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

Sparse Network Optimization for Synchronization

Authors    Regina Sandra Burachik   Alexander C. Kalloniatis   C. Yalçın Kaya Mathematics, UniSA STEM, University of South Australia, Mawson Lakes, S.A. 5095, Australia. E-mail: [email protected] .Joint and Operations Analysis Division, Defence Science and Technology Group, Canberra 2600, Australia. E-mail: [email protected] .Mathematics, UniSA STEM, University of South Australia, Mawson Lakes, S.A. 5095, Australia. E-mail: [email protected] .

Abstract.  We propose new mathematical optimization models for generating sparse dynamical graphs, or networks, that can achieve synchronization. The synchronization phenomenon is studied using the Kuramoto model, defined in terms of the adjacency matrix of the graph and the coupling strength of the network, modelling the so-called coupled oscillators. Besides sparsity, we aim to obtain graphs which have good connectivity properties, resulting in small coupling strength for synchronization. We formulate three mathematical optimization models for this purpose. Our first model is a mixed integer optimization problem, subject to ODE constraints, reminiscent of an optimal control problem. As expected, this problem is computationally very challenging, if not impossible, to solve, not only because it involves binary variables but also some of its variables are functions. The second model is a continuous relaxation of the first one, and the third is a discretization of the second, which is computationally tractable by employing standard optimization software. We design dynamical graphs that synchronize, by solving the relaxed problem and applying a practical algorithm for various graph sizes, with randomly generated intrinsic natural frequencies and initial phase variables. We test robustness of these graphs by carrying out numerical simulations with random data and constructing the expected value of the network’s order parameter and its variance under this random data, as a guide for assessment.

Key words: Optimization, Sparse graphs, Sparse networks, Synchronization, Kuramoto model, Optimal control, Discretization.

AMS subject classifications. Primary 49M20, 49M25  Secondary 92B25

1 Introduction

Spontaneous synchronization happens frequently in nature: common examples are fireflies flashing, crickets chirping, planets orbiting and neurons firing. All these phenomena consist of a set of agents that exhibit a cyclic behaviour: these agents are called oscillators [1]. Two or more oscillators are said to be coupled if they influence each other by some physical or chemical process. In our setting, each oscillator is associated with the node of a graph, and the whole graph represents a network. The result of such interactions is often synchrony, which means that all entities end up with the same frequency after a short period of time. The Kuramoto model [2] was originally motivated whereby a system of coupled oscillating nodes for sufficient coupling lock on to a common frequency, in this case the mean of the ensemble of natural frequencies of the oscillators, despite the variance in those frequencies. For a recent literature review of the model see [3].

In the following we summarise salient aspects of the literature relevant to network optimization. The original model, as in [2], was posed for all-to-all coupling in a complete graph. Since then the pattern of synchronization has been explored for a variety of regular and irregular graphs. At one end, sparse regular graphs are the poorest in synchronization, by which we mean they require high coupling strength to achieve synchronization. For example, synchronization on regular structures such as trees [4] and rings [5, 6] admit a degree of solubility for critical coupling, or bounds thereon. These are in some respects balanced structures: sparse, with uniform degree distributions. However, as they scale in size they require increasingly higher values of coupling to achieve synchrony. Irregular or complex graphs, such as Erdös-Rényi [7], small world [8], and the scale-free Barabasi-Alberts [9] networks synchronize for significantly smaller coupling, with small world generally the worst of these [10]. However, their complexity leads to very uneven distribution of degrees which may incur some cost, for example imbalances in work or communication loads depending on the physical setting of the synchronization process. In our case, our interest in such graphs is for the purpose of synchronized decision-making in organizations [11] where span of control and cognitive overload require sparseness and balanced degree distribution.

In terms of approaches that explicitly use a form of optimization, [12] employed a stochastic hill-climbing approach based on random network link rewirings and acceptance or rejection on the criterion of improving the Kuramoto order parameter rr, to be defined below. Contemporaneously, [13] applied gradient descent to the order parameter to be able to more efficiently compute optimal weighted networks. The authors of [14] follow an evolution process similar to that of [12], but, to avoid hubs in the network, the hill-climbing trajectory only accepts a rewiring that preserves the initial degree distribution. In the same work, the authors propose the use of an additional energy-like function in the selection process.

A Markov Monte Carlo replica exchange approach was developed by [15] to optimise network synchronization against noise, but the method becomes computationally demanding beyond N=15N=15 size graphs. In a second order generalization of the Kuramoto model, relevant to electrical power systems, [16] used convex optimization to determine optimal placements of frequencies or link weights within an existing network structure. The examples considered in that work were of similarly small size.

A number of optimization approaches have been applied to a related, but simpler, coupled oscillator model in which the graph Laplacian figures explicitly in the interaction function, and to which the Kuramoto model approximates under linearization. Using the master stability function of [17], the authors of [18] observed that graphs with lower Laplacian spectral ratio exhibited improved synchronization. In this spirit, a stochastic rewiring approach based on improving the spectral ratio was explored by [19, 20] yielding heterogeneous and typically dense graphs. Complementary to the spectral ratio approach is that of expander graphs which have the property that every partition of the nodes into two subsets has a number of boundary edges between them that scales with the size of the smallest subset. Here [21] conducted computational search over instances of graphs with such properties, with results that have excellent expansion and synchronization properties but that increase in degree as the graphs are scaled in size.

To conclude this summary we mention that [22] have provided a construction for the Kuramoto model that yields modified trees – which are therefore sparse and well-balanced – that are provably expanders, with enhanced synchronization. They also showed that a computational search over random variations of the links on the leaf nodes of this construction (maintaining the degree distribution) to find instances with lowest possible spectral ratio did not necessarily give graphs of better synchronization.

For more details on similar techniques, which progressively search for a network so as to promote its fitness for synchronization, see [3, Section 8].

In most of these approaches, however, optimization takes the form of a hill-climbing evolution, where, according to certain rules and merit functions, a sequence of “mutations” of an initial network, is constructed. After a prescribed number of steps (usually stipulated a priori), the last mutation accepted will exhibit some “optimal” properties, in the sense that the merit functions used in the process have the best values among all the previous mutations.

In the present paper, we propose, implement and solve a new optimization model for designing graphs that (i) are sparse, (ii) can achieve synchronization in the context of the Kuramoto model [3] and (iii) have a relatively narrow range of small degrees. Conditions (i) and (iii) are helpful for ease of implementation of a network, as well as for promoting balanced sharing of communication- or work-loads.

We mathematically incorporate these three features into our optimization model. Indeed, our optimization variables are the set of all weighted (connected) undirected networks (expressed as N×NN\times N symmetric matrices with nonnegative coefficients), and a set of NN functions, namely the phases {θi}\{\theta_{i}\} for i=1,,Ni=1,\ldots,N in the Kuramoto system. Our objective function involves a penalty term ensuring sparsity (given by the 1\ell_{1} norm of the matrix) and two other terms promoting the small variance of the degrees (see the second and third terms in the objective function in (6)).

To the authors knowledge, there is no available approach that proposes a mathematical optimization model that performs a “universal search” among all undirected networks which are able to synchronize by means of the Kuramoto model.

The initial version of the optimization model we construct is an infinite-dimensional mixed-integer programming problem, which is in the form of an optimal control problem. By relaxing the integer variables, we obtain a continuous-variable model, which is nonsmooth and still infinite-dimensional, and therefore not tractable yet. We then discretize the problem, so that optimization software can be employed, as it is done for optimal control problems. The discretization produces a large-scale problem, which, in general, prohibits any use of nonsmooth solvers. However, in our particular model, non-smoothness appears in a peculiar form that can be transformed into a smooth, i.e., differentiable, form. We use tricks and techniques from optimization to convert the problem into a smooth one so that powerful differentiable solvers can be employed to get a solution, even though the problem is still a large-scale and non-convex one.

