1 Introduction

We study genericity of controllability and stabilizability of differential algebraic systems described by the equation

$$\begin{aligned} \tfrac{\mathrm {d}}{\mathrm {d}t}(Ex) = Ax + Bu, \end{aligned}$$
(1)

where

$$\begin{aligned} (E,A,B)\in \quad \Sigma _{\ell ,n,m}\ := \ \mathbb {R}^{\ell \times n}\times \mathbb {R}^{\ell \times n}\times \mathbb {R}^{\ell \times m}. \end{aligned}$$

To be precise, we first say what we understand under genericity.

Definition 1.1

[13, p. 28] and [11, p. 50] A set \( \mathbb {V} \subseteq \mathbb {R}^n\) is called an algebraic variety, if there exist finitely many polynomials

$$\begin{aligned} p_1(x_1,\ldots , x_n),\ldots ,p_k(x_1,\ldots , x_n) \in \mathbb {R}[x_1,\ldots , x_n] \end{aligned}$$

such that \(\mathbb {V}\) is the locus of their zeros, i.e.,

$$\begin{aligned} \mathbb {V} = \left\{ x\in \mathbb {R}^n\ \big \vert \, \forall \, i\in \underline{k}: p_i(x)=0 \right\} = \bigcap _{i=1}^k p_i^{-1}(\left\{ 0 \right\} ). \end{aligned}$$
(2)

An algebraic variety \( \mathbb {V}\) is called proper if  \( \mathbb {V} \subsetneq \mathbb {R}^n\). The set of all algebraic varieties in \(\mathbb {R}^n\) is denoted as

$$\begin{aligned} \mathscr {V}_n(\mathbb {R}) := \left\{ \mathbb {V}\subseteq \mathbb {R}^n\,\left| \,\exists \, q_1(\cdot ),\ldots ,q_k(\cdot )\in \mathbb {R}[x_1,\ldots , x_n]: \bigcap _{i=1}^k q_i^{-1}(\left\{ 0 \right\} ) = \mathbb {V}\right. \right\} \end{aligned}$$
(3)

and the set of all proper algebraic varieties as

$$\begin{aligned} \mathscr {V}_n^{\text {prop}}(\mathbb {R}) := \mathscr {V}_n(\mathbb {R}){\setminus }\left\{ \mathbb {R}^n \right\} . \end{aligned}$$
(4)

A set \(S\subseteq \mathbb {R}^n\) is called generic, if there exists a proper algebraic variety \(\mathbb {V}\in \mathscr {V}_n^{\text {prop}}(\mathbb {R})\) so that \(S^c\subseteq \mathbb {V}\). If the algebraic variety \(\mathbb {V}\) is known, then we call S generic with respect to (w.r.t.) \(\mathbb {V}\). \(\diamond \)

“Generic” is not consistently used in the literature. We show in the following proposition that generic as in Definition 1.1 is stronger than containing an open and dense subset with respect to the Euclidean topology.

Proposition 1.2

Any set \(S\subseteq \mathbb {R}^n\) satisfies:

  1. (i)

    If \(\mathbb {V}\subseteq \mathbb {R}^n\) is a proper algebraic variety, then \(\mathbb {V}^c\) is open and dense.

  2. (ii)

    If \(S\subseteq \mathbb {R}^n\) is generic w.r.t. \(\mathbb {V}\in \mathscr {V}_n^{\text {prop}}(\mathbb {R})\), then S contains the open and dense subset \(\mathbb {V}^c\).

  3. (iii)

    There exists an open and dense set \(S\subseteq \mathbb {R}^n\) with \(\lambda ^n(S)<\infty \), and therefore S is neither generic nor is \(S^c\)a proper algebraic variety. In other words, the reverse implications in (i) and (ii) do not hold true.

Proof

(i) We show that \(\mathbb {V}\) is closed. Let \(\mathbb {V}\) be given as in (2). Then, each \(p_i^{-1}(\left\{ 0 \right\} )\) is closed since \(p_i\) is continuous and hence, the claim follows.

