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

Resilient Strong Structural Controllability in Networks
using Leaky Forcing in Graphs

Waseem Abbas W. Abbas is with the Systems Engineering Department at the University of Texas at Dallas, Richardson, TX (Email: [email protected]).
Abstract

This paper studies the problem of selecting input nodes (leaders) to make networks strong structurally controllable despite misbehaving nodes and edges. We utilize a graph-based characterization of network strong structural controllability (SSC) in terms of zero forcing in graphs, which is a dynamic coloring of nodes. We consider three types of misbehaving nodes and edges that disrupt the zero forcing process in graphs, thus, deteriorating the network SSC. Then, we examine a leader selection guaranteeing network SSC by ensuring the accuracy of the zero forcing process, despite kk misbehaving nodes/edges. Our main result shows that a network is resilient to kk misbehaving nodes/edges under one threat model if and only if it is resilient to the same number of failures under the other threat models. Thus, resilience against one threat model implies resilience against the other models. We then discuss the computational aspects of leader selection for resilient SSC and present a numerical evaluation.

Index Terms:
Strong structural controllability, zero forcing sets, resilience in networks, dynamics over graphs.

I Introduction

The ability to control a network of agents through external control signals profoundly depends on the underlying network topology [1, 2, 3, 4, 5]. The control signals are injected into the network via a subset of agents, typically called leaders. Computing a minimum number of leaders and their locations within the network for complete controllability are central issues in the cooperative control of networks. A popular approach for the leader selection problem, which is well-studied, is to characterize network controllability from a graph-theoretic perspective. Then formulate an appropriate vertex selection problem in graphs that can be solved through various combinatorial methods.

To this end, several graph-based characterizations of network controllability have been presented, including those based on graph distances, zero forcing in graphs, dominating sets, matching in graphs, and others (e.g., [6, 7, 8, 9, 10, 11, 12]). The notion of zero forcing in graphs is particularly useful for strong structural controllability (SSC) in networks, which ascertains that for a given set of leaders, if the network is controllable for one set of edge weights, then it is controllable for all feasible edge weights. The zero forcing phenomenon, introduced in [13], is a dynamic coloring of nodes in a graph in which only a subset of nodes are initially colored, causing other nodes to change colors according to some rules. A set of initially colored nodes that eventually color the entire graph is called a zero forcing set of the graph. Interestingly, a leader set for the strong structural controllability of a network is directly related to the idea of a zero forcing set in the underlying network graph [6, 14, 15]. A set of leaders makes a given network strong structurally controllable if the corresponding nodes in the underlying graph constitute a zero forcing set.

In this paper, we consider the leader selection for resilient strong structural controllability, that is, the network remains SSC despite some misbehaving nodes or edges deviating from their normal functionalities. A single misbehaving agent or an edge (due to a fault or an adversary) can severely deteriorate the network’s controllability. Considering the susceptibility of networks to adversarial attacks and failures, the resilient SSC is a crucial paradigm in network control systems. We consider three different abnormal behaviors by nodes (agents) and edges. These misbehaving nodes and edges primarily disrupt the zero forcing process in the underlying network graph, affecting the network’s SSC. In particular, we examine leak nodes (agents that do not follow the rules of the zero forcing process), non-forcing edges (links that again disrupt the zero forcing process), and removable edges (links that are lost and affect interconnections between nodes). We then study the leader set guaranteeing the network SSC despite a given number of such misbehaving nodes/edges. For this, we utilize the idea of leaky forcing, a variant of zero forcing in graphs recently introduced in [16, 17]. Our main result shows that the resilient leader selection against one abnormal behavior implies resilience against the other two abnormal behaviors. The main contributions are below:

  • We study the resilient SSC problem in networks by connecting it to the notion of leaky forcing in graphs.

  • We consider three models of misbehaving nodes/edges in networks, including node leaks, non-forcing edges, and removable edges, and formulate leader selection in networks to guarantee that the network remains SSC despite kk such abnormal nodes/edges (Section III).

  • We show that a given leader set is resilient to kk misbehaving nodes/edges under one model if and only if it is resilient to the same number of misbehaving nodes/edges in the other model (Section IV).

  • Finally, we discuss computational aspects of selecting leaders for the resilient SSC and also present a numerical evaluation (Section V).

There have been works discussing the impact of nodes/edges addition and deletion on the network’s controllability, for instance, [18, 19, 20, 21]. Similarly, some researchers have considered the effects of actuators/input node failures on the overall controllability of systems. Some also propose methods to have a sufficient number of input nodes maintaining network controllability despite such failures  [22, 23, 24, 25]. Our work differs from theirs as we consider node/edge misbehavior models beyond deletion. These misbehaving nodes/edges remain within the network disrupting the underlying graph-theoretic process required to ensure SSC. Our goal is to have leader sets guaranteeing SSC despite the existence of any kk such nodes/edges within the network. For this, we utilize the idea of leaky forcing in graphs.

The rest of the paper is organized as follows: Section II presents preliminary ideas related to network controllability and zero forcing in graphs. Section III explains various threat models and formulates the leader selection problems for the resilient SSC in networks. Section IV presents the main result regarding the equivalence of leader selection for resilient SSC under various threat models. Section V discusses the computational aspects and provide numerical illustrations. Finally, Section VI concludes the paper.

II Preliminaries

We consider a network of agents modeled by an undirected graph G=(V,E)G=(V,E). The node set VV, and the edge set EE, represent agents and interconnections between agents, respectively. The edge between nodes ii and jj is denoted by an unordered pair (i,j)(i,j). The neighborhood of uu is 𝒩i={jV|(i,j)E}\mathcal{N}_{i}=\{j\in V|\;(i,j)\in E\}. The distance between nodes ii and jj, denoted by d(i,j)d(i,j), is the number of edges in the shortest path between them. For a given graph G=(V,E)G=(V,E) with |V|=n|V|=n nodes, we define a family of symmetric matrices (G)\mathcal{M}(G) as below:

(G)={Mn×n|M=M, and for ij,Mij0(i,j)E}.\begin{split}\mathcal{M}(G)=\{M\in\mathbb{R}^{n\times n}\;&|\;M=M^{\top},\text{ and for }i\neq j,\\ &M_{ij}\neq 0\Leftrightarrow(i,j)\in E\}.\end{split} (1)

Note that the location of zero and non-zero entries in every M(G)M\in\mathcal{M}(G) remains the same, and is entirely described by the edge set of GG.

II-A System Model and Graph Controllability

We consider the following leader-follower system defined over a graph G=(V,E)G=(V,E).

