arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb...

15
Global topological control for synchronized dynamics on networks. Giulia Cencetti, 1, 2 Franco Bagnoli, 3, 2 Giorgio Battistelli, 1 Luigi Chisci, 1 Francesca Di Patti, 3, 2 and Duccio Fanelli 3, 2 1 Dipartimento di Ingegneria dell’Informazione, Universit` a di Firenze, Via S. Marta 3, 50139 Florence, Italy 2 INFN Sezione di Firenze, via G. Sansone 1, 50019 Sesto Fiorentino, Italia 3 Universit` a degli Studi di Firenze, Dipartimento di Fisica e Astronomia and CSDC, via G. Sansone 1, 50019 Sesto Fiorentino, Italia A general scheme is proposed and tested to control the symmetry breaking instability of a ho- mogeneous solution of a spatially extended multispecies model, defined on a network. The inherent discreteness of the space makes it possible to act on the topology of the inter-nodes contacts to achieve the desired degree of stabilization, without altering the dynamical parameters of the model. Both symmetric and asymmetric couplings are considered. In this latter setting the web of con- tacts is assumed to be balanced, for the homogeneous equilibrium to exist. The performance of the proposed method are assessed, assuming the Complex Ginzburg-Landau equation as a reference model. In this case, the implemented control allows one to stabilize the synchronous limit cycle, hence time-dependent, uniform solution. A system of coupled real Ginzburg-Landau equations is also investigated to obtain the topological stabilization of a homogeneous and constant fixed point. PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys- tems constituted by many-body interacting entities [1, 2]. This is a widespread observation in nature which fertil- ized in a cross-disciplinary perspective to ideally embrace distinct realms of investigations. Convection instabili- ties in fluid dynamics, weak turbulences and defects are among the examples that testify on the inherent ability of physical systems to yield coherent dynamical behav- iors [3]. Insect swarms and fish schools exemplify the degree of spontaneous coordination that can be reached in ecological applications [4], while rhythms production and the brain functions refer to archetypical illustrations drawn from biology and life science in general [5–10]. The mathematics that underlies patterns formation focuses on the dynamical interplay between reaction and diffu- sion processes. Usually, reaction-diffusion models are de- fined on a regular lattice, either continuous or discrete [11]. In many cases of interest, it is however more nat- ural to schematize the system as a network, bearing a heterogeneous and complex structure [7, 12–15]. To ac- count for the hierarchical organization in multiple nested layers, networks of networks can be also considered [16– 24]. Imagine microscopic fluctuations to shake a station- ary stable, homogeneous equilibrium of the analyed, spa- tially extended, system. Under specific conditions, the imposed fluctuations get self-consistently enhanced by an intrinsic resonance mechanism, which is ultimately trig- gered by the spatial component of the dynamics: instabil- ities seeded by random perturbations are often patterns precursors and eventually materialize in beautiful patchy motifs for the concentration of the mutually interacting species [6, 25–34]. These are the byproduct of the cele- brated Turing instability, named after Alan Turing who first conceived the symmetry breaking mechanism from which the process originates [35]. For reaction-diffusion systems hosted on a graph, the instability typifies a char- acteristic asymptotic segregation in rich (resp. poor) nodes of activators (resp. inhibitors), a non homogeneous attractor that unavoidably emanates from the initially synchronous configuration [7, 36, 37]. In many cases of interest it is however important to op- pose the natural drive to pattern formation, by preserv- ing (or recovering) the synchronized state [38, 39]. Syn- chronization plays indeed a pivotal role in many branches of science: the efficient coordination of a multitude of events is often decisive to have a system operated as a unison orchestra. In an alternating current electric power grid, one needs to match the speed and frequency of any given generator to the other sources of the shared net- work [40–42]. In neuroscience, patterns of synchronous firings are promoted by dedicated neuronal feedbacks. Circadian rhythms are another example that certifies the ubiquitous tendency towards entrainable oscillations as displayed by a vast plethora of biological processes [5, 43]. In computer science, synchronization is customarily re- ferred to as consensus [44], a form of final agreement, sta- tionary or time dependent, which is reached by a crowd of interacting agents. Given these premises it is in general important to de- vise apt control strategies that enable one to stabilize, and possibly preserve, the synchronous regime. The con- trol is classically applied to the reactive component of the dynamics, and ultimately shape the local interaction between constitutive elements [45]. Global, mean field term can be also accommodated for so as to induce the sought behavior. When the dynamics flow on a network, topology matters and does play a prominent role in elic- iting the instability [14, 36]. This observation motivates in the search of alternative control protocols, which leave the reaction part unchanged, while acting on the under- arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016

Transcript of arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb...

Page 1: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

Global topological control for synchronized dynamics on networks.

Giulia Cencetti,1, 2 Franco Bagnoli,3, 2 Giorgio Battistelli,1 Luigi Chisci,1 Francesca Di Patti,3, 2 and Duccio Fanelli3, 2

1Dipartimento di Ingegneria dell’Informazione, Universita di Firenze, Via S. Marta 3, 50139 Florence, Italy2INFN Sezione di Firenze, via G. Sansone 1, 50019 Sesto Fiorentino, Italia

3Universita degli Studi di Firenze, Dipartimento di Fisica e Astronomia and CSDC,via G. Sansone 1, 50019 Sesto Fiorentino, Italia

A general scheme is proposed and tested to control the symmetry breaking instability of a ho-mogeneous solution of a spatially extended multispecies model, defined on a network. The inherentdiscreteness of the space makes it possible to act on the topology of the inter-nodes contacts toachieve the desired degree of stabilization, without altering the dynamical parameters of the model.Both symmetric and asymmetric couplings are considered. In this latter setting the web of con-tacts is assumed to be balanced, for the homogeneous equilibrium to exist. The performance ofthe proposed method are assessed, assuming the Complex Ginzburg-Landau equation as a referencemodel. In this case, the implemented control allows one to stabilize the synchronous limit cycle,hence time-dependent, uniform solution. A system of coupled real Ginzburg-Landau equations isalso investigated to obtain the topological stabilization of a homogeneous and constant fixed point.

PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt

INTRODUCTION

Self-organized collective dynamics may emerge in sys-tems constituted by many-body interacting entities [1, 2].This is a widespread observation in nature which fertil-ized in a cross-disciplinary perspective to ideally embracedistinct realms of investigations. Convection instabili-ties in fluid dynamics, weak turbulences and defects areamong the examples that testify on the inherent abilityof physical systems to yield coherent dynamical behav-iors [3]. Insect swarms and fish schools exemplify thedegree of spontaneous coordination that can be reachedin ecological applications [4], while rhythms productionand the brain functions refer to archetypical illustrationsdrawn from biology and life science in general [5–10]. Themathematics that underlies patterns formation focuseson the dynamical interplay between reaction and diffu-sion processes. Usually, reaction-diffusion models are de-fined on a regular lattice, either continuous or discrete[11]. In many cases of interest, it is however more nat-ural to schematize the system as a network, bearing aheterogeneous and complex structure [7, 12–15]. To ac-count for the hierarchical organization in multiple nestedlayers, networks of networks can be also considered [16–24]. Imagine microscopic fluctuations to shake a station-ary stable, homogeneous equilibrium of the analyed, spa-tially extended, system. Under specific conditions, theimposed fluctuations get self-consistently enhanced by anintrinsic resonance mechanism, which is ultimately trig-gered by the spatial component of the dynamics: instabil-ities seeded by random perturbations are often patternsprecursors and eventually materialize in beautiful patchymotifs for the concentration of the mutually interactingspecies [6, 25–34]. These are the byproduct of the cele-brated Turing instability, named after Alan Turing whofirst conceived the symmetry breaking mechanism from

which the process originates [35]. For reaction-diffusionsystems hosted on a graph, the instability typifies a char-acteristic asymptotic segregation in rich (resp. poor)nodes of activators (resp. inhibitors), a non homogeneousattractor that unavoidably emanates from the initiallysynchronous configuration [7, 36, 37].

In many cases of interest it is however important to op-pose the natural drive to pattern formation, by preserv-ing (or recovering) the synchronized state [38, 39]. Syn-chronization plays indeed a pivotal role in many branchesof science: the efficient coordination of a multitude ofevents is often decisive to have a system operated as aunison orchestra. In an alternating current electric powergrid, one needs to match the speed and frequency of anygiven generator to the other sources of the shared net-work [40–42]. In neuroscience, patterns of synchronousfirings are promoted by dedicated neuronal feedbacks.Circadian rhythms are another example that certifies theubiquitous tendency towards entrainable oscillations asdisplayed by a vast plethora of biological processes [5, 43].In computer science, synchronization is customarily re-ferred to as consensus [44], a form of final agreement, sta-tionary or time dependent, which is reached by a crowdof interacting agents.

Given these premises it is in general important to de-vise apt control strategies that enable one to stabilize,and possibly preserve, the synchronous regime. The con-trol is classically applied to the reactive component ofthe dynamics, and ultimately shape the local interactionbetween constitutive elements [45]. Global, mean fieldterm can be also accommodated for so as to induce thesought behavior. When the dynamics flow on a network,topology matters and does play a prominent role in elic-iting the instability [14, 36]. This observation motivatesin the search of alternative control protocols, which leavethe reaction part unchanged, while acting on the under-

arX

iv:1

608.

0157

2v1

[co

nd-m

at.s

tat-

mec

h] 4

Aug

201

6

Page 2: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

2

lying web of inter-nodes connections [46, 47]. The aim ofthis paper is to contribute along this direction, by propos-ing a viable topological approach which is fully rooted onfirst principles.

To illustrate the proposed method we shall first oper-ate in the framework of the Complex Ginzburg-LandauEquation (CGLE) [48] a prototypical model for nonlin-ear physics, whose applications range from superconduc-tivity, superfluidity and Bose-Einstein condensation toliquid crystals and strings in field theory. The CGLEassumes a population of oscillators, described in termsof their complex amplitude and mutually coupled viaa diffusive-like interaction. This is mathematically de-scribed in terms of a discrete Laplacian operator. TheCGLE admits a uniform fully synchronized solution, thespatially extended replica of the periodic orbit displayedby the system in its a-spatial version, provided the Lapla-cian is balanced (equal incoming and outgoing connec-tivity) [49]. Hereafter, we shall assume that the nodes ofthe network where oscillators lie are initially paired (andthe reaction parameters set) so as to make the systemunstable to externally injected, non homogeneous, per-turbations. The network of connections is then globallyreshaped (keeping the reaction parameters unchanged) toregain the stability of the synchronized, time dependent,solution. We will then move forward to considering a sys-tem of coupled (real) Ginzburg-Landau equations [50],which admits a stationary stable fixed point. Turing-likeinstabilities will be controlled, hence formally impeded,with a supervised intervention targeted to the net of in-terlaced couplings.

The paper is organized as follows. In the next sectionwe will introduce the CGLE and carry out a linear sta-bility analysis to delineate the conditions that make thespatially extended, homogenous limit cycle solution sta-ble. We will in particular elaborate on the remarkabledifferences that arise when the system involves a finiteand discrete collection of interlinked oscillators, as op-posed to the reference case where the population of ele-mentary constituents is made infinite and continuous. InSection III we will provide the mathematical basis for theproposed control method. The approach will be success-fully tested by operating with the CGLE and assuminga symmetric network. In Section IV, we will considera directed, although balanced, network of couplings andextend to this setting the analysis. Finally in Section Vwe will sum up and draw our conclusions. The appendixis devoted to discussing the stabilization of a constant ho-mogeneous fixed point, and proving, also in this respect,the adequacy of the proposed recipe.

COMPLEX GINZBURG-LANDAU EQUATION:LINEAR STABILITY ANALYSIS

Consider an ensemble made of N nonlinear oscillatorsand label with Wi their associated complex amplitude,where i = 1, ..., N . Each individual oscillator obeys to aCGLE which, as will be clarified in the following, com-bines linear and nonlinear (cubic) contributions. In ad-dition, we assume the oscillators to be mutually coupledvia a diffusive-like interaction which is mathematicallyexemplified via the discrete Laplacian operator. To fixideas consider as an example the simplified setting inwhich the oscillators are coupled to nearest neighborsonly, on a one dimensional lattice complemented with pe-riodic boundary conditions. The network of connectionsyields a binary adjacency matrix, termed A: the entryAij is set to one if nodes i and j are paired together, orzero if the link is absent. The associated discrete Lapla-cian operator ∆ results in a circulant matrix with threenon trivial entries per row, namely ∆ii = −2, ∆i,i+1 = 1and ∆i,i−1 = 1. In general, a complex web of inter-nodesconnections yields a heterogenous network, potentiallydirected (Aij 6= Aji) and weighted. In this case, let usdenote with kouti =

∑j Aji (resp. kini =

∑j Aij) the

outgoing (resp. ingoing) connectivity of generic node i.In our study we will deal with symmetric or balancedand directed networks, hence kouti = kini ≡ ki. The el-ements of the Laplacian matrix ∆ are therefore definedas ∆ij = Aij − kiδij , where δij stands for the usual Kro-necker δ.