One should note that a solution to our discrete and smooth optimization model does not represent a graph immediately. A key procedure we propose, Algorithm 1, uses solutions of our discrete and smooth optimization model so as to construct increasingly sparse graphs while keeping the spectral ratio of the Laplacian relatively small. In other words, Algorithm 1 helps us

  • (i)

    find a sparse adjacency matrix of the graph such that the system defined by the equations in (1) below is synchronized, and

  • (ii)

    ensure that the resulting graph exhibits a small spectral ratio (defined in (5) below) and

  • (iii)

    has a possibly small and narrow range of node degrees.

We present extensive numerical experiments by using this algorithm for graphs with N=20N=20 and N=127N=127. We illustrate the fact that, when adding a moderate number of edges, a great improvement in synchronization properties of the graph can be achieved. We also make comparisons with certain graphs from the literature, in particular from [22], where graphs are obtained by appending at the bottom level of a hierarchical tree, additional matchings linking the leaf nodes. The graphs found in this way in [22] enhance synchronization compared to the original tree. We emphasise the key result: that we can generate graphs with optimal sparseness, balanced load, and good synchronization properties starting from a random frequency and initial phase input. Because the graphs we generate show low degree heterogeneity, the resultant synchronization is not so sensitive to the input frequency choices.

The paper is organized as follows. In Section 2, we recall classical concepts and properties of the Kuramoto model and the spectral ratio of a graph. In Section 3 we introduce our mathematical model that promotes synchronization as well as pose its relaxation. In Section 4, we discuss discretization of the time-dependent functions in the model and smoothing. In Section 5 we present our numerical experiments and provide a detailed interpretation of the numerical results. In Section 6, we provide concluding remarks and comment on future work.

2 Preliminaries

In this section we briefly recall the mathematical concepts we will use in our models, including the Tree-expander Construction approach presented in [22].

2.1 The Kuramoto Model

Let A{0,1}N×NA\in\{0,1\}^{N\times N} (i.e., AA is an N×NN\times N matrix with all its entries equal to 0 or 11) and ω=(ω1,,ωN)IRN\omega=(\omega_{1},\ldots,\omega_{N})\in{\rm{I\ \kern-5.39993ptR}}^{N}. The Kuramoto model over an NN-node network can be written as a system of ODEs:

θ˙i(t)=ωiσj=1NAijsin(θi(t)θj(t)),θi(0)=θi,0,i=1,,N,\dot{\theta}_{i}(t)=\omega_{i}-\sigma\sum_{j=1}^{{N}}\,A_{ij}\,\sin(\theta_{i}(t)-\theta_{j}(t)),\qquad\theta_{i}(0)=\theta_{i,0}\,,\ \ i=1,\ldots,N, (1)

where θi:[0,T]IR\theta_{i}:[0,T]\to{\rm{I\ \kern-5.39993ptR}} represents the phase of the iith node, with θ˙i:=dθ/dt\dot{\theta}_{i}:=d\theta/dt, and θ(t):=(θ1(t),,θN(t))\theta(t):=(\theta_{1}(t),\ldots,\theta_{N}(t)) is the phase vector in IRN{\rm{I\ \kern-5.39993ptR}}^{N}. Moreover, ω\omega is the vector of intrinsic natural frequencies, σ\sigma is the coupling strength of the network, θ0:=(θi,1,,θi,N)\theta_{0}:=(\theta_{i,1},\ldots,\theta_{i,N}) is the initial phase vector, and AA is the adjacency matrix of a graph with NN nodes.

It is important to distinguish two types of synchronization. Phase synchronization refers to asymptotic equality of the phases, namely the case when (θi(t)θj(t))0(\theta_{i}(t)-\theta_{j}(t))\to 0 as t,i,j{1,,N}t\to\infty,\ \forall\ i,j\in\{1,\ldots,N\}. This cannot hold for non-identical oscillators, where ωiωj\omega_{i}\neq\omega_{j}. Frequency synchronization, on the other hand, refers to asymptotic equality of the frequencies, namely the case when (θ˙i(t)θ˙j(t))0(\dot{\theta}_{i}(t)-\dot{\theta}_{j}(t))\to 0 as t,i,j{1,,N}t\to\infty,\ \forall\ i,j\in\{1,\ldots,N\}.

Except in certain regular graphs, when coupling is sufficient for frequency synchronization to set in, the corresponding phases tend to be clustered close to each other, namely that (θi(t)θj(t))ε(\theta_{i}(t)-\theta_{j}(t))\leq\varepsilon for all t>tt>t_{\ell} and all i,j{1,,N}i,j\in\{1,\ldots,N\}, with some ε>0\varepsilon>0 and sufficiently large t>0t_{\ell}>0, which is referred to as (full) phase locking. The higher the coupling is, the smaller the discrepancies between the oscillator phases will be. In the computational setting that we have, we consider the finite time horizon [0,T][0,T] with TtT\geq t_{\ell}. This synchronization, however, usually requires dense matrices AA. Viewing links as costly (in resource, in communications, in load), we seek to keep the topology of the graph as simple as possible. Therefore, it is relevant to find sparse matrices AA for which synchronization is achieved.

2.2 The order parameter

The so-called order parameter r(t)r(t) is a function of time tt defined as

r(t)=1N|j=1Neiθj(t)|,r(t)=\frac{1}{N}\,\left|\sum_{j=1}^{N}e^{i\,\theta_{j}(t)}\right|\,, (2)

where |||\cdot| denotes the magnitude of a complex number, with i=1i=\sqrt{-1}. It is not difficult to derive, from (2), that

r2(t)=1N2[N+2i<jNcos(θi(t)θj(t))],r^{2}(t)=\frac{1}{N^{2}}\,\left[N+2\,\sum_{i<j}^{N}\,\cos(\theta_{i}(t)-\theta_{j}(t))\right]\,, (3)

where the usage of ii as an index this time should be clear from the context. As a measure of synchronicity, usually the following long-term average r¯\overline{r} of rr is used:

r¯(T)=1TTthTthTr(t)𝑑t,\overline{r}(T)=\frac{1}{T-T_{th}}\,\int_{T_{th}}^{T}r(t)\,dt\,, (4)

where T>0T>0 is a fixed large number and TthT_{th} is some fraction of the time-series describing the transient behaviour.

2.3 Spectral Ratio

Given a graph GG with NN nodes, call did_{i} the degree of node ii. Collecting all these coordinates in a single vector d=(d1,,dN)INNd=(d_{1},\ldots,d_{N})\in{\rm{I\ \kern-5.39993ptN}}^{N} produces the degree vector dd of GG. If AA is the adjacency matrix of GG, then the Laplacian LL of GG is defined as

L:=diag(d)A,L:=\operatorname*{diag}(d)-A\,,

where diag(d)\operatorname*{diag}(d) is the diagonal matrix with did_{i} its iith diagonal element. Let λ2\lambda_{2} and λmax\lambda_{\max} be the second smallest and the largest eigenvalues of LL, respectively. Recall that λ20\lambda_{2}\geq 0. Moreover, if the graph is connected, then λ2>0\lambda_{2}>0. For a connected graph, we denote by

QL:=λmaxλ2,Q_{L}:=\dfrac{\lambda_{\max}}{\lambda_{2}}\,, (5)

the spectral ratio of the graph GG. As alluded in the introduction, for a somewhat different class of dynamical systems, which the Kuramoto model approximates when linearised, graphs with small QLQ_{L} seem to result in highly synchronised graphs, in that the graph is connected and σ\sigma is small [18]. For the Kuramoto model such an effect has been observed in [22].

2.4 Tree–Expander construction

In [22], Taylor, Kalloniatis and Hoek propose a construction procedure for graphs which appends at the bottom level of a hierarchical tree additional matchings linking the leaf nodes. Recall that if the number of levels in a hierarchical tree is mm, then the total number of nodes in the tree is p=2m1p=2^{m}-1. Following the procedure given in [22] we have constructed a graph for m=7m=7, i.e., p=127p=127. We show in Section 5.2.2 a 158-edge graph we have obtained as well as a 158-edge that was obtained in [22], and report their QLQ_{L} values. Intuitively, an expander graph is a graph in which every subset of the vertices that is not “too large” has a “large” boundary. Intuitively, this property favours connectivity. The authors show in [22] that the graph they construct is an expander graph. The resulting graphs have significantly improved synchronization properties, and are sparse, of nearly uniform degree distribution and small QLQ_{L} (although, for other random leaf node matchings of the tree, the expander construction does not give the smallest possible QLQ_{L}). They therefore provide a reasonable benchmark to compare the results of our optimization. In Section 5.2.2 we compare the properties of these expander graphs with those obtained through our optimization approach described in Algorithm 1.