It remains to prove that \(\mathbb {V}^c\) is dense. Seeking a contradiction, suppose that \(\mathbb {V}^c\) is not dense or, equivalently, \(\mathbb {V}\) has at least one inner point. Then, \(\lambda ^n(\mathbb {V})>0\) and \(\mathbb {V}\) cannot be a proper algebraic variety by Proposition A.3.

(ii) This is an immediate consequence of (i).

(iii) Let \(\varphi : \mathbb {N}\rightarrow \mathbb {Q}^n\) be a bijection and set

$$\begin{aligned} S := \bigcup _{i\in \mathbb {N}} \left\{ q \in \mathbb {R}^n \left| \ \left\| q- \varphi (i) \right\| < 42^{-i} \right. \right\} \supseteq \mathbb {Q}^n. \end{aligned}$$

Then, \(S\subseteq \mathbb {R}^n\) is open, dense and satisfies

$$\begin{aligned} \lambda ^n(S) \le \sum _{i\in \mathbb {N}}\frac{2^n}{42^{i\,n}}\le 2^n\sum _{i\in \mathbb {N}}\frac{1}{42^i} = 2^n \frac{42}{41}<\infty . \end{aligned}$$

Now, Lemma A.7 shows that S is not generic.

If \(S^c=\mathbb {V}\) were an algebraic variety, then S is generic w.r.t. \(\mathbb {V}\) which is a contradiction. This completes the proof. \(\square \)

To characterize genericity in terms of the Zariski topology, recall [11, p. 50] that the latter is defined by the property that all closed sets are the algebraic varieties. The Zariski topology is strictly coarser than the Euclidean topology and we have: A set \(S\subseteq \mathbb {R}^n\) is generic if, and only if, S contains a nonempty Zariski open set.

This approach was used by Belur and Shankar in their investigations of genericity of impulse controllable systems (see [3, Section 3]). Since they consider differential-algebraic equations described by differential operator matrices and hence an infinite dimensional vector space, they need to extend the definition of generic sets to this space using the limit topology of the Zariski topology. This is not necessary in our setup.

In the special case that (1) is an ordinary differential equation, that is \(\ell =n\) and \(E=I\), Lee and Markus [10] proved that that the set of all controllable systems is open and dense w.r.t. the Euclidean topology. Wonham [12, Thm. 1.3] showed in the first edition of his monograph that the set of all controllable systems is generic.

Recently, it has been shown that linear, time-invariant port-Hamiltonian systems are generically controllable; see [9].

When it comes to differential-algebraic equations, then to the best of our knowledge there are only two contributions known where open and dense subsets of controllable systems are investigated. Banaszuk and Przyłuski [1] consider an algebraic criterion—which they do not justify analytically and which is not related to any concept of controllability—and give a sufficient condition so that the set of systems satisfying the algebraic criterion contains an open and dense subset. The second contribution is by Belur and Shankar [3].

Their main interest is on polynomial systems, and if specialized to matrix pencils, they derive a characterization of genericity of impulse controllability. Other concepts are not studied.

The basis of our approach is the algebraic characterizations of various concepts of controllability and stabilizability of differential-algebraic equations; this is well known and summarized in Propositions 2.1 and 3.1. We characterize—in terms of the formats \(\ell \), n, and m of (1)—when these controllability and stabilizability concept hold generically. This is the content of Theorems 2.3 and 3.3. The proofs of these two main results are based on methods from algebraic geometry, tailored for our purposes and relegated to “Appendix 1”, and some results on ranks of special matrices are presented in “Appendix 2”.

2 Controllability

There are various controllability concepts for differential-algebraic equations (1) such as freely initializable (also called ‘controllable at infinity’), impulse controllable, completely controllable, behavioral controllable, and strongly controllable. Their definitions and their algebraic characterizations are given in the next proposition. To state this, we need to say what a solution of (1) is. We consider the behavior of (1) given by