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

where x(t)nx(t)\in\mathbb{R}^{n} is the system state, u(t)mu(t)\in\mathbb{R}^{m} is the external input, M(G)M\in\mathcal{M}(G) (as in (1)) is the system matrix, and Bn×mB\in\mathbb{R}^{n\times m} is the input matrix describing the leader (input) nodes. If V={v1,v2,,vn}V=\{v_{1},v_{2},\cdots,v_{n}\}, and V={1,2,,m}VV^{\prime}=\{\ell_{1},\ell_{2},\cdots,\ell_{m}\}\subseteq V is a set of leader nodes, then we define BB as follows:

[B]ij={1if vi=j,0otherwise.[B]_{ij}=\left\{\begin{array}[]{lll}1&\text{if }v_{i}=\ell_{j},\\ 0&\text{otherwise.}\end{array}\right. (3)

We observe that the family of matrices (G)\mathcal{M}(G) describes a number of system matrices defined over GG, including the Laplacian and adjacency matrices.

For a given graph GG, system matrix M(G)M\in\mathcal{M}(G), and input matrix BB, the system in (2) is called controllable if there exists an input to drive the system from an arbitrary initial state x(t0)x(t_{0}) to an arbitrary final state x(tf)x(t_{f}). In this case, we say that (M,B)(M,B) is a controllable pair. A pair (M,B)(M,B) is controllable if and only if the rank of the controllability matrix Γ(M,B)\Gamma(M,B), defined below, is |V|=n|V|=n (i.e., full rank).

Γ(M,B)=[BMBM2BMn1B].\Gamma(M,B)=\left[\begin{array}[]{lllll}B&MB&M^{2}B&\cdots&M^{n-1}B\\ \end{array}\right]. (4)

Since leader nodes VV^{\prime} define the input matrix BB, we sometimes abuse the notation slightly and use (M,V)(M,V^{\prime}) is controllable to denote that (M,B)(M,B) is a controllable pair.

Definition 1

(Network Strong Structural Controllability) A network G=(V,E)G=(V,E) with a leader set VVV^{\prime}\subseteq V is strong structurally controllable if and only if (M,V)(M,V^{\prime}) is a controllable pair for all M(G)M\in\mathcal{M}(G).

Network strong structural controllability is a stronger notion compared to the (weak) structural controllability, which requires the existence of at least one matrix M(G)M\in\mathcal{M}(G) for which (M,V)(M,V^{\prime}) is a controllable pair.

II-B Network Controllability and Zero Forcing in Graphs

A central problem in network controllability is to compute a minimum leader set VVV^{\prime}\subseteq V that makes the network strong structurally controllable (as defined above). The problem is often referred to as the minimum leader selection for network controllability. A graph-theoretic characterization of the minimum leader set rendering the network strong structurally controllable is remarkably useful here. Monshizadeh et al. characterized the minimum leader set for network strong structural controllability in [6] using the notion of zero forcing in graphs, which is related to the dynamic coloring of nodes. We introduce zero forcing ideas below and then state the main result from [6].

Definition 2

(Zero forcing) Given a graph G=(V,E)G=(V,E) whose nodes are initially colored either black or white. Consider the following node color changing rule: If vVv\in V is colored black and has exactly one white neighbor uu, change the color of uu to black. Zero forcing is the application of the above rule until no further color changes are possible.

Definition 3

(Force) A force is an application of the color changing rule due to which the color of a white node uu is changed to black by some black node vv. We say that vv forced uu and denote it by vuv\rightarrow u.

Definition 4

(Input and derived sets) For a graph G=(V,E)G=(V,E) with an initial set of black nodes VVV^{\prime}\subseteq V, the derived set of VV^{\prime}, denoted by 𝒟(V)\mathcal{D}(V^{\prime}), is the set of black nodes obtained after applying the zero forcing rule exhaustively. The initial set of black nodes VV^{\prime} is the set of input nodes.

We note that for a given input set VV^{\prime}, the derived set 𝒟(V)\mathcal{D}(V^{\prime}) is unique [13].

Definition 5

(Zero forcing set (ZFS) and zero forcing number) For a graph G=(V,E)G=(V,E), an input set VVV^{\prime}\subseteq V is a ZFS if 𝒟(V)=V\mathcal{D}(V^{\prime})=V (i.e., all nodes are colored black after the exhaustive application of the zero forcing rule). We denote a ZFS by Z0Z_{0}. The number of nodes in the minimum ZFS is the zero forcing number of the graph and denoted by z0(G)z_{0}(G).

Figure 1 illustrates the zero forcing terms defined above.

Refer to caption
Figure 1: V={v4,v5}V^{\prime}=\{v_{4},v_{5}\} is a ZFS of the graph along with a sequence of forces coloring all nodes black.

A leader set for the strong structural controllability is closely related to the notion of ZFS of the network graph. A direct consequence of Theorems IV.4, IV.8, and Proposition IV.9 in [6] is the following result:

Theorem 2.1

[6] The undirected network G=(V,E)G=(V,E) is strong structurally controllable with a leader set VVV^{\prime}\subseteq V (as in Definition 1) if and only if VV^{\prime} is a ZFS of GG, (i.e., 𝒟(V)=V\mathcal{D}(V^{\prime})=V).

Thus, ZFS in graphs is an important idea from the network controllability perspective and it completely characterizes the leader set for the strong structural controllability of the network. The above results implies that the minimum number of leaders needed for the network strong structural controllability is same as the zero forcing number of the network graph. We note that computing a minimum ZFS and zero forcing number are NP-hrad in general [26]. However, there are several heuristics to compute a small-sized ZFS, for instance, see [27, 28, 29, 30].

III Resilient Strong Structural Controllability (SSC) in Networks

The equivalence between ZFS and leader set for network SSC is significant. The zero forcing process can be disrupted by some nodes/edges abnormal behaviors or edge failures, and consequently, the network SSC can be adversely impacted. For instance, consider the network in Figure 2 with the leader set V={v1,v2}V^{\prime}=\{v_{1},v_{2}\}. If all nodes are normal, the ZF process initiated by the leader set VV^{\prime} will eventually color all nodes in VV (i.e., 𝒟(V)=V\mathcal{D}(V^{\prime})=V), and the network will be strong structurally controllable with VV^{\prime}. However, if v5v_{5} behaves abnormally in the sense that it does not force any other node, then the zero forcing process will be hindered and 𝒟(V)V\mathcal{D}(V^{\prime})\neq V asserting that the network is not strong structurally controllable. Similarly, if we consider all nodes normal, but the edge (v5,v6)(v_{5},v_{6}) is removed, the ZF process will again be disrupted, causing 𝒟(V)V\mathcal{D}(V^{\prime})\neq V, which means the network is not strong structurally controllable with VV^{\prime}.

Refer to caption
(a) V={v1,v2}V^{\prime}=\{v_{1},v_{2}\}
Refer to caption
(b)
Refer to caption
(c)
Figure 2: (a) VV^{\prime} is a ZFS of GG. (b) v5v_{5} is a misbehaving node not forcing any other node. (c) (v5,v6)(v_{5},v_{6}) is a misbehaving edge.