3 Optimization Models

As mentioned in Section 1, our aims are (i) to find a sparse AA, and to achieve (ii) synchronization and (iii) a narrow distribution of degrees. These aims, when realised, are expected to yield a relatively small QLQ_{L}. To obtain such a matrix AA, we propose optimization models with decision variables (A,θ)(A,\theta), where AIRN×NA\in{\rm{I\ \kern-5.39993ptR}}^{N\times N} and θ\theta is the vector function defined in (1). When the graph is assumed to be undirected, the matrix AA is symmetric. Similar models can be posed for the directed case. We fix a random vector ω\omega, and a time tct_{c} after which we impose synchronization among θi\theta_{i}’s.

3.1 A Mixed Integer Optimization Problem

The coefficients of the adjacency matrix are binary, i.e., Aij{0,1}A_{ij}\in\{0,1\}. Let dd be the vector of degrees as defined in Section 2.3. Also define

dmin:=min1iNdianddmax:=max1iNdi.d_{\min}:=\min_{1\leq i\leq N}d_{i}\quad\mbox{and}\quad d_{\max}:=\max_{1\leq i\leq N}d_{i}.

Fix positive parameters c1,c2,c3c_{1},c_{2},c_{3}, tct_{c}, as well as a small ε>0\varepsilon>0. We propose the following optimization model for the problem that we have.

{objective:minA,θc1i,j|Aij|+c2dmaxdmin+c3(dmaxdmin1)subject to the following constraints.integrality:Aij{0,1}i,j,symmetricity:Aij=Ajii,j,oscillator dynamics:θ˙i(t)=ωi+σj=1NAijsin(θj(t)θi(t)),t[0,T],θi(0)=θi,0,i,synchrony:maxttc{0,(ij|θi(t)θj(t)|)ε}=0,i,j.\left\{\begin{array}[]{ll}&\hbox{\em objective}:\\ &\hskip 28.45274pt\displaystyle\min_{A,\theta}\,\,c_{1}\sum_{i,j}|A_{ij}|+c_{2}\,\dfrac{d_{\max}}{d_{\min}}+c_{3}\,\left(\dfrac{d_{\max}}{d_{\min}}-1\right)\\[17.07164pt] &\hbox{\em subject to the following constraints.}\\ &\\ &\hbox{\em integrality}:\ \ A_{ij}\in\{0,1\}\ \ \forall i,j,\\ &\\ &\hbox{\em symmetricity}:\ \ A_{ij}=A_{ji}\ \ \forall i,j,\\ &\\ &\hbox{\em oscillator dynamics}:\\ &\displaystyle\dot{\theta}_{i}(t)=\omega_{i}+\sigma\,\sum_{{j=1}}^{{N}}A_{ij}\sin(\theta_{j}(t)-\theta_{i}(t))\,,\ \forall t\in[0,T]\,,\quad\theta_{i}(0)=\theta_{i,0}\,,\ \ \forall i\,,\\ &\\ &\hbox{\em synchrony}:\ \ \max_{t\geq t_{c}}\left\{0,\displaystyle\left(\sum_{ij}|\theta_{i}(t)-\theta_{j}(t)|\right)-\varepsilon\right\}=0\,,\ \forall i,j\,.\end{array}\right. (6)

The following comments and observations can be made about this model.

  • The first term in the objective function is the 1\ell_{1}-norm of the elements of the matrix AA, which, when minimized, is well-known to promote sparsity of AA. The two subsequent terms represent the overall discrepancy between the degrees of the nodes. It turns out that addition of these terms accelerates computation by 10 to 100 times.

  • We have tried specific software for mixed integer optimization, but due to the large number of variables, the above model has so far been intractable.

  • We use a given value of σ\sigma in the ODE constraints. It is possible to consider σ\sigma as another decision variable, if smaller values of σ\sigma are desired.

3.2 A model with relaxation

Due to the computational complexity, or intractability, of the model in (6), we can consider a relaxation of the model in the following way. First, combine the coefficients of the matrix and the coupling strength σ\sigma into a single aggregated variable

A~ij:=σAij,\widetilde{A}_{ij}:=\sigma A_{ij}\,, (7)

and allow A~ij\widetilde{A}_{ij} to take any nonnegative real value, i.e. A~ij0\widetilde{A}_{ij}\geq 0. Next, we also relax the (integer-valued) degree vector by means of the generalized degree vector d~=(d~1,,d~N)IRN\widetilde{d}=(\widetilde{d}_{1},\ldots,\widetilde{d}_{N})\in{\rm{I\ \kern-5.39993ptR}}^{N} such that

d~i:=j=1NA~ij,\widetilde{d}_{i}:=\sum_{j=1}^{N}\widetilde{A}_{ij}\,, (8)

for i=1,,Ni=1,\ldots,N. Subsequently, the maximum and minimum degree variables we used in the previous model are modified accordingly as

d~min:=min1iNd~iandd~max:=max1iNd~i.\widetilde{d}_{\min}:=\min_{1\leq i\leq N}\widetilde{d}_{i}\quad\mbox{and}\quad\widetilde{d}_{\max}:=\max_{1\leq i\leq N}\widetilde{d}_{i}\,. (9)

The relaxation (7) allows to consider the ODE (1) with no σ\sigma appearing on the right-hand side, and, more importantly, to replace the integrality constraints used in the 1\ell_{1}-model in (6) by non-negativity constraints. The use of these aggregated variables allows us to propose a model which can be more easily implemented. So, fix positive parameters c1,c2c_{1},c_{2} and c3c_{3}, as before, and consider the following optimization problem.

{objective:minA~,θc1i,j|A~ij|+c2d~maxd~min+c3(d~maxd~min1)subject to the following constraints.symmetricity:Aij=Ajii,j,nonnegativity:A~ij0,i,j,oscillator dynamics:θ˙i(t)=ωi+j=1NA~ijsin(θj(t)θi(t)),t[0,T],θi(0)=θi,0,i,synchrony:maxttc{0,(ij|θi(t)θj(t)|)ε}=0,i,j.\left\{\begin{array}[]{ll}&\hbox{\em objective}:\\ &\hskip 28.45274pt\displaystyle\min_{\widetilde{A},\theta}\,\,c_{1}\sum_{i,j}|\widetilde{A}_{ij}|+c_{2}\,\dfrac{\widetilde{d}_{\max}}{\widetilde{d}_{\min}}+c_{3}\,\left(\dfrac{\widetilde{d}_{\max}}{\widetilde{d}_{\min}}-1\right)\\[17.07164pt] &\hbox{\em subject to the following constraints.}\\ &\\ &\hbox{\em symmetricity}:\ \ A_{ij}=A_{ji}\ \ \forall i,j,\\ &\\ &\hbox{\em nonnegativity}:\ \ \widetilde{A}_{ij}\geq 0\,,\ \ \forall i,j,\\ &\\ &\hbox{\em oscillator dynamics}:\\ &\displaystyle\dot{\theta}_{i}(t)=\omega_{i}+\sum_{{j=1}}^{{N}}\widetilde{A}_{ij}\sin(\theta_{j}(t)-\theta_{i}(t))\,,\ \forall t\in[0,T]\,,\quad\theta_{i}(0)=\theta_{i,0}\,,\ \ \forall i\,,\\ &\\ &\hbox{\em synchrony}:\ \ \max_{t\geq t_{c}}\left\{0,\left(\sum_{ij}|\theta_{i}(t)-\theta_{j}(t)|\right)-\varepsilon\right\}=0\,,\ \ \forall i,j\,.\end{array}\right.\\ (10)

Since every term in (8) is multiplied by σ\sigma, we have that

d~maxd~min=dmaxdmin.\dfrac{\widetilde{d}_{\max}}{\widetilde{d}_{\min}}=\dfrac{{d}_{\max}}{{d}_{\min}}\,.