$$\begin{aligned} \mathfrak {B}_{[E,A,B]} := \left\{ \ (x,u)\in {\mathscr {W}^{1,1}_{\text {loc}}(\mathbb {R},\mathbb {R}^n)} \times \mathscr {L}^1_{\text {loc}}(\mathbb {R},\mathbb {R}^m) \ \left| \ \begin{array}{l} Ex \text { is absolutely continuous} \\ \text {and for a.a.}~t\in \mathbb {R} \ :\\ \frac{\text { d}}{\text { d}t}(Ex)(t) = Ax(t) + Bu(t) \end{array} \right. \right\} \end{aligned}$$

where \(\mathscr {L}^1_{\text {loc}}(\mathbb {R},\mathbb {R}^d)\) denotes the set of locally integrable functions \(f:\mathbb {R}\rightarrow {\mathbb {R}}^d\), and \(\mathscr {W}^{1,1}_{\text {loc}}(\mathbb {R},\mathbb {R}^d)\) is the Sobolev space of all functions \(f\in \mathscr {L}^1_{\text {loc}}(\mathbb {R},\mathbb {R}^d)\) with \(f^{(1)}\in \mathscr {L}^1_{\text {loc}}(\mathbb {R},\mathbb {R}^d)\). Note that any \(f\in \mathscr {W}^{1,1}_{\text {loc}}(\mathbb {R},\mathbb {R}^d)\) is continuous.

Controllability of a system (1) is a property of the corresponding behavior \(\mathfrak {B}_{[E,A,B]}\). If systems described by ordinary differential equations are considered, i.e., the special case \(E = I\), then the initial value \(x^0\in \mathbb {R}^n\) can be freely chosen and the problem is to which other points it can be steered in finite time. It is well-known that the system (1) with \(E=I\) is called controllable if, and only if, for any given initial state \(x^0\in \mathbb {R}^n\) and any terminal state \(x^1\in \mathbb {R}^n\), there exists a control \(u\in \mathscr {L}^1_{\text {loc}}(\mathbb {R},\mathbb {R}^m)\) which steers \(x^0\) to \(x^1\) in finite time \(T>0\), more formally, there exists \((x,u)\in \mathfrak {B}_{[E,A,B]}\) such that \(x(0) = x^0\) and \(x(T) = x^1\).

However, if an arbitrary matrix E is allowed in (1), then algebraic constraints are added to the differential equation. So it is unclear as to whether the initial value can be chosen freely. If the latter is the case, then the system is called freely initializable, sometimes also called controllable at infinity.

If (1) is assumed to be freely initializable, then one may ask, whether each initial state can be steered to any final state in finite time. If both conditions are fulfilled, then the system is called completely controllable.

A stronger controllability concept—but also a generalization of the ODE case—is in the behavioral setup the problem as to whether it is always possible to concatenate two given solutions \((x_1,u_1), (x_2,u_2)\in \mathfrak {B}_{[E,A,B]}\) by another solution \((x,u)\in \mathfrak {B}_{[E,A,B]}\) over the time interval [0, T]. Such systems are called behavioral controllable systems.

The concepts of freely initializable and completely controllable systems can be weakened in the sense that the initial and the terminal value are compared with respect to the image of E; for example, the initial condition becomes \(Ex^0=Ex(0)\). These weakened concepts are called impulse controllable and strongly controllable.

The precise definitions and algebraic characterizations are given in the following proposition.

Proposition 2.1

For any \((E,A,B)\in \Sigma _{\ell ,n,m}\), the following controllability definitions associated with the system (1) are algebraic characterized as follows:

$$\begin{aligned} \begin{array}{lcclcl} \text {freely initializable} &{}:=&{}&{} \forall \, x^0\in \mathbb {R}^n~\exists \, (x,u)\in \mathfrak {B}_{[E,A,B]} : \ x(0) = x^0\\ &{}\iff &{}&{} \mathrm {rk}\,[E,B] = \mathrm {rk}\,[E,A,B]\ ;\\ \text {impulse controllable} &{}:=&{}&{} \forall \, x^0\in \mathbb {R}^n~\exists \, (x,u)\in \mathfrak {B}_{[E,A,B]} : \ Ex^0 = Ex(0)\\ &{}\iff &{} &{} \forall \, Z\in \mathbb {R}^{n\times n-\mathrm {rk}\,E}~~~\text {with}~~~\mathrm {im}\,_{\mathbb {R}} Z = \ker _{\mathbb {R}}E \\ &{}&{}&{} : \ \mathrm {rk}\,[E,A,B] = \mathrm {rk}\,[E,AZ,B]\ ;\\ \text {behavioral controllable} &{}:=&{}&{} \forall \, (x_1,u_1),(x_2,u_2)\in \mathfrak {B}_{[E,A,B]} \ \exists \, T>0~\exists \, (x,u)\in \mathfrak {B}_{[E,A,B]}\\ &{}&{}&{}: \ (x,u)(t) = {\left\{ \begin{array}{ll}(x_1,u_1)(t), &{} t<0\\ (x_2,u_2)(t), &{} t>T\end{array}\right. }\\ &{}\iff &{}&{} \forall \, \lambda \in \mathbb {C}\ : \mathrm {rk}\,_{\mathbb {R}(s)}[sE-A,B] = \mathrm {rk}\,_{\mathbb {C}}[\lambda E-A,B]\ ;\\ \text {completely controllable} &{}:=&{}&{} \exists \, T>0~\forall \, x^0,x_T\in \mathbb {R}^n \exists \, (x,u)\in \mathfrak {B}_{[E,A,B]}\\ &{}&{}&{} : \ x(0) = x^0\wedge x(T) = x_T \\ &{}\iff &{}&{} \forall \,\lambda \in \mathbb {C}: \mathrm {rk}\,[E,A,B] = \mathrm {rk}\,[E,B] = \mathrm {rk}\,[\lambda E-A,B] \ ;\\ \text {strongly controllable} &{}:=&{}&{} \exists \, T>0~\forall \, x^0,x_T\in \mathbb {R}^n ~\exists \, (x,u)\in \mathfrak {B}_{[E,A,B]}\\ &{}&{}&{} :\ Ex(0) = Ex^0 \ \wedge \ Ex(T) = Ex_T\\ &{}\iff &{}&{} \forall \,\lambda \in \mathbb {C}~\forall \, Z\in \mathbb {R}^{n\times n-\mathrm {rk}\,E}\ \text {with}~\mathrm {im}\,_{\mathbb {R}} Z = \ker _{\mathbb {R}}E\\ &{}&{}&{} : \ \mathrm {rk}\,[E,A,B] = \mathrm {rk}\,[E,AZ,B] = \mathrm {rk}\,[\lambda E-A,B] \ .\\ \end{array} \end{aligned}$$

Proof

Berger and Reis [4] derive a feedback form and use this as a tool in conjunction with ‘canonical’ representatives of certain equivalence classes to prove all characterizations of controllability in their survey. Note that in their characterization of strongly controllability, the term ‘\(+\mathrm {im}\,_{\mathbb {R}}E\)’ is missing in the first respective line in [4, Cor. 4.3]. \(\square \)

Remark 2.2

The equivalences

$$\begin{aligned} (E,A,B)~\text {str. contr.}\iff (E,A,B)~\text {imp. contr. and beh. contr.} \end{aligned}$$

and

$$\begin{aligned} (E,A,B)~\text {compl. contr.}\iff (E,A,B)~\text {freely initial. and beh. contr.} \end{aligned}$$

are proved in [4, Rem. 4.4]. \(\square \)

Genericity of the different controllability concepts can be characterized in terms of the system dimensions. To this end, we introduce the notation

$$\begin{aligned} S_{\text {controllable}} \ := \ \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \ (1)~\text {is controllable} \right\} \end{aligned}$$

where ‘controllable’ stands for one of the controllability concepts.

