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

Synchronization in dynamical systems coupled via multiple directed networks

Chai Wah Wu
IBM T.J. Watson Research Center
Yorktown Heights, NY 10598, USA
[email protected]
March 24, 2021
Abstract

We study synchronization and consensus in a group of dynamical systems coupled via multiple directed networks. We show that even though the coupling in a single network may not be sufficient to synchronize the systems, combination of multiple networks can contribute to synchronization. We illustrate how the effectiveness of a collection of networks to synchronize the coupled systems depends on the graph topology. In particular, we show that if the graph sum is a directed graph whose reversal contains a spanning directed tree, then the network synchronizes if the coupling is strong enough. This is intuitive as there is a root node that influence every other node via a set of edges where each edge in the set is in one of the networks.

Index Terms:
nonlinear dynamical systems, graph theory

I Introduction

There has been much research activity in studying the synchronization and consensus behavior in networks of coupled dynamical systems [1, 2, 3, 4, 5, 6, 7, 8, 9]. The coupled systems model various types of networks both man-made and naturally occurring, such as social networks, transportation networks, genetic networks and neural networks. In many cases the individual systems or agents are connected via multiple networks. For instance, automobiles are connected via a transportation network, but they can also connected via cell phones (or other vehicle-to-vehicle technologies such as DSRC [10]) in a communication network. Similarly, people can be connected via multiple social networks such as Facebook and LinkedIn. The Erdős-Bacon number is the sum of the distance in the actor and in the mathematical collaboration network.

Prior work on systems coupled via multiple networks include [11] where synchronization is studied in continuous-time dynamical systems coupled via 22 networks with one of the networks coupling delayed state variables and conditions were provided under which the system is synchronizing. Refs. [12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23] also studied synchronization in continuous-time dynamical systems coupled via multiple networks where a local stability approach numerically computing Lyapunov exponents or a mean field approximation is used. In some of these works, networks are studied whose Laplacian matrices are simultaneously diagonalizable. This paper differs from prior work in that 1) we use the Lyapunov function approach which provides a guaranteed synchronization criteria, 2) consider simultaneously triangularizable matrices which is a superset of the simultaneously diagonalizable matrices, 3) consider directed networks 4) and consider extensions to discrete-time systems.

In this paper we study synchronization and consensus in dynamical systems coupled under multiple networks and study how these multiple networks can work together to reach synchronization even though each individual network might not be able to effect synchronization. We consider the case of continuous time systems in Section III. The case of discrete time systems will be studied in Section IV.

II Mathematical preliminaries

We define eie_{i} as the ii-th unit vector and EiE_{i} as the diagonal square matrix such that all entries are 0 except that the ii-th entry on the diagonal is 11 and define ee as the vector of all 11’s and II as the identity matrix. For a Hermitian matrix, the eigenvalues are listed as λ1λ2λn\lambda_{1}\leq\lambda_{2}\cdots\leq\lambda_{n}. We use A0A\succ 0 and A0A\succeq 0 to denote that the matrix AA is positive definite and positive semidefinite respectively111For non-Hermitian matrices, we say AA is positive (semi)definite if A+AHA+A^{H} is positive (semi)definite.. The following simple results generalize the results in [24] and in [25, Lemma 3.1] and prove to be useful when studying coupling via multiple networks:

Lemma II.1

Let ={Ai}{\cal M}=\{A_{i}\} be a set of nn by nn simultaneously triangularizable matrices, i.e. there exists a nonsingular matrix PP which transforms these matrices simultaneously into upper triangular form, i.e. Pi,PAiP1=Ji\exists P\forall i,PA_{i}P^{-1}=J_{i} is upper triangular with corresponding eigenvalues λij\lambda_{ij} on the diagonal of JiJ_{i}. Let {Bi}\{B_{i}\} be a set of mm by mm matrices. Then the nmnm by nmnm matrix C=iAiBiC=\sum_{i}A_{i}\otimes B_{i} has eigenvalues γjk\gamma_{jk} where γjk\gamma_{jk} are the eigenvalues of iλijBi\sum_{i}\lambda_{ij}B_{i} for each jj.

Proof: The matrix (PI)C(P1I)=i(JiBi)(P\otimes I)C(P^{-1}\otimes I)=\sum_{i}(J_{i}\otimes B_{i}) is block upper triangular with diagonal blocks equal to iλijBi\sum_{i}\lambda_{ij}B_{i}. Thus the eigenvalues are the same as the eigenvalues of iλijBi\sum_{i}\lambda_{ij}B_{i} for 1jn1\leq j\leq n.222Since all GiG_{i} share the same set of unit length eigenvectors vjv_{j}, we denote λij\lambda_{ij} as the eigenvalue of BiB_{i} corresponding to the eigenvector vjv_{j}. This convention will used throughout. \Box

It is clear that if {\cal M} satisfies the condition of Lemma II.1, so does {αI}{\cal M}\cup\{\alpha I\} where α\alpha\in{\mathbb{C}}.

Corollary II.1 ([24])

The eigenvalues of IB1+AB2I\otimes B_{1}+A\otimes B_{2} are equal to the eigenvalues of B1+λiB2B_{1}+\lambda_{i}B_{2} for each ii, where λi\lambda_{i} are the eigenvalues of AA.

Corollary II.2

Let pip_{i} be polynomials and AA be a square matrix with eigenvalues λj\lambda_{j}. Then C=ipi(A)BiC=\sum_{i}p_{i}(A)\otimes B_{i} has eigenvalues γjk\gamma_{jk} where γjk\gamma_{jk} are the eigenvalues of ipi(λj)Bi\sum_{i}p_{i}(\lambda_{j})B_{i} for each jj.