However, if V={v1,v2,v7}V^{\prime}=\{v_{1},v_{2},v_{7}\}, then the network remains SSC despite any single misbehaving node (refusing to force other nodes) or an edge. Thus, the network SSC can be preserved even in the presence of misbehaving nodes or edges through some redundant leader nodes selected carefully. Our goal here is to study: how we can guarantee resilient network SSC in the face of some misbehaving nodes and edges that might disrupt the zero forcing process and impact the network SSC.

Next, we will consider three different abnormal node and edge behaviors disrupting the zero forcing process. Then, we present leader selections guaranteeing all nodes in the network get colored due to the zero forcing process despite a certain number of misbehaving nodes and edges, thus achieving the resilient network SSC. Our main result (in Section IV) shows that resilience to one type of misbehaving nodes/edges implies resilience to the other kinds of misbehaving nodes/edges.

III-A Failure Models and Resilience Problems

We consider the following three node and edge misbehaviors that can be caused by the adversarial attack or other abnormality. All of these failures ultimately disrupt the zero forcing process.

1) Leak (non-forcing) nodes: A leak is a node vVv\in V that does not force any other node, i.e., considering vv to be a leak node that is colored black and has exactly one white neighbor, then vv does not force its white neighbor (which it should in case vv was normal). A set of all leaks is the leak set, denoted by LVL\subseteq V.

The term ‘leak node’ is adapted from [16], where such a non-forcing behavior of nodes is introduced. Practically, a leak node can be realized in multiple ways. For instance, if an additional node α\alpha, which is not a part of the original network and is colored white, becomes adjacent to exactly one node, say vv, in the network GG, then vv is unable to force any other node in GG. Figure 3 illustrates this situation.

Refer to caption
Figure 3: vv is a leak node not forcing any other node. Equivalently, an outside node α\alpha becomes adjacent to vv and prevents vv from forcing any node.

Now the resilience problem is to have a (minimal) leader set such that all nodes are colored at the end of the zero forcing process despite \ell leak nodes, which are unknown. For a given \ell, computing such a leader set is referred to as the \ell-leaky forcing set problem [16, 17]. We formally define the leaky derived set and leaky forcing set below:

Definition 6

(Leaky derived set) Given a graph G=(V,E)G=(V,E), input set VV^{\prime}, and a leak set LL, then the set of black nodes obtained after applying the zero forcing rule exhaustively while considering the leaks in LL is the leaky derived set, denoted by 𝒟L(V)\mathcal{D}_{L}(V^{\prime}).

Definition 7

(\ell-leaky forcing set (\ell-LFS)) An input set VVV^{\prime}\subseteq V is an \ell-LFS if for any leak set LVL\subset V with \ell leaks, 𝒟L(V)=V\mathcal{D}_{L}(V^{\prime})=V. In other words, starting with VV^{\prime}, all nodes are colored black by iteratively applying the zero forcing rule with any \ell leaks. The cardinality of the minimum \ell-LFS is the \ell-forcing number of GG, denoted by z(G)z_{\ell}(G).

We note that for =0\ell=0, the \ell-forcing number is same as the zero forcing number. Further, we know [16],

z0(G)z1(G)z|V|(G).z_{0}(G)\;\leq\;z_{1}(G)\;\leq\;\cdots\;\leq\;z_{|V|}(G). (5)

2) Non-forcing edges: Here, we consider edge attacks through which an edge cannot be used by either of its end nodes to force the other end node. We call such an edge as a non-forcing edge. In particular, an edge (u,v)(u,v) is a non-forcing edge if uu cannot force vv, and vv cannot force vv. In this case, the resilience problem is to find the \ell-edge forcing set defined below:

Definition 8

(\ell-Edge forcing set (\ell-EFS)) For a given G=(V,E)G=(V,E) and a positive integer \ell, let EEE_{\ell}\subseteq E be an arbitrary subset of at most \ell non-forcing edges (i.e., |E||E_{\ell}|\leq\ell). An input set VVV^{\prime}\subseteq V is an \ell-EFS if there is a zero forcing process that colors all nodes in VV without using the edges in EE_{\ell} to force nodes.

Figure 4 illustrates the non-forcing edge and 11-EFS. V={v1,v2}V^{\prime}=\{v_{1},v_{2}\} is a ZFS of GG in Figure 4(a). If the edge between v2v_{2} and v3v_{3} is non-forcing, then the derived set consists of only three nodes {v1,v2,v4}\{v_{1},v_{2},v_{4}\} at the end of the zero forcing process. However, if the leader set is V={v1,v2,v7}V^{\prime}=\{v_{1},v_{2},v_{7}\}, then all nodes are colored as a result of the zero forcing process despite any single non-forcing edge.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 4: (a) {v1,v2}\{v_{1},v_{2}\} is a ZFS given all edges are normal. (b) (v2,v3)(v_{2},v_{3}) is a non-forcing edge. (c) {v1,v2,v7}\{v_{1},v_{2},v_{7}\} is a 11-EFS.

3) Removable edges: The third failure model we consider is the one where a maximum of \ell edges are removed from the graph to disrupt the zero forcing process. The corresponding resilience problem is to have enough leaders to guarantee that despite \ell edge removals, the network remains strong structurally controllable, or equivalently all nodes are colored as a result of the zero forcing. In other words, the goal is to find a minimum size \ell-forcing set with removable edges defined below:

Definition 9

\ell-forcing set with removable edges (\ell-FSR) For a given G=(V,E)G=(V,E) and \ell, consider a subgraph G=(V,E)G^{\prime}=(V,E^{\prime}), where EEE^{\prime}\subseteq E and |E||E||E|-|E^{\prime}|\leq\ell. Then, VVV^{\prime}\subseteq V is an \ell-FSR of GG if VV^{\prime} is a ZFS of every such GG^{\prime}. Note that an \ell-FSR must also be a ZFS of GG.

We observe that making an edge non-forcing can be different from removing the edge. For instance, unlike a non-forcing edge, removing an edge can sometimes be useful, as Figure 5 illustrates. If edge (v4,v7)(v_{4},v_{7}) in GG in Figure 5(a) is removed, then V={v1,v2}V^{\prime}=\{v_{1},v_{2}\} is a ZFS of the resulting graph (Figure 5(b)), thus, making the network controllable. However, if (v4,v7)(v_{4},v_{7}) is a non-forcing edge (Figure 5(c)), then {v1,v2}\{v_{1},v_{2}\} is no longer a ZFS of the network.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 5: (a) A graph GG. (b) {v1,v2}\{v_{1},v_{2}\} becomes a ZFS of GG after removing the edge (v4,v7)(v_{4},v_{7}). (c) If (V4,v7)(V_{4},v_{7}) is a non-forcing edge, then {v1,v2}\{v_{1},v_{2}\} is a not a ZFS of GG.

