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

Partial Strong Structural Controllability

Yuan Zhang, Yuanqing Xia Department of Automatic Control, Beijing Institute of Technology, Beijing, China
 Email: {zhangyuan14,xia_yuanqing}@bit.edu.cn
Abstract

This paper introduces a new controllability notion, termed partial strong structural controllability (PSSC), on a structured system whose entries of system matrices are either fixed zero or indeterminate, which naturally extends the conventional strong structural controllability (SSC) and bridges the gap between structural controllability and SSC. Dividing the indeterminate entries into two categories, generic entries and unspecified entries, a system is PSSC, if for almost all values of the generic entries in the parameter space except for a set of measure zero, and any nonzero (complex) values of the unspecified entries, the corresponding system is controllable. We highlight that this notion generalizes the generic property embedded in the conventional structural controllability for single-input systems. We then give algebraic and (bipartite) graph-theoretic necessary and sufficient conditions for single-input systems to be PSSC. Conditions for multi-input systems are subsequently given for a special case. It is shown the established results can induce a new maximum matching based criterion for SSC over the system bipartite graph representations.

keywords:
Structural controllability, strong structural controllability, generic property, maximum matching
journal: Technical Reportmyfootnotemyfootnotefootnotetext: This work was supported in part by the China Postdoctoral Innovative Talent Support Program under Grant BX20200055, the China Postdoctoral Science Foundation under Grant 2020M680016, and the National Natural Science Foundation of China under Grant 62003042. Corresponding author: Y. Xia.

1 Introduction

The past decades have witnessed an explosion of research interest into control and observation of complex networks [1, 2, 3]. This is perhaps because many real-world systems could be naturally modeled as complex dynamic networks, such as social networks, biological networks, traffic networks, as well as the internet [4]. Among the related problems, network controllability has been extensively explored both from a qualitative [5, 6, 7] and quantitative perspective [8, 3, 9].

A well-accepted alternative of the classical Kalman controllability is the notion of structural controllability. This notion was first introduced by Lin [10] by inspecting that controllability is a generic property, in the sense that depending on the structure of the state-space matrices, either the corresponding system realization is controllable for almost all values of the indeterminate entries, or there is no controllable system realization. Various criteria for structural controllability were found subsequently [11, 12]. Structural controllability is promising from a practical view, since it does not require the accurate values of system parameters, thus immune to numerical errors. Moreover, criteria for structural controllability usually are directly linked to some mild interconnection conditions of the graph associated with the system structure [13], making it attractive in analyzing large-scale network systems [1, 14, 15].

A more stringent notion than structural controllability is the strong structural controllability (SSC), which requires the system to be controllable for all (nonzero) values of the indeterminate entries [16]. As argued in [16], although the uncontrollable system realizations are atypical if a system is structurally controllable, there do exist scenarios where the existence of an uncontrollable system realization is prohibited, especially in some critical infrastructures where high-level robustness of the system controllability is required. The first criterion for SSC of single-input systems was given in [16], followed by some graph-theoretic characterizations of SSC of multi-input systems in [17, 18]. [19] provided a constrained-matching based criterion, while [20] related SSC to the zero-forcing set and graph-coloring. Further, SSC of undirected networks was studied in [21]. For a comprehensive comparison of these criteria, see [22].

Refer to caption
Figure 1: Inverted-pendulum system. Figure borrowed from [23].

It is well-accepted that being SSC requires a system to have a much more restrictive structure than being structurally controllable. In fact, it is shown in [24, 25] that, for structural controllability, the ratio of controllable graphs (networks) to the total number of graphs with nn nodes tends to one, as nn\to\infty. By contrast, [26] demonstrates that for SSC, the same ratio approaches zero, provided all nodes have self-loops and the ratio of control nodes to nn converges to zero. This phenomenon indicates that there is an essential gap between structural controllability and SSC. On the other hand, from a generic view, SSC requires that the system realization is controllable subject to all possible interrelations among the (nonzero) indeterminate parameters. In contrast, structural controllability only requires the system realization to be controllable assuming independence among all the indeterminate parameters. Note in practice, it is possible that only a partial subset of indeterminate parameters may be interrelated while the rest are independent of each other. This means, even a system itself is not SSC, the robustness requirement for system controllability might be met. The following example illustrates this.

Example 1 (Motivating example)

Consider the following
  inverted-pendulum system shown in Fig. 1, which also appears in [23, Example 3–6]. Let θ\theta and xx be respectively the angle of the pendulum rod and the location of the cart. Define state variables x1,x2,x3x_{1},x_{2},x_{3}, and x4x_{4} by x1=θx_{1}=\theta, x2=θ˙x_{2}=\dot{\theta}, x3=xx_{3}=x, x4=x˙x_{4}=\dot{x}. According to [23], the linearized state-space equation of this system reads

[x˙1x˙2x˙3x˙4]=[0100(M+m)gMl0000001mgM000]A=[aij][x1x2x3x4]+[01Ml01M]B=[bij]u{\small\left[\begin{array}[]{c}\dot{x}_{1}\\ \dot{x}_{2}\\ \dot{x}_{3}\\ \dot{x}_{4}\\ \end{array}\right]=\underbrace{\left[\begin{array}[]{cccc}0&1&0&0\\ \frac{(M+m)g}{Ml}&0&0&0\\ 0&0&0&1\\ -\frac{mg}{M}&0&0&0\\ \end{array}\right]}_{A=[a_{ij}]}\left[\begin{array}[]{c}x_{1}\\ x_{2}\\ x_{3}\\ x_{4}\\ \end{array}\right]+\underbrace{\left[\begin{array}[]{c}0\\ -\frac{1}{Ml}\\ 0\\ \frac{1}{M}\\ \end{array}\right]}_{B=[b_{ij}]}u}

where mm and MM are the masses of the pendulum bob and the cart, ll is the length of the massless rod, uu is the force imposed on the cart, and gg is the acceleration of gravity. From [16], the zero/nonzero structure of (A,B)(A,B) is not SSC. However, direct calculation shows that the determinant of the controllability matrix of (A,B)(A,B) is a123a34b212(a21b41a41b21)2a_{12}^{3}a_{34}b_{21}^{2}(a_{21}b_{41}-a_{41}b_{21})^{2}. This means, if we regard a21,a41,b41a_{21},a_{41},b_{41}, and b21b_{21} as independent parameters, leading to the usual satisfaction of a21b41a41b210a_{21}b_{41}-a_{41}b_{21}\neq 0, then whatever nonzero values a12a_{12} and a34a_{34} may take, the corresponding system will be controllable. In fact, back to this example, a21b41a41b21=MgM2l0a_{21}b_{41}-a_{41}b_{21}={\footnotesize{\frac{Mg}{M^{2}l}}}\neq 0 is always fulfilled. \square

Motivated by the above observations, this paper proposes a new controllability notion, named partial strong structural controllability (PSSC), trying to bridge the gap between structural controllability and SSC. In this notion, the indeterminate entries of a (structured) system are divided into two categories, generic entries that are assumed to take independent values, and unspecified entries that can take arbitrary nonzero (complex) values. A system is PSSC, if for almost all values of the generic entries, the corresponding system realization is controllable for all nonzero values of the unspecified entries. The main contributions are detailed as follows:

1) We propose a novel controllability notion of PSSC for linear structured systems. This notion naturally extends the conventional SSC, and for single-input systems, inherits and generalizes the generic property characterized by the conventional structural controllability.

2) We give an algebraic and a graph-theoretic necessary and sufficient conditions for single-input systems to be PSSC, the latter of which can be verified in polynomial time.

3) We extend 2) to the multi-input systems subject to the single unspecified entry constraint.

4) Finally, it is shown the established results can provide a new graph-theoretic criterion for SSC of single-input systems involving maximum matchings over the system bipartite graph representations.

Notably, our work provides a unifying viewpoint towards two seemingly different controllability notions: structural controllability and SSC. It is worth noting that a related notion is the perturbation-tolerant structural controllability proposed in [27]. Compared to that notion, PSSC addressed in this paper is an essentially different concept with many different properties. Particularly, PSSC contains the conventional SSC as a special case, while PTSC does not. Furthermore, from a technical view, due to the nonzero constraint of the unspecified entries, the techniques in [27] are not sufficient to characterize PSSC, and it is nontrivial to extend them to obtain the presented criteria here.

The rest of this paper is organized as follows. Section 2 presents the problem formulation and some preliminaries. Section 3 justifies the generic property involved in PSSC and provides necessary and sufficient conditions for single-input systems to be PSSC. Section 4 extends these results to the multi-input case subject to the single unspecified entry constraint. Section 5 discusses the implications of the established results for the existing SSC theory. All proofs of the technical results are given in the appendix.

Notations: ,,{\mathbb{N}},{\mathbb{R}},{\mathbb{C}} denote the sets of natural, real, and complex numbers, respectively. Let ¯=\{0}\bar{\mathbb{R}}={\mathbb{R}}\backslash\{0\}, ¯=\{0}\bar{\mathbb{C}}={\mathbb{C}}\backslash\{0\}, and ¯n\bar{\mathbb{C}}^{n} (resp. ¯n\bar{\mathbb{R}}^{n}) be the set of nn-dimensional complex (real) vectors with each of its entries being nonzero. For nn\in{\mathbb{N}}, JnJ_{n} stands for {1,,n}\{1,...,n\}. For an n×mn\times m matrix MM, M[I1,I2]M[I_{1},I_{2}] denotes the submatrix of MM whose rows are indexed by I1JnI_{1}\subseteq J_{n} and columns by I2JmI_{2}\subseteq J_{m}.

2 Problem formulation and preliminaries

2.1 Structured matrix and generic matrix

We often use a structured matrix to denote the sparsity pattern of a numerical matrix [28]. A structured matrix is a matrix whose entries are chosen from the set {0,,×}\{0,*,\times\}. The * entries and ×\times entries are both called indeterminate entries. We use {0,,×}p×q\{0,*,\times\}^{p\times q} to denote set of all p×qp\times q structured matrices with entries from {0,,×}\{0,*,\times\}. Moreover, for M¯{0,,×}p×q\bar{M}\in\{0,*,\times\}^{p\times q}, 𝐒M¯{\bf S}_{\bar{M}} is defined as

𝐒M¯={Mp×q:Mij0ifM¯ij=×,Mij=0ifM¯ij=0}.{\bf S}_{\bar{M}}\!=\!\{M\in{\mathbb{C}}^{p\times q}:M_{ij}\neq 0\ {\text{if}}\ \bar{M}_{ij}=\times,M_{ij}=0\ {\text{if}}\ \bar{M}_{ij}=0\}.

That is, a * entry can take either zero or nonzero, while a ×\times entry can take only nonzero value. An M𝐒M¯M\in{\bf S}_{\bar{M}} is called a realization of M¯\bar{M}. We also define two structured matrices M¯{0,}p×q\bar{M}^{*}\in\{0,*\}^{p\times q} and M¯×{0,×}p×q\bar{M}^{\times}\in\{0,\times\}^{p\times q} associated with M¯{0,,×}p×q\bar{M}\in\{0,*,\times\}^{p\times q} respectively as

M¯ij={=,ifM¯ij=×=M¯ij, else,M¯ij×={=×,ifM¯ij==M¯ij, else.\bar{M}_{ij}^{*}=\left\{\begin{aligned} &=*,{\text{if}}\ \bar{M}_{ij}=\times\\ &=\bar{M}_{ij},{\text{ else}}\end{aligned}\right.,\ \bar{M}_{ij}^{\times}=\left\{\begin{aligned} &=\times,{\text{if}}\ \bar{M}_{ij}=*\\ &=\bar{M}_{ij},{\text{ else}}.\end{aligned}\right. (1)

That is, M¯\bar{M}^{*} (resp. M¯×\bar{M}^{\times}) is obtained from M¯\bar{M} by replacing all of its indeterminate entries with * (resp. ×\times) entries.

The operation ‘+’ between two structured matrices with the same dimension is an entry-wise addition operation so that every indeterminate entry in the sum appears exactly once in one of the addends. We define a generic realization of M¯\bar{M} as the realization whose indeterminate entries are assigned with independent parameters, and call such a realization a generic matrix without specifying the corresponding structured matrix. The generic rank of a generic matrix (or structured matrix), denoted by grank(){\rm grank}(\cdot), is the maximum rank it achieves as the function of its indeterminate parameters. For a generic matrix MM and a constant matrix NN with the same dimension, M+λNM+\lambda N defines a generic matrix pencil [28], which is a matrix-valued polynomial of the independent parameters in MM and the variable λ\lambda.

2.2 Notion of PSSC

Consider a linear time invariant system as

x˙(t)=Ax(t)+Bu(t),\dot{x}(t)=Ax(t)+Bu(t), (2)

where x(t)nx(t)\in{\mathbb{C}}^{n} is the state variable, u(t)mu(t)\in{\mathbb{C}}^{m} is the input, An×nA\in{\mathbb{C}}^{n\times n} and Bn×mB\in{\mathbb{C}}^{n\times m} are respectively the state transition matrix and input matrix. For description simplicity, by the pair (A,B)(A,B) we refer to a system described by (2) that can be either single-input (i.e., m=1m=1) or multi-input (i.e., m>1m>1), while by (A,b)(A,b) we refer to a single-input system. Note dynamic systems with complex-valued system matrices and state variables are not rare in the literature [29, 30, 31, 32, 33], either for practical utility or for theoretical studies. Following a similar manner as (A,B)(A,B) in the real field [23, Chap. 9], it turns out the existence and uniqueness of solutions of system (2) are guaranteed even in the complex domain.

Denote the controllability matrix of system (2) by 𝒞(A,B){\cal C}(A,B), i.e.,

𝒞(A,B)=[B,AB,,An1B].{\cal C}(A,B)=[B,AB,\cdots,A^{n-1}B].

It is known from [29, 30, 31] that (A,B)(A,B) is controllable, if and only if 𝒞(A,B){\cal C}(A,B) is of full row rank.

Throughout this paper, given A¯{0,,×}n×n\bar{A}\in\{0,*,\times\}^{n\times n}, B¯{0,,×}n×m\bar{B}\in\{0,*,\times\}^{n\times m}, 𝒩×{\cal N}_{\times} denotes the set of positions of all ×\times entries in [A¯,B¯][\bar{A},\bar{B}], i.e., 𝒩×={(i,j):[A¯,B¯]ij=×}{\cal N}_{\times}=\{(i,j):[\bar{A},\bar{B}]_{ij}=\times\}. 𝒩={(i,j):[A¯,B¯]ij=}{\cal N}_{*}=\{(i,j):[\bar{A},\bar{B}]_{ij}=*\} is defined similarly. We use nn_{*} and n×n_{\times} to denote the numbers of * entries and ×\times entries in [A¯,B¯][\bar{A},\bar{B}], respectively. Let p1,,pnp_{1},...,p_{n_{*}} and p¯1,,p¯n×\bar{p}_{1},...,\bar{p}_{n_{\times}} be respectively the parameters for the * entries and ×\times entries (also called indeterminate parameters) in [A¯,B¯][\bar{A},\bar{B}]. Denote p=(p1,,pn)p=(p_{1},...,p_{n_{*}}) and p¯=(p¯1,,p¯n×)\bar{p}=(\bar{p}_{1},...,\bar{p}_{n_{\times}}). The following notions are natural extensions of the original ones from the real field to the complex field.

Definition 1 (Structural controllability, [10])

Given
  A¯{0,,×}n×n\bar{A}\in\{0,*,\times\}^{n\times n}, B¯{0,,×}n×m\bar{B}\in\{0,*,\times\}^{n\times m}, (A¯,B¯)(\bar{A},\bar{B}) is structurally controllable, if there is an (A,B)(A,B) satisfying [A,B]𝐒[A¯,B¯][A,B]\in{\bf S}_{[\bar{A}^{*},\bar{B}^{*}]} that is controllable.

Definition 2 (SSC, [16])

Given A¯{0,,×}n×n\bar{A}\in\{0,*,\times\}^{n\times n}, B¯\bar{B}\in
{0,,×}n×m\{0,*,\times\}^{n\times m}, (A¯,B¯)(\bar{A},\bar{B}) is SSC, if (A,B)(A,B) is controllable for all [A,B]𝐒[A¯×,B¯×][A,B]\in{\bf S}_{[\bar{A}^{\times},\bar{B}^{\times}]}.

Note that controllability is a generic property in the sense if (A¯,B¯)(\bar{A},\bar{B}) is structurally controllable, where A¯{0,}n×n\bar{A}\in\{0,*\}^{n\times n}, B¯{0,}n×m\bar{B}\in\{0,*\}^{n\times m}, then (A,B)(A,B) is controllable for almost all [A,B]𝐒[A¯,B¯][A,B]\in{\bf S}_{[\bar{A},\bar{B}]}. Hereafter, ‘almost all’ means ‘all except for a set of zero Lebesgue measure in the corresponding parameter space’. Motivated by Example 1, as argued in Section 1, the PSSC is formally defined as follows.