The spatially extended CGLE can be hence cast in theform:

d

dtWj = Wj − (1 + ic2)|Wj |2Wj + (1 + ic1)K

∑k

∆jkWk

(1)where c1 and c2 are real parameters, which can be exter-nally assigned. The index j runs from 1 to N , the sizeof the inspected system. Here, K is a suitable parametersetting the coupling strength. We shall begin by consider-ing a symmetric adjacency matrix and postpone to a laterstage the case of a directed, though balanced, network ofcouplings. For pedagogical reasons, let us start by consid-ering a regular lattice, embedded on a Euclidean space ofarbitrary dimension. By performing the continuum limit,i.e assuming the linear distance between neighbor nodesto asymptotically vanish, one can formally replace thediscrete variable Wj(t) (j = 1, 2, · · · , N) with its con-tinuous counterpart W (r, t). Here, W (r, t) ∈ C and ridentifies the space location. Under these conditions, thediscrete operator ∆ transforms into ∇2, the standardLaplacian on a continuous support. For this reason, andwith a slight abuse of language, we shall often employthe adjective spatial to tipify the nature of the coupling,even when the network of oscillators is not necessarily

Page 3: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

3

bound to a physical space.As a preliminary remark we note that WLC(t) = e−ic2t

is a homogenous solution of the CGLE, both in its dis-crete or continuous, spatially extended, versions. Thislatter can be referred to as the limit cycle (LC) solu-tion, since it results from a uniform, fully synchronized,replica of the periodic orbit displayed by the system inits a-spatial (K = 0) version. In the remaining partof this section, we shall determine the stability of theLC solution. We will deal at first with the continuousversion of the model and re-derive for completeness theconditions for the onset of the so called Benjamin-Feirinstability [51, 52]. The peculiarities that stem from as-suming a discrete and heterogeneous web of symmetriccouplings will be also reviewed.

To assess the stability of the LC solution we introducea non homogeneous perturbation, both in phase and am-plitude:

W (r, t) = WLC(t)[1 + ρ(r, t)]eiθ(r,t). (2)

Linearizing around the LC (ρ(r, t) = 0, θ(r, t) = 0) onereadily obtains:

d

dt

[ρθ

]=

[−2 0−2c2 0

] [ρθ

]+K

[1 −c1c1 1

]∇2

[ρθ

]. (3)

To solve the above linear problem we perform a space-time Fourier transform:

ρ(r, t) =∫ ∫

dωdkeiωteik·rρkθ(r, t) =

∫ ∫dωdkeiωteik·rθk.

(4)

A straightforward calculation returns the following con-dition that should be matched as a necessary consistentrequirement for the linear problem to admit a meaningfulsolution:

det

λ+ 2 +Kk2 −Kc1k2

2c2 +Kc1k2 λ+Kk2

= 0 (5)

with λ = iω and k = |k|. The quantity λ hence as-sesses the linear growth rate associated to the k-th mode.Without losing generality we will hereafter set the cou-pling constant to unit (K = 1) and proceed with thecalculation to determine the root of the characteristicpolynomial with largest real part:

λ(k2) = −k2 − 1 +√−c21k4 − 2c1c2k2 + 1. (6)

The perturbation that shakes the homogenous andtime dependent solution WLC(t) gets exponentially mag-nified in the linear regime of the evolution provided the

real part of λ (the dispersion relation λRe) is positive.Notice that λ(0) = λRe(0) = 0, as expected, based ona obvious argument of internal coherence. ExpandingEq. (6) for small k returns λRe ' −(1 + c1c2)k2. Thestability of the synchronized LC solution is therefore lostwhen (1+c1c2) < 0, the standard condition for the onsetof the Benjamin-Feir instability.

We now turn to considering the case of a heterogenous,although symmetric, network of connections among os-cillators. To investigate the conditions that are to be metfor a symmetry breaking instability of the homogeneousLC solution to develop, we proceed in analogy with theabove and set Wj(t) = WLC(t)[1 + ρj(t)]e

iθj(t), with aclear meaning of the symbols. Plugging this latter ex-pression in the CGLE (1) and expanding to the first or-der in the perturbation amount, one obtains the obviousgeneralization of system (3):

d

dt

[ρjθj

]=

[−2 0−2c2 0

] [ρjθj

]+

[1 −c1c1 1

]∑k

∆jk

[ρkθk

].

(7)where we recall that K = 1. For regular lattices, theFourier transform is usually invoked to solve the systemof equations homologous to (7). This amounts to expand-ing the spatial perturbations on a set of planar waves,the eigenfunctions of the continuous Laplacian operator.When the system is instead defined on a network, ananalogous procedure can be employed. To this end, wedefine the eigenvalues and eigenvectors of the discreteLaplacian operator:

∑j

∆ijφ(α)j = Λ(α)φ

(α)i α = 1, ..., N (8)

When the network is undirected, the Laplacian oper-ator is symmetric. Therefore, the eigenvalues Λ(α) arereal and the eigenvectors φ(α) form an orthonormal ba-sis. This condition needs to be relaxed when dealing withthe more general setting of a directed graph, as we shalldiscuss in the second part of the paper[57]. The sym-metric Laplacian matrix ∆ has a single zero eigenvalueΛ(α=1) corresponding to the uniform eigenvector and allother eigenvalues are negative. The indices α are sortedso as to satisfy 0 = Λ(1) > Λ(2) ≥ · · · ≥ Λ(N).

The inhomogeneous perturbations ρj and θj can beexpanded as:

[ρjθj

]=

N∑α=1

[ρ(α)

θ(α)

]eλtφ

(α)j . (9)

By inserting Eq. (9) in Eq. (7) and making use of rela-tion (8) one eventually gets a condition formally equiv-alent to expression (5). As an important difference, the

Page 4: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

4

eigenvalues of the continuous Laplacian, −k2, are re-placed by the discrete (real and negative) quantities Λ(α),the eigenvalues of the discrete Laplacian. Insisting on theanalogy, it is of immediate evidence that the instabilityrises for a CGLE defined on a symmetric network when1 + c1c2 < 0, a dynamical condition identical to thatobtained when operating under the continuous, by defi-nition regular, viewpoint. The quantity Λ(α) constitutesthe analogue of the wavelength for a spatial pattern in asystem defined on a continuous regular lattice. It is thislatter quantity which determines the spatial characteris-tic of the emerging patterns, when the system is definedon a heterogeneous complex support.