Next, we present the main result showing that a leader set resilient to one misbehavior model is also resilient to the other models.

IV Equivalence of Resilient Leader Selection for Various Failure Models

Here, for a given G=(V,E)G=(V,E) and \ell, we show the equivalence between \ell-LFS, \ell-EFS and \ell-FSR. As a result, we show that a leader set VVV^{\prime}\subseteq V ensures resilient controllability against one misbehavior model if and only if it extends resilience to the other two models (discussed above). For instance, a leader set VV^{\prime} that is resilient to \ell non-forcing nodes must also be resilient to \ell non-forcing edges and \ell removable edges simultaneously. We introduce the following terms as in [31, 17].

Definition 10

Consider a graph G=(V,E)G=(V,E), input set VVV^{\prime}\subseteq V and the corresponding derived set 𝒟(V)\mathcal{D}(V^{\prime}), then we define the following terms:

  • A chronological list of forces is a list of forces recorded in the order in which they are performed to construct the derived set.

  • A forcing process FF is a set of forces containing a chronological ordering of forces through which all nodes in VV are colored black (i.e., 𝒟(V)=V\mathcal{D}(V^{\prime})=V).

  • A forcing chain is a sequence of forces vivi+1v_{i}\rightarrow v_{i+1}, for i=1,2,,k1i=1,2,\cdots,k-1. We denote such a forcing sequence by v1v2vk1vkv_{1}\rightarrow v_{2}\rightarrow\cdots\rightarrow v_{k-1}\rightarrow v_{k}.

  • A maximal forcing chain v1v2vk1vkv_{1}\rightarrow v_{2}\rightarrow\cdots\rightarrow v_{k-1}\rightarrow v_{k} is a forcing chain such that v1Vv_{1}\in V^{\prime} and vkv_{k} does not force ant other node in GG.

  • A total forcing set of VV^{\prime}, denoted by (V)\mathcal{F}(V^{\prime}), is a set of all possible forces given an input VV^{\prime}, i.e., vivj(V)v_{i}\rightarrow v_{j}\in\mathcal{F}(V^{\prime}) if there is a forcing process in GG containing vivjv_{i}\rightarrow v_{j}.

  • A total forcing set with leaks LL and input set VV^{\prime}, denoted by L(V)\mathcal{F}_{L}(V^{\prime}), is a set of all possible forces given an input set VV^{\prime} and leaks LL. In other words, if vivjL(V)v_{i}\rightarrow v_{j}\in\mathcal{F}_{L}(V^{\prime}), then viLv_{i}\notin L and there is a forcing process containing the force vivjv_{i}\rightarrow v_{j}.

Next, consider G=(V,E)G=(V,E), a ZFS VVV^{\prime}\subset V, and a forcing process FF with VV^{\prime}. Then, we define the following notations:

F(V)={xyF:yV}.F(V^{\prime})=\{x\rightarrow y\in F:\;y\notin V^{\prime}\}. (6)

Similarly,

F\F(V)={xyF:yV}.F\backslash F(V^{\prime})=\{x\rightarrow y\in F:\;y\in V^{\prime}\}. (7)

We now state some results from [17] that will be used later.

Lemma 4.1

[17] Consider G=(V,E)G=(V,E) and a ZFS VV^{\prime}. Let FF and FF^{\prime} be two forcing processes. Then, (F\F(V~))F(V~)(F\backslash F(\tilde{V}))\cup F^{\prime}(\tilde{V}) is a forcing process with VV^{\prime} for any V~\tilde{V} obtained from VV^{\prime} using FF. Here,

Lemma 4.2

[17] A set VV^{\prime} is a 11-LFS if and only if vVV\forall v\in V\setminus V^{\prime}, there exists xv(V)x\rightarrow v\in\mathcal{F}(V^{\prime}), yv(V)y\rightarrow v\in\mathcal{F}(V^{\prime}), where yxy\notin x.

Theorem 4.3

[17] A set VV^{\prime} is a \ell-LFS if and only if VV^{\prime} is an (1)(\ell-1)-LFS for every set of of 1\ell-1 leaks LL and vVVv\in V\setminus V^{\prime}, there exists xvL(V)x\rightarrow v\in\mathcal{F}_{L}(V^{\prime}), yvL(V)y\rightarrow v\in\mathcal{F}_{L}(V^{\prime}), where yxy\notin x.

IV-A Main Result

Our main result here is to show the following:

Theorem 4.4

Given a graph G=(V,E)G=(V,E), input set VVV^{\prime}\subseteq V, and a positive integer |V|\ell\leq|V|, the following statements are equivalent:

  1. 1.

    VV^{\prime} is an \ell-LFS.

  2. 2.

    VV^{\prime} is an \ell-EFS.

  3. 3.

    VV^{\prime} is an \ell-FSR.

We recall that notions of \ell-LFS, \ell-EFS, and \ell-FSR are explained in Definitions 7, 8, and 9, respectively. To prove Theorem 4.4, we need some intermediate results.

Lemma 4.5

If VV^{\prime} is a 11-LFS of G=(V,E)G=(V,E), then for every edge e=(u,v)Ee=(u,v)\in E, there exists a zero forcing process FeF_{e} that does not use ee, i.e., uvFeu\rightarrow v\notin F_{e} and vuFev\rightarrow u\notin F_{e}.

Proof:

Consider a forcing process FF which contains uvu\rightarrow v. By Lemma 4.2, there exists a forcing process FF^{\prime} such that xvx\rightarrow v, where xux\neq u. It suffices to show that there exists a forcing process FeF_{e} such that xvx\rightarrow v and yuy\rightarrow u, where yvy\neq v. If uVu\in V^{\prime}, then Fe=FF_{e}=F^{\prime}. So, assume uVu\notin V^{\prime}. Now, again consider FF and let V~\tilde{V} be the set of black vertices obtained from VV^{\prime} such that uvu\rightarrow v is valid given V~\tilde{V}, but vV~v\notin\tilde{V}. Note that at this point, uu and all of its neighbors except vv are colored black through FF. Moreover, a node can only force one node. Thus, uu will not force any node in F(V~)F^{\prime}(\tilde{V}), i.e., uqF(V~),qVu\rightarrow q\notin F^{\prime}(\tilde{V}),\;\forall q\in V. Now, consider Fe=(F\F(V~))F(V~)F_{e}=(F\backslash F(\tilde{V}))\cup F^{\prime}(\tilde{V}). Note that uvFeu\rightarrow v\notin F_{e} and vuFev\rightarrow u\notin F_{e}. By Lemma 4.1, FeF_{e} is a forcing process with input VV^{\prime}, which proves the desired claim.  