Theorem 2.3

For each of the controllability concepts defined in Proposition 2.1, the following characterizations hold:

$$\begin{aligned} \begin{array}{llclcl} S_{\text {freely initial.}} &{} \text {is generic} &{}\qquad \Longleftrightarrow &{} \qquad \ell \le n+m\ ; \\ S_{\text {imp. contr.}}&{} \text {is generic} &{}\qquad \Longleftrightarrow &{} \qquad \ell \le n+m\ ; \\ S_{\text {beh. contr.}} &{} \text {is generic} &{}\qquad \Longleftrightarrow &{} \qquad \ell \not = n+m \ ; \\ S_{\text {compl. contr.}}&{} \text {is generic} &{}\qquad \Longleftrightarrow &{} \qquad \ell< n+m \ ; \\ S_{\text {strongly. contr.}} &{} \text {is generic} &{}\qquad \Longleftrightarrow &{} \qquad {\ell <n+m}\ . \end{array} \end{aligned}$$

Proof

We proceed in steps.

Step 1 We show: \(S_{\text {freely initial.}} \) is generic if, and only if, \( \ell \le n+m\).

First note that Proposition 2.1 yields

\(S_{\text {freely initial.}} = \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \,\mathrm {rk}\,[E,A,B] = \mathrm {rk}\,[E,B] \right\} \).

\(\implies \) Let \(\ell >n+m\). By Proposition B.3 (i) and (ii), the sets

$$\begin{aligned} \begin{array}{rcl} S_\mathrm{(ii)} &{} := &{} \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \ \mathrm {rk}\,[E,B] = n+m \right\} ,\\ S_\mathrm{(i)} &{} := &{} \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \ \mathrm {rk}\,[E,A,B] = \min \left\{ \ell , 2n+m \right\} \right\} . \end{array} \end{aligned}$$

are generic sets. Hence, Corollary A.5(ii) implies that \(S_\mathrm{(i)}\cap S_\mathrm{(ii)}\) is a generic set. If \((E,A,B)\in S_\mathrm{(i)}\cap S_\mathrm{(ii)} \), then \(\ell >n+m\) yields

$$\begin{aligned} \mathrm {rk}\,[E,B] = n+m < \min \left\{ \ell , 2n+m \right\} = \mathrm {rk}\,[E,A,B] , \end{aligned}$$

and therefore \(S_\mathrm{(i)}\cap S_\mathrm{(ii)} \subseteq S_{\text {freely initial.}}^c\) and \(S_{\text {freely initial.}}^c\) is generic. Thus, Lemma A.6 shows that \(S_{\text {freely initial.}}\) is not generic.

\(\Longleftarrow \) Since \(\ell \le n+m\) , the sets

$$\begin{aligned} \begin{array}{rcl} S_\mathrm{(ii)} &{} := &{} \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \ \mathrm {rk}\,[E,B] = \ell \right\} ,\\ S_\mathrm{(i)} &{} := &{} \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \ \mathrm {rk}\,[E,A,B] = \ell \right\} \end{array} \end{aligned}$$

are both non-empty and by Proposition B.3 (i) and (ii) they are generic. Now, Corollary A.5(ii) yields that \(S_\mathrm{(i)} \cap S_\mathrm{(ii)}\) is a generic set, and by Remark A.1, \(S_{\text {freely initial.}} \supseteq S_1 \cap S_2\) is generic, too.

Step 2 We show: \(S_{\text {imp. contr.}}\) is generic if, and only if, \( \ell \le n+m\).

We consider the two cases \(\ell \ge n\) and \(\ell <n\).

\(\ell \ge n\): By Proposition B.3 (i) and (iii), the sets

$$\begin{aligned} \begin{array}{rcl} S_\mathrm{(i)} &{} := &{}\left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \,\mathrm {rk}\,[E,A,B] = \min \left\{ \ell ,2n+m \right\} \right\} ,\\ S_\mathrm{(iii)} &{} := &{} \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \,\mathrm {rk}\,E = n \right\} \end{array} \end{aligned}$$

are generic and thus Corollary A.5 implies that \(S_\mathrm{(i)}\cap S_\mathrm{(iii)}\) is also generic. For each \((E,A,B)\in S_2\), we find by the rank-nullity theorem that \(\ker E = \left\{ 0 \right\} \). Hence, Proposition 2.1 yields that

$$\begin{aligned} S_{\text {imp. contr.}}\cap S_\mathrm{(i)}\cap S_\mathrm{(iii )} \ {=} \ \underbrace{\left\{ (E,A,B)\in \Sigma _{n,n,m}\ \big \vert \,\mathrm {rk}\,[E,B] = \min \left\{ \ell ,2n+m \right\} \right\} }_{{=:\widetilde{S}}}{\cap S_\mathrm{(i)}\cap S_\mathrm{(iii)}}. \end{aligned}$$