In Fig. 1 the dispersion relation λRe is plotted versus

−Λ(α)Re , for a specific choice of c1 and c2, so that 1 +

c1c2 < 0. The solid line refers to the continuum setting

(−Λ(α)Re → k2,), while circles are obtained when operating

with the CGLE, hosted on a Watts-Strogatz network [53].As anticipated, the discrete collection of points whichdefines the dispersion relation when a symmetric, finiteand heterogenous network of coupling is accommodatedfor, follows the same profile which applies to the limitingcontinuum setting. In Fig. 2 the temporal evolutionof the system is displayed for a choice of the parameterswhich corresponds to the unstable dispersion relation ofFig. 1. After a given time the synchronized LC solutionis perturbed by insertion of an external source of nonhomogenous disturbance. This latter grows, as predictedby the linear stability analysis, and yields the irregularpatterns displayed for both WRe and |W|2.

Back to Fig. 1, it is however important to realize thatthe instability actually takes place only when at least one

eigenvalue −Λ(α)Re exists in the range where λRe is posi-

tive. If the ensemble of discrete modes, which ultimatelyreflects the topology of the imposed couplings, populatesthe portion of the dispersion relation with λRe < 0, noinstability can develop, even if 1 + c1c2 < 0. Stated dif-ferently the spectral gap ∆(Λ), the difference betweenthe moduli of the two largest eigenvalues of the Lapla-cian operator (∆(Λ) = |Λ(2)| − |Λ(1)| = |Λ(2)|) should belarger than −2(c1c2 + 1)/(1 + c21), the non trivial root ofEq. (6), for the instability to take place.

This observation has been exploited by Nakao inRef. [49] to propose a novel control strategy aimed at sup-pressing the Benjamin-Feir instability and thus preserv-ing the initial synchronized regime for a CGLE definedon a symmetric network support. Imagine to start withan unstable condition, which in turn implies to operatewith a suitable choice for both the reaction parametersand the network specificity. The key idea of Ref. [49] isto randomly rewire the network so as to make the secondeigenvalue progressively more negative. Random movesare accepted or rejected following a Metropolis scheme.The numerical procedure converges to a (globally) modi-fied network which has no eigenvalue in the range where

-$Re(,)

0 1 2 3 4 5 6 7

6R

e

-8

-6

-4

-2

0

FIG. 1: Continuous (solid line) and discrete (blue circles)dispersion relation. Here, c1 = −1.8, c2 = 1.6, K = 1. Thenetwork, composed of 100 nodes, is generated from the Watts-Strogatz method with rewiring probability 0.8.

λRe > 0. The control is topological since it only affectsthe couplings that links the oscillators, without actingon the dynamical parameters c1 and c2. In practice, thediscrete network-like system can be made stable for achoice of the parameter that would drive a Benjamin-Feirinstability in the continuum limit. Building on these in-triguing observations, we will here devise an analyticalapproach that enables us to implement a similar controlprotocol, without resorting to an iterative, numericallysupervised, rewiring. Importantly, the method that weshall introduce here can be successfully extended to thegeneral case where a directed network of connections isassumed to hold. The next section is devoted to dis-cussing the proposed method.

GLOBAL TOPOLOGICAL CONTROL

As outlined in the preceding section, here we aim atdeveloping an apposite control strategy which acts on theglobal network of connections, leaving unchanged the dy-namical parameters of the model. The method that weshall hereafter discuss takes inspiration from the semi-nal work of Nakao [49]. There it was shown that a nu-merically supervised rewiring of the inter-oscillators cou-plings can stabilize the CGLE, thus preserving the con-sensus state. Building on similar grounds, we will providein the following an analytical procedure to achieve thesought stabilization. The proposed method allows one toimmediately generate the controlled matrix of contacts,without involving any iterative scheme, thus freeing from

Page 5: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

5

Time10 20 30 40 50 60

Nod

e nu

mbe

r

20

40

60

80

100-1

-0.5

0

0.5

1

Time10 20 30 40 50 60

Nod

e nu

mbe

r

20

40

60

80

100

0.2

0.4

0.6

0.8

1

1.2

1.4

FIG. 2: Evolution of WRe (upper panel) and |W|2 (lowerpanel) versus time, assuming a uniform LC initial condition.At time τ1 = 15, a non homogeneous perturbation is insertedand the synchronized state is consequently disrupted. Here,c1 = −1.8, c2 = 1.6. The nonlinear oscillators are mutuallylinked via the Watts-Strogatz network, used in depicting thediscrete dispersion relation of Fig. 1.

concerns on the numerical convergence. Starting from acondition of instability, as displayed in Fig. 1, we wish tomodify the spectrum of the Laplacian operator so as toforce the finite and discrete collection of modes to popu-late the negative branch of the dispersion relation λRe.

As emphasized in the previous section, when the net-work is undirected the discrete dispersion relation super-poses to the continuum one (the solid line in Fig. 1). Theinstability localizes on a finite set of modes, those fallingon the positive bump of the curve λRe(k

2). Is it possibleto alter the network topology so as to make the (neg-

atively defined and real) eigenvalues larger in absolutevalue than −2(c1c2 + 1)/(1 + c21), the point where theparabola λRe(k

2) crosses the horizontal axis, so turningnegative? In a figurative sense, we want to slide the dis-crete points of Fig. 1 onto the curve, as beads on a cord,causing them to reach its negative branch. To answerthis question we make use of simple linear algebra tools.

Let us start by defining the N × N matrix Φ whosecolumns are the eigenvectors φ(1), · · · , φ(N) of the Lapla-cian operator ∆. Hence D = Φ−1∆Φ, where D is thediagonal matrix formed by the eigenvalues Λ(1), ...,Λ(N).We then calculate the minimal corrections δΛ(α) (α =1, · · · , N) that need to be imposed to shift the eigen-values Λ(α) on the stable side of the dispersion relation.The computed corrections are then organized in a diag-onal matrix D′, so that D′αα = δΛ(α), for α = 1, · · · , N .As a matter of facts, and to keep the formulation general,δΛ(α) ∈ C. For the case of a symmetric network that weare bound to explore within this Section, the quantitiesδΛ(α) are however real and negative. When the origi-nal Λ(α) falls in the region of stability, the correspondingcorrection δΛ(α) is set to zero.