Proof: The matrix AA is similar to an upper triangular matrix JJ (e.g. the Jordan normal form). If JJ is in upper triangular form, so is J2J^{2} and it is clear from the spectral mapping theorem that if JJ has eigenvalues λi\lambda_{i}, then p(J)p(J) has eigenvalues p(λi)p(\lambda_{i}). This implies that the set of matrices {p1(A),p2(A),,pn(A)}\{p_{1}(A),p_{2}(A),\cdots,p_{n}(A)\} where AA is a fixed matrix and pip_{i} are polynomials can be transformed to an upper triangular form via the same matrix PP and the result follows from Lemma II.1. \Box

It is well known that the set of simultaneously triangularizable matrices corresponds to the set of commuting matrices [26]. Examples of commuting sets of matrices include the set of nn by nn circulant matrices. These matrices are diagonalizable with eigenvectors equal to the columns of the Discrete Fourier Transform matrix. As noted before, the identity matrix II (and its scalar multiple) can always be added to such sets since the identity matrix commutes with any matrix.

In our study, AiA_{i} will be Laplacian matrices of directed graphs333A Laplacian matrix of a directed graph is defined as DAD-A where DD is the diagonal matrix of vertex outdegrees and AA is the adjacency matrix. and Aie=0A_{i}e=0 for all ii. This means that for all vector vv, eve\otimes v is an eigenvector of AiBiA_{i}\otimes B_{i} corresponding to eigenvalue 0, i.e. the eigenvalue 0 has multiplicity at least mm.444Recall that mm is the order of the matrices BiB_{i}. If in addition the graph 𝒢i{\mathcal{G}}_{i} corresponding to AiA_{i} is connected, then 0 is an eigenvalue of AiA_{i} with multiplicity 11. If BiB_{i} is nonsingular, then AiBiA_{i}\otimes B_{i} has a 0 eigenvalue of multiplicity mm.

We ask: when does iAiBi\sum_{i}A_{i}\otimes B_{i} have a zero eigenvalue of multiplicity mm? The answer to this question will be useful later on in our study of synchronization in network of systems coupled via multiple networks. The following results give some conditions under which this question can be answered.

Theorem II.1

If BB is nonsingular, and iAi\sum_{i}A_{i} has a zero eigenvalue of multiplicity 1, then iAiB\sum_{i}A_{i}\otimes B has a zero eigenvalue of multiplicity mm.

Proof: The eigenvalues of iAiB\sum_{i}A_{i}\otimes B are the eigenvalues of λiB\lambda_{i}B where λi\lambda_{i} are the eigenvalues of AA. \Box

Definition II.1

The graph sum of (weighted) graphs G1,G2,GrG_{1},G_{2}\cdots,G_{r} with the same vertex set is a graph with adjacency matrix equal to the sum of the adjacency matrices of GiG_{i}.

For example, Fig. 1 shows two directed graphs with the same vertex set and edges labeled ‘A’ and ‘B’ resp. Under the canonical ordering of the nodes, the Laplacian matrices of these graphs are circulant and thus commuting. Even though each of the graph is not connected, the graph sum is a connected graph and between every pair of nodes there is a directed path between them using either ‘A’ or ‘B’ edges.

Refer to caption

Figure 1: Two directed graphs with the same vertex set superimposed. The edges of the two graphs are labeled ‘A’ and ‘B’ respectively.

Note that the graph sum depends on the ordering of the vertices and is thus not invariant under graph isomorphism, i.e. we consider labeled graphs. If AiA_{i} are the Laplacian matrices of graphs, then iAi\sum_{i}A_{i} is the Laplacian matrix of the graph sum of these graphs and iAi\sum_{i}A_{i} having a zero eigenvalue of multiplicity 1 is equivalent to the reversal555i.e. the directed graph obtained by reversing the orientation of all edges. of the graph sum containing a spanning directed tree [27]. Note that for directed graphs, the Laplacian matrix is not necessarily symmetric.

Lemma II.2

Let LiL_{i} be the Laplacian matrices of graphs 𝒢i{\mathcal{G}}_{i} with the same vertex set such that the graph sum is a directed graph whose reversal contains a spanning directed tree. If LiL_{i} are simultaneously triangularizable with eigenvalues λij\lambda_{ij}, then for exactly one fixed jj is it true that λij=0\lambda_{ij}=0 for all ii. In fact, the index jj corresponds to the zero eigenvector ee.

Proof: The Laplacian matrix of the graph sum is iLi\sum_{i}L_{i}. By the simultaneously triangularizable property, its eigenvalues are iλij\sum_{i}\lambda_{ij} for each jj. By [27], its Laplacian matrix has a 0 eigenvalue of multiplicity 11 and the conclusion follows. \Box

Theorem II.2

Let {Li}\{L_{i}\} be a set of nn by nn simultaneously triangularizable Laplacian matrices, with corresponding eigenvalues λij\lambda_{ij} that are all nonnegative. Let {Bi}\{B_{i}\} be a set of mm by mm nonsingular matrices such that any convex combination of BiB_{i} is nonsingular. If the graph sum of the graph corresponding to {Li}\{L_{i}\} is a directed graph whose reversal contains a spanning directed tree, then the nmnm by nmnm matrix C=iLiBiC=\sum_{i}L_{i}\otimes B_{i} has a zero eigenvalue of multiplicity mm.

Proof: By Lemma II.1, the eigenvalues of CC are given by the eigenvalues iλijBi\sum_{i}\lambda_{ij}B_{i} for each jj. By Lemma II.2 only for one jj is it true that λij=0\lambda_{ij}=0 for all ii, in which case iλijBi=0\sum_{i}\lambda_{ij}B_{i}=0 and we have mm zero eigenvalues. For all other jj there exists nonzero λij\lambda_{ij}’s. Since λij0\lambda_{ij}\geq 0, this can be rescaled to a convex combination of BiB_{i} which is nonsingular by hypothesis. \Box

Examples of nonsingular matrices BiB_{i} for which convex combinations are nonsingular include diagonal positive definite matrices or strictly diagonally dominant matrices whose eigenvalues have positive real parts.