Invoking Lemma A.8 gives that \(S_{\text {imp. contr.}}\) is generic if, and only if, \(\widetilde{S}\) is generic. By Proposition B.3, this is the case if, and only if, \(\min \left\{ \ell ,2n+m \right\} \le \min \left\{ \ell ,n+m \right\} \). This inequality holds true if, and only if, \(\ell \le n+m\).

\(\ell < n\):    Put \(\widehat{S} := \left\{ (E,A,B)\in \Sigma _{n,n,m}\ \big \vert \,\mathrm {rk}\,[E,B] = \ell = \min \left\{ \ell ,n+m \right\} \right\} \), which is generic by Proposition B.3 (ii). Note that both statements

$$\begin{aligned} \forall \,(E,A,B)\in \Sigma _{\ell ,n,m}~\forall \, Z\in \mathbb {R}^{n\times n-\mathrm {rk}\,E}~\text {with}~\mathrm {im}\,_{\mathbb {R}} Z = \ker _{\mathbb {R}}E \ :&\ \mathrm {rk}\,[E,AZ,B]\le \ell ,\\ \forall \,(E,A,B)\in \widehat{S}\qquad \forall \, Z\in \mathbb {R}^{n\times n-\mathrm {rk}\,E}~\text {with}~\mathrm {im}\,_{\mathbb {R}} Z = \ker _{\mathbb {R}}E \ :&\ \mathrm {rk}\,[E,AZ,B]=\ell \end{aligned}$$

hold true. By Proposition 2.1, we find that \(\widehat{S}\cap S_{(i)]}\subseteq S_{\text {imp. contr.}}\) and hence, in view of Corollary A.5(ii) and Remark A.1, the set \(S_{\text {imp. contr.}}\) is generic.

Step 3 Proposition 2.1 yields

$$\begin{aligned} S_{\text {beh. contr.}} = \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\,\left| \ \begin{array}{l} \forall \lambda \in \mathbb {C}:\ \mathrm {rk}\,_{\mathbb {R}(s)}[sE-A,B] \\ = \mathrm {rk}\,_{\mathbb {C}}[\lambda E-A,B] \end{array}\right. \right\} . \end{aligned}$$
(5)

We show: \(S_{\text {beh. contr.}}\) is generic if, and only if, \(\ell \not = n+m\).

From Proposition B.5, we find that the set

$$\begin{aligned} S_{{3}} := \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \ \mathrm {rk}\,_{\mathbb {R}(s)}[sE-A,B] = \min \left\{ \ell ,n+m \right\} \right\} \end{aligned}$$

is a generic set. The equation (5) implies

$$\begin{aligned} S_{\text {beh. contr.}} \cap S_{{3}} = \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \ \forall \lambda \in \mathbb {C}:\mathrm {rk}\,_{\mathbb {C}}[\lambda E-A,B] = \min \left\{ \ell ,n+m \right\} \right\} . \end{aligned}$$

By Lemma A.8, genericity of \(S_{\text {beh. contr.}}\cap S_{{3}}\) is necessary and sufficient for \(S_{\text {beh. contr.}}\) being generic. Now, Proposition B.8 gives: \(S_{\text {beh. contr.}}\cap S_1\) is generic if, and only if, \(\ell \ne n+m\).

Step 4 Since

$$\begin{aligned} S_{\text {strongly contr.}} \ = \ S_{\text {imp. contr.}} \cap S_{\text {beh. contr.}}, \end{aligned}$$

Corollary A.5 (ii) implies that \(S_{\text {strongly contr.}}\) is generic if, and only if, both \(S_{\text {imp. contr.}}\) and \( S_{\text {beh. contr.}}\) are generic. In view of Step 2 and Step 3, this is the case if, and only if, \(\ell < n+m\).