Lemma 4.6

If VV^{\prime} is a 11-EFS, then VV^{\prime} is a 11-LFS

Proof:

Let FF be a forcing process with VV^{\prime} as input nodes. Consider uVu\in V to be be an arbitrary fixed leak. If uu is an end node of a forcing chain, then uu does not force any node. So, we assume that uvFu\rightarrow v\in F for some vv. Since VV^{\prime} is 11-EFS, there exists another forcing process FF^{\prime} such that vv is not forced by uu. Let xvx\rightarrow v, for some xux\neq u. Now, consider FF till the point uvu\rightarrow v becomes valid but vv is not colored black. Let V~\tilde{V} be the set of black nodes till this point. Note that vV~v\notin\tilde{V}. Note that all nodes in 𝒩[u]{v}\mathcal{N}[u]\setminus\{v\} are colored black and vv is the only node that uu can force. Consider Fu=(F\F(V~))F(V~)F_{u}=(F\backslash F(\tilde{V}))\cup F^{\prime}(\tilde{V}). By Lemma 4.1, FuF_{u} is a valid forcing process with input nodes VV^{\prime} coloring all nodes black. Note that uu is not forcing vv in FuF_{u}, and hence uu is not forcing any node in FuF_{u}. Thus, VV^{\prime} is a 11-LFS, which is the desired claim.  

Lemma 4.7

Let VV^{\prime} be an (1)(\ell-1)-EFS, EE_{\ell} be a set of \ell non-forcing edges, and 𝒟(V)\mathcal{D}(V^{\prime}) be the derived set after forcing. If e=(u,v)Ee=(u,v)\in E_{\ell}, then uu and vv can not be white simultaneously. Moreover, if exactly one end node of ee, say uu, is black, then N[u]{v}𝒟(V)N[u]\setminus\{v\}\subseteq\mathcal{D}(V^{\prime}).

Proof:

If both end nodes of eEe\in E_{\ell} are white, then none of the end nodes can force the other end node. Thus, zero forcing behavior of VV^{\prime} does not change even if ee is not a non-forcing edge. So, if we consider EeE_{\ell}\setminus e as the set of non-forcing edges, 𝒟(V)V\mathcal{D}(V^{\prime})\neq V, implying that VV^{\prime} is not an (1)(\ell-1)-EFS, which is a contradiction. Similarly, let e=(u,v)Ee=(u,v)\in E_{\ell} be such that u𝒟(V)u\in\mathcal{D}(V^{\prime}) and v𝒟(V)v\notin\mathcal{D}(V^{\prime}). Assume that xN(u)x\in N(u) is white and xvx\neq v. Since uu has two white neighbors, uu can not force any node (including vv) even if ee is not a non-forcing edge. Again, considering EeE_{\ell}\setminus e as the set of non-forcing edges will give 𝒟(V)V\mathcal{D}(V^{\prime})\neq V. It means VV^{\prime} is not an (1)(\ell-1)-EFS, which is a contradiction.  

Lemma 4.8

Let VV^{\prime} be an (1)(\ell-1)-LFS and LL be a set of \ell leaks. Then L𝒟(V)L\subseteq\mathcal{D}(V^{\prime}), where 𝒟(V)\mathcal{D}(V^{\prime}) is a derived set with leaks. Also, each vLv\in L has at most one white neighbor.

Proof:

Assume vLv\in L is white, i.e., v𝒟(V)v\notin\mathcal{D}(V^{\prime}). A white leak node does not change the zero forcing behavior of the black nodes. So, we consider L=L{v}L^{\prime}=L\setminus\{v\}. Since VV^{\prime} is an (1)(\ell-1)-LFS, so all nodes should be black for any (1)(\ell-1) leaks. However, all nodes are not black, which is a contradiction. For the second part, assume that there is leak node vLv\in L that is colored black and has two white neighbors. A black node with two white neighbors can not force any node, so we consider L=L{v}L^{\prime}=L\setminus\{v\} as a set of (1)(\ell-1) leaks. Since VV^{\prime} is (1)(\ell-1)-LFS, it should color all nodes black with leaks in LL^{\prime}, which is not the case. Hence a contradiction arises, proving the desired claim.  

Next, we show the equivalence between \ell-LFS and \ell-EFS.

Theorem 4.9

For a given \ell and G=(V,E)G=(V,E), VVV^{\prime}\subseteq V is an \ell-LFS if and only if VV^{\prime} is an \ell-EFS.

Proof:

(\ell-LFS \rightarrow\ell-EFS) We will prove using induction on \ell. From Lemma 4.5, if VV^{\prime} is 11-LFS, then it is 11-EFS. So, our induction hypothesis is, if VV^{\prime} is (1)(\ell-1)-LFS, then VV^{\prime} is (1)(\ell-1)-EFS. Now, assume that VV^{\prime} is \ell-LFS. Thus, VV^{\prime} must be (1)(\ell-1)-LFS implying that it is also (1)(\ell-1)-EFS (by our induction hypothesis). Let EE_{\ell} be a set of \ell non-forcing edges. By Lemma 4.7, there is a forcing process FF such that for each e=(u,v)Ee=(u,v)\in E_{\ell}, either both end nodes uu and vv are black, or one end node, say uu, is black with N[u]{v}N[u]\setminus\{v\} also colored black. Let V~\tilde{V} be the set of black nodes with the forcing process FF. Next, for each eEe\in E_{\ell}, we consider one of its black end node as a leak, and denote the set of leaks by LL. There will be at most \ell leaks. We observe that a black colored leak node can be ignored and deleted without altering the zero forcing behavior of the other black nodes. So, we consider G=G{L}G^{\prime}=G\setminus\{L\}. Since GG is \ell-LFS with VV^{\prime}, V~L\tilde{V}\setminus L is a ZFS of GG^{\prime}. As a result, all nodes in VV are colored black while considering EE_{\ell} as non-forcing edges, implying VV^{\prime} is an \ell-EFS.

(\ell-EFS \rightarrow\ell-LFS) Again, we will use induction on \ell. For =1\ell=1, if VV^{\prime} is 11-EFS, then it is 11-LFS (by Lemma 4.6). Our induction hypothesis is that VV^{\prime} is (1)(\ell-1)-EFS implies it to be (1)(\ell-1)-LFS. Now assume VV^{\prime} to be \ell-EFS. It means VV^{\prime} is (1)(\ell-1)-EFS and hence, (1)(\ell-1)-LFS (by the induction hypothesis). Consider LL to be a set of \ell leaks. By Lemma 4.8, these leaks are colored black and each of them has at most one white neighbor. Next, we consider the edge between the leak and its white neighbor as a non-forcing edge. There will be at most \ell such non-forcing edges, which we denote by EE_{\ell}. Since one end node (leak) of each non-forcing edge is colored black and a leak has at most one white neighbor (which is the other end node of the non-forcing edge), we can safely delete the non-forcing edge without affecting the zero forcing behavior of the black node. Thus, we get G=GEG^{\prime}=G\setminus E_{\ell}. Since VV^{\prime} is \ell-EFS, it means VV^{\prime} is a ZFS of GG^{\prime}. Thus, all nodes will be colored black despite \ell leaks. Hence, VV^{\prime} is \ell-LFS, which proves the desired claim.  