Definition 3 (PSSC)

Given A¯{0,,×}n×n\bar{A}\in\{0,*,\times\}^{n\times n}, B¯\bar{B}\in
{0,,×}n×m\{0,*,\times\}^{n\times m}, suppose [A¯,B¯][\bar{A},\bar{B}] is divided into [A¯,B¯]=[A¯,B¯]+[A¯×,B¯×][\bar{A},\bar{B}]\!\!=\!\![\bar{A}_{*},\bar{B}_{*}]+[\bar{A}_{\times},\bar{B}_{\times}], in which A¯{0,}n×n\bar{A}_{*}\in\{0,*\}^{n\times n}, B¯{0,}n×m\bar{B}_{*}\in\{0,*\}^{n\times m}, A¯×{0,×}n×n\bar{A}_{\times}\in\{0,\times\}^{n\times n}, and B¯×{0,×}n×m\bar{B}_{\times}\in\{0,\times\}^{n\times m}. (A¯,B¯)(\bar{A},\bar{B}) is PSSC, if for almost all [A,B]𝐒[A¯,B¯][A_{*},B_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{B}_{*}]}, (A+A×,B+B×)(A_{*}+A_{\times},B_{*}+B_{\times}) is controllable for all [A×,B×]𝐒[A¯×,B¯×][A_{\times},B_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{B}_{\times}]}.

Equivalently, (A¯,B¯)(\bar{A},\bar{B}) is PSSC, if for all pnp\in{\mathbb{C}}^{n_{*}} except a set of zero measure in n{\mathbb{C}}^{n_{*}} and all p¯¯n×\bar{p}\in{\bar{\mathbb{C}}^{n_{\times}}}, the corresponding realization of (A¯,B¯)(\bar{A},\bar{B}) is controllable. PSSC of (A¯,B¯)(\bar{A},\bar{B}) indicates that, if we randomly generate values for the * entries (from a continuous interval), then with probability 11, the corresponding system will be controllable for all nonzero values of the ×\times entries. For this reason, we may also call an * entry a generic entry, and the ×\times one an unspecified entry.

To show PSSC is well-defined, we introduce the concept
  perturbation-tolerant structural controllability (PTSC) from [27], which is defined for a structured system as the property that, for almost all values of its generic entries, there exist no complex values (including zero) for its unspecific entries such that the corresponding system realization is uncontrollable (see Definition 3 of [27] for a precise description). It is proven in [27] that, depending solely on the structure of the structured system, for almost all values of its generic entries, either there exist no values for its unspecific entries that can make the system realization uncontrollable, or there exist such values. Obviously, PTSC is a sufficient condition for the property in Definition 3 to hold. This means, there do exist structured systems satisfying the property in Definition 3.

Further, it is easy to see, if no * entries exist in (A¯,B¯)(\bar{A},\bar{B}), then PSSC collapses to SSC, which explains the term ‘partial’ in this terminology. On the other hand, if no ×\times entries exist in (A¯,B¯)(\bar{A},\bar{B}), then PSSC collapses to structural controllability (this also demonstrates that PSSC is well-defined). In this sense, PSSC bridges the gap between structural controllability and SSC. Moreover, it is apparent that for a given (A¯,B¯)(\bar{A},\bar{B}), SSC implies PSSC, and PSSC implies structural controllability111Note as controllability is a generic property [13], if a structured system is structurally controllable, there must exist a controllable realization of it subject to the constraint that an arbitrary subset of its indeterminate entries are nonzero.; while the inverse direction is not necessarily true. The following Example 2 highlights the differences among these concepts.

Example 2

Consider a single-input structured system as

A¯=[00000000×000],b¯=[×000].\bar{A}=\left[\begin{array}[]{cccc}0&0&0&0\\ &0&0&0\\ 0&\times&0&0\\ &*&0&*\\ \end{array}\right],\bar{b}=\left[\begin{array}[]{c}\times\\ 0\\ 0\\ 0\\ \end{array}\right]. (3)

Let [A~,b~][\tilde{A},\tilde{b}] be a generic realization of [A¯,b¯][\bar{A},\bar{b}] by assigning aija_{ij} (resp. bijb_{ij}) to the (i,j)(i,j)th indeterminate entry of A¯\bar{A} (resp. b¯\bar{b}). We have

det𝒞(A~,b~)=a212a32b114a44(a41a44+a21a42).\det{\cal C}(\tilde{A},\tilde{b})=a^{2}_{21}a_{32}b_{11}^{4}a_{44}(a_{41}a_{44}+a_{21}a_{42}). (4)

Hence, (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable. Moreover, in case
  a41a44+a21a42=0a_{41}a_{44}+a_{21}a_{42}=0, the obtained system will be uncontrollable, indicating that (A¯,b¯)(\bar{A},\bar{b}) is not SSC. However, for all (a21,a41,a42,a44)4(a_{21},a_{41},a_{42},a_{44})\in{\mathbb{C}}^{4} except {(a21,a41,a42,a44)4:a212a44(a41a44+a21a42)=0}\{(a_{21},a_{41},a_{42},a_{44})\in{\mathbb{C}}^{4}:a_{21}^{2}a_{44}(a_{41}a_{44}+a_{21}a_{42})=0\}, whatever nonzero values a32a_{32} and b11b_{11} take, the obtained system is controllable. This means (A¯,b¯)(\bar{A},\bar{b}) is PSSC. \square

Remark 1 (Complex field and real field)

It is noted that in Definition 3, PSSC is defined for (A,B)(A,B) in the complex-filed. As shown in Section 3.1, this can guarantee that if (A¯,b¯)(\bar{A},\bar{b}) is not PSSC, then for almost all values of its * entries, there exist nonzero values for its ×\times entries, such that the corresponding realization is uncontrollable. We may also define PSSC in the real field, by restricting that all of the indeterminate parameters should take real values. However, as explained in Remark 2, it then may happen that even for single-input systems, this property will not solely depend on the combinatorial properties of system structures. By Definition 3 and from the property of algebraic independence [34], it is not hard to see, PSSC in the complex field is sufficient for the same property to hold in the real field.

2.3 Preliminaries

A (directed) graph 𝒢{\cal G} is represented by 𝒢=(V,E){\cal G}=(V,E), where VV is the vertex set and EV×VE\subseteq V\times V is the edge set. If there is a path (i.e., a sequence of successive edges) from vertex vjv_{j} to vertex viv_{i}, we say viv_{i} is reachable from vjv_{j}. A subgraph 𝒢s=(Vs,Es){\cal G}_{s}=(V_{s},E_{s}) of 𝒢\cal G is a graph such that VsVV_{s}\subseteq V and EsEE_{s}\subseteq E. And 𝒢s{\cal G}_{s} is said to be induced by VsV_{s} if Es=(Vs×Vs)EE_{s}=(V_{s}\times V_{s})\cap E. For VsVV_{s}\subseteq V, 𝒢Vs{\cal G}-V_{s} denotes the graph obtained from 𝒢{\cal G} by deleting the vertices in VsV_{s} together with their incident edges (i.e., 𝒢Vs{\cal G}-V_{s} is the subgraph of 𝒢{\cal G} induced by V\VsV\backslash V_{s}).

A bipartite graph is written as 𝒢=(V+,V,E){\cal G}=(V^{+},V^{-},E) with V+V^{+} and VV^{-} being the bipartitions and EE the edges [28]. For a vV+Vv\in V^{+}\cup V^{-}, 𝒩(𝒢,v){\cal N}({\cal G},v) denotes the set of neighbors in 𝒢\cal G, i.e., the set of vertices that are connected with vv by an edge. A matching of 𝒢{\cal G} is a set of its edges, no two of which share a common end vertex. The size (or cardinality) of a matching is the number of edges it contains. A vertex is said to be matched by a matching if it is an end vertex of edges in this matching. A maximum matching of 𝒢\cal G is the matching with the largest size, which value is denoted by mt(𝒢){\rm{mt}}({\cal G}). An edge of 𝒢\cal G is said to be admissible if it is contained in some maximum matching of 𝒢{\cal G}. We say a matching {\cal M} covers 𝒢{\cal G}, if E{\cal M}\cap E forms a maximum matching of 𝒢{\cal G}. If each edge of 𝒢\cal G has a non-negative weight, the maximum (minimum) weight maximum matching is the maximum matching of 𝒢\cal G with the largest (smallest) weight (the weight of a matching is the sum of its edge weights).

Definition 4 (DM-decomposition, [28])

Given a bipartite graph 𝒢=(V+,V,E){\cal G}=(V^{+},V^{-},E), the Dulmage-Mendelsohn decomposition (DM-decomposition) of 𝒢{\cal G} is to decompose 𝒢{\cal G} into subgraphs 𝒢i=(Vi+,Vi,Ei){\cal G}_{i}=(V_{i}^{+},V_{i}^{-},E_{i}) (i=0,1,,d,i=0,1,...,d,\infty, each 𝒢i{\cal G}_{i} is called a DM-component) satisfying:

1) V=i=0ViV^{\star}=\bigcup\nolimits_{i=0}^{\infty}V_{i}^{\star}, ViVj=V_{i}^{\star}\bigcap V_{j}^{\star}=\emptyset for iji\neq j, with =+,\star=+,-; Ei={(v+,v)E:v+Vi+,vVi}E_{i}=\{(v^{+},v^{-})\in E:v^{+}\in V_{i}^{+},v^{-}\in V_{i}^{-}\};

2) mt(𝒢i)=|Vi+|=|Vi|{\rm mt}({\cal G}_{i})=|V_{i}^{+}|=|V_{i}^{-}| for i=1,,di=1,...,d, mt(𝒢0)=|V0+|<|V0|{\rm mt}({\cal G}_{0})=|V^{+}_{0}|<|V^{-}_{0}| if V0+V^{+}_{0}\neq\emptyset, and mt(𝒢)=|V|<|V+|{\rm mt}({\cal G}_{\infty})=|V^{-}_{\infty}|<|V^{+}_{\infty}| if VV^{-}_{\infty}\neq\emptyset; Moreover, each eEie\in E_{i} is admissible in 𝒢i{\cal G}_{i}, i=0,1,,d,i=0,1,...,d,\infty.

3) Eij=E_{ij}=\emptyset unless 0ij0\leq i\leq j\leq\infty, and EijE_{ij}\neq\emptyset only if 0ij0\leq i\leq j\leq\infty, where Eij={(v+,v)E:v+Vi+,vVj}E_{ij}=\{(v^{+},v^{-})\in E:v^{+}\in V^{+}_{i},v^{-}\in V^{-}_{j}\};

4) {\cal M} (\subseteq{\cal E}) is a maximum matching of 𝒢\cal G iif i=0k{\cal M}\subseteq\bigcup\nolimits_{i=0}^{\infty}{\cal E}_{k} and i{\cal M}\cap{\cal E}_{i} is a maximum matching of 𝒢i{\cal G}_{i} for i=0,,i=0,...,\infty;

5) 𝒢{\cal G} cannot be decomposed into more components satisfying conditions 1)-4).

In Definition 4, 𝒢0{\cal G}_{0} (if exists) is called the horizontal tail, and 𝒢{\cal G}_{\infty} the vertical tail. Additionally, if 𝒢{\cal G} contains only one DM-component, then 𝒢{\cal G} is called DM-irreducible.

For an n1×n2n_{1}\times n_{2} matrix MM, the bipartite graph associated with MM is given by (M)=(V+,V,EM){\cal B}(M)=(V^{+},V^{-},E_{M}), where V+={v1+,,vn1+}V^{+}=\{v^{+}_{1},...,v^{+}_{n_{1}}\} (resp. V={v1,,vn2}V^{-}=\{v^{-}_{1},...,v^{-}_{n_{2}}\}) corresponds to the rows (columns) of MM, and EM={(vi+,vj):Mij0,vi+V+,vjV}E_{M}=\{(v^{+}_{i},v^{-}_{j}):M_{ij}\neq 0,v^{+}_{i}\in V^{+},v^{-}_{j}\in V^{-}\}. It is known that, for a generic matrix MM, grank(M)=mt((M)){\rm grank}(M)={\rm mt}({\cal B}(M)) [28]. For Vs+V+V_{s}^{+}\subseteq V^{+} and VsVV_{s}^{-}\subseteq V^{-}, we also write M[Vs+,Vs]M[V_{s}^{+},V_{s}^{-}] as the submatrix of MM whose rows correspond to Vs+V_{s}^{+} and columns to VsV_{s}^{-}, where elements in Vs+V_{s}^{+} and VsV_{s}^{-} are ordered. Note such an expression of submatrices is invariant subject to row and column permutations on MM, that is, upon letting Pn1×n1P\in{\mathbb{R}}^{n_{1}\times n_{1}} and Qn2×n2Q\in{\mathbb{R}}^{n_{2}\times n_{2}} be two permutation matrices and M=PMQM^{\prime}=PMQ, M[Vs+,Vs]M^{\prime}[V^{+}_{s},V^{-}_{s}] is the same as M[Vs+,Vs]M[V^{+}_{s},V^{-}_{s}].

For a structured pair (A¯,B¯)(\bar{A},\bar{B}), its associated graph is given by 𝒢(A¯,B¯)=(X,EX){\cal G}(\bar{A},\bar{B})=(X,E_{X}), where X={x1,,xn,,xn+m}X=\{x_{1},...,x_{n},...,x_{n+m}\}, EX={(xj,xi):[A¯,B¯]ij0}E_{X}=\{(x_{j},x_{i}):[\bar{A},\bar{B}]_{ij}\neq 0\}. A vertex xiX{x_{i}\in X} is said to be input-reachable, if xi{x_{i}} is reachable from at least one vertices of xn+1,,x_{n+1},..., xn+mx_{n+m} in 𝒢(A¯,B¯){\cal G}(\bar{A},\bar{B}).

Lemma 1

[13] Given A¯{0,}n×n\bar{A}\in\{0,*\}^{n\times n}, B¯{0,}n×m\bar{B}\in\{0,*\}^{n\times m}, (A¯,B¯)(\bar{A},\bar{B}) is structurally controllable, if and only if i) every vertex xXx\in X is input-reachable, and ii) grank[A¯,B¯]=n{\rm grank}[\bar{A},\bar{B}]=n.

3 Properties and conditions of PSSC for single-input systems

In this section, we first present some properties of PSSC of single-input systems. Particularly, we show PSSC generalizes the generic property embedded in the conventional structural controllability. We then give a computationally efficient graph-theoretic criterion for PSSC.

3.1 Properties

We first give an algebraic condition for PSSC, in terms of determinants of the controllability matrix of the generic realization of (A¯,b¯)(\bar{A},\bar{b}).

Theorem 1

Given A¯{0,,×}n×n\bar{A}\in\{0,*,\times\}^{n\times n}, b¯{0,,×}n×1\bar{b}\in\{0,*,\times\}^{n\times 1}, let [A~,b~][\tilde{A},\tilde{b}] be a generic realization of [A¯,b¯][\bar{A},\bar{b}], with the parameters for the * entries being p1,,pnp_{1},...,p_{n_{*}} and for the ×\times entries being p¯1,,p¯n×\bar{p}_{1},...,\bar{p}_{n_{\times}}. (A¯,b¯)(\bar{A},\bar{b}) is PSSC, if and only if det𝒞(A~,b~)0\det{\cal C}(\tilde{A},\tilde{b})\neq 0 and has the following form

det𝒞(A~,b~)=f(p1,,pn)i=1n×p¯iri,\det{\cal C}(\tilde{A},\tilde{b})=f(p_{1},...,p_{n_{*}})\prod\nolimits_{i=1}^{n_{\times}}\bar{p}_{i}^{r_{i}}, (5)

where f(p1,,pn)f(p_{1},...,p_{n_{*}}) is a nonzero polynomial (including the constant 11) of p1,,pnp_{1},...,p_{n_{*}}, ri0r_{i}\geq 0.

Although the verification of Theorem 1 is prohibitive for large-scale systems, as computing the determinant of a symbolic matrix has computational complexity increasing exponentially with its dimensions [35], it is theoretically significant in proving some properties of PSSC. Particularly, based on Theorem 1, we have the following two properties of PSSC of single-input systems.

Proposition 1

Given A¯{0,,×}n×n\bar{A}\in\{0,*,\times\}^{n\times n}, b¯{0,,×}n×1\bar{b}\in\{0,*,\times\}^{n\times 1}, let [A¯,b¯][\bar{A}_{*},\bar{b}_{*}] and [A¯×,b¯×][\bar{A}_{\times},\bar{b}_{\times}] be defined in Definition 3. Then, depending on [A¯,b¯][\bar{A},\bar{b}], either for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, (A+A×,b+b×)(A_{*}+A_{\times},b_{*}+b_{\times}) is controllable for every [A×,b×]𝐒[A¯×,b¯×][A_{\times},b_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{b}_{\times}]}, or for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, there is a [A×,b×]𝐒[A¯×,b¯×][A_{\times},b_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{b}_{\times}]} such that (A+A×,b+b×)(A_{*}+A_{\times},b_{*}+b_{\times}) is uncontrollable.

The above proposition reveals for a single-input structured system, if it is not PSSC, then for almost all values of the * entries, there exist nonzero (complex) values for the ×\times entries so that the realization is uncontrollable; otherwise, for almost all values of the * entries, the corresponding realization is controllable for all nonzero values of the ×\times entries. This generalizes the generic property embedded in structural controllability. This proposition also explains the motivation behind the definition of PSSC.