Step 5 Applying Corollary A.5 (ii) to the equality

$$\begin{aligned} S_{\text {compl. contr.}} = S_{\text {freely initial.}}\cap S_{\text {beh. contr.}}, \end{aligned}$$

we find with Step 1 and Step 3 that \(S_{\text {compl. contr.}}\) is generic if, and only if, \(\ell <n+m\). \(\square \)

In the following remark, we stress the observation that linear differential-algebraic systems are either generically controllable or generically not controllable.

Remark 2.4

A closer inspection of the proof of Theorem 2.3 yields that if \(S_{\text {arbitrary~controllability}}\) is not generic, then it is contained in a proper algebraic variety and thus its complement is generic.

3 Stabilizability

In the present section, genericity of stabilizability of DAEs is studied.

In the ODE-case, a system (1) is called stabilizable if, and only if, for each initial value \(x^0\in {\mathbb {R}}^n\) there exists a control \(u\in \mathscr {L}^1_{\text {loc}}(\mathbb {R},\mathbb {R}^m)\) which steers the forced trajectory x in (possibly) infinite time to zero, that is \((x,u)\in \mathfrak {B}_{[I,A,B]}\) so that \(x(0) = x^0\) and \(\lim _{t\rightarrow \infty }\text {ess sup}\, x|{ (t,\infty )|} = 0\).

As for controllability, different generalizing concepts have to be studied for DAEs. Each system that is stabilizable in the ODE-sense is called completely stabilizable. Similar to controllability, this concept is weakened if only Ex(t) is considered—in this case we speak of strong stabilizability.

Finally, a system is called behavioral controllable if, and only if, each \((x,u)\in \mathfrak {B}_{[E,A,B]}\) can be concatenated with some \((x,u)\in \mathfrak {B}_{[E,A,B]}\) which tends to zero as t tends to infinity.

The precise definitions and algebraic characterizations are given in the following proposition. We write \(\mathscr {W}^{1,1}_{\text {loc}}(I,\mathbb {R}^n)\) for the set of all weakly differentiable \(\varphi \in \mathscr {L}^1_\text {loc}({\mathbb {R}},\mathbb {R}^n)\).

Proposition 3.1

For any \((E,A,B)\in \Sigma _{\ell ,n,m}\), the following controllability definitions associated with the system (1) are algebraic characterized as follows:

$$\begin{aligned} \begin{array}{lclcl} \text {compl. stabl.} &{}:=&{} \forall \, x^0\in \mathbb {R}^n~\exists \, (x,u)\in \mathfrak {B}_{[E,A,B]} \ x(0) = x^0 \ \wedge \lim _{t\rightarrow \infty }\text {ess sup}\, x|_{~_{| (t,\infty )}} = 0\\ &{} \iff &{} \forall \,\lambda \in \overline{\mathbb {C}}_{+}\ : \mathrm {rk}\,[E,A,B] = \mathrm {rk}\,[E,B] =\mathrm {rk}\,[\lambda E-A,B];\\ \text {str. stabl.} &{}:=&{} \forall \, x^0\in \mathbb {R}^n~\exists \, (x,u)\in \mathfrak {B}_{[E,A,B]} : Ex(0) = Ex^0 \quad \wedge \lim _{t\rightarrow \infty }Ex(t) = 0\\ &{}\iff &{} \forall \,\lambda \in \overline{\mathbb {C}}_+~\forall \,Z~\text {with}~\mathrm {im}\,Z = \ker E\\ &{}&{} :\ \mathrm {rk}\,[E,A,B] = \mathrm {rk}\,[E,AZ,B] = \mathrm {rk}\,[\lambda E-A,B] ; \\ \text {beh. stabl.} &{}:=&{} \forall \, (x,u)\in \mathfrak {B}_{[E,A,B]}~\exists \, (x_1,u_1) \in \mathfrak {B}_{[E,A,B]}\cap \left( \mathscr {W}^{1,1}_\mathrm{{loc}}(\mathbb {R},\mathbb {R}^n)\times \mathscr {W}^{1,1}_\mathrm{{loc}}(\mathbb {R},\mathbb {R}^m)\right) \\ &{}&{} : \ \left[ \forall \, t<0:(x(t),u(t)) = (x_1(t),u_1(t))\right] \quad \wedge \quad \lim _{t\rightarrow \infty }(x_1(t),u_1(t)) = 0 \\ &{}\iff &{} \forall \,\lambda \in \overline{\mathbb {C}}_+: \mathrm {rk}\,_{\mathbb {R}(s)}[sE-A,B] = \mathrm {rk}\,_{\mathbb {C}}[\lambda E-A,B]. \end{array} \end{aligned}$$