Lemma 4.10

If VV^{\prime} is 11-FSR then it is 11-EFS.

Proof:

By contraposition, let VV^{\prime} be a ZFS that is not a 11-EFS. It means there is an edge, say (u,v)(u,v), such that every forcing process must use the edge, i.e., (u,v)(u,v) must be a forcing edge in any forcing process. Without the loss of generality, we assume uu forces vv. It implies that all the black neighbors of vv have at least two white neighbors and uu has only one white neighbor vv. Thus, by removing edge (u,v)(u,v), vv can not be forced. Hence, VV^{\prime} is not a 11-FSR, which is the desired claim.  

In the following, we show the equivalence of \ell-EFS and \ell-FSR.

Theorem 4.11

VV^{\prime} is \ell-EFS if and only if VV^{\prime} is \ell-FSR.

Proof:

(\ell-EFS \rightarrow \ell-FSR) We first note that if FF is a ZFP with VV^{\prime} as a ZFS, then there are at most n1n-1 edges used in FF. If we remove edges not used in FF to get GG^{\prime}, then VV^{\prime} will still be a ZFS of GG^{\prime}. Thus, if VV^{\prime} is an \ell-EFS, it means for any edge set EEE_{\ell}\subseteq E, where |E||E_{\ell}|\leq\ell, there exist a forcing process with VV^{\prime}, say FF_{\ell}, coloring all nodes black without using edges in EE_{\ell}. Since edges in EE_{\ell} are not used in FF_{\ell}, we can remove them from GG while maintaining VV^{\prime} to be a ZFS of G=(V,EE)G^{\prime}=(V,E\setminus E_{\ell}), i.e., FF_{\ell} is a forcing process of GG^{\prime} with VV^{\prime}, implying VV^{\prime} is an \ell-FSR of GG.

(\ell-FSR \rightarrow \ell-EFS) We will prove using induction on \ell. For =1\ell=1, if VV^{\prime} is 11-FSR, then it is 11-EFS by Lemma 4.10. Thus, we make the induction hypothesis, if VV^{\prime} is (1)(\ell-1)-FSR, then it is (1)(\ell-1)-EFS. Assuming VV^{\prime} to be \ell-FSR, we need to show that VV^{\prime} is \ell-EFS. Let EE_{\ell} be a set of \ell non-forcing edges. Since VV^{\prime} is (1)(\ell-1)-EFS (by the induction hypothesis), we apply forces such that each edge e=(u,v)Ee=(u,v)\in E_{\ell} satisfies one of the two conditions (by Lemma 4.7): (i) Both end nodes of ee are colored black. (ii) If one node, say uu, is black, then all nodes in N[u]{v}N[u]\setminus\{v\} are also colored black. Note that the removal of edges in both cases (i) and (ii) will not change the zero forcing behavior of the black nodes. Thus, we remove these \ell edges. Since VV^{\prime} is \ell-FSR, it means that all nodes will be colored black at the end of the forcing process. Thus, all nodes are colored black in spite of \ell non-forcing edges, i.e., VV^{\prime} is \ell-EFS, which is the desired result.  

A direct corollary of Theorems 4.9 and 4.11 is Theorem 4.4 entailing that resilience to one type of misbehaving nodes/edges implies resilience to other kinds of misbehaving nodes/edges.

V Computation and Numerical Illustration

Computing a minimum \ell-LFS, and therefore, \ell-EFS and \ell-FSR, are NP hard problems (since minimum ZFS is NP-hard). Here, first, we illustrate the characterization of 11-LFS provided in Theorem 4.2. Then, we present a greedy algorithm to compute a small-sized 11-LFS and numerically evaluate it. A greedy algorithm for any \ell can be designed using a similar approach and utilizing the characterization in Theorem 4.3.

For a given input set VV^{\prime}, we define 𝒬(V)\mathcal{Q}(V^{\prime}) to be the set of non-input nodes, each of which can be forced by at least two distinct nodes. More precisely,

𝒬(V)={vVV:xv(V) andyv(V), and xy}.\begin{split}\mathcal{Q}(V^{\prime})=\{v\in V\setminus V^{\prime}:\;\;&\exists\;x\rightarrow v\in\mathcal{F}(V^{\prime})\text{ and}\\ &y\rightarrow v\in\mathcal{F}(V^{\prime}),\text{ and }x\neq y\}.\end{split} (8)

By Theorem 4.2, VV^{\prime} is 11-LFS if and only if 𝒬(V)=VV\mathcal{Q}(V^{\prime})=V\setminus V^{\prime}. Consider the graph in Figure 6, where V={v1,v2,v4,v6}V^{\prime}=\{v_{1},v_{2},v_{4},v_{6}\} is a 11-LFS. To verify this, we need to show that for each vVVv\in V\setminus V^{\prime}, there exist zero forcing processes such that vv is forced by at least two distinct nodes. Table I outlines multiple zero forcing processes. We observe that for each vVVv\in V\setminus V^{\prime}, there exist zero forcing processes where vv is forced by two distinct forcers. For instance, v3v_{3} is forced by v2v_{2} and v8v_{8} in ZFP 1 and ZFP 2, respectively. Similarly, v5v_{5} is forced by v9v_{9} and v3v_{3} in ZFP 1 and ZFP 3, respectively.