Therefore, the terms involving (dmax/dmin)(d_{\max}/d_{\min}) in the model (6) remain unchanged by the relaxation, in that (d~max/d~min)(\widetilde{d}_{\max}/\widetilde{d}_{\min}) still truly means the ratio of the maximum and minimum degrees in the graph.

4 Discretization and Smoothing

As it stands in Section 3.2, the 1\ell_{1}-model is nondifferentiable. On the other hand, it is well-known that a smooth re-formulation of the 1\ell_{1}-model can be obtained by using standard nonlinear programming techniques—see e.g. [23, 24]. This allows us to use smooth optimization solvers for solving the discretized 1\ell_{1}-model.

With the elements A~ij\widetilde{A}_{ij}s of the matrix A~\widetilde{A} interpreted as constant control functions and θi\theta_{i}s interpreted as the state variables, the model in (10) is an optimal control problem. We have discretized the 1\ell_{1}-model, using the Trapezoidal rule, in a fashion similar to the way optimal control problems are discretized; see, for example, [25, 26]. Challenging optimal control problems have successfully been solved using direct discretization previously [28, 30, 27, 29].

Suppose that the ODE in Model (10) is defined over the time horizon [0,T][0,T], with T=tcT=t_{c}. Let

f(θ1(t),,θN(t),A~):=ωi+j=1NA~ijsin(θj(t)θi(t)),t[0,T].f(\theta_{1}(t),\ldots,\theta_{N}(t),\widetilde{A}):=\omega_{i}+\sum_{{j=1}}^{{N}}\widetilde{A}_{ij}\sin(\theta_{j}(t)-\theta_{i}(t)),\quad\forall t\in[0,T]\,. (11)

Consider a regular partition of [0,T][0,T], such that 0=t0<t1<<tM=T0=t_{0}<t_{1}<\ldots<t_{M}=T, with tk+1:=tk+ht_{k+1}:=t_{k}+h and h:=T/Mh:=T/M. Let θik\theta_{i}^{k} be an approximation of θi(tk)\theta_{i}(t_{k}), k=0,1,,M1k=0,1,\ldots,M-1. Then the ODE can be discretized by applying the trapezoidal rule:

θik+1\displaystyle\theta_{i}^{k+1} =\displaystyle= θik+h2(f(θ1k,,θNk,A~)+f(θ1k+1,,θNk+1,A~)),\displaystyle\theta_{i}^{k}+\frac{h}{2}\left(f(\theta_{1}^{k},\ldots,\theta_{N}^{k},\widetilde{A})+f(\theta_{1}^{k+1},\ldots,\theta_{N}^{k+1},\widetilde{A})\right), (12)

k=0,1,,M1k=0,1,\ldots,M-1. After this discretization, and the replacement of the ODE with this discretization, the optimization model in (10) becomes a finite-dimensional optimization problem. With the number of optimization variables being [(N1)NM/2][(N-1)NM/2], the discretized problem is large-scale even with a moderate graph size NN and a moderate partition size MM. Moreover, the objective function of the problem is nondifferentiable, prohibiting efficient use of smooth optimization methods and software.

Techniques from the optimization literature can be employed to transform the nonsmooth objective function into a smooth one as follows. Nonsmoothness of the first term (the 1\ell_{1}-norm) is caused by the moduli, which can be avoided by defining two new variables, Bij1B_{ij}^{1} and Bij2B_{ij}^{2} such that (see e.g. [24])

A~ij:=Bij1Bij2,Bij1,Bij20.\widetilde{A}_{ij}:=B_{ij}^{1}-B_{ij}^{2}\,,\quad B_{ij}^{1},B_{ij}^{2}\geq 0\,. (13)

Then one can show that

|A~ij|=Bij1+Bij2.|\widetilde{A}_{ij}|=B_{ij}^{1}+B_{ij}^{2}\,. (14)

The RHS of the ODEs in (11) then becomes

f(θ1(t),,θN(t),Bij1,Bij2):=ωi+j=1N(Bij1Bij2)sin(θj(t)θi(t)),t[0,T].f(\theta_{1}(t),\ldots,\theta_{N}(t),B_{ij}^{1},B_{ij}^{2}):=\omega_{i}+\sum_{{j=1}}^{{N}}(B_{ij}^{1}-B_{ij}^{2})\,\sin(\theta_{j}(t)-\theta_{i}(t))\,,\quad\forall t\in[0,T]. (15)

To tackle the second and third terms in the objective function, one can define the new optimization variables (see e.g. [23])

α:=d~maxandβ:=d~min.\alpha:=\widetilde{d}_{\max}\quad\mbox{and}\quad\beta:=\widetilde{d}_{\min}\,.

Then

mind~maxd~minminαβ\min\frac{\widetilde{d}_{\max}}{\widetilde{d}_{\min}}\equiv\min\frac{\alpha}{\beta}

with

j=1NA~ijα,for i=1,,N,\displaystyle\sum_{j=1}^{N}\widetilde{A}_{ij}\leq\alpha\,,\quad\mbox{for }i=1,\ldots,N\,, (16)
j=1NA~ijβ,for i=1,,N.\displaystyle\sum_{j=1}^{N}\widetilde{A}_{ij}\geq\beta\,,\quad\mbox{for }i=1,\ldots,N\,. (17)

The synchrony constraint in Model (10) is also nonsmooth, which can be replaced by a smooth version given by

ij(θiMθjM)2ε.\displaystyle\sum_{i\neq j}\left(\theta_{i}^{M}-\theta_{j}^{M}\right)^{2}\leq\varepsilon\,. (18)

Incorporation of (11)–(18) transforms Model (10) into the smooth optimization problem below.

{objective:minB1,B2,θk,α,βc1i,j(Bij1+Bij2)+c2αβ+c3(αβ1)subject to the following constraints.symmetricity:Bij1Bij2=Bji1Bji2,i,j,nonnegativity:Bij1Bij20,i,j,discretization:θik+1=θik+h2(f(θ1k,,θNk,Bij1,Bij2)+f(θ1k+1,,θNk+1,Bij1,Bij2)),k=0,1,,M1,θi0=θi,0,i,synchrony:ij(θiMθjM)2ε,auxiliary:Bij1,Bij20,i,j,j=1N(Bij1Bij2)α,i,j=1N(Bij1Bij2)β,i.\left\{\begin{array}[]{ll}&\hbox{\em objective}:\\ &\displaystyle\min_{B^{1},B^{2},\theta^{k},\alpha,\beta}\,\,c_{1}\sum_{i,j}(B_{ij}^{1}+B_{ij}^{2})+c_{2}\,\dfrac{\alpha}{\beta}+c_{3}\,\left(\dfrac{\alpha}{\beta}-1\right)\\[17.07164pt] &\hbox{\em subject to the following constraints.}\\ &\\ &\hbox{\em symmetricity}:\ \ B_{ij}^{1}-B_{ij}^{2}=B_{ji}^{1}-B_{ji}^{2}\,,\ \ \forall i,j,\\ &\\ &\hbox{\em nonnegativity}:\ \ B_{ij}^{1}-B_{ij}^{2}\geq 0\,,\ \ \forall i,j,\\ &\\ &\hbox{\em discretization}:\\[5.69054pt] &\theta_{i}^{k+1}=\theta_{i}^{k}+\displaystyle\frac{h}{2}\left(f(\theta_{1}^{k},\ldots,\theta_{N}^{k},B_{ij}^{1},B_{ij}^{2})+f(\theta_{1}^{k+1},\ldots,\theta_{N}^{k+1},B_{ij}^{1},B_{ij}^{2})\right),\\[5.69054pt] &\hskip 199.16928ptk=0,1,\ldots,M-1\,,\ \ \theta_{i}^{0}=\theta_{i,0}\,,\ \ \forall i,\\[5.69054pt] &\hbox{\em synchrony}:\\[2.84526pt] &\displaystyle\sum_{i\neq j}\left(\theta_{i}^{M}-\theta_{j}^{M}\right)^{2}\leq\varepsilon\,,\\ &\\ &\hbox{\em auxiliary}:\ \ B_{ij}^{1},B_{ij}^{2}\geq 0\,,\forall i,j,\\ &\\ &\hskip 56.9055pt\sum_{j=1}^{N}(B_{ij}^{1}-B_{ij}^{2})\leq\alpha\,,\ \ \forall i\,,\\ &\\ &\hskip 56.9055pt\sum_{j=1}^{N}(B_{ij}^{1}-B_{ij}^{2})\geq\beta\,,\ \ \forall i\,.\end{array}\right. (19)

For the above problem a standard differentiable optimization software can now be used to find a solution.

