (30) is shown in the proof of [17,Prop. 3.4.(ⅲ)]. The statement (29) for the first coordinate then follows from [30,Lem. 6.1].

    Next we apply Proposition 3.1 to the present situation.

    Lemma 5.13. Let \lambda \in\rho(A). Then for all 0\le\alpha\le1

    \begin{align*} \bigl[e^{\lambda \alpha}\cdot T(1-\alpha)B_\lambda \bigr]_1(s)&= \begin{cases} \lambda \varepsilon_\lambda (1+s)\cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\, v&if \;\;0\le s < \alpha,\\ \lambda \varepsilon_\lambda (1+s)\cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\, v-\varepsilon_\lambda (s)\cdot \Psi v&if \;\;\alpha\le s\le1. \end{cases} \\ \bigl[e^{\lambda \alpha}\cdot T(1-\alpha)B_\lambda \bigr]_2 &=e^\lambda \mathbb{A} R(\lambda e^\lambda ,\mathbb{A})\, v \end{align*}

    Hence the equality in (6) is satisfied for

    \begin{align*} M&=\binom{\Psi v}{0}\in\mathcal{L}\Bigl({\rm L}^p[0,1],{\rm L}^p([0,1],\mathbb{C}^m)\times \mathbb{C}^n\Bigr),\ (M u)(s)=\binom{ u(s)\cdot\Psi v}{0}. \end{align*}

    Proof. Using the explicit representations of Q_\lambda and T(t) given in Lemma 5.12 and the relations (25) we obtain

    \begin{align*} \bigl[e^{\lambda \alpha}\cdot &T(1-\alpha)B_\lambda \bigr]_1(s)=\\ &=e^{\lambda \alpha}\cdot \begin{cases} \lambda \varepsilon_\lambda (1-\alpha+s)\cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\, v&\text{if }0\le s < \alpha,\\ \lambda \mathbb{B}\, V_{s-\alpha}\,\varepsilon_\lambda \cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\, v+\Psi\mathbb{A} R(\lambda e^\lambda ,\mathbb{A}) v&\text{if }\alpha\le s\le1, \end{cases} \\ &=e^{\lambda \alpha}\cdot \begin{cases} \lambda \varepsilon_\lambda (1-\alpha+s)\cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\, v&\text{if }0\le s < \alpha,\\ \bigl(\varepsilon_\lambda (s-\alpha)-1\bigr)\cdot\Psi\,\mathbb{A}R(\lambda e^\lambda ,\mathbb{A})\, v+\Psi\mathbb{A} R(\lambda e^\lambda ,\mathbb{A}) v&\text{if }\alpha\le s\le1, \end{cases} \\ &= \begin{cases} \lambda \varepsilon_\lambda (1+s)\cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\, v&\text{if }0\le s < \alpha,\\ \varepsilon_\lambda (s)\cdot\Psi\bigl(\lambda e^\lambda R(\lambda e^\lambda ,\mathbb{A})-Id\bigr)\, v&\text{if }\alpha\le s\le1. \end{cases} \\ &=\begin{cases} \lambda \varepsilon_\lambda (1+s)\cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\, v&\text{if }0\le s < \alpha,\\ \lambda \varepsilon_\lambda (1+s)\cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\, v-\varepsilon_\lambda (s)\cdot \Psi v&\text{if }\alpha\le s\le1. \end{cases} \end{align*}

    Similarly, for the second coordinate we have

    \begin{align*} \bigl[e^{\lambda \alpha}\cdot T(1-\alpha)B_\lambda \bigr]_2 &=e^{\lambda \alpha}\Bigl(\lambda \Phi_w^+\,V_{1-\alpha}\,\varepsilon_\lambda \cdot\Psi R(\lambda e^\lambda ,\mathbb{A})\,v+\mathbb{A} R(\lambda e^\lambda ,\mathbb{A})\, v\Bigr)\\ &=e^{\lambda \alpha}\Bigl(\bigl(\varepsilon_\lambda (1-\alpha)-1\bigr)\cdot\mathbb{A}R(\lambda e^\lambda ,\mathbb{A})\,v+\mathbb{A} R(\lambda e^\lambda ,\mathbb{A})\, v\Bigr)\\ &=e^\lambda \mathbb{A} R(\lambda e^\lambda ,\mathbb{A})\, v, \end{align*}

    where we used (27).