Refer to caption
Figure 6: V={v1,v2,v4,v6}V^{\prime}=\{v_{1},v_{2},v_{4},v_{6}\} is a 11-LFS of GG.
ZFP 1 ZFP 2 ZFP 3 ZFP 4 ZFP 5 ZFP 6
v1v8v2v3v4v7v6v9v5\begin{array}[]{c}v_{1}\rightarrow v_{8}\\ v_{2}\rightarrow v_{3}\\ v_{4}\rightarrow v_{7}\\ v_{6}\rightarrow v_{9}\rightarrow v_{5}\end{array} v1v8v3v2v4v7v6v9v5\begin{array}[]{c}v_{1}\rightarrow v_{8}\rightarrow v_{3}\\ v_{2}\\ v_{4}\rightarrow v_{7}\\ v_{6}\rightarrow v_{9}\rightarrow v_{5}\end{array} v1v8v2v3v5v4v7v6v9\begin{array}[]{c}v_{1}\rightarrow v_{8}\\ v_{2}\rightarrow v_{3}\rightarrow v_{5}\\ v_{4}\rightarrow v_{7}\\ v_{6}\rightarrow v_{9}\end{array} v1v8v2v3v5v4v6v9v7\begin{array}[]{c}v_{1}\rightarrow v_{8}\\ v_{2}\rightarrow v_{3}\rightarrow v_{5}\\ v_{4}\\ v_{6}\rightarrow v_{9}\rightarrow v_{7}\end{array} v1v2v3v8v4v7v6v9v5\begin{array}[]{c}v_{1}\\ v_{2}\rightarrow v_{3}\rightarrow v_{8}\\ v_{4}\rightarrow v_{7}\\ v_{6}\rightarrow v_{9}\rightarrow v_{5}\end{array} v1v8v3v5v2v9v4v7v6\begin{array}[]{c}v_{1}\rightarrow v_{8}\rightarrow v_{3}\rightarrow v_{5}\\ v_{2}\rightarrow v_{9}\\ v_{4}\rightarrow v_{7}\\ v_{6}\end{array}
TABLE I: Multiple zero forcing processes (ZFP) with a given V={v1,v2,v4,v6}V^{\prime}=\{v_{1},v_{2},v_{4},v_{6}\} for the graph in Figure 6. For each vVVv\in V\setminus V^{\prime}, there are at least two ZFP such that vv is forced by distinct nodes in such ZFP.

Algorithm 1 presents a greedy heuristic to compute a 11-LFS. The main idea is to iteratively include nodes in the leader set VV^{\prime} to maximize the size of 𝒬(V)\mathcal{Q}(V^{\prime}) (as in (8)) until 𝒬(V)=VV\mathcal{Q}(V^{\prime})=V\setminus V^{\prime}. Since every 11-LFS is a ZFS, we initialize VV^{\prime} in Algorithm 1 with a ZFS. As a result of this greedy selection, VV^{\prime} might contain some redundant nodes. Lines 9149-14 in Algorithm 1 remove such redundant nodes to reduce the size of 11-LFS.

 
Algorithm 1: Greedy Heuristics for 1-LFS
 
 1:\;1:\; given: G=(V,E)G=(V,E), |V|=n|V|=n
 2:\;2:\; initialization: V=V^{\prime}=\emptyset
 3:\;3:\; Compute ZFS Z0Z_{\textbf{0}}, and assign V=Z0V^{\prime}=Z_{\textbf{0}}
 4:\;4:\; Compute 𝒬(V)\mathcal{Q}(V^{\prime})
 5:\;5:\; while |𝒬(V)|<n|Z1||\mathcal{Q}(V^{\prime})|<n-|Z_{\textbf{1}}|
 6:\;6:      v=argmaxvV(V𝒬(V))𝒬(V{v})v^{\ast}=\operatorname*{arg\,max}\limits_{v\in V\setminus(V^{\prime}\cup\mathcal{Q}(V^{\prime}))}{\mathcal{Q}(V^{\prime}\cup\{v\})}
 7:\;7:      V=V{v}V^{\prime}=V^{\prime}\cup\{v^{\ast}\}
 8:\;8:\; end while
       -------- removing redundancies --------
 9:\;9:Z=VZ=V^{\prime}
 10:\;10:\; for all vZv\in Z
 11:\;11:      if |𝒬(V{v})|=n|V||\mathcal{Q}(V^{\prime}\setminus\{v\})|=n-|V^{\prime}| - 11
 12:\;12:            V=V{v}V^{\prime}=V^{\prime}\setminus\{v\}
 13:\;13:      end if
 14:\;14:\; end for
 15:\;15:\; return VV^{\prime}
 

Figure 7 compares the greedy heuristic with the optimal solution for Erdös-Rényi (ER) and Barabási-Albert (BA) random graphs with n=20n=20 nodes. In ER graphs, any two nodes are adjacent with a probability pp. BA graphs are obtained by attaching a new node (one at a time) to an existing graph through mm edges using a preferential attachment model. The optimal solution is computed using an exhaustive search. Also, each point on the plots averages 1515 randomly generated instances. Figures 7(a) and (c) plot the size of 11-LFS as a function of mm in BA graphs, and as a function of pp in ER graphs, respectively. Figures 7(b) and (d) plot the time taken by the greedy and optimal solutions to compute 11-LFS in BA and ER graphs, respectively. We observe that greedy and optimal solutions are very close, and the difference between the two solutions reduces as the graphs become dense. However, the difference in time to compute greedy and optimal solutions is orders of magnitude, even in small-sized graphs. We can design a similar greedy heuristic to compute an \ell-LFS for >1\ell>1; however, the time complexity will increase significantly with increasing \ell, and even the greedy heuristic will become inefficient for higher \ell. Thus, more efficient heuristics are needed to compute \ell-LFS for large \ell values.

223344556677889910101144771010131316161919mmNo. of nodes in 11-LFSoptimalgreedy
(a) BA
2233445566778899101004004008008001,2001{,}2001,6001{,}6002,0002{,}000time (sec.)No. of nodes in 11-LFSoptimalgreedy
(b) BA
0.20.20.30.30.40.40.50.50.60.60.70.71144771010131316161919ppNo. of nodes in 11-LFSoptimalgreedy
(c) ER
0.20.20.30.30.40.40.50.50.60.60.70.704004008008001,2001{,}2001,6001{,}6002,0002{,}000time (sec.)No. of nodes in 11-LFSoptimalgreedy
(d) ER
Figure 7: Comparison of optimal and greedy heuristic (Algorithm 1) for the computation of 11-LFS in Erdös-Rényi (ER) and Barabási-Albert (BA) random graphs.

VI Conclusion

We studied the problem of maintaining SSC in networks despite misbehaving nodes and edges. We considered different models of misbehaving nodes/edges aiming to disrupt the zero forcing process in graphs, consequently deteriorating the network SSC. We showed that resilience to one type of misbehaving nodes/edges is possible if and only if resilience to other threat types is ensured. In other words, for a given leader set, if a network is strong structurally controllable despite kk misbehaving nodes/edges under one threat model, then the network remains to be strong structurally controllable in the presence of the same number of misbehaving nodes/edges under the other threat models. We also discussed the leader selection guaranteeing resilient SSC. Resilience is expensive in the context of SSC because many extra leaders are needed to nullify the effect of a small number of misbehaving nodes and edges.

In the future, we aim to address this problem by safeguarding a subset of nodes and edges, i.e., making them less susceptible to faults and adversarial attacks, and then analyzing how the number of leaders for resilient SSC is impacted in the presence of such ‘trusted’ nodes and edges. Since it would be expensive or infeasible to make every node/edge trusted, we would need to select them in the network carefully. Another interesting direction in this domain requiring more exploration is designing efficient heuristics to compute leaders for resilient SSC.