5 Numerical Implementation and Experiments

5.1 Numerical implementation

We consider a numerical implementation of the smooth and finite-dimensional model (19) derived in Section 4. We choose the initial values for the vector of intrinsic frequencies ωi\omega_{i} from the normal distribution with mean 22 and variance 1/21/2, namely ωi𝒩(2,1/2)\omega_{i}\sim{\cal N}(2,1/2) and the initial conditions θi(0)\theta_{i}(0), i=1,,Ni=1,\ldots,N, are drawn from the uniform distribution over the interval [π/2,π/2][-\pi/2,\pi/2], i=1,,Ni=1,\ldots,N.

In solving the discretized and re-formulated optimization model (19), we have paired up the optimization modelling language AMPL [31] and the finite-dimensional optimization software Ipopt [32]. The AMPL–Ipopt suite was run on a 13-inch 2018 model MacBook Pro, with the operating system macOS Mojave (version 10.14.6), the processor 2.7 GHz Intel Core i7 and the memory 16 GB 2133 MHz LPDDR3. The Ipopt options tol=1e-8, acceptable_tol=1e-8 and linear_solver=ma57 were provided within an AMPL code.

We have used the following choice of the penalty parameters in model (19): c1=1c_{1}=1, c2=c3=100c_{2}=c_{3}=100. As mentioned before, the use of the terms with c2c_{2} and c3c_{3} accelerated convergence by 10–100 times, compared with the case when c2=c3=0c_{2}=c_{3}=0 and the Euclidean distance ij(d~(i)d~(j))2\sum_{ij}(\widetilde{d}(i)-\widetilde{d}(j))^{2} is penalized instead.

A numerical solution of (19) yields nonnegative real values for A~ij=Bij1Bij2\widetilde{A}_{ij}=B_{ij}^{1}-B_{ij}^{2}. It should be noted that A~ij/σ\widetilde{A}_{ij}/\sigma for some σ>0\sigma>0 does not necessarily result in Aij{0,1}{A}_{ij}\in\{0,1\}. Therefore, we implement the following practical prototype algorithm in constructing a sparse binary adjacency matrix AA from the real matrix A~\widetilde{A}.

Prototype Algorithm 

Step 1

(Initialization) Using some ωi𝒩(2,1/2)\omega_{i}\in{\cal N}(2,1/2), θi(0)𝒰[π/2,π/2]\theta_{i}(0)\in{\cal U}[-\pi/2,\pi/2], i=1,,Ni=1,\ldots,N, and integer M>0M>0, find a solution A~ij\widetilde{A}_{ij} to the relaxed model (19). Set ε>0\varepsilon>0 small, η0=0\eta_{0}=0, T=2000T=2000 and tc=1500t_{c}=1500. Set the coupling strength σ>0\sigma>0 large. Set k=0k=0.

Step 2

(Generate adjacency matrix A(k)A^{(k)}) Let  Aij(k):=sgn(max{0,A~ijηk}){A}_{ij}^{(k)}:=\operatorname*{sgn}(\max\{0,\widetilde{A}_{ij}-\eta_{k}\}),  for all i,j=1,,Ni,j=1,\ldots,N.

Step 3

(Solve Kuramoto dynamics) Solve (1) over t[0,T]t\in[0,T], with σ\sigma, ωi\omega_{i} and θi(0)\theta_{i}(0) from Step 1 and A(k)A^{(k)} from Step 2.

Step 4

(Test synchrony to potentially increase sparsity of A(k)A^{(k)}) If  maxtctTij|θ˙i(t)θ˙j(t)|ε\max_{t_{c}\leq t\leq T}\sum_{ij}|\dot{\theta}_{i}(t)-\dot{\theta}_{j}(t)|\leq\varepsilon ,  then choose ηk+1>ηk\eta_{k+1}>\eta_{k}, set k:=k+1k:=k+1 and go to Step 2.

Step 5

(Stopping) If k1k\geq 1, then declare “Algorithm successful,” return the adjacency matrix A=A(k1)A=A^{(k-1)}. Otherwise, declare “Algorithm unsuccessful.” Stop.

In Step 2 of the algorithm above, we define the entries Aij(k):=sgn(max{0,A~ijηk}){A}_{ij}^{(k)}:=\operatorname*{sgn}(\max\{0,\widetilde{A}_{ij}-\eta_{k}\}) of A(k)A^{(k)} in each iteration kk. Note that, with ηk>0\eta_{k}>0, some of the entries A~ijηk\widetilde{A}_{ij}-\eta_{k} may become negative, but after applying the max\max operator one gets Aij(k){0,1}{A}_{ij}^{(k)}\in\{0,1\}. In Step 4 of the algorithm, first, synchrony is tested with η0=0\eta_{0}=0, and then ηk\eta_{k} is increased in each iteration kk (i.e., ηk+1>ηk\eta_{k+1}>\eta_{k}) so as to increase the number of 0-entries in A(k+1)A^{(k+1)}, i.e., to obtain a sparser A(k+1)A^{(k+1)}.

Note that the prototype algorithm above is deemed unsuccessful (in Step 5) only in the case when k=0k=0 and the (binary) adjacency matrix A(0)A^{(0)} that is generated by Step 2 (with η0=0\eta_{0}=0) does not produce synchrony. In the numerical experiments we have performed (those reported in this paper and those not reported), the prototype algorithm (as well as Algorithm 1 given below) was always successful.

Step 3 and the synchrony verification in Step 4 of the prototype algorithm may be replaced by the verification of a connectivity property, as connected graphs do synchronize. One such property is the strict positivity of the second smallest eigenvalue of the Laplacian of the graph. Alternatively, one may also replace the same parts of the prototype algorithm by the requirement of a “suitably small” value of QLQ_{L} in Step 4, so as to have a “more desirable” connectivity property of the graph. Implementation of the latter criterion leads to a slightly different algorithm from the prototype as described below.

Algorithm 1
Step 1

(Initialization) Using some ωi𝒩(2,1/2)\omega_{i}\in{\cal N}(2,1/2), θi(0)𝒰[π/2,π/2]\theta_{i}(0)\in{\cal U}[-\pi/2,\pi/2], i=1,,Ni=1,\ldots,N, and integer M>0M>0, find a solution A~ij\widetilde{A}_{ij} to the relaxed model (19). Set η0>0\eta_{0}>0 small. Set k=0k=0.

Step 2

(Generate adjacency matrix A(k)A^{(k)}) Same as in Prototype Algorithm.

Step 3

(Potentially increase sparsity of A(k)A^{(k)} while keeping spectral ratio small) While   QLQ_{L}  is “small enough,” choose ηk+1>ηk\eta_{k+1}>\eta_{k}, set k:=k+1k:=k+1 and go to Step 2.

Step 4

(Stopping) Same as Step 5 of Prototype Algorithm.

Although QLQ_{L} has not been included in the optimization model directly, it is computed in Step 3 of Algorithm 1 (exogenously) for each new graph generated in Step 2. We have observed in the numerical experiments that, when k=0k=0, QLQ_{L} is in general small enough, however the matrix is not sufficiently sparse. As the sparsity of AA is increased, QLQ_{L} is in general also increased. Then, if QLQ_{L} is not “small enough” any longer, we stop and take the latest graph for which QLQ_{L} is acceptable. In other words, we increase kk in Algorithm 1 until sparsity and the value of QLQ_{L} are both acceptable.