III Continuous-time systems

For the continuous-time case, consider the state equation:

x˙=(f(x1,t)f(xn,t))k=1r(Gk(t)Dk(t))x\dot{x}=\left(\begin{array}[]{c}f(x_{1},t)\\ \vdots\\ f(x_{n},t)\end{array}\right)-\sum_{k=1}^{r}(G_{k}(t)\otimes D_{k}(t))x (1)

where x=(x1,,xn)x=(x_{1},\dots,x_{n}) and Gk(t)G_{k}(t) and Dk(t)D_{k}(t) are square matrices for each kk and tt. There are nn identical systems described by xi˙=f(xi,t)\dot{x_{i}}=f(x_{i},t) and each xix_{i} is the mm-dimensional state vector of the ii-th system. The coupling matrices Gk(t)G_{k}(t) have zero row sums and describe rr different coupling networks coupling these systems together, i.e. the ijij-th entry of the matrix Gk(t)G_{k}(t) is nonzero if there is a coupling from the jj-th system to the ii-th system via the kk-th network. The matrix Dk(t)D_{k}(t) describes the coupling between a pair of systems in the kk-th network and is the same between any pair of systems in the kk-th network. We are interested in the case where the DkD_{k} are not all equal, as otherwise it reverts to the single network case with the network being the graph sum. We say the system is globally synchronizing if xixj0\|x_{i}-x_{j}\|\rightarrow 0 as tt\rightarrow\infty for all ii,jj and all initial conditions. One important special case is when Dk=EkD_{k}=E_{k}, which corresponds to the case where the kk-th network only couples the kk-th component of the state vectors and the internetwork connections occur solely due to the dynamics of the individual systems.

Definition III.1

𝒲\cal W is the class of symmetric matrices that have zero row sums and nonpositive off-diagonal elements.

The following result provides sufficient conditions under which the system in Eq. (1) is globally synchronizing. and follows directly from the synchronization theorems in [28, 29, 30].

Theorem III.1

Let P(t)P(t) be a time-varying matrix and V0V\succ 0 be a symmetric matrix such that (yz)TV(f(y,t)P(t)yf(z,t)+P(t)z)cyz2(y-z)^{T}V(f(y,t)-P(t)y-f(z,t)+P(t)z)\leq-c\|y-z\|^{2} for some c>0c>0. Then the network in Eq. (1) is globally synchronizing if there exists an irreducible matrix U𝒲U\in\cal W such that for all tt

(UV)(k=1rGk(t)Dk(t)IP(t))0(U\otimes V)\left(\sum_{k=1}^{r}G_{k}(t)\otimes D_{k}(t)-I\otimes P(t)\right)\succeq 0 (2)

We next look at various special cases of Theorem III.1.

III-A The case P=ξIP=\xi I

Note that the matrix P(t)P(t) can be thought of as a stabilizing (time-varying) negative linear state feedback such that x˙=f(x,t)P(t)x\dot{x}=f(x,t)-P(t)x is asymptotically stable. Using Theorems II.2, III.1 and choosing P=ξIP=\xi I, we get:

Corollary III.1

Let {Li}\{L_{i}\} be a set of commuting normal Laplacian matrices where the graph sum’s reversal contains a spanning directed tree. If DkD_{k} are normal, ff is Lipschitz continuous with Lipschitz constant cc, Gk=ξLkG_{k}=\xi L_{k} and matrices in conv(DkD_{k}) have eigenvalues with positive real part, then Eq. (1) is synchronizing if ξcRe(λL)Re(λD)\xi\geq\frac{c}{Re(\lambda_{L})Re(\lambda_{D})} where conv(DkD_{k}) is the convex hull of DkD_{k}, λL\lambda_{L} is the nonzero eigenvalue of LiL_{i} with minimal real part and λD\lambda_{D} is the eigenvalue of matrices in conv(DkD_{k}) with minimal real part.

This result can be shown by noting that eve\otimes v are exactly the eigenvectors corresponding to zero eigenvalues of both UVU\otimes V and k=1rGk(t)Dk(t)\sum_{k=1}^{r}G_{k}(t)\otimes D_{k}(t). Note that by hypothesis Re(λD)>0Re(\lambda_{D})>0 as the convex hull is compact. Corollary III.1 is intuitive in the sense that if the graph sum’s reversal contains a spanning directed tree, there is at least one node that influences every other node via edges where each edge is in at least one of the networks. Even though each of the network might not be connected by themselves, synchronization can occur via the graph sum if the coupling is strong enough.

The converse is true as well in the following sense. If GkG_{k} and DkD_{k} do not depend on tt and the graph sum does not contain a spanning directed tree, then there exists two nodes in the graph sum that do not influence each other and for general dynamical systems, especially chaotic systems, it is not possible for the systems at these 2 nodes to synchronize to each other.

III-B x˙=f(x,t)ξkDk(t)\dot{x}=f(x,t)-\xi\sum_{k}D_{k}(t) is asymptotically stable

Consider the case where x˙=f(x,t)ξkDk(t)\dot{x}=f(x,t)-\xi\sum_{k}D_{k}(t) is asymptotically stable for some ξ>0\xi>0. In this case, we can choose P(t)=ξkDk(t)P(t)=\xi\sum_{k}D_{k}(t). Some examples are when P(t)P(t) is a diagonal matrix666In several prior studies PP is chosen to be the identity matrix. for all tt and Di(t)=ci(t)EiD_{i}(t)=c_{i}(t)E_{i}. In this case Eq. (2) simplifies to (UV)(k=1r(Gk(t)ξI)Dk(t))0(U\otimes V)\left(\sum_{k=1}^{r}(G_{k}(t)-\xi I)\otimes D_{k}(t)\right)\succeq 0. If in addition VDk0VD_{k}\succeq 0, such as when VV is a diagonal matrix and positive definite and DkD_{k} are diagonal and positive semidefinite matrices, then Eq. (2) is satisfied if U(Gk(t)ξI)0U(G_{k}(t)-\xi I)\succeq 0 for all k=1,,rk=1,\cdots,r.

Thus we have shown the following:

Theorem III.2

Let V0V\succ 0 be some symmetric matrix such that (yz)TV(f(y,t)ξkDkyf(z,t)+ξkDkz)cyz2(y-z)^{T}V(f(y,t)-\xi\sum_{k}D_{k}y-f(z,t)+\xi\sum_{k}D_{k}z)\leq-c\|y-z\|^{2} for some c,ξ>0c,\xi>0. Suppose that VDk0VD_{k}\succeq 0 for all kk. Then the network in Eq. (1) is globally synchronizing if there exists an irreducible matrix U𝒲U\in\cal W such that U(Gk(t)ξI)0U(G_{k}(t)-\xi I)\succeq 0 for all tt and all k=1,,rk=1,\cdots,r.