Proof

All characterizations are proved in the survey article by Berger and Reis [4, Cor. 4.3]. \(\square \)

Remark 3.2

The equivalences

$$\begin{aligned} (E,A,B)~\text {compl. stabl.}\iff (E,A,B)~\text {beh. stab. and freely init.} \end{aligned}$$

and

$$\begin{aligned} (E,A,B)~\text {str. stabl.}\iff (E,A,B)~\text {beh. stab. and str. contr.} \end{aligned}$$

are proved in [4, Rem. 4.5]. \(\diamond \)

We now show how genericity of the different stability concepts can be characterized in terms of the matrix dimensions. To this end, we introduce the notion

$$\begin{aligned} S_{\text {stabliizable}} \ := \ \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \, (1)~\text {stabilizable} \right\} \end{aligned}$$

where ‘stabilizable’ stands for one of the stability concepts.

Theorem 3.3

For each of the three stabilizability concepts from Proposition 3.1, the following characterizations hold:

$$\begin{aligned} \begin{array}{llclcl} S_{\text {beh. stab.}}&{} \text {is generic} &{}\qquad \Longleftrightarrow &{} \qquad \ell \not = n+m\ ;\\ S_{\text {compl. stab.}} &{} \text {is generic} &{}\qquad \Longleftrightarrow &{} \qquad \ell<n+m\ ;\\ S_{\text {strongly stab.}}&{} \text {is generic} &{}\qquad \Longleftrightarrow &{} \qquad \ell < n+m\ . \end{array} \end{aligned}$$

Proof

Step 1 We show: \(S_{\text {beh. stab.}}\) is generic if, and only if, \(\ell \not = n+m\).

Proposition B.5 yields that the set

$$\begin{aligned} S_1 := \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \,\mathrm {rk}\,_{\mathbb {R}(s)}[sE-A,B] = \min \left\{ \ell ,n+m \right\} \right\} \end{aligned}$$

is generic. By Lemma A.8, genericity of

$$\begin{aligned} S_2 := \left\{ (E,A,B)\in \Sigma _{\ell ,n,m}\ \big \vert \,\forall \,\lambda \in \overline{\mathbb {C}}_+: \mathrm {rk}\,_{\mathbb {C}}[\lambda E-A,B] = \min \left\{ \ell ,n+m \right\} \right\} \end{aligned}$$

is a necessary and sufficient condition for genericity of \(S_{\text {beh. stab.}}\). By Proposition B.9\(S_2\) is generic if, and only if, \(\ell \ne n+m\).

Step 2 Remark 3.2 yields that

$$\begin{aligned} S_{\text {compl. stabl.}} = S_{\text {beh. stabl.}}\cap S_{\text {freely init.}} \end{aligned}$$

Thus, Corollary A.5 (ii) together with Theorem 2.3 gives that \(S_{\text {compl. stabl.}}\) is generic if, and only if, \(\ell <n+m\).

Step 3 Applying Corollary A.5 (ii) and Theorem 2.3 to the equality

$$\begin{aligned} S_{\text {str. stabl.}} = S_{\text {beh. stabl.}}\cap S_{\text {str. contr.}} \end{aligned}$$

yields that \(S_{\text {str. stabl.}}\) is generic if, and only if \(\ell <n+m\). \(\square \)

As for controllability, we would like to emphasize that linear differential-algebraic systems are either generically controllable or generically not controllable.

Remark 3.4

A closer inspection of the proof of Theorem 3.3 yields as in Remark 2.4 that \(S_{\text {stabilizable}}\) is either generic or contained in a proper algebraic variety.