Note that a solution of the relaxed model (19) depends on the random selection of the vectors ω\omega and θ(0)\theta(0). Moreover, even if we fix the choice for ω\omega and θ(0)\theta(0), each run of the AMPL-Ipopt suite with the same ω\omega and θ(0)\theta(0) is observed to generate a different (presumably locally optimal) solution. This is something expected because of the combinatorial nature of the underlying model (6). So, the solution obtained with Algorithm 1, or Prototype Algorithm, will be different with each A~\widetilde{A} found by solving model (19). Prototype Algorithm is designed to promote sparsity of the graph, and Algorithm 1 produces graphs with a relatively small spectral ratio. In the numerical experiments we describe in the next section, we have always used Algorithm 1. This is because we wanted to make sure our resulting graphs had a relatively small spectral ratio, and the matrices produced by it had a high level of sparsity.

Remark 1

In applications, there might be a need to find out a small set of the so-called vital nodes [33] that play a critical role, e.g., in propagating information and/or maintaining network connectivity. With this in mind, define a set of links in a connected network to be critical if their deletion produces a network which is either disconnected, or it has unacceptable connectivity properties such as a large QLQ_{L}. In Step 3 of Algorithm 1, we discard an adjacency matrix if it produces a disconnected graph, or if QLQ_{L} is too large. Therefore, Step 3 in Algorithm 1 can ultimately be also used to identify a set of critical links. However, this potential use of Algorithm 1 is not investigated here as it is outside the scope of the current paper.

5.2 Numerical experiments

In Subsections 5.2.1 and 5.2.2, Problems with 20 and 127 nodes, respectively, are solved. For simplicity in the computational modelling, we set tc:=T=20t_{c}:=T=20 in each of these two problems, where [0,T][0,T] is the domain, or time horizon, of θ\theta as in (1). This “terminal” choice of tct_{c} indeed produces the desired synchronization, way earlier. We also use a moderate grid size, M=10M=10, for the time horizon so as to make computations faster.

The reason we have chosen N=127N=127 is that this number is of the form N=2m1N=2^{m}-1, with m=7m=7, and therefore the 127-node graphs can be compared with those obtained in [22]. In the latter reference, graphs with good connectivity properties are constructed by combining a binary tree with an expander graph, as also briefly explained in Section 2.4. Our aim in choosing N=127N=127 is simply to be able to compare the connectivity properties of constructions such as those in [22] with those we obtain through Algorithm 1.

For each of the 20- and 127-node problems, we proceed as follows. First we obtain an acceptable (binary) adjacency matrix AA as an output of Algorithm 1. We randomly generate 30 sets of the vectors ω\omega and θ(0)\theta(0), whose components are taken from the distributions 𝒩(2,1/2){\cal N}(2,1/2) and 𝒰[π/2,π/2]{\cal U}[-\pi/2,\pi/2], respectively. For each of the 30 randomly generated sets of ω\omega and θ(0)\theta(0), we solve the Kuramoto system for a sequence of (fixed) values of σ\sigma. For this purpose we have used Matlab, Release 2019b. In order to solve (1), we implemented the solver ode15s, effective for stiff ODEs, with the absolute and relative tolerances of 10610^{-6}.

Now, with the discrete solution of the phase vector θ(t)\theta(t) for a fixed value of σ\sigma at hand, Equations (3)–(4) are used to obtain the corresponding r¯\bar{r}. For numerical integration we have used Simpson’s rule. Then using the 30 different sequences of r¯\bar{r} obtained for each of the 30 random data, the expected value E[r¯]\operatorname*{\mbox{E}}[\bar{r}] and the variance Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] of r¯\bar{r} are computed and plotted against the coupling strength σ\sigma. We also provide for each graph a histogram of the degrees of the nodes.

The variance plot can be used as a measure of robustness of the connectivity of the graph for ω\omega and θ(0)\theta(0) coming from distributions that were also chosen in Algorithm 1. The plot provides a quick picture of what coupling strength values a network will be resilient, i.e. will be insensitive, to possible changes in ω\omega and θ(0)\theta(0).

The adjacency matrices of the graphs mentioned in what follows can be accessed at the URL in [34].

Refer to caption

(a) E[r¯]\operatorname*{\mbox{E}}[\bar{r}] vs σ\sigma.

Refer to caption

(b) Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] vs σ\sigma.

Figure 1: N=20N=20: Mean and variance of the averaged order parameters of the graphs G1G_{1}G4G_{4} obtained by Algorithm 1.

5.2.1 20-node networks

We used Algorithm 1 for N=20N=20. We have chosen some representative instances that illustrate the effectiveness of our approach. Figure 1 corresponds to the graphs, denoted GiG_{i}, i=1,,4i=1,\ldots,4, that we obtained by using Algorithm 1.

In Figure 1(a) we display the plots of the expected values E[r¯]\operatorname*{\mbox{E}}[\bar{r}] of r¯\bar{r} vs the coupling strength σ\sigma, shown by solid curves which are colour-coded for each GiG_{i}, i=1,,4i=1,\ldots,4. The dashed curves, also colour-coded, represent the standard deviation from each of the E[r¯]\operatorname*{\mbox{E}}[\bar{r}] curves. Figure 1(b), on the other hand, depicts the variance Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] of r¯\bar{r}, colour-coded in the same way as in Figure 1(a). These plots indicate for which values of the coupling strength one can achieve connectivity, and thus synchrony, with a degree of robustness. Figure 2 displays the degree histogram of each graph.

G1G_{1}: Refer to caption G2G_{2}: Refer to caption

G3G_{3}: Refer to caption G4G_{4}: Refer to caption

Figure 2: N=20N=20: Degree distributions of the graphs G1G_{1}G4G_{4} obtained by Algorithm 1.

Figure 1(a) and 1(b) suggest that the graph G1G_{1}, with 25 edges and QL=16.5Q_{L}=16.5, would likely achieve a robust synchrony for σ1.2\sigma\geq 1.2, since, for these values, E[r¯]\operatorname*{\mbox{E}}[\bar{r}] is near 1.01.0 and Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] is relatively small. The graph G2G_{2} has just one more edge than G1G_{1}: With 26 edges, it has a better value of QLQ_{L}, namely QL=13Q_{L}=13. Figure 2(a) shows that its E[r¯]\operatorname*{\mbox{E}}[\bar{r}] curve comes near the value 1.01.0 more rapidly than that of G1G_{1}. Moreover, one can observe in Figure 2(b) that its Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] curve decays relatively more quickly. As can be seen in Figure 2, the degrees of the nodes of both graphs G1G_{1} and G2G_{2} range between 2 and 4, but a great majority of the nodes of G2G_{2} are of degree 3 or 4.

Figure 2 also depicts that G3G_{3} is a cubic graph, i.e. each of its nodes has degree 3. Cubic graphs are desirable in many practical applications because of their balanced workload sharing nature. We note that G3G_{3} has 30 edges but QL=11.3Q_{L}=11.3 is significantly smaller than those for G1G1 and G2G2. In the plots in Figure 1, we observe that E[r¯]\operatorname*{\mbox{E}}[\bar{r}] approaches 1.01.0 and Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] (after making the usual peak) dies down much more rapidly than the those curves for G1G1 and G2G2, pointing to far more robustness in its synchronization.

The graph G4G_{4} has 31 edges, and an even better value of QL=9.3Q_{L}=9.3. The node degrees in this case range from 3 to 4, which can also be regarded as sharing workloads more evenly compared to G1G_{1} and G2G_{2}. The “performance indicators” of the graph given in Figures 1 are even more impressive than before, in terms of the criteria discussed for the previous graphs. Robustness of its synchronization is more evident with its Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] being much smaller than those for G1G_{1}, G2G_{2} and G3G_{3} with larger (synchronizing) values of σ\sigma.

Refer to caption

(a) E[r¯]\operatorname*{\mbox{E}}[\bar{r}] vs σ\sigma.

Refer to caption

(b) Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] vs σ\sigma.

Figure 3: N=127N=127: Mean and variance of the averaged order parameters of the graphs G5G_{5}G10G_{10} obtained by Algorithm 1.