For U=LKU=L_{K}, the Laplacian matrix of the complete graph, the proof of Theorem 3 in [29] can be used to show the following:

Theorem III.3

Let V0V\succ 0 be some symmetric matrix such that (yz)TV(f(y,t)ξkDkyf(z,t)+ξkDkz)cyz2(y-z)^{T}V(f(y,t)-\xi\sum_{k}D_{k}y-f(z,t)+\xi\sum_{k}D_{k}z)\leq-c\|y-z\|^{2} for some c,ξ>0c,\xi>0. Suppose that VDk0VD_{k}\succeq 0 and Gk(t)G_{k}(t) has zero row sums and zero column sums such that Gk(t)+GkT(t)G_{k}(t)+G_{k}^{T}(t) has a simple zero eigenvalue. Then the network in Eq. (1) is globally synchronizing if λ2(12(Gk(t)+GkT(t)))ξ\lambda_{2}\left(\frac{1}{2}\left(G_{k}(t)+G^{T}_{k}(t)\right)\right)\geq\xi for all tt and all k=1,,rk=1,\cdots,r.

Proof: U=LK=nIJU=L_{K}=nI-J, where JJ is the nn by nn matrix of all 11’s. Then U(Gk(t)I)=nGk(t)LKU(G_{k}(t)-I)=nG_{k}(t)-L_{K}. Now nGk(t)LK0nG_{k}(t)-L_{K}\geq 0 if all eigenvalues of 12n(Gk(t)+GkT(t))LK\frac{1}{2}n\left(G_{k}(t)+G^{T}_{k}(t)\right)-L_{K} are nonnegative. For the vector ee, (12n(Gk(t)+GkT(t))LK)e=0\left(\frac{1}{2}n\left(G_{k}(t)+G^{T}_{k}(t)\right)-L_{K}\right)e=0. For a vector xx orthogonal to ee, LKx=nxL_{K}x=nx and xT(12(Gk(t)+GkT(t)))xxTxx^{T}\left(\frac{1}{2}\left(G_{k}(t)+G^{T}_{k}(t)\right)\right)x\geq x^{T}x and thus the proof is complete. \Box

The interpretation here is that even though the coupling matrix DkD_{k} corresponding to an individual network GkG_{k} is not sufficient to drive ff to stability, the combination of all DkD_{k} is sufficient stabilizing feedback. The consequence is that even though a single coupled network GkG_{k} with coupling matrix DkD_{k} is not sufficient to achieve synchronization, the combination of multiple networks is.

Next, define Ξ\Xi as the set of real numbers ξ\xi such that there exists an irreducible matrix U𝒲U\in{\cal W} satisfying U(Gk(t)ξI)0U(G_{k}(t)-\xi I)\succeq 0 for all kk and tt. Let ξM\xi_{M} be the supremum of Ξ\Xi.

Corollary III.2

Let V0V\succ 0 be some symmetric matrix such that (yz)TV(f(y,t)ξMkDkyf(z,t)+ξMkDkz)cyz2(y-z)^{T}V(f(y,t)-\xi_{M}\sum_{k}D_{k}y-f(z,t)+\xi_{M}\sum_{k}D_{k}z)\leq-c\|y-z\|^{2} for some c>0c>0 and that VDk0VD_{k}\succeq 0 for all kk. Then Eq. (1) is globally synchronizing.

The quantity ξM\xi_{M} allows us to compare different sets of graphs. If one set of graphs has a higher ξM\xi_{M} than another set, then this suggests that the first set of graphs can synchronize the system with less coupling (as expressed by DkD_{k}).

For time-invariant GkG_{k} (or Gk(t)G_{k}(t) belongs to a finite fixed set of matrices for all tt), we can use a bisection approach to compute ξM\xi_{M} [31]. Let QQ be an nn by n1n-1 matrix whose columns form an orthonormal basis of ee^{\perp}. Consider for a fixed ξ\xi the following feasibility semidefinite programming (SDP) problem:

Find U=UTU=U^{T} such that k\forall k U(GkξI)0U(G_{k}-\xi I)\succeq 0, Ue=0Ue=0 ij\forall i\neq j Uij0U_{ij}\leq 0 and QTUQIQ^{T}UQ\succeq I.

Using a bisection search to repeatedly solve the SDP problem above allows us to compute ξM\xi_{M}.

Data: GkG_{k}, ϵ\epsilon, lower and upper bound lblb, ubub
Result: ξM\xi_{M}
ξ12(lb+ub)\xi\leftarrow\frac{1}{2}(lb+ub);
while |ublb|>ϵ|ub-lb|>\epsilon do
       if SDP problem is infeasible for ξ\xi then
             ubξub\leftarrow\xi;
            
      else
             lbξlb\leftarrow\xi;
            
       end if
      ξ12(lb+ub)\xi\leftarrow\frac{1}{2}(lb+ub);
      
end while
ξMξ\xi_{M}\leftarrow\xi;
Algorithm 1 Compute ξM\xi_{M}

For example, consider the following set of Laplacian matrices GkG_{k}

G1=(1010002011103111103100011)G_{1}=\begin{pmatrix}1&0&-1&0&0\\ 0&2&0&-1&-1\\ -1&0&3&-1&-1\\ -1&-1&0&3&-1\\ 0&0&0&-1&1\end{pmatrix}
G2=(2100101010002111001010113)G_{2}=\begin{pmatrix}2&-1&0&0&-1\\ 0&1&0&-1&0\\ 0&0&2&-1&-1\\ -1&0&0&1&0\\ -1&0&-1&-1&3\end{pmatrix}
G3=(1100012100011000011011103)G_{3}=\begin{pmatrix}1&-1&0&0&0\\ -1&2&-1&0&0\\ 0&-1&1&0&0\\ 0&0&-1&1&0\\ -1&-1&-1&0&3\end{pmatrix}

corresponding to the directed graphs in Fig. 2.