The next step of the procedure is to perform the fol-lowing transformation ∆′ = ΦD′Φ−1 and define the con-trolled matrix ∆c = ∆ + ∆′. By construction the eigen-values of ∆c (with the only exception of the zero eigen-value, α = 1) are smaller than 2(c1c2 + 1)/(1 + c21).Assume for a moment that ∆c can be interpreted asa Laplacian operator. Hence, the obvious conclusion isthat we have generated a modified adjacency matrix Ac,hidden inside ∆c, which should engender a negative dis-persion relation (when employed in the CGLE, at fixedc1 and c2), thus preserving the stability of the synchro-nized configuration. Before concluding in this respect,one needs to prove that ∆c is indeed a Laplacian ma-trix, namely that (i) the entries of the matrix are realand that (ii) every column of the matrix sums up to zero,∑i(∆c)ij = 0 ∀j). In the following we set down to prove

the above properties (for both the symmetric and asym-metric settings), before turning to provide a numericalvalidation of the devised control procedure. As an im-portant complement, we will also show that symmetryand balancedness are perpetuated from ∆ to ∆c.

The elements of ∆c are real. The generic entriesof ∆′ can be written as:

∆′il =∑j

ΦijD′jj(Φ

−1)jl. (10)

In the symmetric case, the eigenvalues and their corre-sponding corrections are real. Also the eigenvectors havereal entries (vectors in RN ). Hence, the elements of D′,Φ, Φ−1 are real and, consequently, ∆′ij ∈ R. The undi-rected case is more complicated to handle. Let us bringinto evidence the real and imaginary parts of every ele-

Page 6: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

6

ment of Eq. (10). It is immediate to see that the imagi-nary part of ∆′il reads:

(∆′Im)il =

=∑j

(D′Im)jj [(ΦRe)ij(Φ−1Re)jl − (ΦIm)ij(Φ

−1Im)jl]+

+∑j

(D′Re)jj [(ΦRe)ij(Φ−1Im)jl + (ΦIm)ij(Φ

−1Re)jl].

(11)

To match condition (∆′Im)il = 0, both terms on theright hand side of Eq. (11) should be zero. To provethis fact let us begin by recalling that the eigenvalues ofa real asymmetric matrix either are real or are complexand come in conjugate pairs. Consider first the lattercase and label with α and β the generic pair of conjugateeigenvalues. By definition ∆φ(α) = Λ(α)φ(α). Takingthe complex conjugate yields ∆(φ(α))∗ = (Λ(α))∗(φ(α))∗

where (·)∗ stands for the complex conjugate and whereuse has been made of the condition ∆ = ∆∗. Recallingthat (Λ(α))∗ = Λ(β) we can immediately conclude that

(φ(α))∗ is an eigenvector of ∆ relative to the eigenvalue

Λ(β) and thus φ(β) = (φ(α))∗. Hence

(ΦRe)iα = (ΦRe)iβ(ΦIm)iα = −(ΦIm)iβ

(12)

for every i and (α, β). Consider now the equation

(φ−1)(α)∆ = Λ(α)(φ−1)(α) (13)

with (φ−1)(α) α-th row of Φ−1. Proceeding in analogywith the above, one eventually gets:

(Φ−1Re)αl = (Φ−1Re)βl(Φ−1Im)αl = −(Φ−1Im)βl.

(14)

Let us go back to Eq. (11). Performing the summationon j = α and j = β, using Eq. (12) and Eq. (14) andthe fact that the corrections D′αα and D′ββ are complex

conjugated as the original eigenvalues Λ(α) and Λ(β) are,we finally conclude that the terms of the sums in Eq. (11)cancel in pairs.

Consider now the case of a real eigenvalue Λ(k). Hence,by definition, (D′Im)kk = 0, since, in this case, the sta-bilization can be solely achieved by acting on the realpart of the eigenvalue (see next Section). To prove that(∆Im)il = 0 we need therefore to focus on the secondterm of Eq. (11), with j = k. Without losing generality

(up to a constant scaling factor) φ(k)Im = 0: the eigen-

vector associated to Λ(k) is hence real. The ik entriesof matrix Φ are indeed the elements of φ(k) and, forthis reason, (ΦIm)ik = 0 ∀i. To conclude the reason-ing and eventually prove that (∆′Im)il = 0, one needs toshow that (ΦRe)ik(Φ−1Im)kl = 0. This is in fact the case:

(φ−1)(k) is the left eigenvector of matrix ∆, relative tothe real eigenvalue Λ(k). Reasoning as above, one cantake (φ−1)(k) to be real and thus (Φ−1Im)kl = 0. Then,summing up, (∆′Im)il = 0 ∀i, l.

Each column of ∆c sums up to zero. Consider∑i

∆′il =∑ij

ΦijD′jj(Φ

−1)jl =

=∑j

(Φ−1)jlD′jj

∑i

Φij .(15)

Observe that

Λ(α)φ(α)i =

∑j

∆ijφ(α)j =

∑j

(Aij − δijkj)φ(α)j =

=∑j

Aijφ(α)j − kiφ(α)i =

=∑j

Aijφ(α)j −

∑l

Aliφ(α)i

(16)

Summing over i one obtains:

Λ(α)∑i

φ(α)i =

∑ij

Aijφ(α)j −

∑li

Aliφ(α)i = 0, (17)

thus the sum of elements of each Laplacian’s eigenvector(corresponding to an eigenvalue different from zero) isidentically equal to zero. This observation can be usedto conclude the proof. In fact, in Eq. (15),

∑i Φij is

equal to zero for all j associated to Λ(j) 6= 0. On theother hand, when j corresponds to Λ(j) = 0, D′jj = 0,as no correction is in this limiting case needed. Hence,∑i(∆′)il = 0 ∀l which in turn implies

∑i(∆c)il = 0 ∀l.

If ∆ is symmetric, then also ∆c is. It is enough toprove that the corrections ∆′ are bound to be symmetric.Indeed, when ∆ is symmetric, matrix Φ is orthogonal[58] (Φ−1 = ΦT ). Hence:

∆′il =∑j

ΦijD′jj(Φ

−1)jl =

=∑j

ΦijD′jjΦlj =

=∑j

ΦljD′jj(Φ

−1)ji = ∆′li

(18)

which ends the proof.

If ∆ is balanced, then also ∆c is. The Laplacianis termed balanced when, for every node, the ingoingconnectivity equals the outgoing one (kini = kouti ∀i), i.e.,if and only if each Laplacian row sums up to zero [59]. Itis then sufficient to prove

∑l ∆′il = 0. First of all, recall

that the columns of matrix Φ are the right eigenvectors

Page 7: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

7

of ∆, while the rows of Φ−1 are the left eigenvectors,namely:

∆Φ = ΦDΦ−1∆ = DΦ−1

(19)

By definition of Laplacian, the uniform vector 1 is theleft eigenvector of ∆ corresponding to Λ(1) = 0:

1T∆ = 0 ⇒∑i

∆ij = 0. (20)

If ∆ is balanced then also the right eigenvector corre-sponding to Λ(1) = 0 is equal to 1, hence:

∆1 = 0 ⇒∑j

∆ij = 0. (21)

By controlling the network of connections we modify theeigenvalues of the Laplacian operator, while keeping theeigenvectors unchanged. As a consequence, vectors 1 and1T are still solutions of the right and left eigenvalue prob-lems. Moreover, the modified Laplacian ∆c has still zeroamong the eigenvalues, thus:

∆c1 = 0 ⇒∑j(∆c)ij = 0

1T∆c = 0 ⇒∑i(∆c)ij = 0

(22)

which proves the claim.

In the remaining part of this section we will test theproposed control scheme assuming a symmetric matrixof inter-nodes couplings. In the next Section we willturn to discussing the more general case of a directed,although balanced, adjacency matrix. To demonstratethe adequacy of the technique, we will assume the set-ting depicted in Fig. 1: the parameters (c1, c2) and theunderlying network of contact are chosen so as to makethe system unstable to external non homogeneous per-turbations. By rewiring the network following the strat-egy outlined above we obtain the dispersion relation rep-resented in Fig. 3. The circles stand for the discretedispersion relation and populate the negative portion ofthe continuous curve: the instability has been hence re-moved, by solely acting on the topology of the graph.This latter was initially assumed of the binary type: theentries of the adjacency matrix are therefore a collectionof zeros and ones. The elements of the controlled ma-trix are still characterized by a bimodal distribution, asdisplayed in Fig. 4. Each element of the controlled ad-jacency matrix (Ac)ij takes a value close to the initialentry Aij . In practice, the control returns a local ad-justment of the weights, strong (resp. weak) couplingsbeing preserved under the imposed rewiring. Interest-ingly, negative coupling constants appear as a result ofthe continuous smoothing of the peak initially localizedin zero. Inhibitory interactions should be hence at playfor an effective stabilization of the dynamics.

-$Re(,)

0 1 2 3 4 5 6 7

6R

e

-8

-6

-4

-2

0

FIG. 3: Circles show the dispersion relation obtained for thecontrolled adjacency matrix. The solid line refers to the dis-persion relation in the continuum limit. Parameters are setas in Fig. 1.

To provide a numerical validation of our conclusion,we evolved for a transient the CGLE assuming the orig-inal, unstable and binary, adjacency matrix. When theimposed perturbation has grown to become significant,we instantaneously switched to the controlled Laplacian.As shown in Fig. 5, the perturbation fades progressivelyaway and the synchronous dynamics is eventually re-stored. The proposed control scheme was originally de-vised to contrast the onset of instability and, as such,targeted to the linear regime of the evolution. As demon-strated in Fig. 5, the method proves however effective instabilizing the system also at relatively large time, whennonlinearities are at play.

CONTROLLING THE INSTABILITY ONBALANCED DIRECTED NETWORKS

Let us now turn to considering the case of a CGLEdefined on a directed, heterogeneous although balanced(for each node the sum of incoming weights coincides withthe sum of outgoing weights), network. Before discussingthe application of the control technique introduced inthe previous section, we will review the conditions thatdetermine the emergence of instability.

When reaction-diffusion systems are placed on di-rected, hence asymmetric graphs, patterns can develop,even if they are formally impeded on a symmetric, contin-uum or discrete, spatial support. Directionality mattersand proves indeed fundamental in shaping the emergingpatterns. The conditions for the asymmetry driven in-

Page 8: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

8

FIG. 4: Main panel: distribution of the elements of the ad-jacency matrix before (large dark bins) and after (light smallbins) the control. Initially the distribution displays two peakslocalized in 0 and 1, reflecting the choice of a binary matrixof contacts. The controlled adjacency matrix is still bimodal,but the peaks are now smoothed out. Importantly, negativeconnections, pointing to inhibitory loops, should be accommo-dated for in the rewired weighed network. Inset: the elementsof the controlled adjacency matrix (Ac)ij are plotted vs. theoriginal adjacency matrix Aij and a clear correlation is dis-played. The control manifests as a rather local modificationof the weights, strong (resp. weak) couplings being preservedunder the imposed rewiring.

stability, reminiscent of a Turing like mechanism, for amulti-species reaction diffusion model evolving on a di-rected graph have been discussed in Ref. [14]. In this lat-ter case, the perturbation acts on a homogeneous fixedpoint, a time independent equilibrium for the reactiondynamics. In Ref. [54] the analysis has been extendedto the setting where the unperturbed homogeneous solu-tion is a LC and thus depends explicitly on time. In thefollowing, for the sake of consistency, we will go throughthe analysis of Ref. [54] to eventually obtain the condi-tions that instigate the topological instability of a time-dependent solution of the LC type.

By perturbing the WLC(t) as discussed in the firstSection, one eventually ends up with the self-consistentcondition:

det

[−2 + Λ(α) − λ −c1Λ(α)

−2c2 + c1Λ(α) Λ(α) − λ

]= 0 (23)

which is equivalent to det (Jα − λI2) = 0 with:

Jα =

(−2 + Λ(α) −c1Λ(α)

−2c2 + c1Λ(α) Λ(α)

).