G5G_{5}: Refer to caption G6G_{6}: Refer to caption
G7G_{7}: Refer to caption G8G_{8}: Refer to caption
G9G_{9}: Refer to caption G10G_{10}: Refer to caption

Figure 4: N=127N=127: Degree distributions of the graphs G5G_{5}G10G_{10} obtained by Algorithm 1.

5.2.2 127-node networks

To understand the influence of the penalizing terms in the relaxed and discretized Problem (19), we implemented Algorithm 1 by solving first Problem (19) using only the 1\ell_{1}-objective and ignoring all other terms which penalize a wide distribution of the degrees. This produced a graph with 178 edges, QL300Q_{L}\approx 300, and degrees ranging from 1 to 13. This degree distribution is obviously not desirable.

Including the terms in (19) penalizing wider degree distributions across the graph nodes, i.e., setting c2=c3=100c_{2}=c_{3}=100, Algorithm 1 has produced more desirable graphs, e.g., the graphs G6G_{6} and G7G_{7}, the features E[r¯]\operatorname*{\mbox{E}}[\bar{r}], Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] and the degree histograms of which are depicted in Figures 3 and 4. The generated graphs are still sparse and have much better spectral ratios QLQ_{L}.

As mentioned in Section 2.4, we have also constructed a 127-node graph, namely G5G_{5}, using the tree–expander approach introduced and described in [22]. All of G5G_{5}, G6G_{6} and G7G_{7} show somewhat similar/close characteristics in terms of the E[r¯]\operatorname*{\mbox{E}}[\bar{r}] and Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] curves, although the graph featured in Figure 4 clearly has a much better degree distribution; its nodes have either degree 2 or degree 3. A closer examination of the E[r¯]\operatorname*{\mbox{E}}[\bar{r}] and Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] curves suggests that G7G_{7} will synchronize more robustly than G5G_{5} and G6G_{6}, since its E[r¯]\operatorname*{\mbox{E}}[\bar{r}] curve is above those of G5G_{5} and G6G_{6} and its Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] curve is below those of G5G_{5} and G6G_{6}, for σ1.4\sigma\geq 1.4.

The tree–expander construction of a graph that is introduced and described in [22] allows/requires choices amongst many options; so it is possible to get quite different graphs by this construction. A different graph constructed using an alternative sequence of choices is reported in [22], the features of which can also be seen in Figure 3, labelled G8G_{8}. This graph has better E[r¯]\operatorname*{\mbox{E}}[\bar{r}] and Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] curves, in that the E[r¯]\operatorname*{\mbox{E}}[\bar{r}] curve approaches 1.0 relatively more quickly and the peak of Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] is smaller as well as it dying down more quickly, compared with the previous three graphs discussed, namely G5G_{5}, G6G_{6} and G7G_{7}.

A desirable/useful characteristic of a graph would presumably be the following: (i) E[r¯]\operatorname*{\mbox{E}}[\bar{r}] is almost “S-shaped” and approaches 1.0 rapidly and (ii) Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] peaks early and dies down quickly. Such a characteristic seems to be attainable by adding more edges, via an investigation using our optimization approach in Algorithm 1. In what follows, we consider solutions which have more edges, namely the graphs with 180 and 240 edges, respectively.

One might then argue that by increasing the number of edges the sparsity might be compromised. This would really depend on a particular application and what (number of edges) is practically meant by a sparse graph. Define the sparsity of an undirected graph as the number of edges divided by (N(N1)/2)(N(N-1)/2). In the previously discussed 127-node graphs the sparsity is 98.03% (G5G_{5} and G8G_{8}, 158 edges), 97.93% (G6G_{6}, 166 edges), 97.91% (G7G_{7}, 167 edges). In the next two example graphs, the sparsity is 97.75% (180 edges) and 97.00% (240 edges), respectively. In all of the 127-node examples presented in this paper, the sparsity might be regarded to be quite close to one another (depending on a particular application, of course), which in this case is around 97–98%.

Figure 3 features the optimized graph G9G_{9} found by using Algorithm 1, which has 180 edges, instead of the 158–167 edges the previous graphs have. Although G9G_{9} has more edges than the previous graphs, this may not necessarily be regarded to be compromising sparsity (as discussed above). Now, both of the E[r¯]\operatorname*{\mbox{E}}[\bar{r}] and Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] curves have markedly better behaviour: The peak of Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] is shifted to the left and E[r¯]\operatorname*{\mbox{E}}[\bar{r}] approaches 1.0 markedly more rapidly. One might point to the rather wider range of degrees compared to the previous graphs; however, again depending on the application, this “slightly” wider range might be acceptable.

Figure 3 also features the optimized graph G10G_{10}, again found by using Algorithm 1, which has 240 edges. Although 240 edges sounds to be much bigger than 160 edges, the sparsity of G10G_{10} is worse only by 1 percentage point, which might again be acceptable in some practical situations. The behaviour of the graph shown by the E[r¯]\operatorname*{\mbox{E}}[\bar{r}] and Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] curves is rather impressive: The curve of E[r¯]\operatorname*{\mbox{E}}[\bar{r}] is almost S-shaped, settling very near 1.0 quite early. The peak of Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] has been pushed to the far left and Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] dies down to almost zero around σ=1.0\sigma=1.0 where the graph synchronize very well. These point to a desirable robust synchrony of the graph.

5.2.3 Computational effort

minimum average maximum
NN CPU time [s] CPU time [s] CPU time [s] Success rate
20 3 5 8 17/30 (57%)
30 10 18 54 23/30 (77%)
50 64 139 365 14/30 (47%)
127 2038 4419 7828 23/30 (77%)
Table 1: CPU times (in seconds) and success rates in solving model (19) with various graph sizes NN.

We have run AMPL–Ipopt suite to solve the optimization model (19), with randomly generated data, 30 times, with the graph sizes N=20N=20, 30, 50, and 127, on the computer the specifications of which we provide in Section 5.1. As can be seen from Table 1, the computational suite was successful in about slightly more than half the time it was run. As can also be seen, the elapsed computational time grew exponentially with the graph size, as expected. Despite the exponential computational complexity, (locally) optimized graphs can be obtained in a reasonable amount of time. For example, a locally optimal 127-node graph can be designed in about one to two hours’ time.

Once A~\widetilde{A} is found by solving model (19) and an adjacency matrix AA is constructed by Algorithm 1, the expected value E[r¯]\operatorname*{\mbox{E}}[\bar{r}] and the variance Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}] of r¯\bar{r} are computed and plotted against the coupling strength σ\sigma. This computation also incurs a comparable CPU time, in addition to the CPU times reported in Table 1. As we mention in Section 5.2, we use the Matlab solver ode15s, effective for stiff ODEs, to solve  (1), which can take a long time depending on the randomly generated data. In the case of N=20N=20, the CPU times for the graphs G1G_{1} and G4G_{4} are about 6.2 and 4.3 minutes, respectively. Given the fact that G4G_{4} is more robust (although less sparse) than G1G_{1}, it is fair to say that these computations are expected to take shorter time for G1G_{1} than those for G4G_{4}. We observe a similar situation for the N=127N=127 case: the CPU times for the graphs G5G_{5} and G10G_{10} are about 180 and 70 minutes, respectively.

6 Conclusion and Future Work

We have proposed an optimization algorithm (Algorithm 1) for designing sparse undirected networks, i.e., graphs, which synchronize well. The algorithms involve solution of a discretized relaxed mathematical model which maximize sparsity and minimize the spread of degrees (in some sense) at the same time. Synchrony of the network is posed as a constraint by means of the well-known Kuramoto system of coupled oscillators. By means of carefully chosen examples, from small to large scale, we illustrate the working of our algorithm and optimization model. The outcome is a method to generate sparse well balanced graphs with good synchronization from a random set of frequencies and initial phases. These graphs retain these properties when other frequency choices are used. In other words, in the resultant graphs there is not a strong correlation between the frequencies and degrees of the nodes (with a cubic generated in one case) underlying this insensitivity.