    We note that by [17,Prop. 3.5] the states of the controlled flow at time t\ge0 are given by the first coordinate of the states in our "extended" state space X={\rm L}^p([0,1],\mathbb{C}^m)\times\mathbb{C}^n. For this reason we also need to compute the first coordinate of T(1)^k\binom{\Psi g}{0}.

    Lemma 5.14. We have

    \left(T(1)\tbinom fd \right)(s)= \begin{pmatrix} \Psi\,\Phi_w^+\, V_s&\Psi\\\kern10pt \Phi_w^+\,V_1&Id \end{pmatrix} \binom{f}{d} =\binom{\mathbb{B}\,V_s f+\Psi d}{\Phi_w^+\, V_1 f +d},

    where the operator V_s\in\mathcal{L}\bigl({\rm L}^p([0,1],\mathbb{C}^m),{\rm W}^{1,p}([0,1],\mathbb{C}^m)\bigr) is defined in (31). Moreover, for k\in\mathbb{N}_1 we have

    \left[T(1)^k\tbinom{\Psi g}0\right]_1 (s) =\Psi(\mathbb{A} V_s +\delta_1)^{k-1}\mathbb{A}\,V_sg =(\mathbb{B} V_s+\delta_1)^{k-1}\mathbb{B}\Psi\,V_sg. (32)

    Proof. The formula for T(1) follows immediately from Lemma 5.12.(ⅱ). Since \Psi\mathbb{A} = \mathbb{B}\Psi, it suffices to show the second equality in (32). Obviously this equation holds for k=1. To verify it for k>1 we note that by (25) the matrix \Psi is left invertible with left inverse \Phi^-. Hence, we obtain

    \left[T(1)\tbinom fd\right]_2=\Phi^-\delta_1 \left[T(1)\tbinom fd\right]_1.

    If \binom fd\in{\rm{rg}} T(1) we can write f=\Psi h and the previous equation implies

    \begin{align*} \left[T(1)\tbinom fd\right]_1 (s) =\left[T(1)\tbinom {\Psi h}{\delta_1h}\right]_1(s) =\mathbb{B} V_s\Psi h+\Psi\delta_1 h =(\mathbb{B} V_s +\delta_1)f. \end{align*}

    Now assume that (32) holds for some k\ge1. Then for \binom fd=T(1)^k\binom{\Psi g}{0}\in{\rm{rg}} T(1) we conclude

    \begin{align*} \left[T(1)^{k+1}\tbinom{\Psi g}0\right]_1(s) &=\left[T(1)\cdot T(1)^{k}\tbinom{\Psi g}0\right](s)\\ &=(\mathbb{B} V_s +\delta_1)\cdot(\mathbb{B} V_s+\delta_1)^{k-1}\,\mathbb{B}\Psi\,V_sg\\ &=(\mathbb{B} V_s+\delta_1)^{k}\,\mathbb{B}\Psi\,V_s g. \end{align*}

    The previous two lemmas together with Corollary 3.2 imply the following result.

    Corollary 5.15. For l\in\mathbb{N}_2 and u\in{\rm L}^{p}[0,l] we have

    \begin{align} \bigl[\mathcal{B}_l^{BC} u\bigr]_1(s) &=\Psi\biggl( u_0\otimes v+\sum\limits_{k=1}^{l-1}\bigl(\mathbb{A} V_s+\delta_1\bigr)^{k-1}\, V_s(u_{k}\otimes\mathbb{A} v)\biggr) \\ &=u_0\otimes \Psi v+\sum\limits_{k=1}^{l-1}\bigl(\mathbb{B} V_s+\delta_1\bigr)^{k-1}\,V_s(u_{k}\otimes\mathbb{B}\Psi\, v) \end{align} (33)

    where u_k\in{\rm L}^p[0,1] is defined as in (11).