Remark 2 (PSSC in real field)

From Theorem 1, it is easy to see, if (A¯,b¯)(\bar{A},\bar{b}) is PSSC, then for almost all pnp\in{\mathbb{R}}^{n_{*}} and all p¯¯n×\bar{p}\in\bar{\mathbb{R}}^{n_{\times}}, the corresponding realization is controllable. However, if (A¯,b¯)(\bar{A},\bar{b}) is not PSSC, it is possible that for pp in a full-dimensional semi-algebraic set 𝒫{\cal P}^{*} of n{\mathbb{R}}^{n_{*}} (i.e., a subset of n{\mathbb{R}}^{n_{*}} defined by a finite sequence of polynomial equations and inequalities, or their finite unions [34]), the corresponding realization of (A¯,b¯)(\bar{A},\bar{b}) is controllable for all p¯¯n×\bar{p}\in\bar{\mathbb{R}}^{n_{\times}}; while for pp in the semi-algebraic set n\𝒫{\mathbb{R}}^{n_{*}}\backslash{\cal P}^{*} of full dimension, there is a p¯¯n×\bar{p}\in\bar{\mathbb{R}}^{n_{\times}} that makes the corresponding realization uncontrollable. This is because, as argued in the proof of Theorem 1, det𝒞(A~,b~)\det{\cal C}(\tilde{A},\tilde{b}) then has a form of (15), which does not necessarily have real solutions for p¯i\bar{p}_{i}. Therefore, the study of PSSC in the complex filed is not only of theoretical significance itself, as it characterizes how certain combinatorial properties of the system structure affect the robustness of controllability (at least for single-input systems), but also of practical significance for the robustness evaluation of controllability in the real filed, particularly when only the system structure is available but the semi-algebraic subset of its parameters.

Proposition 2

Let (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) be the structured system obtained from (A¯,b¯)(\bar{A},\bar{b}) by preserving its (i,j)(i,j)th ×\times entry and replacing the remaining indeterminate entries with * entries, for π(i,j)𝒩×\pi\doteq(i,j)\in{\cal N}_{\times}, i.e., [A¯π,b¯π]lk=[\bar{A}^{\pi},\bar{b}^{\pi}]_{lk}=* if [A¯,b¯]lk0[\bar{A},\bar{b}]_{lk}\neq 0 and (l,k)π(l,k)\neq\pi, and [A¯π,b¯π]lk=[A¯,b¯]lk[\bar{A}^{\pi},\bar{b}^{\pi}]_{lk}=[\bar{A},\bar{b}]_{lk} otherwise. Then, (A¯,b¯)(\bar{A},\bar{b}) is PSSC, if and only if for each π𝒩×\pi\in{\cal N}_{\times}, (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) is PSSC.

The above proposition is beneficial in deriving conditions for PSSC, as it actually transforms the problem of verifying PSSC of (A¯,b¯)(\bar{A},\bar{b}) to n×n_{\times} subproblems of verifying PSSC of (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) for each π𝒩×\pi\in{\cal N}_{\times}.

3.2 Necessary and sufficient conditions

In this subsection, we give testable necessary and sufficient conditions for single-input systems to be PSSC. Owing to Proposition 2, in the following, we first focus on conditions of PSSC for (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}), i.e. systems that contain only one ×\times entry, and then on the general systems. Recall from the PBH test [29, 31], an uncontrollable mode for system (A,B)(A,B) is a λ\lambda\in{\mathbb{C}} such that rank([AλI,B])<n{\rm rank}([A-\lambda I,B])<n. Inspired by [16, 27], we shall respectively give the conditions for the nonexistence of zero uncontrollable and nonzero uncontrollable modes. Particularly, the following lemma, relating the ‘structure’ of a vector in the left null space of a given matrix to the full row rank of its submatrices, is fundamental to our subsequent derivations.

Lemma 2 (Lemma 7 of [36])

Given Mn1×n2M\in{\mathbb{C}}^{n_{1}\times n_{2}}, let TT consist of rr linearly independent row vectors that span the left null space of MM. Then, for any KJn1K\subseteq J_{n_{1}}, T[Jr,K]T[J_{r},K] is of full row rank, if and only if M[Jn1\K,Jn2]M[J_{n_{1}}\backslash K,J_{n_{2}}] is of full row rank.

Proposition 3

Suppose (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable and [A¯,b¯][\bar{A},\bar{b}] contains only one ×\times entry, with its position being (i,j)(i,j). Moreover, let [A¯,b¯][\bar{A},\bar{b}] be divided into [A¯,b¯]+[A¯×,b¯×][\bar{A}_{*},\bar{b}_{*}]+[\bar{A}_{\times},\bar{b}_{\times}] in the way described in Definition 3. For almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, there exist no [A×,b×]𝐒[A¯×,b¯×][A_{\times},b_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{b}_{\times}]} and nonzero vector qnq\in{\mathbb{C}}^{n} that satisfy q[A+A×,b+b×]=0q^{\intercal}[A_{*}+A_{\times},b_{*}+b_{\times}]=0, if and only if one of the following conditions holds:

a1) grank([A¯,b¯][Jn,Jn+1\{j}])=n{\rm grank}([\bar{A},\bar{b}][J_{n},J_{n+1}\backslash\{j\}])=n;

a2) For each k𝒩jk\in{\cal N}_{*}^{j}, grank([A¯,b¯][Jn\{k},Jn+1\{j}])=n2{\rm grank}([\bar{A},\bar{b}][J_{n}\backslash\{k\},J_{n+1}\backslash\{j\}])=n-2, in which 𝒩j={kJn\{i}:[A¯,b¯]kj=}{\cal N}_{*}^{j}=\{k\in J_{n}\backslash\{i\}:[\bar{A},\bar{b}]_{kj}=*\};222If 𝒩j={\cal N}_{*}^{j}=\emptyset, this condition is set to be satisfied (the same below).

a3) grank([A¯,b¯][Jn\{i},Jn+1\{j}])=n2{\rm grank}([\bar{A},\bar{b}][J_{n}\backslash\{i\},J_{n+1}\backslash\{j\}])=n-2.

Proposition 3 gives an algebraic condition for the nonexistence of zero uncontrollable modes. The following proposition gives an equivalent bipartite graph form of Proposition 3. Recall the bipartite graph ([A¯,b¯])=(V+,V,E[A¯,b¯]){\cal B}([\bar{A},\bar{b}])=(V^{+},V^{-},E_{[\bar{A},\bar{b}]}) associated with [A¯,b¯][\bar{A},\bar{b}] is defined in Section 2.3, where V+={v1+,,vn+}V^{+}=\{v_{1}^{+},...,v_{n}^{+}\}, V={v1,,vn+1}V^{-}=\{v_{1}^{-},...,v_{n+1}^{-}\}.

Proposition 4

Under the same setting in Proposition 3, conditions a1)-a3) are respectively equivalent to

b1) ([A¯,b¯]){\cal B}([\bar{A},\bar{b}]) contains a maximum matching that does not match vjv_{j}^{-};

b2) Either 𝒩(([A¯,b¯]),vj)={vi+}{\cal N}({\cal B}([\bar{A},\bar{b}]),v_{j}^{-})=\{v_{i}^{+}\} or for each vk+v_{k}^{+}\in
 𝒩(([A¯,b¯]),vj)\{vi+}{\cal N}({\cal B}([\bar{A},\bar{b}]),v_{j}^{-})\backslash\{v_{i}^{+}\}, mt(([A¯,b¯]){vk+,vj})=n2{\rm mt}({\cal B}([\bar{A},\bar{b}])-\{v_{k}^{+},v_{j}^{-}\})=n-2;

b3) mt(([A¯,b¯]){vi+,vj})=n2{\rm mt}({\cal B}([\bar{A},\bar{b}])-\{v_{i}^{+},v_{j}^{-}\})=n-2.

We are now presenting the conditions for nonzero uncontrollable modes. Again, suppose (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable and [A¯,b¯][\bar{A},\bar{b}] contains only one ×\times entry, with its position being (i,j)(i,j). For a generic realization (A~,b~)(\tilde{A},\tilde{b}) of (A¯,b¯)(\bar{A},\bar{b}), define a generic matrix pencil Hλ=[A~λI,b~]H_{\lambda}=[\tilde{A}-\lambda I,\tilde{b}], and let HλjcHλ[Jn,Jn+1\{j}]H^{j_{c}}_{\lambda}\doteq H_{\lambda}[J_{n},J_{n+1}\backslash\{j\}] and HλjHλ[Jn,{j}]H^{j}_{\lambda}\doteq H_{\lambda}[J_{n},\{j\}]. Let (Hλ)=(V+,V,EHλ){\cal B}(H_{\lambda})=(V^{+},V^{-},E_{H_{\lambda}}) be the bipartite graph associated with HλH_{\lambda}, defined as follows: V+={v1+,,vn+}V^{+}=\{v^{+}_{1},...,v^{+}_{n}\}, V={v1,,vn+1}V^{-}=\{v^{-}_{1},...,v^{-}_{n+1}\}, and EHλ=EIE[A¯,b¯]E_{H_{\lambda}}=E_{I}\cup E_{[\bar{A},\bar{b}]}, in which EI={(vk+,vk):k=1,,n}E_{I}=\{(v^{+}_{k},v^{-}_{k}):k=1,...,n\} and E[A¯,b¯]={(vk+,vl):[A¯,b¯]kl0}E_{[\bar{A},\bar{b}]}=\{(v^{+}_{k},v^{-}_{l}):[\bar{A},\bar{b}]_{kl}\neq 0\}. An edge ee is called a λ\lambda-edge if eEIe\in{E}_{I}, and a self-loop if eEIE[A¯,b¯]e\in E_{I}\cap E_{[\bar{A},\bar{b}]}. (Hλjc){\cal B}(H_{\lambda}^{j_{c}}) is the subgraph of (Hλ){\cal B}(H_{\lambda}) induced by V+V\{vj}V^{+}\cup V^{-}\backslash\{v^{-}_{j}\}.

Let 𝒢kjc=(Vk+,Vk,Ek){\cal G}^{j_{c}}_{k}=(V^{+}_{k},V^{-}_{k},E_{k}) (k=0,,d,k=0,...,d,\infty) be the DM-components of (Hλjc){\cal B}(H_{\lambda}^{j_{c}}). From [27, Lem 4], as (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable, we have mt((Hλjc))=n{\rm mt}({\cal B}(H_{\lambda}^{j_{c}}))=n for j=1,,n+1j=1,...,n+1. Hence, 𝒢0jc=𝒢jc={\cal G}^{j_{c}}_{0}={\cal G}^{j_{c}}_{\infty}=\emptyset, jJn+1\forall j\in J_{n+1}. By the correspondence between a matrix and its associated bipartite graph, the DM-decomposition of (Hλjc){\cal B}(H_{\lambda}^{j_{c}}) corresponds to that the matrix HλjcH_{\lambda}^{j_{c}} is transformed into the block-triangular form via two n×nn\times n permutation matrices PP and QQ [28]:

PHλjcQ=[M1jc(λ)M1djc(λ)000Mdjc(λ)]Mλjc,PH_{\lambda}^{j_{c}}Q=\left[\begin{array}[]{ccc}M^{j_{c}}_{1}(\lambda)&\cdots&M^{j_{c}}_{1d}(\lambda)\\ 0&\ddots&\vdots\\ 0&0&M^{j_{c}}_{d}(\lambda)\\ \end{array}\right]\doteq M_{\lambda}^{j_{c}}, (6)

where the submatrix Mkjc(λ)=Hλjc[Vk+,Vk]M^{j_{c}}_{k}(\lambda)=H_{\lambda}^{j_{c}}[V^{+}_{k},V^{-}_{k}] corresponds to 𝒢k{\cal G}_{k} (k=1,,dk=1,...,d). Accordingly, let MλjPHλjM_{\lambda}^{j}\doteq PH_{\lambda}^{j}.

For each k{1,,d}k\in\{1,...,d\}, let γmin(𝒢kjc)\gamma_{\min}({\cal G}^{j_{c}}_{k}) and γmax(𝒢kjc)\gamma_{\max}({\cal G}^{j_{c}}_{k}) be respectively the minimum and maximum numbers of λ\lambda-edges contained in a matching over all maximum matchings of 𝒢kjc{\cal G}^{j_{c}}_{k}. We borrow the boolean function γnz()\gamma_{nz}(\cdot) for 𝒢kjc{\cal G}_{k}^{j_{c}} from [27], which is defined as

γnz(𝒢kjc)={1ifγmax(𝒢kjc)γmin(𝒢kjc)>0or𝒢kjccontains a self-loop0otherwise.\gamma_{\rm nz}({\cal G}_{k}^{j_{\rm c}})=\begin{cases}1&{\begin{array}[]{c}\text{if}\ \gamma_{\max}({\cal G}_{k}^{j_{\rm c}})-\gamma_{\min}({\cal G}_{k}^{j_{\rm c}})>0\\ \text{or}\ {\cal G}_{k}^{j_{\rm c}}\ {\text{contains a self-loop}}\end{array}}\\ 0&\text{otherwise}.\end{cases} (7)

The following lemma explains the motivation of introducing γnz()\gamma_{\rm nz}(\cdot).

Lemma 3 (Lemma 9 of [27])

With notations as above,
 detMkjc(λ)\det M_{k}^{j_{c}}(\lambda) (k{1,,d}k\in\{1,...,d\}) generically has nonzero roots for λ\lambda, if and only if γnz(𝒢kjc)=1\gamma_{\rm nz}({\cal G}_{k}^{j_{\rm c}})=1.

Next, define the set

Ωj={k{1,,d}:γnz(𝒢kjc)=1}.{\Omega}_{j}=\{k\in\{1,...,d\}:{\gamma}_{\rm nz}({\cal G}^{j_{\rm c}}_{k})=1\}. (8)

Moreover, for a vertex vl+V+v^{+}_{l}\in V^{+}, define a set ΩjlΩj\Omega^{l}_{j}\subseteq\Omega_{j} as

Ωjl={kΩj:𝒢kjc is not covered by at leastone maximum matching of (Hλjc){vl+}}.\begin{array}[]{c}\Omega^{l}_{j}=\left\{k\in\Omega_{j}:{\cal G}_{k}^{j_{c}}\ {\text{ is not covered by at least}}\right.\\ \left.{\text{one maximum matching of }}\ {\cal B}(H_{\lambda}^{j_{c}})-\{v^{+}_{l}\}\right\}\end{array}. (9)

It is clear Ωjl=\Omega_{j}^{l}=\emptyset implies that every maximum matching of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v^{+}_{l}\} covers kΩj𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}}{\cal G}_{k}^{j_{c}}. As we shall see, Ωjl\Omega_{j}^{l} is the set of indices of all DM-components associated with which detMkjc(λ)\det M_{k}^{j_{c}}(\lambda) has a nonzero root zz, such that Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v_{l}^{+}\},V^{-}] is of full row rank (MzjcM_{z}^{j_{c}} is obtained by substituting λ=z\lambda=z into MλjcM_{\lambda}^{j_{c}}). Hereafter, by saying MλjcM_{\lambda}^{j_{c}} (or its sub-matrices) satisfies certain properties, we mean these properties are satisfied for almost all values of the corresponding indeterminate parameters (i.e., they are satisfied generically).

Lemma 4

Let MλjcM_{\lambda}^{j_{c}} be defined in (6). Assume Ωj\Omega_{j}\neq\emptyset. The following properties are true:

1) Given a vl+V+v_{l}^{+}\in V^{+}, Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{l}\},V^{-}] is generically row rank deficient for all z{λ\{0}:detMλjc=0}z\in\{\lambda\in{\mathbb{C}}\backslash\{0\}:\det M_{\lambda}^{j_{c}}=0\}, if and only if Ωjl=\Omega_{j}^{l}=\emptyset;333Please note, since MzjcM_{z}^{j_{c}} does not contain the column corresponding to vjv_{j}^{-}, Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v_{l}^{+}\},V^{-}] is the same as Mzjc[V+\{vl+},V\{vj}]M_{z}^{j_{c}}[V^{+}\backslash\{v_{l}^{+}\},V^{-}\backslash\{v_{j}^{-}\}]. The same case holds for other matrices.

2) Suppose Ωji\Omega_{j}^{i}\neq\emptyset. For all nonzero zz making
Mzjc[V+\{vi+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{i}\},V^{-}] of full row rank and detMzjc=0\det M_{z}^{j_{c}}=0 simultaneously, Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{l}\},V^{-}] is not of full row rank for a given vl+V+\{vi+}v_{l}^{+}\in V^{+}\backslash\{v_{i}^{+}\}, if and only if every maximum matching of B(Hλjc){vl+}B(H_{\lambda}^{j_{c}})-\{v_{l}^{+}\} covers kΩji𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}^{i}}{\cal G}_{k}^{j_{c}}.

Proposition 5

Suppose (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable and [A¯,b¯][\bar{A},\bar{b}] contains only one ×\times entry, with its position being (i,j)(i,j). Let [A¯,b¯][\bar{A},\bar{b}] be divided into [A¯,b¯]+[A¯×,b¯×][\bar{A}_{*},\bar{b}_{*}]+[\bar{A}_{\times},\bar{b}_{\times}] in the way described in Definition 3. For almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, there exist no [A×,b×]𝐒[A¯×,b¯×][A_{\times},b_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{b}_{\times}]}, nonzero complex number zz and nonzero vector qnq\in{\mathbb{C}}^{n} that satisfy q[A+A×zI,b+b×]=0q^{\intercal}[A_{*}+A_{\times}-zI,b_{*}+b_{\times}]=0, if and only if one of the following conditions holds:

c1) Ωji=\Omega_{j}^{i}=\emptyset;

c2) Ωji\Omega_{j}^{i}\neq\emptyset, iji\neq j, and for each vl+𝒩((Hλ),vj)\{vi+}v_{l}^{+}\in{\cal N}({\cal B}(H_{\lambda}),v_{j}^{-})\backslash\{v_{i}^{+}\}, every maximum matching of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v^{+}_{l}\} covers kΩji𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}^{i}}{\cal G}_{k}^{j_{c}}.444If 𝒩((Hλ),vj)\{vi+}={\cal N}({\cal B}(H_{\lambda}),v_{j}^{-})\backslash\{v_{i}^{+}\}=\emptyset, the third item is automatically satisfied by property 4) of Definition 4 (the same below).