Recall that for an asymmetric network, the Laplacian

eigenvalues Λ(α) are complex. Furthermore, Λ(α)Re < 0,

Time20 40 60 80 100 120

Nod

e nu

mbe

r

20

40

60

80

100

0.2

0.4

0.6

0.8

1

1.2

1.4

FIG. 5: |W|2 vs. time. The system assumes initially a binarymatrix of connections and it is unstable to external non ho-mogenous perturbation. At time τ1 = 10 the LC is perturbed,and the injected disturbances grow, yielding the expected lossof synchronization. At time τ2 = 60 the adjacency matrix isinstantaneously controlled, according to the scheme explainedin the main body of the paper. The perturbation is then re-absorbed and the consensus state recovered.

since the spectrum of the Laplacian matrix falls in the lefthalf of the complex plane, according to the Gerschgorintheorem [55]. Simple calculations yield:

(trJα)Re = −2 + 2Λ(α)Re

(trJα)Im = 2Λ(α)Im

(detJα)Re = −2Λ(α)Re + (Λ

(α)Re )2 − (Λ

(α)Im)2

− 2c1c2Λ(α)Re + c21

[(Λ

(α)Re )2 − (Λ

(α)Im)2

](detJα)Im = −2Λ

(α)Im + 2(1 + c21)Λ

(α)Re Λ

(α)Im

− 2c1c2Λ(α)ImΛ

(α)Im

(24)

with a clear meaning of the chosen notation.

From Eq. (23), one gets:

λ =1

2[(trJα)Re + γ] +

1

2[(trJα)Im + δ] i (25)

where:

γ =

√a+√a2 + b2

2(26)

Page 9: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

9

δ = sgn(b)

√−a+

√a2 + b2

2(27)

and:

a = [(trJα)Re]2 − [(trJα)Im]2 − 4(detJα)Re

b = 2(trJα)Re(trJα)Im − 4(detJα)Im.(28)

As discussed in Ref. [14, 54], diffusion driven instabili-ties arise also when tr(Jα)Re < 0, as opposed to what ithappens when the system evolves on a symmetric spatialsupport. In fact, λRe > 0 if:

|(trJα)Re| 6

√a+√a2 + b2

2(29)

a condition that can be met for tr(Jα)Re < 0, if the net-work of interactions is made directed and, consequently,an imaginary component of the Laplacian spectrum isaccommodated for. A straightforward, though lengthy,calculation allows one to derive the following compactcondition for the topology instability to occurr:

S2(Λ(α)Re ) ≤ S1(Λ

(α)Re )

[Λ(α)Im

]2(30)

where

S2(ΛαRe) = C2,4(Λ(α)Re )4 − C2,3(Λ

(α)Re )3 + C2,2(Λ

(α)Re )2

− C2,1Λ(α)Re

S1(ΛαRe) = C1,2(ΛαRe)2 − C1,1Λ

(α)Re + C1,0

(31)

with

C2,4 = 1 + c21

C2,3 = 4 + 2c1c2 + 2c21

C2,2 = 5 + 4c1c2 + c21

C2,1 = 2 + 2c1c2

C1,2 = c41 + c21

C1,1 = 2c31c2 + 2c21

C1,0 = c21(1 + c22) .

(32)

Notice that Eq. (30) reduces to S2(Λ(α)Re ) ≤ 0 when

dropping the imaginary components of Λ(α), or, equiva-lently, when assuming a symmetric network of couplings.

Expanding the solution for small Λ(α)Re , assumed as a con-

tinuum variable, one readily gets 1 + c1c2 < 0, i.e. thestandard condition for the Benjamin-Feir instability on asymmetric support.

-$Re(,)

0 1 2 3 4 5 6 7

6R

e

-8

-6

-4

-2

0

2

4

FIG. 6: The dispersion relation λRe as a function of −Λ(α)Re .

The solid line stands for the continuum dispersion relation.The (blue online) circles are obtained for the CGLE defined ona NW balanced network, with p = 0.27. The (red online) starsrepresent the dispersion relation obtained for the controlledmatrix of couplings. Here, c1 = 3 and c2 = 2.4224.

To gain insight into the above analysis, we generate adirected and balanced network, via a suitable modifica-tion of the Newman-Watts (NW) algorithm [56]. We be-gin from a substrate L-regular ring made of N nodes andadd, on average, NLp long-range directed links. Here,p ∈ [0, 1] is a probability that quantifies the amount ofintroduced long-range links. To keep the network bal-anced, the insertion of a long-range link stemming fromnode i is followed by a fixed number (3 is our arbitrarychoice) of additional links to form a loop that closes oni [14].

In Fig. 6 the dispersion relation λRe is plotted as a

function of −Λ(α)Re . The black solid line refers to the lim-

iting case of a symmetric (and continuum) support: thereaction parameters (c1, c2) are chosen so as to preventthe instability to develop since λRe < 0. The (blue on-line) circles refer instead to the directed case: the pointsabandon the solid curve and lift above zero, signaling atopology driven instability of the uniform LC solution.

In Fig. 7 the same situation is illustrated in the refer-

ence plane (Λ(α)Re ,Λ

(α)Im). Once the reaction parameters c1

and c2 have been assigned, one can calculate the coeffi-cients C1,q(q = 0, 1, 2) and C2,q(q = 0, ..., 4) via Eqs. (32).The inequality (30) allows us to draw the domain of in-stability, depicted as a shaded region in Fig. 7. Eacheigenvalue (blue circles) of the discrete Laplacian cor-

responds to a localized point in the plane (Λ(α)Re ,Λ

(α)Im).

Page 10: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

10

$Re(,)

-10 -8 -6 -4 -2 0 2

$Im(,

)

-5

0

5

FIG. 7: Eigenvalues in the complex plane (Λ(α)Re ,Λ

(α)Im). The

blue circles represent the eigenvalues of the initial Laplacian∆, while the red stars are the eigenvalues of ∆c. The shadedarea represents the instability region obtained from Eq. (30).The eigenvalues in this region correspond to the unstablemodes, characterized by λRe > 0, in Fig. 6.

The instability develops when at least one non-null eigen-value enters the shaded region. For an undirected graph,the points are distributed on the (horizontal) axis, thusoutside the region deputed to the instability. Whenthe graph turns asymmetric the imaginary componentof Λ(α) promotes an instability, which bears a direct im-print of the network topology. As usual, the instabilitywill eventually unfold complex patterns, in the nonlinearregime of the evolution.

Starting from this setting, and to restore the synchro-nization, one can rewire the network connections, accord-ing to the control procedure outlined in the precedingSection [60]. In this case, one needs to operate in the

complex plane (Λ(α)Re ,Λ

(α)Im), and act simultaneously on

the imaginary component of Λ(α), to force the eigenval-ues outside the region of instability [61] . In other wordsthe elements of the diagonal matrix D′ which encodes forthe imposed shifts are, in general, complex. For the caseat hands, the spectrum of the controlled Laplacian oper-ator is displayed in Figs. 6 and 7 with (red online) stars.The dispersion relation λRe is now consistently negative,reflecting the fact that stabilization has been enforcedinto the model. Similarly, stars populate the domain ofstability in Fig. 7 without invading the shaded portion ofthe plane.