    Using this explicit representation of the controllability map we now compute the exact reachability space for the control problem given in (28).

    Corollary 5.16. If t\ge\min\{m,n\}=:l then the exact reachability space of the controlled flow with dynamic boundary conditions (28) is given by7

    7Here we define {\rm W}^{0,p}[0,1]:={\rm L}^p[0,1].

    \begin{align*} \bigl[e\mathcal{R}_t^{BC}\bigr]_1 &\subseteq \left\{ \Psi\sum\limits_{k=0}^{l}\left(u_{k}\otimes\mathbb{A}^k\, v\right):u_k\in{\rm W}^{k,p}[0,1]\;\; for \;\;0\le k\le l \right\}\\ &= \left\{ \sum\limits_{k=0}^{l}\left(u_{k}\otimes\mathbb{B}^k\Psi\, v\right):u_k\in{\rm W}^{k,p}[0,1]\;\; for \;\;0\le k\le l \right\}. \end{align*}

    Proof. The equality of the two sets on the right-hand-side follows immediately from (25). To show the inclusion in the second set we combine Corollaries 3.3 and 5.15. First observe, that for the operators \mathbb{B}, V_s, and \delta_1 we have

    \mathbb{B} V_s f = V_s \mathbb{B} f, \;\; \mathbb{B} \delta_1 f = \delta_1\mathbb{B} f,\;\; \delta_1 V_s f = V_1f

    for every f\in{\rm L}^p([0,1],\mathbb{C}^m) while

    \delta_1^k f = \delta_1f =f(1)\;\;\text{for } k\ge 1.

    So, when expanding (\mathbb{B} V_s+\delta_1)^{k-1}V_s we can rearrange the terms to obtain expressions of the form

    \alpha_i \mathbb{B}^i V_{s_1}\cdots V_{s_{i+1}}, 0\le i\le k-1,

    where \alpha_i are scalar coefficients and s_j\in\{s,1\}, 1\le j\le i+1. Next, for arbitrary u\in{\rm L}^p[0,1] and 0\le k\le l we have

    V_{s_1}\cdots V_{s_k} u \in {\rm W}^{k,p}[0,1], \;\; s_j\in\{s,1\}, 1\le j\le k.

    Combining these facts we obtain the desired result by considering (33) for all u\in{\rm L}^{p}[0,l].

    By the previous Corollary we immediately obtain the following result which improves [17,Thm. 3.10] and shows that \bigl[a\mathcal{R}_t^{\rm{BC}}\bigr]_1 is constant for t\ge\min\{m,n\}=:l.

    Corollary 5.17. If t\ge\min\{m,n\}=:l then the approximate controllability space of the controlled flow with dynamic boundary conditions (28) is given by

    \begin{align*} \bigl[a\mathcal{R}_t^{BC}\bigr]_1 &={\rm L}^p[0,1]\otimes{\rm{span}}\left\{\Psi v,\mathbb{B}\,\Psi v,\ldots,\mathbb{B}^{l-1}\Psi v\right\}\\ &={\rm L}^p[0,1]\otimes\Psi{\rm{span}}\left\{v,\mathbb{A} v,\ldots,\mathbb{A}^{l-1}v\right\}. \end{align*}

    In the same manner as before we also obtain the following result on positive controllability.