Proposition 5 gives the necessary and sufficient condition for the nonexistence of nonzero uncontrollable modes. Combining Propositions 2, 4, and 5 yields a necessary and sufficient condition for general single-input systems to be PSSC.

Theorem 2

Given A¯{0,,×}n×n\bar{A}\in\{0,*,\times\}^{n\times n}, b¯{0,,×}n×1\bar{b}\in\{0,*,\times\}^{n\times 1}, (A¯,b¯)(\bar{A},\bar{b}) is PSSC, if and only if (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable, and for each π(i,j)𝒩×\pi\doteq(i,j)\in{\cal N}_{\times}, the following two conditions hold for the system (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}), recalling (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) is defined in Proposition 2:

1) condition b1), condition b2), or condition b3) holds;

2) condition c1) or condition c2) holds.

We present some examples to illustrate Theorem 2.

Example 3

Consider the system (A¯,b¯)(\bar{A},\bar{b}) in Example 2. In this system, 𝒩×={(3,2),(1,5)}{\cal N}_{\times}=\{(3,2),(1,5)\}. For π=(3,2)\pi=(3,2), the associated ([A¯π,b¯π]){\cal B}([\bar{A}^{\pi},\bar{b}^{\pi}]), (Hλ){\cal B}(H_{\lambda}), and (Hλjc){\cal B}(H_{\lambda}^{j_{c}}) (j=2j=2), as well as the DM-decomposition of (Hλjc){\cal B}(H_{\lambda}^{j_{c}}), are given respectively in Figs. 2(a), 2(b), and 2(c). From Fig. 2(a), condition b2) is fulfilled, as the bipartite graph ([A¯π,b¯π]){v4+,v2}{\cal B}([\bar{A}^{\pi},\bar{b}^{\pi}])-\{v_{4}^{+},v_{2}^{-}\} has a maximum matching with size 22. From Fig. 2(c), Ω2={2}\Omega_{2}=\{2\} and Ω23=\Omega_{2}^{3}=\emptyset, implying condition c1) is fulfilled. Similarly, for π=(1,5)\pi=(1,5), the associated ([A¯π,b¯π]){\cal B}([\bar{A}^{\pi},\bar{b}^{\pi}]) and (Hλjc){\cal B}(H_{\lambda}^{j_{c}}) (j=5j=5), as well as its DM-decomposition, are given respectively in Figs. 2(a) and 2(d). Fig. 2(a) shows 𝒩(([A¯π,b¯π]),v5)\{v1+}={\cal N}({\cal B}([\bar{A}^{\pi},\bar{b}^{\pi}]),v_{5}^{-})\backslash\{v_{1}^{+}\}=\emptyset, meaning condition b2) is satisfied. Fig. 2(d) indicates Ω5={2}\Omega_{5}=\{2\} and Ω51={2}\Omega_{5}^{1}=\{2\}, meanwhile, 𝒩((Hλ),v5)\{v1+}={\cal N}({\cal B}(H_{\lambda}),v_{5}^{-})\backslash\{v_{1}^{+}\}=\emptyset. This means condition c2) is satisfied. As a consequence, (A¯,b¯)(\bar{A},\bar{b}) is PSSC, which is consistent with the analysis in Example 2. Further, suppose we change (A¯,b¯)(\bar{A},\bar{b}) to

A¯=[0000000000×0×],b¯=[000].\bar{A}=\left[\begin{array}[]{cccc}0&0&0&0\\ &0&0&0\\ 0&*&0&0\\ \times&*&0&\times\\ \end{array}\right],\bar{b}=\left[\begin{array}[]{c}*\\ 0\\ 0\\ 0\\ \end{array}\right]. (10)

Then, 𝒩×={(4,1),(4,4)}{\cal N}_{\times}=\{(4,1),(4,4)\}. For each π𝒩×\pi\in{\cal N}_{\times}, ([A¯π,b¯π]){\cal B}([\bar{A}^{\pi},\bar{b}^{\pi}]) and the corresponding (Hλ){\cal B}(H_{\lambda}) are of the same form as Figs. 2(a) and 2(b), respectively. For π=(4,4)\pi=(4,4), ([A¯π,b¯π]){\cal B}([\bar{A}^{\pi},\bar{b}^{\pi}]) satisfies condition b2), with the corresponding (Hλjc){\cal B}(H_{\lambda}^{j_{c}}) (j=4)(j=4) and its DM-decomposition given in Fig. 2(e). It turns out that, Ωj=Ωji={3}\Omega_{j}=\Omega_{j}^{i}=\{3\}, and meanwhile, i=ji=j. Therefore, neither condition c1) nor c2) is fulfilled, meaning (A¯,b¯)(\bar{A},\bar{b}) in (10) is not PSSC. This can be validated by using Theorem 1. Alternatively, we can obtain the same conclusion by inspecting that, for π=(4,1)\pi=(4,1), the corresponding (Hλjc){\cal B}(H_{\lambda}^{j_{c}}) (j=1j=1; see Fig. 2(f)) does not satisfy condition c1) or c2). See, associated with (Hλjc){\cal B}(H_{\lambda}^{j_{c}}), Ωji={1}\Omega_{j}^{i}=\{1\}, and there is a maximum matching of (Hλjc){v4+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{4}^{+}\} that does not cover 𝒢1jc{\cal G}_{1}^{j_{c}}. \square

Refer to caption
(a) ([A¯π,b¯π]){\cal B}({[\bar{A}^{\pi},\bar{b}^{\pi}]})
Refer to caption
(b) (Hλ){\cal B}(H_{\lambda})
Refer to caption
(c) (Hλjc)(j=2){\cal B}(H_{\lambda}^{j_{c}})\ (j=2)
Refer to caption
(d) (Hλjc)(j=5){\cal B}(H_{\lambda}^{j_{c}})\ (j=5)
Refer to caption
(e) (Hλjc)(j=4){\cal B}(H_{\lambda}^{j_{c}})\ (j=4)
Refer to caption
(f) (Hλjc)(j=1){\cal B}(H_{\lambda}^{j_{c}})\ (j=1)
Figure 2: Illustration of Theorem 2 on systems (3) and (10). Bold edges represent self-loops, and dotted edges represent λ\lambda-edges that are not self-loops.
Example 4 (Example 1 continuing)

The system in Example 1 corresponds to the following structured matrices

A¯=[0×00000000×000],b¯=[00].\bar{A}=\left[\begin{array}[]{cccc}0&\times&0&0\\ &0&0&0\\ 0&0&0&\times\\ &0&0&0\\ \end{array}\right],\bar{b}=\left[\begin{array}[]{c}0\\ \\ 0\\ \\ \end{array}\right]. (11)

Following the similar process to Example 3, it turns out for π=(1,2)\pi=(1,2) and π=(3,4)\pi=(3,4), the corresponding systems (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) both satisfy conditions b2) and c1). Hence, system (10) is PSSC, consistent with the analysis in Example 1. \square

3.3 Efficient verification of the proposed conditions

Conditions b1), b2), and b3) can be verified directly via maximum matching computations on the associated bipartite graphs. As for conditions c1) and c2), first, γ(𝒢kjc)\gamma({\cal G}_{k}^{j_{c}}) can be determined, as argued in [27], via computing the minimum/maximum weight maximum matchings. Specifically, if we assign weight 11 to each λ\lambda-edge of 𝒢kjc{\cal G}_{k}^{j_{c}} and weight 0 to the other edges, then by definition, γmax(𝒢kjc)\gamma_{\max}({\cal G}_{k}^{j_{c}}) equals the maximum weight of all maximum matchings of 𝒢kjc{\cal G}_{k}^{j_{c}}, and γmin(𝒢kjc)\gamma_{\min}({\cal G}_{k}^{j_{c}}) equals the minimum weight over all maximum matchings of 𝒢kjc{\cal G}_{k}^{j_{c}}. After determining Ωj\Omega_{j}, for a given kΩjk\in\Omega_{j}, assign weight to the edge ee of (Hλjc){vi+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{i}^{+}\} as follows

w(e)={1ifeEk\{(vi+,vl):(vi+,vl)EHλ}0otherwise.w(e)=\begin{cases}1&{\text{if}}\ e\in E_{k}\backslash\{(v_{i}^{+},v^{-}_{l}):(v_{i}^{+},v^{-}_{l})\in E_{H_{\lambda}}\}\\ 0&\text{otherwise}.\end{cases} (12)

It is not hard to see, kΩjik\in\Omega_{j}^{i}, if and only if the minimum weight over all maximum matchings of the weighted (Hλjc){vi+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{i}^{+}\} is less than |Vk+||V_{k}^{+}|. Indeed, if such a minimum weight is less than |Vk+||V_{k}^{+}|, then there must be a maximum matching of (Hλjc){vi+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{i}^{+}\} that does not cover 𝒢kjc{\cal G}_{k}^{j_{c}}. On the other hand, if such a minimum weight is equal to |Vk+||V_{k}^{+}|, then each maximum matching of (Hλjc){vi+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{i}^{+}\} should cover 𝒢kjc{\cal G}_{k}^{j_{c}}. In such a manner, we can determine the set Ωji\Omega_{j}^{i}. Afterwards, if iji\neq j and Ωji0\Omega_{j}^{i}\neq 0, for a given vl+𝒩((Hλ),vj)\{vi+}v_{l}^{+}\in{\cal N}({\cal B}(H_{\lambda}),v_{j}^{-})\backslash\{v_{i}^{+}\}, to determine whether the third item of condition c2) is fulfilled, we can adopt a similar manner to the preceding scenario, that is, assigning weight to the edge ee of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v^{+}_{l}\} as follows

w(e)={1ife(kΩjiEk)\{(vl+,vl):(vl+,vl)EHλ}0otherwise.w(e)\!=\!\begin{cases}1&{\text{if}}\ e\in\left(\bigcup\nolimits_{k\in\Omega_{j}^{i}}E_{k}\right)\backslash\{(v_{l}^{+},v^{-}_{l^{\prime}}):(v_{l}^{+},v^{-}_{l^{\prime}})\in E_{H_{\lambda}}\}\\ 0&\text{otherwise}.\end{cases} (13)

Then, similarly, it follows that, every maximum matching of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v^{+}_{l}\} covers kΩji𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}^{i}}{\cal G}_{k}^{j_{c}}, if and only if the minimum weight over all maximum matchings of the weighted (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v^{+}_{l}\} equals |kΩjiVk+||\bigcup\nolimits_{k\in\Omega_{j}^{i}}V^{+}_{k}|. It is remarkable that determining whether Ωji=\Omega_{j}^{i}=\emptyset can be implemented in a similar manner, i.e., by replacing Ωji\Omega_{j}^{i} in (13) with Ωj\Omega_{j}.

Let us figure out the computational complexity of the above procedure. Note that determining the maximum matching of a bipartite graph with |V||V| vertices and |E||E| edges incurs O(|V||E|)O(\sqrt{|V|}|E|) time via the Hopcroft-Karp algorithm, and there are algorithms computing the maximum weighted matching in O(|V|3)O(|V|^{3}) [37]. In addition, DM-decomposition has the same complexity as finding a maximum matching [28]. Verifying whether (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable can invoke the strongly-connected component decomposition and maximum matching algorithms, which incurs O(n2.5)O(n^{2.5}). Moreover, as analyzed above, for each (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}), conditions b1), b2) and b3) can be verified in O(nn0.5|EX|)O(n3.5)O(n\cdot n^{0.5}\cdot|E_{X}|)\to O(n^{3.5}) time, and conditions c1) and c2) can be checked in time complexity at most

O(n2.5)DMdecomposition+O(n3)findingΩj+O(n4)determiningΩji+O(n4)checkingc2)\underbrace{O(n^{2.5})}_{{\rm DM-decomposition}}+\underbrace{O(n^{3})}_{{\rm finding}\ \Omega_{j}}+\underbrace{O(n^{4})}_{{\rm determining}\ \Omega_{j}^{i}}+\underbrace{O(n^{4})}_{\rm checking\ c2)}

\to O(n4)O(n^{4}). To sum up, since there are |𝒩×||{\cal N}_{\times}| ×\times entries, the total time complexity of Theorem 2 is at most O(|𝒩×|n4)O(|{\cal N}_{\times}|n^{4}).

4 A special case for multi-input systems

In this section, we consider PSSC for a special case in multi-input systems, that is, when there is only one ×\times entry in [A¯,B¯][\bar{A},\bar{B}]. Because of the property revealed in Proposition 2, such a case is enough to obtain necessary and sufficient conditions of PSSC for general single-input systems. However, a similar property does not hold for multi-input systems, for which the general case might need further inspection beyond the scope of this paper.

In the rest of this section, recall [A¯,B¯][\bar{A},\bar{B}] is divided into [A¯,B¯]=[A¯,B¯]+[A¯×,B¯×][\bar{A},\bar{B}]=[\bar{A}_{*},\bar{B}_{*}]+[\bar{A}_{\times},\bar{B}_{\times}] in the way described in Definition 3. First, the result below indicates the similar generic property in Proposition 1 still holds for multi-input systems with a single ×\times entry.

Proposition 6

For a multi-input system (A¯,B¯)(\bar{A},\bar{B}), assume that there is only one ×\times entry in [A¯,B¯][\bar{A},\bar{B}]. Then, either for almost all [A,B]𝐒[A¯,B¯][A_{*},B_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{B}_{*}]}, (A+A×,B+B×)(A_{*}+A_{\times},B_{*}+B_{\times}) is controllable for each [A×,B×]𝐒[A¯×,B¯×][A_{\times},B_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{B}_{\times}]}, or for almost all [A,B]𝐒[A¯,B¯][A_{*},B_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{B}_{*}]}, there is a [A×,B×]𝐒[A¯×,B¯×][A_{\times},B_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{B}_{\times}]} such that (A+A×,B+B×)(A_{*}+A_{\times},B_{*}+B_{\times}) is uncontrollable.

The generic property presented above is characterized by PSSC of (A¯,B¯)(\bar{A},\bar{B}). Next, similar to Proposition 4, the following proposition gives the necessary and sufficient condition for the nonexistence of zero uncontrollable modes.

Proposition 7

Suppose (A¯,B¯)(\bar{A},\bar{B}) is structurally controllable and [A¯,B¯][\bar{A},\bar{B}] contains only one ×\times in its (i,j)(i,j)th position. For almost all [A,B]𝐒[A¯,B¯][A_{*},B_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{B}_{*}]}, there exist no [A×,B×]𝐒[A¯×,B¯×][A_{\times},B_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{B}_{\times}]} and nonzero vector qnq\in{\mathbb{C}}^{n} that satisfy q[A+A×,B+B×]=0q^{\intercal}[A_{*}+A_{\times},B_{*}+B_{\times}]=0, if and only if one of the following conditions holds

d1) ([A¯,B¯]){\cal B}([\bar{A},\bar{B}]) contains a maximum matching that does not match vjv_{j}^{-};

d2) For each vk+𝒩(([A¯,B¯]),vj)\{vi+}v_{k}^{+}\in{\cal N}({\cal B}([\bar{A},\bar{B}]),v_{j}^{-})\backslash\{v_{i}^{+}\} (if exists),
mt(([A¯,B¯]){vk+,vj})=n2{\rm mt}({\cal B}([\bar{A},\bar{B}])-\{v_{k}^{+},v_{j}^{-}\})=n-2;

d3) mt(([A¯,B¯]){vi+,vj})=n2{\rm mt}({\cal B}([\bar{A},\bar{B}])-\{v_{i}^{+},v_{j}^{-}\})=n-2.

Suppose (A¯,B¯)(\bar{A},\bar{B}) is structurally controllable and [A¯,B¯][\bar{A},\bar{B}] contains only one ×\times in its (i,j)(i,j)th position. Let [A~,B~][\tilde{A},\tilde{B}] be a generic realization of [A¯,B¯][\bar{A},\bar{B}]. Define a generic matrix pencil Hλ=[A~λI,B~]H_{\lambda}=[\tilde{A}-\lambda I,\tilde{B}], and let Hλjc=Hλ[Jn,Jn+1\{j}]H_{\lambda}^{j_{c}}=H_{\lambda}[J_{n},J_{n+1}\backslash\{j\}]. Let (Hλ)=(V+,V,EHλ){\cal B}(H_{\lambda})=(V^{+},V^{-},E_{H_{\lambda}}) and (Hλjc){\cal B}(H^{j_{c}}_{\lambda}) be the bipartite graphs associated with HλH_{\lambda} and HλjcH^{j_{c}}_{\lambda}, respectively, defined in the same way as in Section 3. Note compared with the single-input case, the essential difference is that |V\{vj}|=n+m1|V+||V^{-}\backslash\{v_{j}^{-}\}|=n+m-1\geq|V^{+}|, which results in that there are horizontal or vertical tails in DM-decomposing (Hλjc){\cal B}(H^{j_{c}}_{\lambda}). Owing to the structural controllability of (A¯,B¯)(\bar{A},\bar{B}), a trivial extension of [27, Lem 4] shows mt((Hλjc))=n{\rm mt}({\cal B}(H_{\lambda}^{j_{c}}))=n. Consequently, by Definition 4, there is only a horizontal tail in the DM-decomposition of (Hλjc){\cal B}(H^{j_{c}}_{\lambda}) (m>1m>1). Let 𝒢kjc=(Vk+,Vk,Ek){\cal G}_{k}^{j_{c}}=(V_{k}^{+},V_{k}^{-},E_{k}) (k=0,1,,dk=0,1,...,d) be the DM-components of (Hλjc){\cal B}(H^{j_{c}}_{\lambda}). The following intermediate result is crucial for the subsequent derivations.

Lemma 5

If m>1m>1, there is generically no nonzero λ\lambda that can make Hλjc[V0+,V0]H^{j_{c}}_{\lambda}[V^{+}_{0},V^{-}_{0}] row rank deficient.

Moreover, associated with (Hλjc){\cal B}(H_{\lambda}^{j_{c}}) and 𝒢kjc{\cal G}_{k}^{j_{c}} (k=0,1,,dk=0,1,...,d), let Ωj\Omega_{j} and Ωjl\Omega_{j}^{l} (vl+V+v_{l}^{+}\in V^{+}) be defined in the same way as (8) and (9), respectively. Particularly, Lemma 5 implies Hλjc[V0+,V0]H^{j_{c}}_{\lambda}[V^{+}_{0},V^{-}_{0}] would contribute no nonzero λ\lambda that can make HλjcH^{j_{c}}_{\lambda} row rank deficient (thus 0Ωj0\notin{\Omega_{j}}). We have the following proposition, providing a necessary and sufficient condition for the nonexistence of nonzero uncontrollable modes.

Proposition 8

Suppose (A¯,B¯)(\bar{A},\bar{B}) is structurally controllable and [A¯,B¯][\bar{A},\bar{B}] contains only one ×\times entry in its (i,j)(i,j)th position. For almost all [A,B]𝐒[A¯,B¯][A_{*},B_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{B}_{*}]}, there exist no [A×,B×]𝐒[A¯×,B¯×][A_{\times},B_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{B}_{\times}]}, nonzero complex number zz and nonzero vector qnq\in{\mathbb{C}}^{n} that satisfy q[A+A×zI,B+B×]=0q^{\intercal}[A_{*}+A_{\times}-zI,B_{*}+B_{\times}]=0, if and only if one of the following conditions holds

e1) Ωji=\Omega_{j}^{i}=\emptyset;

e2) Ωji\Omega_{j}^{i}\neq\emptyset, iji\neq j, and for each vl+𝒩((Hλ),vj)\{vi+}v_{l}^{+}\in{\cal N}({\cal B}(H_{\lambda}),v_{j}^{-})\backslash\{v_{i}^{+}\}, every maximum matching of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v^{+}_{l}\} covers kΩji𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}^{i}}{\cal G}_{k}^{j_{c}}.

Combining Propositions 7 and 8 yields a necessary and sufficient condition for PSSC of (A¯,B¯)(\bar{A},\bar{B}) with a single ×\times entry.

Theorem 3

Suppose (A¯,B¯)(\bar{A},\bar{B}) contains only one ×\times entry in its (i,j)(i,j)th position. (A¯,B¯)(\bar{A},\bar{B}) is PSSC, if and only if: i) (A¯,B¯)(\bar{A},\bar{B}) is structurally controllable, ii) Condition d1), d2) or d3) holds, and iii) Condition e1) or e2) holds.

Similar to the single-input case, Theorem 3 can be verified in polynomial time mainly via the (weighed) maximum matching computations. Specifically, following a similar manner to Section 3.3, it can be found the total complexity of Theorem 3 is at most O(n(n+m)3)O(n(n+m)^{3}).

Remark 3

Although presented in an analogous form to the single-input case, results in this section are not simple extensions of the previous section. As shown in our derivations, since 𝒞(A,B){\cal C}(A,B) is not square, the proof for genericity needs to consider multiple n×nn\times n submatrices of 𝒞(A,B){\cal C}(A,B). Besides, as HλjcH_{\lambda}^{j_{c}} is no longer square, we have to consider the horizontal tail of the DM-components of (Hλjc){\cal B}(H_{\lambda}^{j_{c}}).

While Theorem 3 is devoted to the single ×\times entry case, it can provide some necessary conditions for PSSC of more general cases. Specifically, it is easy to see, for (A¯,B¯)(\bar{A},\bar{B}) to be PSSC, by preserving arbitrary one of its ×\times entries and changing the remaining ×\times entries to *, the obtained structured system should be PSSC, i.e., satisfying the conditions in Theorem 3; otherwise, (A¯,B¯)(\bar{A},\bar{B}) cannot be PSSC by Proposition 6.

Example 5

Consider (A¯,B¯)(\bar{A},\bar{B}) as

A¯=[000000000000000000×0000000000],B¯=[00000000],\bar{A}=\left[\begin{array}[]{cccccc}0&0&0&0&0&0\\ 0&*&*&0&0&0\\ 0&0&0&*&*&0\\ 0&0&0&0&\times&0\\ 0&0&0&0&0&*\\ 0&*&0&0&0&*\\ \end{array}\right],\bar{B}=\left[\begin{array}[]{cc}*&0\\ &0\\ 0&0\\ 0&0\\ 0&*\\ 0&*\\ \end{array}\right], (14)

which is structurally controllable. The associated ([A¯,B¯]){\cal B}([\bar{A},\bar{B}]) and (Hλjc){\cal B}(H_{\lambda}^{j_{c}}) (and its DM-decomposition, j=5j=5) are given in Figs. 3(a) and 3(b). From them, we know ([A¯,B¯]){v3+,v5}{\cal B}([\bar{A},\bar{B}])-\{v^{+}_{3},v^{-}_{5}\} has a maximum matching with size 44, and Ωj=\Omega_{j}=\emptyset. This means the conditions of Theorem 3 are satisfied. Hence, (A¯,B¯)(\bar{A},\bar{B}) is PSSC. Further, it can be found that, by replacing arbitrary one of the indeterminate entries of (A¯,B¯)(\bar{A},\bar{B}) with a ×\times entry and changing the remaining ones to *, the obtained structured system is still PSSC. This is consistent with the fact that (A¯,B¯)(\bar{A},\bar{B}) is actually SSC (c.f. [18, Theo 4]). \square

Refer to caption
(a) ([A¯,B¯]){\cal B}({[\bar{A},\bar{B}]})
Refer to caption
(b) (Hλjc)(j=5){\cal B}(H_{\lambda}^{j_{c}})\ (j=5)
Figure 3: Illustration of Theorem 3 on systems (14). The indications of different edge styles are the same as Fig. 2.

5 Implications for the existing SSC theory

In this section, we point out the proposed PSSC criteria in the special case can provide new graph-theoretic conditions for SSC, even restricted to the real field. Further, we demonstrate the new conditions can provide a classification of edges (or indeterminate entries) with respect to system controllability.

As mentioned earlier, when there is no * entry in [A¯,b¯][\bar{A},\bar{b}], Theorem 2 collapses to the criterion for SSC (in the complex field). With this idea, the following corollary provides a new necessary and sufficient condition for SSC in terms of (weighted) maximum matchings over the system bipartite graph representation.

Corollary 1

Given A¯{0,×}n×n\bar{A}\in\{0,\times\}^{n\times n}, b¯{0,×}n×1\bar{b}\in\{0,\times\}^{n\times 1}, (A¯,b¯)(\bar{A},\bar{b}) is SSC in the real field (i.e., (A,b)(A,b) is controllable for all real-valued [A,b]𝐒[A¯,b¯][A,b]\in{\bf S}_{[\bar{A},\bar{b}]}), if and only if (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable, and for each π𝒩×\pi\in{\cal N}_{\times}, (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) satisfies: i) at least one of conditions b1), b2) or b3) holds; and ii) condition c1) or condition c2) holds.

Although it is not hard to expect that, necessary and sufficient conditions for SSC in the real field and the complex field should have the same form (c.f. [17]), we have provided a self-contained proof for the corollary above from the developed PSSC theory (see the Appendix). Compared to [19, Theo 5], Corollary 1 is not appealing in terms of computational complexity. Nevertheless, since Corollary 1 is an entry-wise criterion, it does provide some deep insight into the role of each edge of 𝒢(A¯,b¯){\cal G}(\bar{A},\bar{b}) in system controllability. For description convenience, for [A¯,b¯]{0,×}n×(n+1)[\bar{A},\bar{b}]\in\{0,\times\}^{n\times(n+1)}, given π𝒩×\pi\in{\cal N}_{\times}, let p¯π\bar{p}_{\pi} be the parameter for the π\pith ×\times entry of [A¯,b¯][\bar{A},\bar{b}], and p¯πc\bar{p}_{\pi^{c}} the vector consisting of parameters for the remaining ×\times entries. Given that (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable, for any π=(i,j)𝒩×\pi=(i,j)\in{\cal N}_{\times}, depending on what conditions in Theorem 2 are satisfied for (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}), the edge (xj,xi)(x_{j},x_{i}) can be classified into:

  • 1.

    Critical edge: there is a nonzero (complex) value for p¯π\bar{p}_{\pi} making the corresponding realization uncontrollable, for almost all p¯πc¯n×1\bar{p}_{\pi^{c}}\in\bar{\mathbb{C}}^{n_{\times}-1}, if (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) does not satisfy condition 1) or condition 2) of Theorem 2;

  • 2.

    Stable edge: there is no nonzero value for p¯π\bar{p}_{\pi} that can make the corresponding realization uncontrollable, for almost all p¯πc¯n×1\bar{p}_{\pi^{c}}\in\bar{\mathbb{C}}^{n_{\times}-1}, if (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) satisfies conditions 1) and 2) of Theorem 2;

The above classification rule is immediate from the definitions of PSSC and Theorem 2.

Example 6

Consider (A¯,b¯)(\bar{A},\bar{b}) that is obtained from system (2) by replacing all its indeterminate entries with ×\times entries. Corollary 1 yields that (A¯,b¯)(\bar{A},\bar{b}) is not SSC. Further, a byproduct of Corollary 1 on this system is the following classifications for its edges (which can be obtained from Example 3): (a) critical edges: (x1,x2)(x_{1},x_{2}), (x1,x4)(x_{1},x_{4}), (x2,x4)(x_{2},x_{4}) and (x4,x4)(x_{4},x_{4}); (b) stable edges: (x2,x3)(x_{2},x_{3}), and (x5,x1)(x_{5},x_{1}). It is easy to validate the above assertion by noting that the determinant of a generic realization of (A¯,b¯)(\bar{A},\bar{b}) is exactly the right-hand side of (4). \square

6 Conclusions

In this paper, a new controllability notion, named PSSC, has been proposed for linear systems, with the aim to extend the existing SSC and bridge the gap between structural controllability and SSC. Algebraic and bipartite graph-theoretic necessary and sufficient conditions are given for single-input systems to be PSSC, the latter of which can be verified efficiently. Extension to the multi-input case on a special case is also given. Further, it is demonstrated the established results for PSSC in the single-input case could provide a new graph-theoretic criterion for the conventional SSC. In the future, we plan to extend the previous results to the general multi-input case, and to investigate PSSC in the real field systematically. It is also interesting to enumerate the number of PSSC networks statistically.

Appendix: Proofs of the technical results

Proof of Theorem 1: Sufficiency: Let p(p1,,pn)p\doteq(p_{1},...,p_{n_{*}}). Since f(p)f(p) is a nonzero polynomial, pnp\in{\mathbb{C}}^{n_{*}} making f(p)=0f(p)=0 has zero measure in n{\mathbb{C}}^{n_{*}}. Therefore, in case p¯i0\bar{p}_{i}\neq 0 (i=1,,n×i=1,...,n_{\times}) and f(p)0f(p)\neq 0, it follows f(p)i=1n×p¯iri0f(p)\prod\nolimits_{i=1}^{n_{\times}}\bar{p}_{i}^{r_{i}}\neq 0, indicating the corresponding realization is always controllable.

Necessity: The necessity of det𝒞(A~,b~)0\det{\cal C}(\tilde{A},\tilde{b})\neq 0 is obvious. If det𝒞(A~,b~)0\det{\cal C}(\tilde{A},\tilde{b})\neq 0 but the remaining condition is not satisfied, then there is a p¯i\bar{p}_{i} (1in×1\leq i\leq n_{\times}) that exists in two different monomials with different degrees (including zero) for p¯i\bar{p}_{i} (the degree of p¯i\bar{p}_{i} is the exponent of p¯i\bar{p}_{i}) in det𝒞(A~,b~)\det{\cal C}(\tilde{A},\tilde{b}). Let p¯=(p¯1,,p¯i1,p¯i+1,,p¯n×)\bar{p}^{\prime}=(\bar{p}_{1},...,\bar{p}_{i-1},\bar{p}_{i+1},...,\bar{p}_{n_{\times}}) and p=(p1,,pn)p¯p^{\prime}=(p_{1},...,p_{n_{*}})\cup\bar{p}^{\prime}. In this case, write det𝒞(A~,b~)\det{\cal C}(\tilde{A},\tilde{b}) as a polynomial of p¯i\bar{p}_{i} as

det𝒞(A~,b~)=frp¯ir++f1p¯i+f0g(p;p¯i),\det{\cal C}(\tilde{A},\tilde{b})=f_{r}\bar{p}_{i}^{r}+\cdots+f_{1}\bar{p}_{i}+f_{0}\doteq g(p^{\prime};\bar{p}_{i}), (15)

in which the coefficients fjf_{j} (j=0,,rj=0,...,r) are polynomials of pp^{\prime}, and frf_{r} as well as another fjf_{j} (0j<r0\leq j<r) is not identically zero. Consider the set 𝒫1={(p1,,pn)n:p¯n×1,s.t.fr0,fj0,k=1,kin×p¯k0}{\cal P}_{1}=\{(p_{1},...,p_{n_{*}})\in{\mathbb{C}}^{n_{*}}:\exists\bar{p}^{\prime}\in{\mathbb{C}}^{n_{\times}-1},{\rm{s.t.}}\ f_{r}\neq 0,f_{j}\neq 0,\prod\nolimits_{k=1,k\neq i}^{n_{\times}}\bar{p}_{k}\neq 0\}. Obviously, the complement of 𝒫1{\cal P}_{1} in n{\mathbb{C}}^{n_{*}} has zero measure, as {pn+n×1:fr0,fj0,k=1,kin×p¯k0}\{p^{\prime}\in{\mathbb{C}}^{n_{*}+n_{\times}-1}:f_{r}\neq 0,f_{j}\neq 0,\prod\nolimits_{k=1,k\neq i}^{n_{\times}}\bar{p}_{k}\neq 0\} has full dimension in n+n×1{\mathbb{C}}^{n_{*}+n_{\times}-1}. Note in case fr0f_{r}\neq 0 and fj0f_{j}\neq 0, g(p;p¯i)g(p^{\prime};\bar{p}_{i}) has at least one nonzero root for p¯i\bar{p}_{i} (as otherwise g(p;p¯i)=frp¯irg(p^{\prime};\bar{p}_{i})=f_{r}\bar{p}_{i}^{r}). Therefore, for all (p1,,pn)𝒫1(p_{1},...,p_{n_{*}})\in{\cal P}_{1}, there exists (p¯1,,p¯n×)¯n×(\bar{p}_{1},...,\bar{p}_{n_{\times}})\in\bar{\mathbb{C}}^{n_{\times}} such that g(p;p¯i)=0g(p^{\prime};\bar{p}_{i})=0, making the obtained realization uncontrollable.

Proof of Proposition 1: The statement follows directly from the proof of Theorem 1. To be specific, for a given (A¯,b¯)(\bar{A},\bar{b}), the first case emerges if (A¯,b¯)(\bar{A},\bar{b}) is PSSC, while the second case emerges if (A¯,b¯)(\bar{A},\bar{b}) is not PSSC.

Proof of Proposition 2: By Theorem 1, if (A¯,b¯)(\bar{A},\bar{b}) is PSSC, then for its generic realization (A~,b~)(\tilde{A},\tilde{b}), det𝒞(A~,b~)\det{\cal C}(\tilde{A},\tilde{b}) has the form of (5). Let [A~π,b~π][\tilde{A}^{\pi},\tilde{b}^{\pi}] be the generic realization of [A¯π,b¯π][\bar{A}^{\pi},\bar{b}^{\pi}]. It is easy to see, for every π𝒩×\pi\in{\cal N}_{\times}, det𝒞(A~π,b~π)\det{\cal C}(\tilde{A}^{\pi},\tilde{b}^{\pi}) then has the form of (5), which indicates (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) is PSSC.

On the other hand, suppose (A¯π,b¯π)(\bar{A}^{\pi},\bar{b}^{\pi}) is PSSC, π(i,j)𝒩×\forall\pi\doteq(i,j)\in{\cal N}_{\times}. Let p¯π\bar{p}_{\pi} be the parameter for the (i,j)(i,j)th entry of [A¯,b¯][\bar{A},\bar{b}]. By Theorem 1, for each π𝒩×\pi\in{\cal N}_{\times}, det𝒞(A~π,b~π)\det{\cal C}(\tilde{A}^{\pi},\tilde{b}^{\pi}) has a factor p¯πrπ\bar{p}_{\pi}^{r_{\pi}} for some rπ0r_{\pi}\geq 0, and any other factor containing p¯π\bar{p}_{\pi} does not exist. Consequently, det𝒞(A¯,b¯)\det{\cal C}(\bar{A},\bar{b}) must have a form of (5). This means, (A¯,b¯)(\bar{A},\bar{b}) is PSSC.