References

  • [1] F. Pasqualetti, S. Zampieri, and F. Bullo, “Controllability metrics, limitations and algorithms for complex networks,” IEEE Transactions on Control of Network Systems, vol. 1, no. 1, pp. 40–52, 2014.
  • [2] A. Clark, B. Alomair, L. Bushnell, and R. Poovendran, “Submodularity in input node selection for networked linear systems: Efficient algorithms for performance and controllability,” IEEE Control Systems Magazine, vol. 37, no. 6, pp. 52–74, 2017.
  • [3] T. H. Summers, F. L. Cortesi, and J. Lygeros, “On submodularity and controllability in complex dynamical networks,” IEEE Transactions on Control of Network Systems, vol. 3, no. 1, pp. 91–101, 2015.
  • [4] J. Li, X. Chen, S. Pequito, G. J. Pappas, and V. M. Preciado, “On the structural target controllability of undirected networks,” IEEE Transactions on Automatic Control, vol. 66, no. 10, pp. 4836–4843, 2020.
  • [5] J. Ruths and D. Ruths, “Control profiles of complex networks,” Science, vol. 343, no. 6177, pp. 1373–1376, 2014.
  • [6] N. Monshizadeh, S. Zhang, and M. K. Camlibel, “Zero forcing sets and controllability of dynamical systems defined on graphs,” IEEE Transactions on Automatic Control, vol. 59, pp. 2562–2567, 2014.
  • [7] A. Chapman and M. Mesbahi, “On strong structural controllability of networked systems: A constrained matching approach.” in American Control Conference (ACC), 2013, pp. 6126–6131.
  • [8] S. Zhang, M. Cao, and M. K. Camlibel, “Upper and lower bounds for controllable subspaces of networks of diffusively coupled agents,” IEEE Transactions on Automatic control, vol. 59, pp. 745–750, 2013.
  • [9] J. Jia, H. L. Trentelman, W. Baar, and M. K. Camlibel, “Strong structural controllability of systems on colored graphs,” IEEE Transactions on Automatic Control, vol. 65, no. 10, pp. 3977–3990, 2019.
  • [10] A. Yazıcıoğlu, W. Abbas, and M. Egerstedt, “Graph distances and controllability of networks,” IEEE Transactions on Automatic Control, vol. 61, no. 12, pp. 4125–4130, 2016.
  • [11] H. J. Van Waarde, M. K. Camlibel, and H. L. Trentelman, “A distance-based approach to strong target control of dynamical networks,” IEEE Transactions on Automatic Control, vol. 62, pp. 6266–6277, 2017.
  • [12] J. Jia, H. J. Van Waarde, H. L. Trentelman, and M. K. Camlibel, “A unifying framework for strong structural controllability,” IEEE Transactions on Automatic Control, vol. 66, no. 1, pp. 391–398, 2020.
  • [13] AIM Minimum Rank Special Graphs Work Group, “Zero forcing sets and the minimum rank of graphs,” Linear Algebra and its Applications, vol. 428, no. 7, pp. 1628–1648, 2008.
  • [14] S. S. Mousavi, M. Haeri, and M. Mesbahi, “On the structural and strong structural controllability of undirected networks,” IEEE Transactions on Automatic Control, vol. 63, no. 7, pp. 2234–2241, 2017.
  • [15] M. Trefois and J.-C. Delvenne, “Zero forcing number, constrained matchings and strong structural controllability,” Linear Algebra and its Applications, vol. 484, pp. 199–218, 2015.
  • [16] S. Dillman and F. Kenter, “Leaky forcing: a new variation of zero forcing,” arXiv preprint arXiv:1910.00168, 2019.
  • [17] J. S. Alameda, J. Kritschgau, N. Warnberg, and M. Young, “On leaky forcing and resilience,” Discrete Applied Mathematics, vol. 306, pp. 32–45, 2022.
  • [18] M. Pirani, A. Mitra, and S. Sundaram, “A survey of graph-theoretic approaches for analyzing the resilience of networked control systems,” arXiv preprint arXiv:2205.12498, 2022.
  • [19] S. S. Mousavi, M. Haeri, and M. Mesbahi, “Robust strong structural controllability of networks with respect to edge additions and deletions,” in American Control Conference (ACC), 2017, pp. 5007–5012.
  • [20] Y. Zhang and T. Zhou, “Minimal structural perturbations for controllability of a networked system: Complexities and approximations,” International Journal of Robust and Nonlinear Control, vol. 29, no. 12, pp. 4191–4208, 2019.
  • [21] W. Abbas, M. Shabbir, Y. Yazıcıoğlu, and X. Koutsoukos, “Edge augmentation with controllability constraints in directed laplacian networks,” IEEE Control Systems Letters, vol. 6, pp. 1106–1111, 2021.
  • [22] S. Pequito, G. Ramos, S. Kar, A. P. Aguiar, and J. Ramos, “The robust minimal controllability problem,” Automatica, vol. 82, pp. 261–268, 2017.
  • [23] C.-L. Pu, W.-J. Pei, and A. Michaelson, “Robustness analysis of network controllability,” Physica A: Statistical Mechanics and its Applications, vol. 391, no. 18, pp. 4420–4425, 2012.
  • [24] G. Ramos, D. Silvestre, and C. Silvestre, “The robust minimal controllability and observability problem,” International Journal of Robust and Nonlinear Control, vol. 31, no. 10, pp. 5033–5044, 2021.
  • [25] Y. Zhang, “Controllability robustness under actuator failures: Complexities and algorithms,” arXiv preprint arXiv:1812.07745, 2018.
  • [26] A. Aazami, “Hardness results and approximation algorithms for some problems on graphs,” Ph.D. dissertation, University of Waterloo, 2008.
  • [27] B. Brimkov, C. C. Fast, and I. V. Hicks, “Computational approaches for zero forcing and related problems,” European Journal of Operational Research, vol. 273, no. 3, pp. 889–903, 2019.
  • [28] B. Brimkov, D. Mikesell, and I. V. Hicks, “Improved computational approaches and heuristics for zero forcing,” INFORMS Journal on Computing, 2021.
  • [29] A. Agra, J. O. Cerdeira, and C. Requejo, “A computational comparison of compact MILP formulations for the zero forcing number,” Discrete Applied Mathematics, vol. 269, pp. 169–183, 2019.
  • [30] W. Abbas, M. Shabbir, A. Y. Yazicioğlu, and X. Koutsoukos, “Leader selection for strong structural controllability in networks using zero forcing sets,” in American Control Conference, 2022, pp. 1444–1449.
  • [31] F. Barioli, W. Barrett, S. M. Fallat, H. T. Hall, L. Hogben, B. Shader, P. Van Den Driessche, and H. Van Der Holst, “Zero forcing parameters and minimum rank problems,” Linear Algebra and its Applications, vol. 433, no. 2, pp. 401–411, 2010.