    Corollary 5.18. The approximate positive controllability space of the controlled flow with dynamic boundary conditions (28) is given by

    \begin{align*} \bigl[a^+\mathcal{R}^{BC}\bigr]_1 &={\rm L}^p[0,1]\otimes\overline{\rm{co}}\left\{\mathbb{B}^k\Psi v : k\in\mathbb{N}_0\right\}\\ &={\rm L}^p[0,1]\otimes\Psi\,\overline{\rm{co}}\left\{\mathbb{A}^k v: k\in\mathbb{N}_0\right\}. \end{align*}

    Using a new characterization of admissible boundary control operators (see Proposition 3.1) we are able to describe explicitly the exact reachability space of the abstract boundary control system \sum\nolimits_{{\rm{BC}}} {\left( {{A_m},B,Q} \right)}, cf. (1). Moreover, this approach allows also to determine the positive reachability space obtained considering only positive control functions. Our results generalize and improve the ones obtained in the former works [7,16,17] where only approximate controllability or positive controllability under quite restrictive assumptions are studied.

    [1] Perturbation of analytic semigroups and applications to partial differential equations. J. Evol. Equ. (2016) 1-26.
    [2] W. Arendt, A. Grabosch, G. Greiner, U. Groh, H. P. Lotz, U. Moustakas, R. Nagel, F. Neubrander and U. Schlotterbeck, One-Parameter Semigroups of Positive Operators, vol. 1184 of Lecture Notes in Mathematics, Springer-Verlag, Berlin, 1986. doi: 10.1007/BFb0074922
    [3] W. Arendt, C. J. K. Batty, M. Hieber and F. Neubrander, Vector-Valued Laplace Transforms and Cauchy Problems, vol. 96 of Monographs in Mathematics, 2nd edition, Birkhäuser/Springer Basel AG, Basel, 2011. doi: 10.1007/978-3-0348-0087-7
    [4] Asymptotic state lumping in transport and diffusion problems on networks with applications to population problems. Math. Models Methods Appl. Sci. (2016) 26: 215-247.
    [5] Gas flow in pipeline networks. Netw. Heterog. Media (2006) 1: 41-56.
    [6] A. Bátkai, M. Kramar Fijavž and A. Rhandi, Positive Operator Semigroups: From Finite to Infinite Dimensions, vol. 257 of Operator Theory: Advances and Applications, Birkhäuser Basel, 2017. doi: 10.1007/978-3-319-42813-0
    [7] Approximate positive controllability of positive boundary control systems. Positivity (2014) 18: 375-393.
    [8] Flows on networks: Recent results and perspectives. EMS Surv. Math. Sci. (2014) 1: 47-111.
    [9] Traffic flow on a road network. SIAM J. Math. Anal. (2005) 36: 1862-1886.
    [10] R. F. Curtain and H. Zwart, An Introduction to Infinite-Dimensional Linear Systems Theory, vol. 21 of Texts in Applied Mathematics, Springer-Verlag, New York, 1995. doi: 10.1007/978-1-4612-4224-6
    [11] C. D'Apice, S. Göttlich, M. Herty and B. Piccoli, Modeling, Simulation, and Optimization of Supply Chains, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2010. doi: 10.1137/1.9780898717600
    [12] Semigroups for flows in infinite networks. Semigroup Forum (2008) 76: 341-356.
    [13] The semigroup approach to transport processes in networks. Phys. D (2010) 239: 1416-1421.
    [14] Spectral theory and generator property for one-sided coupled operator matrices. Semigroup Forum (1999) 58: 267-295.
    [15] Generator property and stability for generalized difference operators. J. Evol. Equ. (2013) 13: 311-334.
    [16] Vertex control of flows in networks. Netw. Heterog. Media (2008) 3: 709-722.
    [17] Maximal controllability for boundary control problems. Appl. Math. Optim. (2010) 62: 205-227.
    [18] K. -J. Engel and R. Nagel, One-Parameter Semigroups for Linear Evolution Equations, vol. 194 of Graduate Texts in Mathematics, Springer-Verlag, New York, 2000. doi: 10.1007/b97696
    [19] M. Garavello and B. Piccoli, Traffic Flow on Networks, vol. 1 of AIMS Series on Applied Mathematics, American Institute of Mathematical Sciences (AIMS), Springfield, MO, 2006.
    [20] Perturbing the boundary conditions of a generator. Houston J. Math. (1987) 13: 213-229.
    [21] Optimal control for traffic flow networks. J. Optim. Theory Appl. (2005) 126: 589-616.
    [22] Flow control in gas networks: Exact controllability to a given demand. Math. Methods Appl. Sci. (2011) 34: 745-757.
    [23] Modeling, simulation, and optimization of traffic flow networks. SIAM J. Sci. Comput. (2003) 25: 1066-1087.
    [24] A mathematical model of traffic flow on a network of unidirectional roads. SIAM J. Math. Anal. (1995) 26: 999-1017.
    [25] Difference operators as semigroup generators. Semigroup Forum (2010) 81: 461-482.
    [26] Spectral properties and asymptotic periodicity of flows in networks. Math. Z. (2005) 249: 139-162.
    [27] On the modelling and stabilization of flows in networks of open canals. SIAM J. Control Optim. (2002) 41: 164-180.
    [28] Exact boundary controllability of unsteady flows in a network of open canals. Math. Nachr. (2005) 278: 278-289.
    [29] D. Mugnolo, Semigroup Methods for Evolution Equations on Networks, Understanding Complex Systems, Springer, Cham, 2014. doi: 10.1007/978-3-319-04621-1
    [30] Flows in networks with dynamic ramification nodes. J. Evol. Equ. (2005) 5: 441-463.
    [31] Admissibility of unbounded control operators. SIAM J. Control Optim. (1989) 27: 527-545.
  • This article has been cited by:

    1. S. G. Pyatkov, Solvability of initial boundary value problems for non-autonomous evolution equations, 2020, 20, 1424-3199, 39, 10.1007/s00028-019-00516-6
    2. Marjeta Kramar Fijavž, Delio Mugnolo, Serge Nicaise, Linear hyperbolic systems on networks: well-posedness and qualitative properties, 2021, 27, 1292-8119, 7, 10.1051/cocv/2020091
    3. L. Arlotti, B. Lods, An L^{p}-Approach to the Well-Posedness of Transport Equations Associated with a Regular Field: Part I, 2019, 16, 1660-5446, 10.1007/s00009-019-1425-8
    4. Yassine El Gantouh, Said Hadd, Abdelaziz Rhandi, Approximate controllability of network systems, 2021, 10, 2163-2480, 749, 10.3934/eect.2020091
    5. Yassine El Gantouh, Said Hadd, Well-posedness and approximate controllability of neutral network systems, 2021, 16, 1556-1801, 569, 10.3934/nhm.2021018
    6. L. Arlotti, B. Lods, An L^{p}-Approach to the Well-Posedness of Transport Equations Associated to a Regular Field: Part II, 2019, 16, 1660-5446, 10.1007/s00009-019-1426-7
    7. Marjeta Kramar Fijavž, Aleksandra Puchalska, Semigroups for dynamical processes on metric graphs, 2020, 378, 1364-503X, 20190619, 10.1098/rsta.2019.0619
    8. Christian Budde, Bálint Farkas, A Desch–Schappacher perturbation theorem for bi‐continuous semigroups, 2020, 293, 0025-584X, 1053, 10.1002/mana.201800534
    9. Mohamed El Azzouzi, Abdellah Lourini, Mohamed Laabissi, Null Controllability of Networks Systems, 2022, 1079-2724, 10.1007/s10883-022-09623-z
    10. S. G. Pyatkov, On Some Inverse Problems for First Order Operator-Differential Equations, 2019, 60, 0037-4466, 140, 10.1134/S0037446619010154
    11. Christian Budde, Positive Desch-Schappacher perturbations of bi-continuous semigroups on AM-spaces, 2021, 87, 0001-6969, 571, 10.14232/actasm-021-914-5
    12. Yacine Chitour, Sébastien Fueyo, Guilherme Mazanti, Mario Sigalotti, Approximate and exact controllability criteria for linear one-dimensional hyperbolic systems, 2024, 190, 01676911, 105834, 10.1016/j.sysconle.2024.105834
  • Reader Comments
  • © 2017 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4710) PDF downloads(188) Cited by(12)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog