Partial Strong Structural Controllability
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 matching1 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].

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 nodes tends to one, as . 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 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 and be respectively the angle of the pendulum rod and the location of the cart. Define state variables , and by , , , . According to [23], the linearized state-space equation of this system reads
where and are the masses of the pendulum bob and the cart, is the length of the massless rod, is the force imposed on the cart, and is the acceleration of gravity. From [16], the zero/nonzero structure of is not SSC. However, direct calculation shows that the determinant of the controllability matrix of is . This means, if we regard , and as independent parameters, leading to the usual satisfaction of , then whatever nonzero values and may take, the corresponding system will be controllable. In fact, back to this example, is always fulfilled.
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: denote the sets of natural, real, and complex numbers, respectively. Let , , and (resp. ) be the set of -dimensional complex (real) vectors with each of its entries being nonzero. For , stands for . For an matrix , denotes the submatrix of whose rows are indexed by and columns by .
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 . The entries and entries are both called indeterminate entries. We use to denote set of all structured matrices with entries from . Moreover, for , is defined as
That is, a entry can take either zero or nonzero, while a entry can take only nonzero value. An is called a realization of . We also define two structured matrices and associated with respectively as
(1) |
That is, (resp. ) is obtained from by replacing all of its indeterminate entries with (resp. ) 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 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 , is the maximum rank it achieves as the function of its indeterminate parameters. For a generic matrix and a constant matrix with the same dimension, defines a generic matrix pencil [28], which is a matrix-valued polynomial of the independent parameters in and the variable .
2.2 Notion of PSSC
Consider a linear time invariant system as
(2) |
where is the state variable, is the input, and are respectively the state transition matrix and input matrix. For description simplicity, by the pair we refer to a system described by (2) that can be either single-input (i.e., ) or multi-input (i.e., ), while by 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 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 , i.e.,
It is known from [29, 30, 31] that is controllable, if and only if is of full row rank.
Throughout this paper, given , , denotes the set of positions of all entries in , i.e., . is defined similarly. We use and to denote the numbers of entries and entries in , respectively. Let and be respectively the parameters for the entries and entries (also called indeterminate parameters) in . Denote and . The following notions are natural extensions of the original ones from the real field to the complex field.
Definition 1 (Structural controllability, [10])
Given
, , is structurally controllable, if there is an satisfying that is controllable.
Definition 2 (SSC, [16])
Given ,
, is SSC, if is controllable for all .
Note that controllability is a generic property in the sense if is structurally controllable, where , , then is controllable for almost all . 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 ,
, suppose is divided into , in which , , , and . is PSSC, if for almost all , is controllable for all .
Equivalently, is PSSC, if for all except a set of zero measure in and all , the corresponding realization of is controllable. PSSC of indicates that, if we randomly generate values for the entries (from a continuous interval), then with probability , the corresponding system will be controllable for all nonzero values of the entries. For this reason, we may also call an entry a generic entry, and the 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 , then PSSC collapses to SSC, which explains the term ‘partial’ in this terminology. On the other hand, if no entries exist in , 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 , 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
(3) |
Let be a generic realization of by assigning (resp. ) to the th indeterminate entry of (resp. ). We have
(4) |
Hence, is structurally controllable. Moreover, in case
, the obtained system will be uncontrollable, indicating that is not SSC. However, for all except , whatever nonzero values and take, the obtained system is controllable. This means is PSSC.
Remark 1 (Complex field and real field)
It is noted that in Definition 3, PSSC is defined for in the complex-filed. As shown in Section 3.1, this can guarantee that if is not PSSC, then for almost all values of its entries, there exist nonzero values for its 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 is represented by , where is the vertex set and is the edge set. If there is a path (i.e., a sequence of successive edges) from vertex to vertex , we say is reachable from . A subgraph of is a graph such that and . And is said to be induced by if . For , denotes the graph obtained from by deleting the vertices in together with their incident edges (i.e., is the subgraph of induced by ).
A bipartite graph is written as with and being the bipartitions and the edges [28]. For a , denotes the set of neighbors in , i.e., the set of vertices that are connected with by an edge. A matching of 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 is the matching with the largest size, which value is denoted by . An edge of is said to be admissible if it is contained in some maximum matching of . We say a matching covers , if forms a maximum matching of . If each edge of has a non-negative weight, the maximum (minimum) weight maximum matching is the maximum matching of 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 , the Dulmage-Mendelsohn decomposition (DM-decomposition) of is to decompose into subgraphs (, each is called a DM-component) satisfying:
1) , for , with ; ;
2) for , if , and if ; Moreover, each is admissible in , .
3) unless , and only if , where ;
4) () is a maximum matching of iif and is a maximum matching of for ;
5) cannot be decomposed into more components satisfying conditions 1)-4).
In Definition 4, (if exists) is called the horizontal tail, and the vertical tail. Additionally, if contains only one DM-component, then is called DM-irreducible.
For an matrix , the bipartite graph associated with is given by , where (resp. ) corresponds to the rows (columns) of , and . It is known that, for a generic matrix , [28]. For and , we also write as the submatrix of whose rows correspond to and columns to , where elements in and are ordered. Note such an expression of submatrices is invariant subject to row and column permutations on , that is, upon letting and be two permutation matrices and , is the same as .
For a structured pair , its associated graph is given by , where , . A vertex is said to be input-reachable, if is reachable from at least one vertices of in .
Lemma 1
[13] Given , , is structurally controllable, if and only if i) every vertex is input-reachable, and ii) .
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 .
Theorem 1
Given , , let be a generic realization of , with the parameters for the entries being and for the entries being . is PSSC, if and only if and has the following form
(5) |
where is a nonzero polynomial (including the constant ) of , .
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 , , let and be defined in Definition 3. Then, depending on , either for almost all , is controllable for every , or for almost all , there is a such that 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 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 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 is PSSC, then for almost all and all , the corresponding realization is controllable. However, if is not PSSC, it is possible that for in a full-dimensional semi-algebraic set of (i.e., a subset of defined by a finite sequence of polynomial equations and inequalities, or their finite unions [34]), the corresponding realization of is controllable for all ; while for in the semi-algebraic set of full dimension, there is a that makes the corresponding realization uncontrollable. This is because, as argued in the proof of Theorem 1, then has a form of (15), which does not necessarily have real solutions for . 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 be the structured system obtained from by preserving its th entry and replacing the remaining indeterminate entries with entries, for , i.e., if and , and otherwise. Then, is PSSC, if and only if for each , is PSSC.
The above proposition is beneficial in deriving conditions for PSSC, as it actually transforms the problem of verifying PSSC of to subproblems of verifying PSSC of for each .
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 , i.e. systems that contain only one entry, and then on the general systems. Recall from the PBH test [29, 31], an uncontrollable mode for system is a such that . 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 , let consist of linearly independent row vectors that span the left null space of . Then, for any , is of full row rank, if and only if is of full row rank.
Proposition 3
Suppose is structurally controllable and contains only one entry, with its position being . Moreover, let be divided into in the way described in Definition 3. For almost all , there exist no and nonzero vector that satisfy , if and only if one of the following conditions holds:
a1) ;
a2) For each , , in which ;222If , this condition is set to be satisfied (the same below).
a3) .
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 associated with is defined in Section 2.3, where , .
Proposition 4
Under the same setting in Proposition 3, conditions a1)-a3) are respectively equivalent to
b1) contains a maximum matching that does not match ;
b2) Either or for each
, ;
b3) .
We are now presenting the conditions for nonzero uncontrollable modes. Again, suppose is structurally controllable and contains only one entry, with its position being . For a generic realization of , define a generic matrix pencil , and let and . Let be the bipartite graph associated with , defined as follows: , , and , in which and . An edge is called a -edge if , and a self-loop if . is the subgraph of induced by .
Let () be the DM-components of . From [27, Lem 4], as is structurally controllable, we have for . Hence, , . By the correspondence between a matrix and its associated bipartite graph, the DM-decomposition of corresponds to that the matrix is transformed into the block-triangular form via two permutation matrices and [28]:
(6) |
where the submatrix corresponds to (). Accordingly, let .
For each , let and be respectively the minimum and maximum numbers of -edges contained in a matching over all maximum matchings of . We borrow the boolean function for from [27], which is defined as
(7) |
The following lemma explains the motivation of introducing .
Lemma 3 (Lemma 9 of [27])
With notations as above,
() generically has nonzero roots for , if and only if .
Next, define the set
(8) |
Moreover, for a vertex , define a set as
(9) |
It is clear implies that every maximum matching of covers . As we shall see, is the set of indices of all DM-components associated with which has a nonzero root , such that is of full row rank ( is obtained by substituting into ). Hereafter, by saying (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 be defined in (6). Assume . The following properties are true:
1) Given a , is generically row rank deficient for all , if and only if ;333Please note, since does not contain the column corresponding to , is the same as . The same case holds for other matrices.
2) Suppose . For all nonzero making
of full row rank and simultaneously, is not of full row rank for a given , if and only if every maximum matching of covers .
Proposition 5
Suppose is structurally controllable and contains only one entry, with its position being . Let be divided into in the way described in Definition 3. For almost all , there exist no , nonzero complex number and nonzero vector that satisfy , if and only if one of the following conditions holds:
c1) ;
c2) , , and for each , every maximum matching of covers .444If , 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 , , is PSSC, if and only if is structurally controllable, and for each , the following two conditions hold for the system , recalling 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 in Example 2. In this system, . For , the associated , , and (), as well as the DM-decomposition of , are given respectively in Figs. 2(a), 2(b), and 2(c). From Fig. 2(a), condition b2) is fulfilled, as the bipartite graph has a maximum matching with size . From Fig. 2(c), and , implying condition c1) is fulfilled. Similarly, for , the associated and (), as well as its DM-decomposition, are given respectively in Figs. 2(a) and 2(d). Fig. 2(a) shows , meaning condition b2) is satisfied. Fig. 2(d) indicates and , meanwhile, . This means condition c2) is satisfied. As a consequence, is PSSC, which is consistent with the analysis in Example 2. Further, suppose we change to
(10) |
Then, . For each , and the corresponding are of the same form as Figs. 2(a) and 2(b), respectively. For , satisfies condition b2), with the corresponding and its DM-decomposition given in Fig. 2(e). It turns out that, , and meanwhile, . Therefore, neither condition c1) nor c2) is fulfilled, meaning in (10) is not PSSC. This can be validated by using Theorem 1. Alternatively, we can obtain the same conclusion by inspecting that, for , the corresponding (; see Fig. 2(f)) does not satisfy condition c1) or c2). See, associated with , , and there is a maximum matching of that does not cover .






Example 4 (Example 1 continuing)
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, can be determined, as argued in [27], via computing the minimum/maximum weight maximum matchings. Specifically, if we assign weight to each -edge of and weight to the other edges, then by definition, equals the maximum weight of all maximum matchings of , and equals the minimum weight over all maximum matchings of . After determining , for a given , assign weight to the edge of as follows
(12) |
It is not hard to see, , if and only if the minimum weight over all maximum matchings of the weighted is less than . Indeed, if such a minimum weight is less than , then there must be a maximum matching of that does not cover . On the other hand, if such a minimum weight is equal to , then each maximum matching of should cover . In such a manner, we can determine the set . Afterwards, if and , for a given , 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 of as follows
(13) |
Then, similarly, it follows that, every maximum matching of covers , if and only if the minimum weight over all maximum matchings of the weighted equals . It is remarkable that determining whether can be implemented in a similar manner, i.e., by replacing in (13) with .
Let us figure out the computational complexity of the above procedure. Note that determining the maximum matching of a bipartite graph with vertices and edges incurs time via the Hopcroft-Karp algorithm, and there are algorithms computing the maximum weighted matching in [37]. In addition, DM-decomposition has the same complexity as finding a maximum matching [28]. Verifying whether is structurally controllable can invoke the strongly-connected component decomposition and maximum matching algorithms, which incurs . Moreover, as analyzed above, for each , conditions b1), b2) and b3) can be verified in time, and conditions c1) and c2) can be checked in time complexity at most
. To sum up, since there are entries, the total time complexity of Theorem 2 is at most .
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 entry in . 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 is divided into 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 entry.
Proposition 6
For a multi-input system , assume that there is only one entry in . Then, either for almost all , is controllable for each , or for almost all , there is a such that is uncontrollable.
The generic property presented above is characterized by PSSC of . Next, similar to Proposition 4, the following proposition gives the necessary and sufficient condition for the nonexistence of zero uncontrollable modes.
Proposition 7
Suppose is structurally controllable and contains only one in its th position. For almost all , there exist no and nonzero vector that satisfy , if and only if one of the following conditions holds
d1) contains a maximum matching that does not match ;
d2) For each (if exists),
;
d3) .
Suppose is structurally controllable and contains only one in its th position. Let be a generic realization of . Define a generic matrix pencil , and let . Let and be the bipartite graphs associated with and , respectively, defined in the same way as in Section 3. Note compared with the single-input case, the essential difference is that , which results in that there are horizontal or vertical tails in DM-decomposing . Owing to the structural controllability of , a trivial extension of [27, Lem 4] shows . Consequently, by Definition 4, there is only a horizontal tail in the DM-decomposition of (). Let () be the DM-components of . The following intermediate result is crucial for the subsequent derivations.
Lemma 5
If , there is generically no nonzero that can make row rank deficient.
Moreover, associated with and (), let and () be defined in the same way as (8) and (9), respectively. Particularly, Lemma 5 implies would contribute no nonzero that can make row rank deficient (thus ). We have the following proposition, providing a necessary and sufficient condition for the nonexistence of nonzero uncontrollable modes.
Proposition 8
Suppose is structurally controllable and contains only one entry in its th position. For almost all , there exist no , nonzero complex number and nonzero vector that satisfy , if and only if one of the following conditions holds
e1) ;
e2) , , and for each , every maximum matching of covers .
Combining Propositions 7 and 8 yields a necessary and sufficient condition for PSSC of with a single entry.
Theorem 3
Suppose contains only one entry in its th position. is PSSC, if and only if: i) 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 .
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 is not square, the proof for genericity needs to consider multiple submatrices of . Besides, as is no longer square, we have to consider the horizontal tail of the DM-components of .
While Theorem 3 is devoted to the single entry case, it can provide some necessary conditions for PSSC of more general cases. Specifically, it is easy to see, for to be PSSC, by preserving arbitrary one of its entries and changing the remaining entries to , the obtained structured system should be PSSC, i.e., satisfying the conditions in Theorem 3; otherwise, cannot be PSSC by Proposition 6.
Example 5
Consider as
(14) |
which is structurally controllable. The associated and (and its DM-decomposition, ) are given in Figs. 3(a) and 3(b). From them, we know has a maximum matching with size , and . This means the conditions of Theorem 3 are satisfied. Hence, is PSSC. Further, it can be found that, by replacing arbitrary one of the indeterminate entries of with a entry and changing the remaining ones to , the obtained structured system is still PSSC. This is consistent with the fact that is actually SSC (c.f. [18, Theo 4]).


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 , 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 , , is SSC in the real field (i.e., is controllable for all real-valued ), if and only if is structurally controllable, and for each , 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 in system controllability. For description convenience, for , given , let be the parameter for the th entry of , and the vector consisting of parameters for the remaining entries. Given that is structurally controllable, for any , depending on what conditions in Theorem 2 are satisfied for , the edge can be classified into:
-
1.
Critical edge: there is a nonzero (complex) value for making the corresponding realization uncontrollable, for almost all , if does not satisfy condition 1) or condition 2) of Theorem 2;
-
2.
Stable edge: there is no nonzero value for that can make the corresponding realization uncontrollable, for almost all , if 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 that is obtained from system (2) by replacing all its indeterminate entries with entries. Corollary 1 yields that 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: , , and ; (b) stable edges: , and . It is easy to validate the above assertion by noting that the determinant of a generic realization of is exactly the right-hand side of (4).
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 . Since is a nonzero polynomial, making has zero measure in . Therefore, in case () and , it follows , indicating the corresponding realization is always controllable.
Necessity: The necessity of is obvious. If but the remaining condition is not satisfied, then there is a () that exists in two different monomials with different degrees (including zero) for (the degree of is the exponent of ) in . Let and . In this case, write as a polynomial of as
(15) |
in which the coefficients () are polynomials of , and as well as another () is not identically zero. Consider the set . Obviously, the complement of in has zero measure, as has full dimension in . Note in case and , has at least one nonzero root for (as otherwise ). Therefore, for all , there exists such that , 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 , the first case emerges if is PSSC, while the second case emerges if is not PSSC.
Proof of Proposition 2: By Theorem 1, if is PSSC, then for its generic realization , has the form of (5). Let be the generic realization of . It is easy to see, for every , then has the form of (5), which indicates is PSSC.
On the other hand, suppose is PSSC, . Let be the parameter for the th entry of . By Theorem 1, for each , has a factor for some , and any other factor containing does not exist. Consequently, must have a form of (5). This means, is PSSC.
Proof of Proposition 3: Sufficiency: The sufficiency of condition a1) is obvious, as in this case for almost all , . Suppose condition a2) is fulfilled. We only need to consider the case where condition a1) is not fulfilled. In this case, as is structurally controllable, by Lemma 1, must hold. Then, for almost all , . Next, consider two cases: i) and ii) , i.e., . Let be in the left null space of . Note is unique up to scaling. By Lemma 2, we generically have for each ; additionally, in case i), and in case ii). Hence, in case i), for almost all and all , it holds
In case ii), for almost all and all , we have
indicating is always uncontrollable, which is excluded by the structural controllability of .
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 , following the similar arguments to those for condition a2), upon letting be in the left null space of , we have , for almost all . Then, for all , it holds
where (a) is due to the structural controllability of .
Necessity: Suppose none of the three conditions is satisfied. Then, by the structural controllability of and Lemma 1, it must hold that , and there exists at least one so that and . Under these conditions, for almost all , we have . Upon letting be a nonzero vector in the left null space of , Lemma 2 yields that and generically hold. Assign
(16) |
where the inequality is due to the fact is independent of ( is uniquely determined by up to scaling). We therefore have
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).
Lemma 6 (Lemma 9 of [27])
Let be an generic matrix over the variables , and be an constant matrix whose entries are either or , where each row of , as well as each column, has at most one . Let be the bipartite graph associated with the generic matrix pencil (defined in a way similar to ). Let be the set of variables of that appear in the th column of . If is DM-irreducible, then every nonzero root of (if exists) cannot be independent of (if nonempty), for each .
Proof of Lemma 4 counting: Remember for any , has a maximum matching with size , as otherwise it cannot hold that . We first prove property 1), and then property 2).
Necessity in property 1): Suppose . Then, there is some , so that a matching of with size exists that does not cover . Suppose is not matched by . As , Lemma 3 yields generically has nonzero roots for , and let be one of such roots. Due to the block-triangular structure of , satisfies . Note as is DM-irreducible, from Lemma 6, cannot be independent of the indeterminate parameters in the column of corresponding to (note if and , then ).
On the other hand, for any does not share a common factor with except for the power of . By [38, Lem 2], this implies for any does not share a common nonzero root for with , generically. Therefore, denoting where , if is of full row rank, then will be (due to its block-triangular structure). Noting
is not identically zero (because of the existence of ; note , as otherwise every maximum matching of will cover ), any nonzero root of is independent of the parameters in the column of corresponding to . This means, is generically of full row rank, so is .
Sufficiency in property 1) : Suppose . Due to the block-triangular structure of and from Lemma 3, any nonzero root of , denoted by , must be a nonzero root of for some . Consider an arbitrary maximum matching of . Upon letting be the set of vertices in that are matched by , we have . Let , . From [28, Prop 2.1.3],
(17) | ||||
where is the signature associated with . Since each maximum matching of covers , for any , but , as otherwise will contain a maximum matching of that does not cover , meaning , where denotes the set of edges between and in , , . Note
due to the block-triangular structure of and . Hence, . Since this property holds for any maximum matching of , is row rank deficient.
Necessity in property 2): Suppose there is a maximum matching of that does not cover for some (thus does not cover ). By the definition of , there exists a maximum matching of that does not cover . In this case, following the similar argument to the proof of necessity in property 1), any nonzero root of , denoted by , is also a root of , while making and of full row rank simultaneously.
Sufficiency in property 2): From the proof of property 1), we know a nonzero makes of full row rank while , if and only if is a nonzero root of for some satisfying the property that, is not covered by some maximum matching of (referred to as condition f)). Indeed, following the necessity in property 1), we know a satisfying condition f) is sufficient; and following the similar arguments to the proof for sufficiency in property 1), we know condition f) for is necessary. By definition, the set of such a satisfying condition f) is exactly . Again, following the similar manner to the proof for sufficiency in property 1) by replacing with in (17), it turns out is row rank deficient if every maximum matching of covers .
Proof of Proposition 5: In the following, we consider the generic realization of . Recall . Let be the parameter in the th entry of . For , suppose corresponds to the th row of after the row permutation on .
Sufficiency: Suppose condition c1) is fulfilled. If , by Lemma 3, there is generically no making row rank deficient, due to the block-triangular structure of . Noting from (6), is generically of full rank for all nonzero . This means is sufficient. Now consider but . In this case, consider an arbitrary that makes row rank deficient. As is structurally controllable, generically has rank (otherwise will be row rank deficient). Hence, upon letting be a nonzero vector in the left null space of , is unique up to scaling. From property 1) of Lemma 4, is row rank deficient (recalling ). By Lemma 2, we have . Hence, for any nonzero value of ,
where is due to the controllability of and that is independent of . Therefore, condition c1) is sufficient.
Suppose condition c2) is fulfilled. In this case, for those nonzero roots of that make row rank deficient, following a similar argument to the above, we can obtain that there is no and ( and ) making . For the nonzero root of that makes of full row rank, let be a nonzero vector in the left null space of ( is unique up to scaling). Then, according to property 2) of Lemma 4, it turns out that and for all . We therefore have
where (a) is due to as . Hence, condition c2) is also sufficient.
Necessity: Suppose neither condition c1), nor condition c2) holds. Then, either i): , and the third item of condition c2) does not hold, or ii): , , and the third item of condition 2) holds. In case i), suppose the third item of condition 2) does not hold for a vertex . Considering the generic realization , by property 2) of Lemma 4, we known there is some nonzero making row rank deficient, while and are of full row rank. Note it generically holds , as otherwise , contradicting the structural controllability of . Let be a nonzero vector spanning the left null space of . Then, by Lemma 2, and . Upon letting , we have
which leads to , by substituting , and noting is invertible. That is, by assigning
(18) |
we can make . Note if , is independent of and (as contains a free parameter); if , in case contains more than one nonzero items, at least one is independent of and . Hence, in all these circumstances, it is assured .
As for case ii), following similar arguments to case i), upon letting be in the left null space of for some nonzero making while of full row rank, we have and for all satisfying . Hence, if , then , leading to . Since is a generic realization of , in both cases, we know for almost all , there exist , nonzero and satisfying .
Proof of Proposition 6: The case where is not structurally controllable is trivial. Now assume structural controllability of , and consider its generic realization . Let be the parameter for the unique entry in , and the collection of parameters for the remaining indeterminate entries. Let be the greatest common divisor among all determinants of the submatrices of . According to [38, Lem 2], for almost all (recalling is the number of entries), the determinants of all the submatrices of share a common zero for , if and only if the leading degree for in is no less than one. Therefore, if , where and is a nonzero polynomial of , then for almost all satisfying and all , is of full row rank, indicating that the corresponding realization is controllable. Otherwise, if contains two monomials with different degrees for , then for almost all , there exists a nonzero solution satisfying (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 is the horizontal tail, from [28, Corollary 2.2.23], for each , . From [27, Lem 9], for all , every maximum matching of corresponds to a nonzero term in the determinant of a submatrix of that cannot be cancelled out by other terms. Therefore, supposing there is a nonzero value, denoted by , such that is of row rank deficient, should depend only on the free parameters in , . Applying this across , it turns out that is independent of the free parameters in each column of , causing a contradiction.
Proof of Proposition 8: From Lemmas 3 and 5, if is a nonzero root of for some , then also makes 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 makes
row rank deficient for a given , if and only if makes row rank deficient. Having observed this, it can be found Lemma 4 still holds for the rectangular matrix (by changing to the row-rank deficient of ). 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 has the form , under the proposed conditions, where is a generic realization of with the indeterminate parameters being . It is then obvious that for all , the corresponding system realization is controllable.
Necessity: We prove the necessity by contradiction. Let be the parameter for the th entry of for a , and the vector consisting of the parameters for the remaining entries (except ). The necessity of structural controllability of is obvious. Now assume is structurally controllable. Suppose i) is not fulfilled for a . From the proof for necessity of Proposition 3, for almost all , there exists nonzero real (expressed in (16), where is chosen to be real), such that the corresponding real system realization is uncontrollable. This means is not SSC in the real field.
Furthermore, suppose ii) is not fulfilled for a . We first consider the case and the third item of condition c2) does not hold for some . Following the proof for necessity of Proposition 5, in this case, to construct a real uncontrollable realization of , we only need to demonstrate in (18) can be real (thus the vector therein can be real too). To this end, recall in (18), needs to satisfy: requirement 1) makes row rank deficient, and requirement 2) makes and of full row rank. Notice from the proof for necessity of Lemma 4, if is a real nonzero root of for some , where is such that there is a maximum matching of not covering , then generically satisfies the previous two requirements. Let us assume that vertex of is not matched by , and corresponds to a nonzero entry of . As , there is a maximum matching of not covering . Similarly, assume that vertex of is not matched by , and corresponds to a nonzero entry of . Let and be respectively the set of vertices in and in that are matched by a matching . Denote and , where or .
Since is DM-irreducible, by Lemma 6, does not have a fixed nonzero root for as varies. Note also for any fixed , is both irreducible (i.e., cannot be factored as two polynomials with smaller degrees) and multi-affine in (i.e., every variable in occurs with degree or in every term). Hence, can be written as , where and are polynomials of and , with being the collection of indeterminate parameters in except . Based on the above arguments, the proof for necessity of Lemma 4 indicates that the following constraints suffice to satisfy requirements 1) and 2):
(19) | ||||
(20) | ||||
(21) | ||||
(22) | ||||
(23) | ||||
(24) |
Indeed, (19) ensures leading to , while (22), (23) and (24) ensure that requirement 2) is fulfilled (due to the block-diagonal structure of ). Now consider and as nonzero real numbers. If , the constraints (20)-(24) do not not involve . Then, for almost all nonzero real numbers of and , there is a satisfying (19)-(24). If , then satisfying (19) depends generically on and (by Lemma 6), thus not independent of , while satisfying (23) is independent of . This means, again, for almost all nonzero real numbers of and , there is satisfying (19)-(24). Therefore, for almost all , we can always find such that requirements 1) and 2) are satisfied. Then, after determining according to (18), where and are both real, the corresponding real system realization will be uncontrollable, meaning that is not SSC in the real field. For the case where , , and the third item of condition 2) holds, we can adopt a similar argument to construct a real uncontrollable realization of . 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.