We have illustrated that our algorithm is successful in designing sparse synchronizing graphs with a relatively narrow distribution of degrees and small Laplacian eigenvalue ratio QLQ_{L}. By allowing more edges in the graph, but without sacrificing the sparsity much, we managed to obtain impressive synchronization behaviour in terms of the expected order parameter curve E[r¯]\operatorname*{\mbox{E}}[\bar{r}] and its variance Var[r¯]\operatorname*{\mbox{Var}}[\bar{r}]. These graphs synchronize in a robust manner in that for relatively small synchronizing values of r¯\bar{r} the variance of r¯\bar{r} under randomized data is small.

As pointed out in Remark 1, Algorithm 1 might be used as a valuable tool in identifying the so-called vital nodes and the critical set of links of the optimized sparse networks.

We have not explicitly incorporated the spectral ratio QLQ_{L} into our optimization model. However, the numerical experiments have demonstrated that the additional terms involving the minimum and maximum degrees in the objective function promote smaller QLQ_{L}. Having said this, a measure of QLQ_{L} might still be directly/explicitly included in a future model by using the Rayleigh quotient and the representation properties of the Laplacian using quadratic functions, such as the one in [35, Equation (14)]. Similar expressions that can be investigated can be found in [36, Remark 4.2].

Another interesting feature of a future optimization model would be constraints imposed on the degrees which the nodes of a graph should have. This would result in more targeted design of networks with certain desired load distributions.

Acknowledgments

The authors would like to acknowledge valuable discussions with Subhra Dey at the initiation of this project and with Richard Taylor during its later stages. This research was a collaboration under the auspices of the Modelling Complex Warfighting initiative between the Commonwealth of Australia (represented by the Defence Science and Technology Group) and the University of South Australia through a Defence Science Partnerships agreement.

References

  • [1] D. B. Forger, Biological Clocks, Rhythms, and Oscillations, (The MIT Press, Cambridge, Massachusetts, 2017)
  • [2] Y. Kuramoto, Chemical Oscillations, Waves, and Turbulence (Springer, Berlin, 1984)
  • [3] F. A. Rodrigues, T. K. D. M. Peron, P. Ji, and J. Kurths, The Kuramoto model in complex networks, Phys. Rep. 610, 1–98 (2016)
  • [4] A. H. Dekker and R Taylor, Synchronization properties of trees in the Kuramoto model, SIAM J. Appl. Dyn. Syst. 12, 596-–617 (2013)
  • [5] J. A. Rogge and D. Aeyels, Stability of phase locking in a ring of unidirectionally coupled oscillators, J. Phys. A: Math. Gen. 37, 11135–11148 (2004)
  • [6] J Ochab, and P. F. Gora, Synchronization of coupled oscillators in a local one-dimensional Kuramoto model, Acta Physica Polonica B Proceedings Supplement, 3, 453–462 (2010)
  • [7] T. Ichinomiya, Frequency synchronization in a random oscillator network, Phys. Rev. E 70(2), 026116 (2004)
  • [8] H. Hong, M. Y. Choi, and Beom Jun Kim, Synchronization on small-world networks, Phys. Rev. E 65, 026139 (2002)
  • [9] E. Oh, D.-S. Lee, B. Kahng, and D. Kim, Synchronization transition of heterogeneously coupled oscillators on scale-free networks, Phys. Rev. E 75, 011104 (2007)
  • [10] A. H. Dekker, Studying organisational topology with simple computational models, J. Artif. Soc. Simul. 10, 6 (2007)
  • [11] A. C. Kalloniatis, T. A. McLennan-Smith, D. O. Roberts, Modelling distributed decision-making in command and control using stochastic network synchronisation, European Journal of Operational Research 284, 588–603 (2020)
  • [12] M. Brede, Local versus global synchronization in networks of non-identical Kuramoto oscillators, Eur. Phys. J. B. 62, 87 (2008)
  • [13] T. Tanaka and T. Aoyagi, Optimal weighted networks of phase oscillators for synchronization, Phys. Rev. E. 78, 046210 (2008)
  • [14] D. Kelly and G. A. Gottwald, On the topology of synchrony optimized networks of a Kuramoto-model with non-identical oscillators, Chaos 21, 025110 (2011)
  • [15] T. Yanagita and A. S. Mikhailov, Design of oscillator networks with enhanced synchronization tolerance against noise, Phys. Rev. E. 85, 056206 (2012)
  • [16] M. Fazlyab, F. Doerfler and V. M. Preciado, Optimal network design for synchronization of coupled oscillators, Automatica, 84, 181 (2017)
  • [17] L. M. Pecora and T. L. Carroll, Master stability functions for synchronizes coupled systems, Phys. Rev. Lett. 80, 2109 (1998)
  • [18] M. Barahona and L. M. Pecora, Synchronization in small world systems, Phys. Rev. Lett. 89, 054101 (2002)
  • [19] L. Donetti, P. I. Hurtado, and M. A. Munoz, Entangled networks, synchronization, and optimal network topology, Phys. Rev. Lett. 95, 188701 (2005)
  • [20] L. Donetti, F. Neri, and M. A. Munoz, Optimal network topologies: expanders, cages, Ramanujan graphs, entangled networks and all that, J. Stat. Mech. P08007 (2006)
  • [21] E. Estrada, S. Gago, and G. Caporossi, Design of highly synchronizable and robust networks Automatica, 46, 1835 (2010)
  • [22] R. Taylor, A. Kalloniatis, K. Hoek, Organisational hierarchy constructions with easy Kuramoto synchronisation J.Phys.A: Math and Theor, doi.10.1088/1751-8121/ab69a3
  • [23] J. Nocedal and S. Wright, Numerical Optimization, (Springer, New York, 2006)
  • [24] G. Vossen and H. Maurer, On L1L^{1}-minimization in optimal control and applications to robotics. Optimal Control Applications and Methods. 27, 301–321 (2006)
  • [25] C. Y. Kaya, Inexact restoration for Runge-Kutta discretization of optimal control problems, SIAM J. Numer. Anal. 48(4), 1492–1517 (2010)
  • [26] C. Y. Kaya and J. M. Martínez, Euler discretization for inexact restoration and optimal control, J. Optim. Theory Appl. 134, 191–206 (2007)
  • [27] W. Alt, C. Y. Kaya, and C. Schneider, Dualization and discretization of linear-quadratic control problems with bang–bang solutions, EURO J Comput. Optim. 4, 47–77 (2016)
  • [28] N. Banihashemi and C. Y. Kaya, Inexact restoration for Euler discretization of box-constrained optimal control problems, J. Optim. Theory Appl. 156, 726–760 (2013)
  • [29] R. S. Burachik, C. Y. Kaya, and S. N. Majeed, A duality approach for solving control-constrained linear-quadratic optimal control problems, SIAM J. Control Optim. 52, 1771–1782 (2014).
  • [30] C. Y. Kaya and H. Maurer, A numerical method for nonconvex multi-objective optimal control problems, Comput. Optim. Appl. 57(3), 685–702 (2014)
  • [31] R. Fourer, D. M. Gay, and B. W. Kernighan, AMPL: A Modeling Language for Mathematical Programming, 2nd ed. (Brooks/Cole Publishing Company/Cengage Learning, 2003)
  • [32] A. Wächter and L. T. Biegler, On the implementation of a primal-dual interior point filter line search algorithm for large-scale nonlinear programming. Math. Progr. 106, 25–57 (2006)
  • [33] L. Lü, D. Chen, X. L. Ren, Q. M. Zhang, Y. C. Zhang, T. Zhou, Vital nodes identification in complex networks, Phys. Rep. 650, 1–63 (2016)
  • [34] R. S. Burachik, A. C. Kalloniatis, and C. Y. Kaya, Ancillary files for the preprint arXiv:2006.00428v1. URL: https://arxiv.org/src/2006.00428v1/anc. (2020)
  • [35] B. Mohar, Some applications of Laplace eigenvalues of graphs, in Graph Symmetry: Algebraic Methods and Applications, edited by G. Hahn and G. Sabidussi, NATO ASI Series C 497 (Kluwer, Dordrecht, Boston), 225–275 (1997)
  • [36] M. Fazlyab, F. Dörfler, and V. Preciado, Optimal network design for synchronization of coupled oscillators, Automatica, 84(C), 181–189 (2017)