Proof of Proposition 3: Sufficiency: The sufficiency of condition a1) is obvious, as in this case for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, rank([A,b][Jn,Jn+1\{j}])=n{\rm rank}([A_{*},b_{*}][J_{n},J_{n+1}\backslash\{j\}])=n. Suppose condition a2) is fulfilled. We only need to consider the case where condition a1) is not fulfilled. In this case, as (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable, by Lemma 1, grank([A¯,b¯][Jn,Jn+1\{j}])=n1{\rm grank}([\bar{A},\bar{b}][J_{n},J_{n+1}\backslash\{j\}])=n-1 must hold. Then, for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, rank([A,b][Jn,{\rm rank}([A_{*},b_{*}][J_{n}, Jn+1\{j}])=n1J_{n+1}\backslash\{j\}])=n-1. Next, consider two cases: i) grank([A¯,b¯][Jn\{i},{\rm grank}([\bar{A},\bar{b}][J_{n}\backslash\{i\}, Jn+1\{j}])=n1J_{n+1}\backslash\{j\}])=n-1 and ii) grank([A¯,b¯][Jn\{i},Jn+1\{j}])<n1{\rm grank}([\bar{A},\bar{b}][J_{n}\backslash\{i\},J_{n+1}\backslash\{j\}])<n-1, i.e., grank([A¯,b¯][Jn\{i},Jn+1\{j}])=n2{\rm grank}([\bar{A},\bar{b}][J_{n}\backslash\{i\},J_{n+1}\backslash\{j\}])=n-2. Let qnq\in{\mathbb{C}}^{n} (q0)(q\neq 0) be in the left null space of [A,b][Jn,Jn+1\{j}][A_{*},b_{*}][J_{n},J_{n+1}\backslash\{j\}]. Note qq is unique up to scaling. By Lemma 2, we generically have qk=0q_{k}=0 for each k𝒩jk\in{\cal N}_{*}^{j}; additionally, qi0q_{i}\neq 0 in case i), and qi=0q_{i}=0 in case ii). Hence, in case i), for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]} and all [A×,b×]𝐒[A¯×,b¯×][A_{\times},b_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{b}_{\times}]}, it holds

q[A+A×,b+b×][Jn,{j}]\displaystyle q^{\intercal}[A_{*}+A_{\times},b_{*}+b_{\times}][J_{n},\{j\}]
=k𝒩jqk[A,b]kj+qi[A×,b×]ij=qi[A×,b×]ij0.\displaystyle=\sum\nolimits_{k\in{\cal N}_{*}^{j}}q_{k}[A_{*},b_{*}]_{kj}+q_{i}[A_{\times},b_{\times}]_{ij}=q_{i}[A_{\times},b_{\times}]_{ij}\neq 0.

In case ii), for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]} and all [A×,b×]𝐒[A¯×,b¯×][A_{\times},b_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{b}_{\times}]}, we have

q[A+A×,b+b×][Jn,{j}]=k=1nqk[A,b]kj=0,\displaystyle q^{\intercal}[A_{*}+A_{\times},b_{*}+b_{\times}][J_{n},\{j\}]=\sum\nolimits_{k=1}^{n}q_{k}[A_{*},b_{*}]_{kj}=0,

indicating (A+A×,b+b×)(A_{*}+A_{\times},b_{*}+b_{\times}) is always uncontrollable, which is excluded by the structural controllability of (A¯,b¯)(\bar{A},\bar{b}).

To show the sufficiency of condition a3), we only need to consider the case where condition a1) is not satisfied. In this case, for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, following the similar arguments to those for condition a2), upon letting qq be in the left null space of [A,b][Jn,Jn+1\{j}][A_{*},b_{*}][J_{n},J_{n+1}\backslash\{j\}], we have qi=0q_{i}=0, for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}. Then, for all [A×,b×]𝐒[A¯×,b¯×][A_{\times},b_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{b}_{\times}]}, it holds

q[A+A×,b+b×][Jn,{j}]=kJn,kiqk[A,b]kj(a)0,\displaystyle q^{\intercal}[A_{*}+A_{\times},b_{*}+b_{\times}][J_{n},\{j\}]=\sum\nolimits_{k\in J_{n},k\neq i}q_{k}[A_{*},b_{*}]_{kj}\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny(a)}}}}{{\neq}}}0,

where (a) is due to the structural controllability of (A¯,b¯)(\bar{A},\bar{b}).

Necessity: Suppose none of the three conditions is satisfied. Then, by the structural controllability of (A¯,b¯)(\bar{A},\bar{b}) and Lemma 1, it must hold that grank([A¯,b¯][Jn,Jn+1\{j}])=grank([A¯,b¯][Jn\{i},{\rm grank}([\bar{A},\bar{b}][J_{n},J_{n+1}\backslash\{j\}])={\rm grank}([\bar{A},\bar{b}][J_{n}\backslash\{i\}, Jn+1\{j}])=n1J_{n+1}\backslash\{j\}])=n-1, and there exists at least one kJn\{i}k\in J_{n}\backslash\{i\} so that [A¯,b¯]kj=[\bar{A},\bar{b}]_{kj}=* and grank([A¯,b¯][Jn\{k},{\rm grank}([\bar{A},\bar{b}][J_{n}\backslash\{k\}, Jn+1\{j}])=n1J_{n+1}\backslash\{j\}])=n-1. Under these conditions, for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, we have rank([A,b][Jn,Jn+1\{j}])=n1{\rm rank}([A_{*},b_{*}][J_{n},J_{n+1}\backslash\{j\}])=n-1. Upon letting qq be a nonzero vector in the left null space of [A,b][Jn,Jn+1\{j}][A_{*},b_{*}][J_{n},J_{n+1}\backslash\{j\}], Lemma 2 yields that qi0q_{i}\neq 0 and qk0q_{k}\neq 0 generically hold. Assign

[A×,b×]ij=1/qikJn,kiqk[A,b]kj0,[A_{\times},b_{\times}]_{ij}=-1/{q_{i}}\sum\nolimits_{k\in J_{n},k\neq i}q_{k}[A_{*},b_{*}]_{kj}\neq 0, (16)

where the inequality is due to the fact [A,b]kj[A_{*},b_{*}]_{kj} is independent of qq (qq is uniquely determined by [A,b][Jn,Jn+1\{j}][A_{*},b_{*}][J_{n},J_{n+1}\backslash\{j\}] up to scaling). We therefore have

q[A+A×,b+b×][Jn,{j}]\displaystyle q^{\intercal}[A_{*}+A_{\times},b_{*}+b_{\times}][J_{n},\{j\}]
=kJn,kiqk[A,b]kj+qi[A×,b×]ij=0.\displaystyle=\sum\nolimits_{k\in J_{n},k\neq i}q_{k}[A_{*},b_{*}]_{kj}+q_{i}[A_{\times},b_{\times}]_{ij}=0.

This proves the necessity.

Proof of Proposition 4: The equivalence is obvious since the generic rank of a structured matrix equals the size of a maximum matching of its associated bipartite graph (see Section 2.3).

Proof of Lemma 4: To prove Lemma 4, we need the following auxiliary results.

Lemma 6 (Lemma 9 of [27])

Let MM be an n×nn\times n generic matrix over the variables t1,,trt_{1},...,t_{r}, and EE be an n×nn\times n constant matrix whose entries are either 0 or 11, where each row of EE, as well as each column, has at most one 11. Let (Pλ){\cal B}(P_{\lambda}) be the bipartite graph associated with the generic matrix pencil PλMλEP_{\lambda}\doteq M-\lambda E (defined in a way similar to (Hλ){\cal B}(H_{\lambda})). Let 𝒯i{\cal T}_{i} be the set of variables of t1,,trt_{1},...,t_{r} that appear in the iith column of PλP_{\lambda}. If (Pλ){\cal B}(P_{\lambda}) is DM-irreducible, then every nonzero root of det(Pλ)\det(P_{\lambda}) (if exists) cannot be independent of 𝒯i{\cal T}_{i} (if nonempty), for each i{1,,n}i\in\{1,...,n\}.

Proof of Lemma 4 counting: Remember for any vl+V+v_{l}^{+}\in V^{+}, (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v^{+}_{l}\} has a maximum matching with size n1n-1, as otherwise it cannot hold that mt((Hλjc))=n{\rm mt}({\cal B}(H_{\lambda}^{j_{c}}))=n. We first prove property 1), and then property 2).

Necessity in property 1): Suppose Ωjl\Omega_{j}^{l}\neq\emptyset. Then, there is some kΩjk^{\prime}\in\Omega_{j}, so that a matching \cal M of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{l}^{+}\} with size n1n-1 exists that does not cover 𝒢kjc{\cal G}^{j_{c}}_{k^{\prime}}. Suppose vk′′Vkv^{-}_{k^{\prime\prime}}\in V^{-}_{k^{\prime}} is not matched by Ek{\cal M}\cap E_{k^{\prime}}. As kΩjk^{\prime}\in\Omega_{j}, Lemma 3 yields detMkjc(λ)\det M_{k^{\prime}}^{j_{c}}(\lambda) generically has nonzero roots for λ\lambda, and let zz be one of such roots. Due to the block-triangular structure of MλjcM_{\lambda}^{j_{c}}, zz satisfies detMzjc=0\det M_{z}^{j_{c}}=0. Note as 𝒢kjc{\cal G}_{k^{\prime}}^{j_{c}} is DM-irreducible, from Lemma 6, zz cannot be independent of the indeterminate parameters in the column of Mkjc(λ)M^{j_{c}}_{k^{\prime}}(\lambda) corresponding to vk′′v_{k^{\prime\prime}}^{-} (note if |Vk+|=1|V^{+}_{k^{\prime}}|=1 and Mkjc(λ)=λM^{j_{c}}_{k^{\prime}}(\lambda)=\lambda, then kΩjk^{\prime}\notin\Omega_{j}).

On the other hand, Mtjc(λ)M_{t}^{j_{c}}(\lambda) for any t<kt<k^{\prime} does not share a common factor with Mkjc(λ)M_{k^{\prime}}^{j_{c}}(\lambda) except for the power of λ\lambda. By [38, Lem 2], this implies detMtjc(λ)\det M_{t}^{j_{c}}(\lambda) for any t<kt<k^{\prime} does not share a common nonzero root for λ\lambda with detMkjc(λ)\det M_{k^{\prime}}^{j_{c}}(\lambda), generically. Therefore, denoting Vkd=VkVk+1VdV^{\star}_{k^{\prime}d}=V^{\star}_{k^{\prime}}\cup V^{\star}_{k^{\prime}+1}\cup\cdots V^{\star}_{d} where =+,\star=+,-, if Mzjc[Vkd+\{vl+},Vkd]M_{z}^{j_{c}}[V^{+}_{k^{\prime}d}\backslash\{v_{l}^{+}\},V^{-}_{k^{\prime}d}] is of full row rank, then Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v_{l}^{+}\},V^{-}] will be (due to its block-triangular structure). Noting
det(Mλjc[Vkd+\{vl+},Vkd\{vk′′}])\det(M_{\lambda}^{j_{c}}[V^{+}_{k^{\prime}d}\backslash\{v^{+}_{l}\},V^{-}_{k^{\prime}d}\backslash\{v^{-}_{k^{\prime\prime}}\}]) is not identically zero (because of the existence of \cal M; note vl+Vkd+v_{l}^{+}\in V^{+}_{k^{\prime}d}, as otherwise every maximum matching of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{l}^{+}\} will cover 𝒢kjc{\cal G}^{j_{c}}_{k^{\prime}}), any nonzero root of det(Mλjc[Vkd+\{vl+},Vkd\{vk′′}])\det(M_{\lambda}^{j_{c}}[V^{+}_{k^{\prime}d}\backslash\{v^{+}_{l}\},V^{-}_{k^{\prime}d}\backslash\{v^{-}_{k^{\prime\prime}}\}]) is independent of the parameters in the column of Mkjc(λ)M_{k^{\prime}}^{j_{c}}(\lambda) corresponding to vk′′v_{k^{\prime\prime}}^{-}. This means, Mzjc[Vkd+\{vl+},Vkd]M_{z}^{j_{c}}[V^{+}_{k^{\prime}d}\backslash\{v_{l}^{+}\},V^{-}_{k^{\prime}d}] is generically of full row rank, so is Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{l}\},V^{-}].

Sufficiency in property 1) : Suppose Ωjl=\Omega_{j}^{l}=\emptyset. Due to the block-triangular structure of MλjcM_{\lambda}^{j_{c}} and from Lemma 3, any nonzero root of detMλjc\det M_{\lambda}^{j_{c}}, denoted by zz, must be a nonzero root of detMkjc(λ)\det M_{k^{\prime}}^{j_{c}}(\lambda) for some kΩjk^{\prime}\in\Omega_{j}. Consider an arbitrary maximum matching \cal M of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{l}^{+}\}. Upon letting VV^{-}_{\cal M} be the set of vertices in VV^{-} that are matched by \cal M, we have kΩjVkV\bigcup\nolimits_{k\in\Omega_{j}}V^{-}_{k}\subseteq V^{-}_{\cal M}. Let R1kΩjVk+R_{1}\doteq\bigcup\nolimits_{k\in\Omega_{j}}V_{k}^{+}, R2V+\{vl+}R_{2}\doteq V^{+}\backslash\{v_{l}^{+}\}. From [28, Prop 2.1.3],

detMzjc[R2,V]=\displaystyle\det M_{z}^{j_{c}}[R_{2},{V}^{-}_{\cal M}]= (17)
JV,|J|=|R1|sgn(R1,J)detMzjc[R1,J]detMzjc[R2\R1,V\J],\displaystyle\sum\limits_{J\subseteq V^{-}_{\cal M},|J|=|R_{1}|}{\rm sgn}(R_{1},J)\cdot{\det}M_{z}^{j_{c}}[R_{1},J]\cdot\det M_{z}^{j_{c}}[R_{2}\backslash R_{1},V^{-}_{\cal M}\backslash J],

where sgn(R1,J)=±1{\rm sgn}(R_{1},J)=\pm 1 is the signature associated with (R1,J)(R_{1},J). Since each maximum matching of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{l}^{+}\} covers kΩj𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}}{\cal G}_{k}^{j_{c}}, detMλjc[R1,J]detMλjc[R2\R1,V\J]{\det}M_{\lambda}^{j_{c}}[R_{1},J]\cdot\det M_{\lambda}^{j_{c}}[R_{2}\backslash R_{1},V^{-}_{\cal M}\backslash J] 0\equiv 0 for any JVJ\subseteq V_{\cal M}^{-}, |J|=|R1||J|=|R_{1}| but JkΩjVkJ\neq\bigcup\nolimits_{k\in\Omega_{j}}V_{k}^{-}, as otherwise E(R1,J)E(R2\R1,E(R_{1},J)\cup E(R_{2}\backslash R_{1}, V\J)V^{-}_{\cal M}\backslash J) will contain a maximum matching of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{l}^{+}\} that does not cover kΩj𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}}{\cal G}_{k}^{j_{c}}, meaning Ωjl\Omega_{j}^{l}\neq\emptyset, where E(Vs+,Vs)E(V^{+}_{s},V^{-}_{s}) denotes the set of edges between Vs+V^{+}_{s} and VsV^{-}_{s} in (Hλjc){\cal B}(H_{\lambda}^{j_{c}}), Vs+V+V^{+}_{s}\subseteq V^{+}, VsVV^{-}_{s}\subseteq V^{-}. Note detMzjc[R1,kΩjVk]=\det M_{z}^{j_{c}}[R_{1},\bigcup\nolimits_{k\in\Omega_{j}}V_{k}^{-}]=
±kΩjdetMkjc(z)=0\pm\prod\nolimits_{k\in\Omega_{j}}\det M_{k}^{j_{c}}(z)=0 due to the block-triangular structure of MλjcM_{\lambda}^{j_{c}} and detMkjc(z)=0\det M_{k^{\prime}}^{j_{c}}(z)=0. Hence, detMzjc[R2,V]=0\det M_{z}^{j_{c}}[R_{2},{V}^{-}_{\cal M}]=0. Since this property holds for any maximum matching of (Hλjc)vl+{\cal B}(H_{\lambda}^{j_{c}})-v_{l}^{+}, Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v_{l}^{+}\},V^{-}] is row rank deficient.

Necessity in property 2): Suppose there is a maximum matching of B(Hλjc){vl+}B(H_{\lambda}^{j_{c}})-\{v_{l}^{+}\} that does not cover 𝒢kjc{\cal G}_{k^{\prime}}^{j_{c}} for some kΩjik^{\prime}\in\Omega_{j}^{i} (thus does not cover kΩji𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}^{i}}{\cal G}_{k}^{j_{c}}). By the definition of Ωji\Omega_{j}^{i}, there exists a maximum matching of B(Hλjc){vi+}B(H_{\lambda}^{j_{c}})-\{v_{i}^{+}\} that does not cover 𝒢kjc{\cal G}^{j_{c}}_{k^{\prime}}. In this case, following the similar argument to the proof of necessity in property 1), any nonzero root of detMkjc(λ)\det M_{k^{\prime}}^{j_{c}}(\lambda), denoted by zz, is also a root of detMλjc\det M_{\lambda}^{j_{c}}, while making Mzjc[V+\{vi+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{i}\},V^{-}] and Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{l}\},V^{-}] of full row rank simultaneously.