Refer to captionRefer to captionRefer to caption

Figure 2: Directed graphs with Laplacian matrices G1G_{1}, G2G_{2}, and G3G_{3}.

Running Algorithm 1 using the software package CVX [32, 33] to solve the SDP problem, we find that ξM=0.838\xi_{M}=0.838. For real zero sums matrices GkG_{k}, the value of ξM\xi_{M} is finite and an upper and lower bound can be explicitly computed (which are needed to initialize the bisection search in Algorithm 1).

Theorem III.4

If all GkG_{k} are all real zero row sums matrices, then ξM\xi_{M} exists and is lower bounded by ξMminkminx0,xexTGkxxTx\xi_{M}\geq\min_{k}\min_{x\neq 0,x\perp e}\frac{x^{T}G_{k}x}{x^{T}x}.

Proof: The proof of Theorem 2 in [31] shows that for UU equal to the Laplacian matrix of the complete graph, U(GμI)0U(G-\mu I)\succeq 0, where μ=minx0,xexTGx\mu=\min_{x\neq 0,x\perp e}x^{T}Gx and the conclusion follows. \Box

Theorem III.5

If all GkG_{k} are real matrices with nonpositive off-diagonal elements and zero row and column sums, then ξM0\xi_{M}\geq 0. If in addition Gk+GkTG_{k}+G_{k}^{T} are all irreducible, then ξM>0\xi_{M}>0.

Proof: Follows from Corollary 2 in [31] and Theorem III.4. \Box

Note that Theorem III.5 implies that if the directed graphs whose Laplacian matrices are GkG_{k} are balanced777A directed graph is balanced if the indegree of each vertex is equal to its outdegree., then ξM0\xi_{M}\geq 0. If they are balanced and strongly connected888Balanced graphs are weakly connected if and only if they are strongly connected., then ξM>0\xi_{M}>0.

Definition III.2

For a real matrix GG with zero row sums, define μ2(G)\mu_{2}(G) as the minimum of Re(λ)Re(\lambda) among all eigenvalues λ\lambda not corresponding to the eigenvector ee.

Theorem III.6

If GkG_{k} are all real matrices with zero row sums, then ξMminkμ2(Gk)\xi_{M}\leq\min_{k}\mu_{2}(G_{k}).

Proof: Similar to the proof of Theorem 3 in [34], let λk\lambda_{k} be an eigenvalue of GkG_{k} with eigenvector vkev_{k}\neq e. Let U𝒲U\in{\cal W} be such that U(GkμI)0U(G_{k}-\mu I)\succeq 0 for all kk for some real number μ\mu. Since the kernel of UU is spanned by ee, vkv_{k} is not in the kernel of UU. Since (GkμI)vk=(λKμ)vk(G_{k}-\mu I)v_{k}=\left(\lambda_{K}-\mu\right)v_{k}, this implies that vkU(GkμI)v=(λKμ)vkUvkv_{k}^{*}U(G_{k}-\mu I)v=\left(\lambda_{K}-\mu\right)v_{k}^{*}Uv_{k}. Since U(GkμI)0U(G_{k}-\mu I)\succeq 0 this means that Re(vkU(GkμI)vk)0Re(v_{k}^{*}U(G_{k}-\mu I)v_{k})\geq 0. Since UU is symmetric positive semidefinite and vkv_{k} is not in the kernel of UU, vkUv>0v_{k}^{*}Uv>0. This means that Re(λk)μ0Re(\lambda_{k})-\mu\geq 0 for all kk. \Box

It is clear that ξM({G1,Gr})mini{ξM(G1),,ξM(Gr)}\xi_{M}(\{G_{1},\cdots G_{r}\})\leq\min_{i}\{\xi_{M}(G_{1}),\cdots,\xi_{M}(G_{r})\}. For the 3 graphs in the above example, 0.838=ξM({G1,Gr})<mini{ξM(G1),,ξM(Gr)}=0.8520.838=\xi_{M}(\{G_{1},\cdots G_{r}\})<\min_{i}\{\xi_{M}(G_{1}),\cdots,\xi_{M}(G_{r})\}=0.852. One interesting question to ask is what are the conditions for which ξM({G1,Gr})=mini{ξM(G1),,ξM(Gr)}\xi_{M}(\{G_{1},\cdots G_{r}\})=\min_{i}\{\xi_{M}(G_{1}),\cdots,\xi_{M}(G_{r})\}.The following result gives one such condition.

Theorem III.7

If GkG_{k} are all real normal matrices with zero row sums, then ξM=minkminx0,xexTGkxxTx=minkμ2(Gk)\xi_{M}=\min_{k}\min_{x\neq 0,x\perp e}\frac{x^{T}G_{k}x}{x^{T}x}=\min_{k}\mu_{2}(G_{k}).

Proof: Follows from the fact that for a real normal zero row sum matrix GG, ξM(G)=minx0,xexTGx=μ2(G)\xi_{M}(G)=\min_{x\neq 0,x\perp e}x^{T}Gx=\mu_{2}(G) with a fixed U=LK𝒲U=L_{K}\in{\cal W} [31] and thus ξMminkμ2(Gk)\xi_{M}\geq\min_{k}\mu_{2}(G_{k}). \Box

III-C Coupled linear systems

Consider now the case where f(x,t)=Ax+bf(x,t)=Ax+b for some matrix AA and vector bb and the coupling does not vary with time. In this case, the state equation is described by x˙=(IA)x+k=1r(GkDk)x+(Ib)\dot{x}=(I\otimes A)x+\sum_{k=1}^{r}(G_{k}\otimes D_{k})x+(I\otimes b). For the case r=1r=1, the eigenvalues of IA+GDI\otimes A+G\otimes D are equal to A+λjDA+\lambda_{j}D where λj\lambda_{j} are the eigenvalues of GG. If the matrices GkG_{k} can be simultaneously transformed to upper triangular form, then by Lemma II.1 the eigenvalues of (IA)+k=1r(GkDk)(I\otimes A)+\sum_{k=1}^{r}(G_{k}\otimes D_{k}) are equal to the eigenvalues of A+kλkjDkA+\sum_{k}\lambda_{kj}D_{k}, where λkj\lambda_{kj} are the eigenvalues of GkG_{k}. Thus we can study the stability of this linear system by looking at the eigenvalues of A+kλkjDkA+\sum_{k}\lambda_{kj}D_{k} for each set of eigenvalues {λ1j,,λrj}\{\lambda_{1j},\cdots,\lambda_{rj}\}. Extending the approach in [24], we can study a map ϕ:r\phi:{\mathbb{C}}^{r}\rightarrow{\mathbb{R}} which maps {λ1j,,λrj}\{\lambda_{1j},\cdots,\lambda_{rj}\} to the most positive real part of the eigenvalues of A+kλkjDkA+\sum_{k}\lambda_{kj}D_{k}. Define 𝒮={x:ϕ(x)<0}{\cal S}=\{x:\phi(x)<0\}. Then the network is synchronizing if all sets of eigenvalues {λ1j,,λrj}\{\lambda_{1j},\cdots,\lambda_{rj}\} lie in 𝒮{\cal S}, except for the set {0,,0}\{0,\cdots,0\} corresponding to the eigenvectors ee for each GkG_{k}. The shape and size of 𝒮{\cal S} provide insight on how the graph topologies affect the synchronizability of the networked system.

IV Discrete-time systems

Consider a network of coupled discrete-time systems with the following state equations:

x(p+1)=(x1(p+1)xn(p+1))=(IkGk(p)Dk(p))(f(x1(p),p)f(xn(p),p))+u(p)=(IkGk(p)Dk(p))F(x(p),p)+u(p)\begin{array}[]{lcl}&&x(p+1)=\left(\begin{array}[]{c}x_{1}(p+1)\\ \vdots\\ x_{n}(p+1)\end{array}\right)\\ &&=\left(I-\sum_{k}G_{k}(p)\otimes D_{k}(p)\right)\left(\begin{array}[]{c}f(x_{1}(p),p)\\ \vdots\\ f(x_{n}(p),p)\end{array}\right)+u(p)\\ &&=\left(I-\sum_{k}G_{k}(p)\otimes D_{k}(p)\right)F(x(p),p)+u(p)\end{array} (3)
Theorem IV.1

Consider the network of coupled discrete-time systems with state equation Eq. (LABEL:eqn:discrete-time) where Gk(p)G_{k}(p) are normal commuting nn by nn matrices with zero row sums for each kk and pp. Let VV be a symmetric positive definite matrix such that (f(z,p)f(z~,p))TV(f(z,p)f(z~,p))c(zz~)TV(zz~)\left(f(z,p)-f(\tilde{z},p)\right)^{T}V\left(f(z,p)-f(\tilde{z},p)\right)\leq c(z-\tilde{z})^{T}V(z-\tilde{z}) for some c>0c>0 and all pp, zz, z~\tilde{z}. Let |ul(p)um(p)|0|u_{l}(p)-u_{m}(p)|\rightarrow 0 for all ll, mm as pp\rightarrow\infty and let VV be decomposed999For instance via the Cholesky decomposition. as V=CTCV=C^{T}C. Then the coupled systems synchronize, i.e. xl(p)xm(p)x_{l}(p)\rightarrow x_{m}(p) for all ll,mm as pp\rightarrow\infty if for each pp and for each λkiSk\lambda_{ki}\in S_{k}

IkλkiCDk(p)C12<1c\left\|I-\sum_{k}\lambda_{ki}CD_{k}(p)C^{-1}\right\|_{2}<\frac{1}{\sqrt{c}} (4)

where SkS_{k} are the eigenvalues of GkG_{k} not corresponding to ee.

Proof: First consider the case V=IV=I, i.e., f(x,p)f(x,p) is Lipschitz continuous in xx with Lipschitz constant c\sqrt{c}. Let W=IkGkDkW=I-\sum_{k}G_{k}\otimes D_{k}. Since Gk(t)G_{k}(t) is normal and simultaneously diagonalizable, Gk(t)G_{k}(t) has an orthonormal set of eigenvectors viv_{i} with e=v1e=v_{1}. Denote 𝒜\cal A as the subspace of vectors of the form eve\otimes v. Since Gke=0G_{k}e=0 it follows that 𝒜{\cal A} is in the kernel of GkDkG_{k}\otimes D_{k}. Let us denote the subspace orthogonal to 𝒜{\cal A} by {\cal B}. A vector x(p)x(p) is decomposed as x(p)=y(p)+z(p)x(p)=y(p)+z(p) where y(p)𝒜y(p)\in{\cal A} and z(p)z(p)\in{\cal B}. By hypothesis F(x(p),p)F(y(p),p)cz(p)\|F(x(p),p)-F(y(p),p)\|\leq c\|z(p)\|. We can decompose F(x(p),p)F(y(p),p)F(x(p),p)-F(y(p),p) as a(p)+b(p)a(p)+b(p) where a(p)𝒜a(p)\in{\cal A} and b(p)b(p)\in{\cal B}. Note that F(y(p),p)F(y(p),p) and a(p)a(p) are in 𝒜{\cal A} and therefore (GkDk)F(y(p),p)=(GkDk)a(p)=0(G_{k}\otimes D_{k})F(y(p),p)=(G_{k}\otimes D_{k})a(p)=0.

x(p+1)=WFk(x(p),p)+u(p)\displaystyle x(p+1)=WF_{k}(x(p),p)+u(p)
=\displaystyle= a(p)+Fk(y(p),p)+Wb(p)+u(p)\displaystyle a(p)+F_{k}(y(p),p)+Wb(p)+u(p)

Since a(p)b(p)a(p)\bot b(p), b(p)F(x(p),p)F(y(p),p)cz(p)\|b(p)\|\leq\|F(x(p),p)-F(y(p),p)\|\leq c\|z(p)\|. Since b(p)b(p)\in{\cal B}, it is of the form b(p)=i>1,jαivifj(p)b(p)=\sum_{i>1,j}\alpha_{i}v_{i}\otimes f_{j}(p).