Sufficiency in property 2): From the proof of property 1), we know a nonzero zz makes Mzjc[V+\{vi+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{i}\},V^{-}] of full row rank while detMzjc=0\det M_{z}^{j_{c}}=0, if and only if zz is a nonzero root of detMkjc(λ)\det M^{j_{c}}_{k^{\prime}}(\lambda) for some kΩjk^{\prime}\in\Omega_{j} satisfying the property that, 𝒢kjc{\cal G}_{k^{\prime}}^{j_{c}} is not covered by some maximum matching of B(Hλjc){vi+}B(H_{\lambda}^{j_{c}})-\{v_{i}^{+}\} (referred to as condition f)). Indeed, following the necessity in property 1), we know a zz satisfying condition f) is sufficient; and following the similar arguments to the proof for sufficiency in property 1), we know condition f) for zz is necessary. By definition, the set of such a kk^{\prime} satisfying condition f) is exactly Ωji\Omega_{j}^{i}. Again, following the similar manner to the proof for sufficiency in property 1) by replacing Mzjc[R1,J]M_{z}^{j_{c}}[R_{1},J] with Mzjc[kΩjiVk+,J]M_{z}^{j_{c}}[\bigcup\nolimits_{k\in\Omega_{j}^{i}}V_{k}^{+},J] in (17), it turns out Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{l}\},V^{-}] is row rank deficient if every maximum matching of B(Hλjc){vl+}B(H_{\lambda}^{j_{c}})-\{v_{l}^{+}\} covers kΩji𝒢kjc\bigcup\nolimits_{k\in\Omega_{j}^{i}}{\cal G}_{k}^{j_{c}}.

Proof of Proposition 5: In the following, we consider the generic realization (A~,b~)(\tilde{A},\tilde{b}) of (A¯,b¯)(\bar{A},\bar{b}). Recall Hλ=[A~λI,b~]H_{\lambda}=[\tilde{A}-\lambda I,\tilde{b}]. Let p¯ij(0)\bar{p}_{ij}(\neq 0) be the parameter in the (i,j)(i,j)th entry of [A~,b~][\tilde{A},\tilde{b}]. For vl+V+v_{l}^{+}\in{V^{+}}, suppose vl+v_{l}^{+} corresponds to the σ(l)\sigma(l)th row of MλjcM_{\lambda}^{j_{c}} after the row permutation PP on HλjcH_{\lambda}^{j_{c}}.

Sufficiency: Suppose condition c1) is fulfilled. If Ωj=\Omega_{j}=\emptyset, by Lemma 3, there is generically no z\{0}z\in{\mathbb{C}}\backslash\{0\} making MzjcM_{z}^{j_{c}} row rank deficient, due to the block-triangular structure of MλjcM_{\lambda}^{j_{c}}. Noting Hλjc=P1MλjcQ1H_{\lambda}^{j_{c}}=P^{-1}M_{\lambda}^{j_{c}}Q^{-1} from (6), HzjcH_{z}^{j_{c}} is generically of full rank for all nonzero zz. This means Ωj=\Omega_{j}=\emptyset is sufficient. Now consider Ωj\Omega_{j}\neq\emptyset but Ωji=\Omega^{i}_{j}=\emptyset. In this case, consider an arbitrary z\{0}z\in{\mathbb{C}\backslash\{0\}} that makes HzjcH^{j_{c}}_{z} row rank deficient. As (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable, HzjcH_{z}^{j_{c}} generically has rank n1n-1 (otherwise [Hzjc,Hzj][H_{z}^{j_{c}},H_{z}^{j}] will be row rank deficient). Hence, upon letting qq be a nonzero vector in the left null space of HzjcH_{z}^{j_{c}}, qq is unique up to scaling. From property 1) of Lemma 4, Hzjc[V+\{vi+},V]H_{z}^{j_{c}}[V^{+}\backslash\{v_{i}^{+}\},V^{-}] is row rank deficient (recalling Hλjc=P1MλjcQ1H_{\lambda}^{j_{c}}=P^{-1}M_{\lambda}^{j_{c}}Q^{-1}). By Lemma 2, we have qi=0q_{i}=0. Hence, for any nonzero value of p¯ij\bar{p}_{ij},

qHzj=k=1,kinqk[Hzj]k(a)0,q^{\intercal}H_{z}^{j}=\sum\nolimits_{k=1,k\neq i}^{n}q_{k}[H_{z}^{j}]_{k}\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny(a)}}}}{{\neq}}}0,

where (a)(a) is due to the controllability of (A~,b~)(\tilde{A},\tilde{b}) and that qHzjq^{\intercal}H_{z}^{j} is independent of p¯ij\bar{p}_{ij}. Therefore, condition c1) is sufficient.

Suppose condition c2) is fulfilled. In this case, for those nonzero roots zz of detHλjc\det H_{\lambda}^{j_{c}} that make Hzjc[V+\{vi+},V]H_{z}^{j_{c}}[V^{+}\backslash\{v_{i}^{+}\},V^{-}] row rank deficient, following a similar argument to the above, we can obtain that there is no p¯ij\bar{p}_{ij} and qq (p¯ij0\bar{p}_{ij}\neq 0 and qn\{0}q\in{\mathbb{C}}^{n}\backslash\{0\}) making qHzj=0q^{\intercal}H_{z}^{j}=0. For the nonzero root zz of detHλjc\det H_{\lambda}^{j_{c}} that makes Hzjc[V+\{vi+},V]H_{z}^{j_{c}}[V^{+}\backslash\{v_{i}^{+}\},V^{-}] of full row rank, let qq be a nonzero vector in the left null space of HzjcH_{z}^{j_{c}} (qq is unique up to scaling). Then, according to property 2) of Lemma 4, it turns out that qi0q_{i}\neq 0 and qk=0q_{k}=0 for all k{l:[Hzj]l0}\{i}k\in\{l:[H_{z}^{j}]_{l}\neq 0\}\backslash\{i\}. We therefore have

qHzj=k=1nqk[Hzj]k=(a)qip¯ij0,q^{\intercal}H_{z}^{j}=\sum\nolimits_{k=1}^{n}q_{k}[H_{z}^{j}]_{k}\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny(a)}}}}{{=}}}q_{i}\bar{p}_{ij}\neq 0,

where (a) is due to [Hzj]i=p¯ij[H_{z}^{j}]_{i}=\bar{p}_{ij} as jij\neq i. Hence, condition c2) is also sufficient.

Necessity: Suppose neither condition c1), nor condition c2) holds. Then, either i): Ωji\Omega_{j}^{i}\neq\emptyset, and the third item of condition c2) does not hold, or ii): Ωji\Omega_{j}^{i}\neq\emptyset, i=ji=j, and the third item of condition 2) holds. In case i), suppose the third item of condition 2) does not hold for a vertex vl+𝒩((Hλ),vj)\{vi+}v_{l^{\prime}}^{+}\in{\cal N}({\cal B}(H_{\lambda}),v_{j}^{-})\backslash\{v_{i}^{+}\}. Considering the generic realization (A~,b~)(\tilde{A},\tilde{b}), by property 2) of Lemma 4, we known there is some nonzero zz making MzjcM_{z}^{j_{c}} row rank deficient, while Mzjc[V+\{vi+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{i}\},V^{-}] and Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v^{+}_{l^{\prime}}\},V^{-}] are of full row rank. Note it generically holds rank(Mzjc)=n1{\rm rank}(M_{z}^{j_{c}})=n-1, as otherwise rank([Mzjc,Mzj])<n{\rm rank}([M_{z}^{j_{c}},M_{z}^{j}])<n, contradicting the structural controllability of (A¯,b¯)(\bar{A},\bar{b}). Let q^\hat{q} be a nonzero vector spanning the left null space of MzjcM_{z}^{j_{c}}. Then, by Lemma 2, q^σ(i)0\hat{q}_{\sigma(i)}\neq 0 and q^σ(l)0\hat{q}_{\sigma(l^{\prime})}\neq 0. Upon letting [Mzj]σ(i)=1/q^σ(i)k=1,kσ(i)nq^k[Mzj]k[M_{z}^{j}]_{\sigma(i)}=-1/\hat{q}_{\sigma(i)}\sum\nolimits_{k=1,k\neq\sigma(i)}^{n}\hat{q}_{k}[M_{z}^{j}]_{k}, we have

q^Mzj=q^σ(i)[Mzj]σ(i)+k=1,kσ(i)nq^k[Mzj]k=0,\hat{q}M_{z}^{j}=\hat{q}_{\sigma(i)}[M_{z}^{j}]_{\sigma(i)}+\sum\nolimits_{k=1,k\neq\sigma(i)}^{n}\hat{q}_{k}[M_{z}^{j}]_{k}=0,

which leads to q^P[Hzjc,Hzj]=0\hat{q}P[H_{z}^{j_{c}},H_{z}^{j}]=0, by substituting Mλjc=PHλjcQM_{\lambda}^{j_{c}}=PH_{\lambda}^{j_{c}}Q, Mλj=PHλjcM^{j}_{\lambda}=PH_{\lambda}^{j_{c}} and noting QQ is invertible. That is, by assigning

p¯ij={z1/q^σ(i)k=1,kσ(i)nq^k[Mzj]kifi=j1/q^σ(i)k=1,kσ(i)nq^k[Mzj]kifij,\bar{p}_{ij}=\begin{cases}z-1/\hat{q}_{\sigma(i)}\sum\nolimits_{k=1,k\neq\sigma(i)}^{n}\hat{q}_{k}[M_{z}^{j}]_{k}&\text{if}\ i=j\\ -1/\hat{q}_{\sigma(i)}\sum\nolimits_{k=1,k\neq\sigma(i)}^{n}\hat{q}_{k}[M_{z}^{j}]_{k}&\text{if}\ i\neq j,\end{cases} (18)

we can make q^P[A~zI,b~]=0\hat{q}P[\tilde{A}-zI,\tilde{b}]=0. Note if i=ji=j, [Mzj]σ(l)[M_{z}^{j}]_{\sigma(l^{\prime})} is independent of q^\hat{q} and zz (as [Mzj]σ(l)[M_{z}^{j}]_{\sigma(l^{\prime})} contains a free parameter); if iji\neq j, in case k=1,kσ(i)nq^k[Mzj]k\sum\nolimits_{k=1,k\neq\sigma(i)}^{n}\hat{q}_{k}[M_{z}^{j}]_{k} contains more than one nonzero items, at least one [Mzj]k[M_{z}^{j}]_{k} is independent of q^\hat{q} and zz. Hence, in all these circumstances, it is assured p¯ij0\bar{p}_{ij}\neq 0.

As for case ii), following similar arguments to case i), upon letting q^0\hat{q}\neq 0 be in the left null space of MzjcM_{z}^{j_{c}} for some nonzero zz making detMzjc=0\det M_{z}^{j_{c}}=0 while Mzjc[V+\{vi+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v_{i}^{+}\},V^{-}] of full row rank, we have q^σ(i)0\hat{q}_{\sigma(i)}\neq 0 and q^σ(k)=0\hat{q}_{\sigma(k)}=0 for all kk satisfying vk+𝒩((Hλ),vj)\{vi+}v_{k}^{+}\in{\cal N}({\cal B}(H_{\lambda}),v_{j}^{-})\backslash\{v_{i}^{+}\}. Hence, if pii=z0p_{ii}=z\neq 0, then q^Mzj=q^σ(i)(piiz)=0\hat{q}M_{z}^{j}=\hat{q}_{\sigma(i)}(p_{ii}-z)=0, leading to q^P[Hzjc,Hzj]=0\hat{q}P[H_{z}^{j_{c}},H_{z}^{j}]\!=\!0. Since (A~,b~)(\tilde{A},\tilde{b}) is a generic realization of (A¯,b¯)(\bar{A},\bar{b}), in both cases, we know for almost all [A,b]𝐒[A¯,b¯][A_{*},b_{*}]\in{\bf S}_{[\bar{A}_{*},\bar{b}_{*}]}, there exist [A×,b×]𝐒[A¯×,b¯×][A_{\times},b_{\times}]\in{\bf S}_{[\bar{A}_{\times},\bar{b}_{\times}]}, nonzero zz and qnq\in{\mathbb{C}}^{n} satisfying q[A+A×zI,b+b×]=0q^{\intercal}[A_{*}+A_{\times}-zI,b_{*}+b_{\times}]=0.

Proof of Theorem 2: The result comes immediately from the PBH test and Propositions 2, 4 and 5.

Proof of Proposition 6: The case where (A¯,B¯)(\bar{A},\bar{B}) is not structurally controllable is trivial. Now assume structural controllability of (A¯,B¯)(\bar{A},\bar{B}), and consider its generic realization (A~,B~)(\tilde{A},\tilde{B}). Let p¯{\bar{p}} be the parameter for the unique ×\times entry in [A~,B~][\tilde{A},\tilde{B}], and pp the collection of parameters for the remaining indeterminate entries. Let Γ(p,p¯)\Gamma(p,{\bar{p}}) be the greatest common divisor among all determinants of the n×nn\times n submatrices of 𝒞(A~,B~){\cal C}(\tilde{A},\tilde{B}). According to [38, Lem 2], for almost all pnp\in{\mathbb{C}}^{n_{*}} (recalling nn^{*} is the number of * entries), the determinants of all the n×nn\times n submatrices of 𝒞(A~,B~){\cal C}(\tilde{A},\tilde{B}) share a common zero for p¯{\bar{p}}, if and only if the leading degree for p¯{\bar{p}} in Γ(p,p¯)\Gamma(p,{\bar{p}}) is no less than one. Therefore, if Γ(p,p¯)=P(p)p¯r\Gamma(p,{\bar{p}})=P(p){\bar{p}}^{r}, where r0r\geq 0 and P(p)P(p) is a nonzero polynomial of pp, then for almost all pnp\in{\mathbb{C}}^{n} satisfying P(p)0P(p)\neq 0 and all p¯0{\bar{p}}\neq 0, 𝒞(A~,B~){\cal C}(\tilde{A},\tilde{B}) is of full row rank, indicating that the corresponding realization is controllable. Otherwise, if Γ(p,p¯)\Gamma(p,{\bar{p}}) contains two monomials with different degrees for p¯{\bar{p}}, then for almost all pnp\in{\mathbb{C}}^{n_{*}}, there exists a nonzero solution p¯{\bar{p}} satisfying Γ(p,p¯)=0\Gamma(p,{\bar{p}})=0 (see the proof of Proposition 1), making the corresponding realization uncontrollable.

Proof of Proposition 7: Note Lemma 2 is applicable to rectangular matrices. This means the proof can be completed in the same manner as in Propositions 3 and 4, which thus is omitted.

Proof of Lemma 5: Since 𝒢0jc{\cal G}_{0}^{j_{c}} is the horizontal tail, from [28, Corollary 2.2.23], for each vlV0v_{l}^{-}\in V_{0}^{-}, mt(𝒢0jc{vl})=|V0+|{\rm mt}({\cal G}_{0}^{j_{c}}-\{v_{l}^{-}\})=|V_{0}^{+}|. From [27, Lem 9], for all vlV0v_{l}^{-}\in V_{0}^{-}, every maximum matching of 𝒢0jc{\cal G}_{0}^{j_{c}} corresponds to a nonzero term in the determinant of a |V0+|×|V0+||V_{0}^{+}|\times|V_{0}^{+}| submatrix of Hλjc[V0+,V0]H^{j_{c}}_{\lambda}[V^{+}_{0},V^{-}_{0}] that cannot be cancelled out by other terms. Therefore, supposing there is a nonzero value, denoted by zz, such that Hzjc[V0+,V0]H^{j_{c}}_{z}[V^{+}_{0},V^{-}_{0}] is of row rank deficient, zz should depend only on the free parameters in Hλjc[V0+,V0\{vl}]H^{j_{c}}_{\lambda}[V^{+}_{0},V^{-}_{0}\backslash\{v_{l}^{-}\}], vlV0\forall v_{l}^{-}\in V_{0}^{-}. Applying this across vlV0v_{l}^{-}\in V_{0}^{-}, it turns out that zz is independent of the free parameters in each column of HλjcH_{\lambda}^{j_{c}}, causing a contradiction.

Proof of Proposition 8: From Lemmas 3 and 5, if zz is a nonzero root of detHλjc[Vk+,Vk]\det H_{\lambda}^{j_{c}}[V_{k}^{+},V_{k}^{-}] for some kΩjk\in\Omega_{j}, then zz also makes HλjcH_{\lambda}^{j_{c}} row rank deficient owing to the block-triangular structure of its DM-decomposition; and vice versa. With these results, it can be proved easily that, a nonzero zz makes
Hλjc[V+\{vl+},V]H_{\lambda}^{j_{c}}[V^{+}\backslash\{v_{l}^{+}\},V^{-}] row rank deficient for a given vl+V+v_{l}^{+}\in V^{+}, if and only if zz makes Hλjc[V+\(V0+{vl+}),V\V0]H_{\lambda}^{j_{c}}[V^{+}\backslash(V^{+}_{0}\cup\{v_{l}^{+}\}),V^{-}\backslash V^{-}_{0}] row rank deficient. Having observed this, it can be found Lemma 4 still holds for the rectangular matrix HλjcH_{\lambda}^{j_{c}} (by changing detMλjc=0\det M_{\lambda}^{j_{c}}=0 to the row-rank deficient of HλjcH_{\lambda}^{j_{c}}). Hence, the proof can be completed in the similar manner to that for the single-input case, i.e., the proof for Proposition 5. The details are omitted due to their similarities.