Wb(p)=i>1,jαijvifji>1,j,kαijλkiviDkfj\displaystyle Wb(p)=\sum_{i>1,j}\alpha_{ij}v_{i}\otimes f_{j}-\sum_{i>1,j,k}\alpha_{ij}\lambda_{ki}v_{i}\otimes D_{k}f_{j}
=i>1,jαijvi(IkλkiDk)fj\displaystyle=\sum_{i>1,j}\alpha_{ij}v_{i}\otimes\left(I-\sum_{k}\lambda_{ki}D_{k}\right)f_{j}

By hypothesis this implies that

Wb(p)2=i>1,j|αij|2(IkλkiDk)fj2\displaystyle\left\|Wb(p)\right\|^{2}=\sum_{i>1,j}|\alpha_{ij}|^{2}\left\|\left(I-\sum_{k}\lambda_{ki}D_{k}\right)f_{j}\right\|^{2}
<\displaystyle< i>1,j|αij|2fj2/cz(p)2\displaystyle\sum_{i>1,j}|\alpha_{ij}|^{2}\|f_{j}\|^{2}/c\leq\|z(p)\|^{2}

Let w(p)w(p) be the orthogonal projection of u(p)u(p) onto {\cal B}. Since ak(p)+Fk(y(p),p)a_{k}(p)+F_{k}(y(p),p) is in 𝒜{\cal A}, z(p+1)z(p+1) is the orthogonal projection of Wb(p)+u(p)Wb(p)+u(p) onto {\cal B} and thus z(p+1)Wb(p)+w(p)\|z(p+1)\|\leq\|Wb(p)\|+\|w(p)\|. The above shows that there exists β<1\beta<1 such that Wb(p)βz(p)\|Wb(p)\|\leq\beta\|z(p)\|. Thus z(p+1)βz(p)+w(p)\|z(p+1)\|\leq\beta\|z(p)\|+\|w(p)\|. Note that k(GkDk)u(p)0\sum_{k}(G_{k}\otimes D_{k})u(p)\rightarrow 0 as pp\rightarrow\infty and thus w(p)0w(p)\rightarrow 0 as pp\rightarrow\infty. This implies that z(p)0z(p)\rightarrow 0 as pp\rightarrow\infty and x(p)x(p) approaches the synchronization manifold 𝒜{\cal A}.

For general VV, let CTCC^{T}C be the decomposition of VV. Applying the state transformation x~=(IC)x\tilde{x}=(I\otimes C)x, we get x~(p+1)=k(IGkCDkC1)(f~k(x~1(p),p)f~k(x~n(p),p))+u~(p)\tilde{x}(p+1)=\sum_{k}(I-G_{k}\otimes CD_{k}C^{-1})\left(\begin{array}[]{c}\tilde{f}_{k}(\tilde{x}_{1}(p),p)\\ \vdots\\ \tilde{f}_{k}(\tilde{x}_{n}(p),p)\end{array}\right)+\tilde{u}(p) where f~(x~i,p)=Cf(C1x~i,p)\tilde{f}(\tilde{x}_{i},p)=Cf(C^{-1}\tilde{x}_{i},p). We can then apply the result from the V=IV=I case and the conclusion follows. \Box

Theorem IV.1 implies that if Gk=ξLkG_{k}=\xi L_{k}, LkL_{k} are commuting normal Laplacian matrices with the graph sum’s reversal containing a spanning directed tree and ξ\xi is small enough, then the discrete system synchronizes. If V=IV=I and Dk(p)D_{k}(p) are also normal commuting matrices, then Eq. (4) can be simplified to: |1kλkiμkj|1c\left|1-\sum_{k}\lambda_{ki}\mu_{kj}\right|\leq\frac{1}{\sqrt{c}} for all λkiSk\lambda_{ki}\in S_{k} where μkj\mu_{kj} are the eigenvalues of DkD_{k}.

V Conclusions

We study coupled dynamical systems coupled via multiple directed networks and provide criteria under which they synchronize for both continuous-time systems and discrete-time systems. In general, under suitable conditions, two conclusions can be drawn from these synchronization criteria. First, if the coupling between individual systems (as expressed by DkD_{k}) are strong enough, then the graph sum of all the networks jointly containing a spanning directed tree is sufficient to synchronize the network, even though a single network may not be connected. This is an intuitive conclusion as it implies that there is a node that influences every other node via a set of edges where each edge in the set belongs to at least one of the networks. Secondly, even if the coupling of a single network DkD_{k} is not sufficient to synchronize the systems, but the sum kDk\sum_{k}D_{k} does, then the synchronization can still occur with multiple networks if each of the network is well connected, for instance, if they are all strongly connected and balanced.