Proof of Corollary 1: Sufficiency: From Theorems 1 and 2, we know det𝒞(A~,b~)\det{\cal C}(\tilde{A},\tilde{b}) has the form det𝒞(A~,b~)=i=1n×p¯iri\det{\cal C}(\tilde{A},\tilde{b})=\prod\nolimits_{i=1}^{n_{\times}}\bar{p}_{i}^{r_{i}}, ri0r_{i}\geq 0 under the proposed conditions, where [A~,b~][\tilde{A},\tilde{b}] is a generic realization of [A¯,b¯][\bar{A},\bar{b}] with the indeterminate parameters being p¯1,,p¯n×\bar{p}_{1},...,\bar{p}_{n_{\times}}. It is then obvious that for all (p¯1,,p¯n×)¯n×(\bar{p}_{1},...,\bar{p}_{n_{\times}})\in\bar{\mathbb{R}}^{n_{\times}}, the corresponding system realization is controllable.

Necessity: We prove the necessity by contradiction. Let p¯π\bar{p}_{\pi} be the parameter for the π\pith ×\times entry of [A¯,b¯][\bar{A},\bar{b}] for a π𝒩×\pi\in{\cal N}_{\times}, and p¯πc\bar{p}_{\pi^{c}} the vector consisting of the parameters for the remaining ×\times entries (except p¯π\bar{p}_{\pi}). The necessity of structural controllability of (A¯,b¯)(\bar{A},\bar{b}) is obvious. Now assume (A¯,b¯)(\bar{A},\bar{b}) is structurally controllable. Suppose i) is not fulfilled for a π=(i,j)𝒩×\pi=(i,j)\in{\cal N}_{\times}. From the proof for necessity of Proposition 3, for almost all p¯πc¯n×1\bar{p}_{\pi^{c}}\in\bar{\mathbb{R}}^{n_{\times}-1}, there exists nonzero real p¯π\bar{p}_{\pi} (expressed in (16), where qq is chosen to be real), such that the corresponding real system realization is uncontrollable. This means (A¯,b¯)(\bar{A},\bar{b}) is not SSC in the real field.

Furthermore, suppose ii) is not fulfilled for a π=(i,j)𝒩×\pi=(i,j)\in{\cal N}_{\times}. We first consider the case Ωji\Omega_{j}^{i}\neq\emptyset and the third item of condition c2) does not hold for some vl+N((Hλ),vj)\{vi+}v_{l^{\prime}}^{+}\in N({\cal B}(H_{\lambda}),v_{j}^{-})\backslash\{v_{i}^{+}\}. Following the proof for necessity of Proposition 5, in this case, to construct a real uncontrollable realization of (A¯,b¯)(\bar{A},\bar{b}), we only need to demonstrate zz in (18) can be real (thus the vector q^\hat{q} therein can be real too). To this end, recall in (18), zz needs to satisfy: requirement 1) zz makes MzjcM_{z}^{j_{c}} row rank deficient, and requirement 2) zz makes Mzjc[V+\{vi+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v_{i}^{+}\},V^{-}] and Mzjc[V+\{vl+},V]M_{z}^{j_{c}}[V^{+}\backslash\{v_{l^{\prime}}^{+}\},V^{-}] of full row rank. Notice from the proof for necessity of Lemma 4, if zz is a real nonzero root of detMkjc(λ)\det M_{k^{\prime}}^{j_{c}}(\lambda) for some kk^{\prime}, where kΩjik^{\prime}\in\Omega_{j}^{i} is such that there is a maximum matching 2{\cal M}_{2} of (Hλjc){vl+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{l^{\prime}}^{+}\} not covering 𝒢kjc{\cal G}^{j_{c}}_{k^{\prime}}, then zz generically satisfies the previous two requirements. Let us assume that vertex vj2v^{-}_{j_{2}} of 𝒢kjc{\cal G}^{j_{c}}_{k^{\prime}} is not matched by 2Ek{\cal M}_{2}\cap E_{k^{\prime}}, and π2=(i2,j2)\pi_{2}=(i_{2},j_{2}) corresponds to a nonzero entry of Mkjc(λ)M_{k^{\prime}}^{j_{c}}(\lambda). As kΩjik^{\prime}\in\Omega_{j}^{i}, there is a maximum matching 1{\cal M}_{1} of (Hλjc){vi+}{\cal B}(H_{\lambda}^{j_{c}})-\{v_{i}^{+}\} not covering 𝒢kjc{\cal G}^{j_{c}}_{k^{\prime}}. Similarly, assume that vertex vj1v^{-}_{j_{1}} of 𝒢kjc{\cal G}^{j_{c}}_{k^{\prime}} is not matched by 1Ek{\cal M}_{1}\cap E_{k^{\prime}}, and π1=(i1,j1)\pi_{1}=(i_{1},j_{1}) corresponds to a nonzero entry of Mkjc(λ)M_{k^{\prime}}^{j_{c}}(\lambda). Let R()R({\cal M}) and C()C({\cal M}) be respectively the set of vertices in V+V^{+} and in VV^{-} that are matched by a matching {\cal M}. Denote V1(k1)=V1Vk1V^{\star}_{1(k^{\prime}-1)}=V^{\star}_{1}\cup\cdots V^{\star}_{k^{\prime}-1} and Vkd=VkVdV^{\star}_{k^{\prime}d}=V^{\star}_{k^{\prime}}\cup\cdots V^{\star}_{d}, where =+\star=+ or -.

Since 𝒢kjc{\cal G}^{j_{c}}_{k^{\prime}} is DM-irreducible, by Lemma 6, detMkjc(λ)\det M_{k^{\prime}}^{j_{c}}(\lambda) does not have a fixed nonzero root for λ\lambda as p¯πc\bar{p}_{\pi^{c}} varies. Note also for any fixed z¯z\in\bar{\mathbb{R}}, detMkjc(z)\det M_{k^{\prime}}^{j_{c}}(z) is both irreducible (i.e., cannot be factored as two polynomials with smaller degrees) and multi-affine in p¯πc\bar{p}_{\pi^{c}} (i.e., every variable in p¯πc\bar{p}_{\pi^{c}} occurs with degree 0 or 11 in every term). Hence, detMkjc(λ)\det M_{k^{\prime}}^{j_{c}}(\lambda) can be written as p¯π1f1(λ,p¯π1c)+f0(λ,p¯π1c)\bar{p}_{\pi_{1}}f_{1}(\lambda,\bar{p}_{\pi^{c^{\prime}}_{1}})+f_{0}(\lambda,\bar{p}_{\pi^{c^{\prime}}_{1}}), where f1(λ,p¯π1c)f_{1}(\lambda,\bar{p}_{\pi^{c^{\prime}}_{1}}) and f0(λ,p¯π1c)f_{0}(\lambda,\bar{p}_{\pi^{c^{\prime}}_{1}}) are polynomials of λ\lambda and p¯π1c\bar{p}_{\pi^{c^{\prime}}_{1}}, with p¯π1c\bar{p}_{\pi^{c^{\prime}}_{1}} being the collection of indeterminate parameters in Mkjc(λ)M_{k^{\prime}}^{j_{c}}(\lambda) except p¯π1\bar{p}_{\pi_{1}}. Based on the above arguments, the proof for necessity of Lemma 4 indicates that the following constraints suffice to satisfy requirements 1) and 2):

p¯π1f1(z,p¯π1c)+f0(z,p¯π1c)\displaystyle\bar{p}_{\pi_{1}}f_{1}(z,\bar{p}_{\pi^{c^{\prime}}_{1}})+f_{0}(z,\bar{p}_{\pi^{c^{\prime}}_{1}}) =0\displaystyle=0 (19)
f1(z,p¯π1c)\displaystyle f_{1}(z,\bar{p}_{\pi^{c^{\prime}}_{1}}) 0\displaystyle\neq 0 (20)
f0(z,p¯π1c)\displaystyle f_{0}(z,\bar{p}_{\pi^{c^{\prime}}_{1}}) 0\displaystyle\neq 0 (21)
detMzjc[R(1)Vkd+,C(1)Vkd]\displaystyle\det M_{z}^{j_{c}}[R({\cal M}_{1})\cap V^{+}_{k^{\prime}d},C({\cal M}_{1})\cap V^{-}_{k^{\prime}d}] 0\displaystyle\neq 0 (22)
detMzjc[R(2)Vkd+,C(2)Vkd]\displaystyle\det M_{z}^{j_{c}}[R({\cal M}_{2})\cap V^{+}_{k^{\prime}d},C({\cal M}_{2})\cap V^{-}_{k^{\prime}d}] 0\displaystyle\neq 0 (23)
detMzjc[V1(k1)+,Vl(k1)]\displaystyle\det M_{z}^{j_{c}}[V^{+}_{1(k^{\prime}-1)},V^{-}_{l(k^{\prime}-1)}] 0.\displaystyle\neq 0. (24)

Indeed, (19) ensures detMkjc(z)=0\det M_{k^{\prime}}^{j_{c}}(z)=0 leading to detMzjc=0\det M_{z}^{j_{c}}=0, while (22), (23) and (24) ensure that requirement 2) is fulfilled (due to the block-diagonal structure of MzjcM_{z}^{j_{c}}). Now consider zz and p¯πc\{p¯π1}\bar{p}_{\pi^{c}}\backslash\{\bar{p}_{\pi_{1}}\} as nonzero real numbers. If j2=j1j_{2}=j_{1}, the constraints (20)-(24) do not not involve p¯π1\bar{p}_{\pi_{1}}. Then, for almost all nonzero real numbers of zz and p¯πc\{p¯π1}\bar{p}_{\pi^{c}}\backslash\{\bar{p}_{\pi_{1}}\}, there is a p¯π1¯\bar{p}_{\pi_{1}}\in{\bar{\mathbb{R}}} satisfying (19)-(24). If j2j1j_{2}\neq j_{1}, then p¯π1\bar{p}_{\pi_{1}} satisfying (19) depends generically on zz and p¯π1c\bar{p}_{\pi^{c^{\prime}}_{1}} (by Lemma 6), thus not independent of p¯π2{\bar{p}}_{\pi_{2}}, while p¯π1\bar{p}_{\pi_{1}} satisfying (23) is independent of p¯π2\bar{p}_{\pi_{2}}. This means, again, for almost all nonzero real numbers of zz and p¯πc\{p¯π1}\bar{p}_{\pi^{c}}\backslash\{\bar{p}_{\pi_{1}}\}, there is p¯π1¯\bar{p}_{\pi_{1}}\in{\bar{\mathbb{R}}} satisfying (19)-(24). Therefore, for almost all z¯z\in\bar{\mathbb{R}}, we can always find p¯πc¯n×1\bar{p}_{\pi^{c}}\in\bar{\mathbb{R}}^{n_{\times}-1} such that requirements 1) and 2) are satisfied. Then, after determining p¯π\bar{p}_{\pi} according to (18), where q^\hat{q} and zz are both real, the corresponding real system realization will be uncontrollable, meaning that (A¯,b¯)(\bar{A},\bar{b}) is not SSC in the real field. For the case where Ωji\Omega_{j}^{i}\neq\emptyset, i=ji=j, and the third item of condition 2) holds, we can adopt a similar argument to construct a real uncontrollable realization of (A¯,b¯)(\bar{A},\bar{b}). This proves the necessity.

References

References

  • [1] Y. Y. Liu, J. J. Slotine, A. L. Barabasi, Controllability of complex networks, Nature 48 (7346) (2011) 167–173.
  • [2] Y.-Y. Liu, J.-J. Slotine, A.-L. Barabási, Observability of complex systems, Proceedings of the National Academy of Sciences 110 (7) (2013) 2460–2465.
  • [3] G. Yan, J. Ren, Y.-C. Lai, C.-H. Lai, B. Li, Controlling complex networks: How much energy is needed?, Physical review letters 108 (21) (2012) 218703.
  • [4] A.-L. Barabási, R. Albert, Emergence of scaling in random networks, science 286 (5439) (1999) 509–512.
  • [5] J. Klamka, Controllability of linear dynamical systems, Contrib. Theory Differ. Equ 1 (1963) 189–213.
  • [6] C. O. Aguilar, B. Gharesifard, Graph controllability classes for the laplacian leader-follower dynamics, IEEE Transactions on Automatic Control 60 (6) (2015) 1611–1623.
  • [7] Y. Zhang, T. Zhou, Controllability analysis for a networked dynamic system with autonomous subsystems, IEEE Transactions on Automatic Control 48 (7) (2017) 3408–3415.
  • [8] P. Müller, H. Weber, Analysis and optimization of certain qualities of controllability and observability for linear dynamical systems, Automatica 8 (3) (1972) 237–246.
  • [9] F. Pasqualetti, S. Zampieri, F. Bullo, Controllability metrics, limitations and algorithms for complex networks, IEEE Transactions on Control of Network Systems 1 (1) (2014) 40–52.
  • [10] C. T. Lin, Structural controllability, IEEE Transactions on Automatic Control 19 (3) (1974) 201–208.
  • [11] R. Shields, J. Pearson, Structural controllability of multiinput linear systems, IEEE Transactions on Automatic control 21 (2) (1976) 203–212.
  • [12] H. Mayeda, On structural controllability theorem, IEEE Transactions on Automatic Control 26 (3) (1981) 795–798.
  • [13] J. M. Dion, C. Commault, J. Van DerWoude, Generic properties and control of linear structured systems: a survey, Automatica 39 (2003) 1125–1144.
  • [14] J. F. Carvalho, S. Pequito, A. P. Aguiar, S. Kar, K. H. Johansson, Composability and controllability of structural linear time-invariant systems: Distributed verification, Automatica 78 (2017) 123–134.
  • [15] Y. Zhang, T. Zhou, Structural controllability of an NDS with LFT parameterized subsystems, IEEE Transactions on Automatic Control 64 (12) (2019) 4920–4935.
  • [16] H. Mayeda, T. Yamada, Strong structural controllability, SIAM Journal on Control and Optimization 17 (1) (1979) 123–138.
  • [17] C. Bowden, W. Holderbaum, V. M. Becerra, Strong structural controllability and the multilink inverted pendulum, IEEE transactions on automatic control 57 (11) (2012) 2891–2896.
  • [18] J. C. Jarczyk, F. Svaricek, B. Alt, Strong structural controllability of linear systems revisited, in: IEEE Conference on Decision and Control and European Control Conference, IEEE, 2011, pp. 1213–1218.
  • [19] A. Chapman, M. Mesbahi, On strong structural controllability of networked systems: A constrained matching approach, in: 2013 American Control Conference, IEEE, 2013, pp. 6126–6131.
  • [20] N. Monshizadeh, S. Zhang, M. K. Camlibel, Zero forcing sets and controllability of dynamical systems defined on graphs, IEEE Transactions on Automatic Control 59 (9) (2014) 2562–2567.
  • [21] S. S. Mousavi, M. Haeri, M. Mesbahi, On the structural and strong structural controllability of undirected networks, IEEE Transactions on Automatic Control 63 (7) (2018) 2234–2241.
  • [22] J. Jia, H. J. van Waarde, H. L. Trentelman, M. K. Camlibel, A unifying framework for strong structural controllability, IEEE Transactions on Automatic Control 66 (1) (2020) 391–398.
  • [23] K. Ogata, Y. Yang, Modern Control Engineering, Vol. 4, Prentice-Hall, 2002.
  • [24] C. Godsil, Controllable subsets in graphs, Annals of Combinatorics 16 (4) (2012) 733–744.
  • [25] S. O’Rourke, B. Touri, On a conjecture of godsil concerning controllable random graphs, SIAM Journal on Control and Optimization 54 (6) (2016) 3347–3378.
  • [26] T. Menara, G. Bianchin, M. Innocenti, F. Pasqualetti, On the number of strongly structurally controllable networks, in: 2017 American Control Conference, IEEE, 2017, pp. 340–345.
  • [27] Y. Zhang, Y. Xia, G. Wang, J. Zhang, Perturbation-tolerant structural controllability for linear systems, arXiv preprint arXiv:2105.00968 (2021)  .
  • [28] K. Murota, Matrices and Matroids for Systems Analysis, Springer Science Business Media, 2009.
  • [29] R. Eising, Between controllable and uncontrollable, Systems & Control Letters 48 (5) (1984) 263–264.
  • [30] M. Gu, New methods for estimating the distance to uncontrollability, SIAM Journal on Matrix Analysis and Applications 21 (3) (2000) 989–1003.
  • [31] E. Mengi, On the estimation of the distance to uncontrollability for higher order systems, SIAM journal on matrix analysis and applications 30 (1) (2008) 154–172.
  • [32] E. Wu, X. Yang, Adaptive synchronization of coupled nonidentical chaotic systems with complex variables and stochastic perturbations, Nonlinear Dynamics 84 (1) (2016) 261–269.
  • [33] Z. Lin, L. Wang, Z. Han, M. Fu, Distributed formation control of multi-agent systems using complex laplacian, IEEE Transactions on Automatic Control 59 (7) (2014) 1765–1777.
  • [34] D. S. Dummit, R. M. Foote, Abstract Algebra, Vol. 3, Wiley Hoboken, 2004.
  • [35] A. Shpilka, A. Yehudayoff, et al., Arithmetic circuits: A survey of recent results and open questions, Foundations and Trends® in Theoretical Computer Science 5 (3–4) (2010) 207–388.
  • [36] Y. Zhang, Y. Xia, H. Zhang, G. Wang, L. Dai, On the generic low-rank matrix completion, arXiv preprint arXiv:2102.11490 (2021)  .
  • [37] D. B. West, Introduction to Graph Theory, Prentice hall, 2001.
  • [38] K. S. Lu, J. N. Wei, Rational function matrices and structural controllability and observability, IET Control Theory and Applications 138 (4) (1991) 388–394.