References

  • [1] “Special issue on chaos synchronization and control: theory and applications.” IEEE Transactions on Circuits and Systems–I: Fundamental Theory and Applications, vol. 44, no. 10, 1997.
  • [2] “Focus issue: control and synchronization of chaos.” Chaos, vol. 7, no. 4, 1997.
  • [3] “Special issue on chaos control and synchronization.” International Bifurcation and Chaos, vol. 10, no. 4, 2000.
  • [4] “Focus issue: control and synchronization in chaotic dynamical systems.” Chaos, vol. 13, no. 1, 2003.
  • [5] C. W. Wu, Synchronization in Complex Networks of Nonlinear Dynamical Systems. World Scientific, 2007.
  • [6] “Focus issue: synchronization in complex networks.” Chaos, vol. 18, no. 3, 2008.
  • [7] L. Kocarev, ed., Consensus and Synchronization in Complex Networks. Understanding Complex Systems, Springer, 2013.
  • [8] S. Baldi and P. Frasca, “Leaderless synchronization of heterogeneous oscillators by adaptively learning the group model,” IEEE Transactions on Automatic Control, vol. 65, no. 1, pp. 412–418, 2020.
  • [9] A. Rahnama and P. J. Antsaklis, “Learning-based event-triggered control for synchronization of passive multiagent systems under attack,” IEEE Transactions on Automatic Control, vol. 65, no. 10, pp. 4170–4185, 2020.
  • [10] K. Abboud, H. A. Omar, and W. Zhuang, “Interworking of DSRC and cellular network technologies for V2X communications: A survey,” in VANET ’04: Proceedings of the 1st ACM international workshop on Vehicular ad hoc networks, vol. 65, pp. 9457–9470, 2016.
  • [11] C. W. Wu, “Synchronization in arrays of coupled nonlinear systems with delay and nonreciprocal time-varying coupling,” IEEE Transactions on Circuits and Systems–II, vol. 53, no. 5, pp. 282–286, 2005.
  • [12] F. Sorrentino, “Synchronization of hypernetworks of coupled dynamical systems,” New Journal of Physics, vol. 14, p. 033035, 2012.
  • [13] L. V. Gambuzza, M. Frasca, and J. Gómez-Gardeñes, “Intra-layer synchronization in multiplex networks,” EPL, vol. 110, p. 20010, 2015.
  • [14] C. I. del Genio, J. Gómez-Gardeñes, I. Bonamassa, and S. Boccaletti, “Synchronization in networks with multiple interaction layers,” Science Advances, vol. 2, no. 11, p. e1601679, 2016.
  • [15] R. Sevilla-Escoboza, I. Sendiña-Nadal, I. Leyva, R. Gutiérrez, J. M. Buldú, and S. Boccaletti, “Inter-layer synchronization in multiplex networks of identical layers,” Chaos, vol. 26, p. 065304, 2016.
  • [16] I. Leyva, I. Sendiña-Nadal, R. Sevilla-Escoboza, V. Vera-Avila, P. Chholak, and S. Boccaletti, “Relay synchronization in multiplex networks,” Scientific reports, vol. 8, no. 1, pp. 1–11, 2018.
  • [17] S. Majhi, D. Ghosh, and J. Kurths, “Emergence of synchronization in multiplex networks of mobile Rössler oscillators,” Physical Review E, vol. 99, no. 1, p. 012308, 2019.
  • [18] L. Tang, X. Wu, J. Lü, J.-a. Lu, and R. M. D’Souza, “Master stability functions for complete, intralayer, and interlayer synchronization in multiplex networks of coupled rössler oscillators,” Physical Review E, vol. 99, no. 1, p. 012304, 2019.
  • [19] K. A. Blaha, K. Huang, F. Della Rossa, L. Pecora, M. Hossein-Zadeh, and F. Sorrentino, “Cluster synchronization in multilayer networks: A fully analog experiment with LC oscillators with physically dissimilar coupling,” Physical Review Letters, vol. 122, no. 1, p. 014101, 2019.
  • [20] B. K. Bera, S. Rakshit, and D. Ghosh, “Intralayer synchronization in neuronal multiplex network,” The European Physical Journal Special Topics, vol. 228, no. 11, pp. 2441–2454, 2019.
  • [21] F. Della Rossa, L. Pecora, K. Blaha, A. Shirin, I. Klickstein, and F. Sorrentino, “Symmetries and cluster synchronization in multilayer networks,” Nature communications, vol. 11, no. 1, pp. 1–17, 2020.
  • [22] A. Kumar, S. Jalan, and A. D. Kachhvah, “Interlayer adaptation-induced explosive synchronization in multiplex networks,” Physical Review Research, vol. 2, p. 023259, 2020.
  • [23] D. A. Burbano-L., S. Yaghouti, C. Petrarca, M. de Magistris, and M. di Bernardo, “Synchronization in multiplex networks of Chua’s circuits: Theory and experiments,” IEEE Transactions on Circuits and Systems I:Regular Papers, vol. 67, no. 3, pp. 927–938, 2020.
  • [24] C. W. Wu and L. O. Chua, “Application of Kronecker products to the analysis of systems with uniform linear coupling,” IEEE Transactions on Circuits and Systems–I: Fundamental Theory and Applications, vol. 42, no. 10, pp. 775–778, 1995.
  • [25] C. W. Wu, Synchronization in coupled chaotic circuits and systems. World Scientific, 2002.
  • [26] R. A. Horn and C. R. Johnson, Matrix analysis. Cambridge University Press, 1985.
  • [27] C. W. Wu, “On bounds of extremal eigenvalues of irreducible and mm-reducible matrices,” Linear Algebra and Its Applications, vol. 402, pp. 29–45, 2005.
  • [28] C. W. Wu and L. O. Chua, “Synchronization in an array of linearly coupled dynamical systems,” IEEE Transactions on Circuits and Systems–I: Fundamental Theory and Applications, vol. 42, no. 8, pp. 430–447, 1995.
  • [29] C. W. Wu, “Perturbation of coupling matrices and its effect on the synchronizability in arrays of coupled chaotic circuits,” Physics Letters A, vol. 319, pp. 495–503, 2003.
  • [30] C. W. Wu, “Synchronization in networks of nonlinear dynamical systems coupled via a directed graph,” Nonlinearity, vol. 18, pp. 1057–1064, 2005.
  • [31] C. W. Wu, “On a matrix inequality and its application to the synchronization in coupled chaotic systems,” in Complex Computing-Networks:Brain-like and Wave-oriented Electrodynamic Algorithms, vol. 104 of Springer Proceedings in Physics, pp. 279–288, 2006.
  • [32] M. Grant and S. Boyd, “CVX: Matlab software for disciplined convex programming, version 2.1.” http://cvxr.com/cvx, Mar. 2014.
  • [33] M. Grant and S. Boyd, “Graph implementations for nonsmooth convex programs,” in Recent Advances in Learning and Control (V. Blondel, S. Boyd, and H. Kimura, eds.), Lecture Notes in Control and Information Sciences, pp. 95–110, Springer-Verlag Limited, 2008. http://stanford.edu/~boyd/graph_dcp.html.
  • [34] C. W. Wu, “Synchronization in coupled arrays of chaotic oscillators with nonreciprocal coupling,” IEEE Transactions on Circuits and Systems–I: Fundamental Theory and Applications, vol. 50, no. 2, pp. 294–297, 2003.