As for the preceding case, the initial adjacency ma-trix is assumed binary. The elements of the controlled

FIG. 8: Main panel: distribution of the elements of the di-rected NW adjacency matrix before (large dark bins) and af-ter (light small bins) the control. The distribution displaysinitially two peaks at 0 and 1. The controlled adjacency ma-trix is still bimodal, but the peaks broaden. Importantly, thecoupling constant takes also negative values: inhibitory loopsshould be accommodated for in the rewired weigthed network.Inset: the elements of the controlled adjacency matrix (Ac)ijare plotted vs. the original adjacency matrix Aij . Also for thedirected case, a clear correlation between the two is observed.The control induces a rather local modification of the cou-plings, strong (resp. weak) couplings being preserved underthe imposed rewiring.

matrix still display a bimodal distribution (see Fig. 8):inhibitory coupling are at play as for the case of a sym-metric support.

To conclude this section we provide a numerical vali-dation of the implemented method. In Fig. 9 we initiallyevolve the perturbation assuming the unstable anddirected adjacency matrix. Then, when the perturbationhas evolved in a nonlinear quasi-wave, the Laplacian isinstantaneously mutated into its controlled counterpart.The perturbation damps and the system regains theinitial homogenous consensus state. We again remarkthat the control is also effective when acted far from thelinear regime of the evolution, when nonlinearities arepresumably playing a role.

As a final mandatory remark, we emphasize that thedeveloped control strategy holds in general, beyond theapplication to the CGLE here considered for purely peda-gogical reasons. Indeed, a formally identical scheme canbe applied to stabilizing homogenous time-independentfixed points, so preventing the classical Turing-like routeto patterns to eventually take place. This extension isdiscussed for completeness in the Appendix , by employ-ing an ad hoc multispecies framework which takes inspi-ration from the Ginzburg-Landau reference model.

Page 11: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

11

Time2 4 6 8 10 12 14 16 18 20

Nod

e nu

mbe

r

20

40

60

80

100

0.5

1

1.5

2

2.5

3

3.5

4

FIG. 9: |W| vs. time. The system assumes an initially bi-nary (directed and balanced) matrix of connections, as inFig. 6, and it is unstable to external non homogenous per-turbation. At time τ1 = 5 the LC is perturbed, and theinjected disturbances develop, yielding a loss of synchroniza-tion, as predicted by the linear stability analysis. At timeτ2 = 7 the adjacency matrix is instantaneously controlled, asfollows the devised scheme. The perturbation is consequentlyre-absorbed and the synchronized configuration recovered.

CONCLUSIONS

Patterns are ubiquitous in nature and arise in differ-ent contexts, ranging from chemistry to physics, passingthrough biology and life sciences. The paradigmatic ap-proach to pattern formation deals with a set of reaction-diffusion equations: an initial homogenous equilibrium,constant or time-dependent, can turn unstable via a sym-metry breaking instability, instigated by the external in-jection of a non homogenous disturbance. A non triv-ial interplay between reaction and diffusion terms, firstimagined by Alan Turing in his seminal paper on mor-phogenesis, is ultimately responsible for the growth ofthe imposed perturbation. This event takes place forspecific choices of the parameter setting and preludesthe outbreak of the fully developed patterns. When thereaction-diffusion system is hosted on a network support,the inherent discreteness and the enforced degree of im-posed asymmetry matter in determining the conditionsthat make the route to patterns possible. The vital rolewhich is played by the topology of the underlying net-works of contacts can be efficiently exploited to controlthe instability and so contrast the drive to pattern for-mation. In this paper we have elaborated along theselines by devising a suitable control strategy that enforces

stabilization, via a supervised redefinition of the inter-nodes couplings. The idea is to modify the spectrum ofthe Laplacian by altering the matrix of connections so asto confine the active modes outside the region of insta-bility. The method builds on the work of Nakao [49] whonumerically showed that an effective stabilization can beachieved by link-rewiring. As in Ref. [49], the ComplexGinzburg-Landau equation has been here assumed as areference model, to provide a probing test for the newlyproposed approach. In this case the control stabilizes thesynchronous limit cycle uniform solution. A multispeciessystem that couples together two real Ginzburg-Landauequations has been also considered, to assess the perfor-mance of the method in presence of a homogeneous sta-tionary stable fixed point. When the adjacency matrix issymmetric, the discrete points that constitute the unsta-ble portion of the dispersion relation are moved along thecontinuum parabola which embodies the characteristic ofstability in the idealized continuum limit. Conversely,when the connections are asymmetric, though balanced,different strategies can be implemented to achieve thesought stabilization. One can in general act on the imag-inary and real components of the spectrum of the Lapla-cian operator, integrating such independent moves as de-sired. Numerical checks confirmed the effectiveness of theproposed scheme. We recall however that a homogenous,fixed or time-dependent solution for the system to be con-trolled, is assumed to exist, which in turn implies dealingwith a specific choice for the inter-nodes couplings. Themethod can be however extended so as to account forthe stabilization of non homogenous equilibria, as we willreport in a separate paper.

Appendix - Stabilizing a homogenous fixed points:multispecies (real) Ginzburg-Landau equations

This Appendix aims at testing the proposed con-trol method for a multispecies model that undergoes aTuring-like instability. More specifically, we shall con-sider a reaction-diffusion model, hosted on a network,either symmetric or directed (and balanced). Diffusionis governed by the discrete Laplacian operator as intro-duced in the main body of the paper. The model admitsa stationary stable homogeneous fixed point. This lat-ter can turn unstable as follows the injection of a nonhomogeneous perturbation. In analogy with the discus-sion carried out for the CGLE, we will consider in a firstplace the usual setting, the instability resulting from thereactive component of the dynamics (on a symmetricsupport). Then we will turn to examine the case of atopology-driven instability (directed network). In idealcontinuity with the above, we shall assume a specificreaction-diffusion system, consisting of a pair of coupledreal Ginzburg-Landau equations, one for each interactingspecies, x and y. In formulae:

Page 12: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

12

-$Re(,)

0 10 20 30 40 50

6R

e

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

FIG. 10: The dispersion relation λRe as a function of −Λ(α)Re .

The solid line stands for the continuous case, while symbolsrefer to the discrete (symmetric) case. Circles (blue online)represent the initial setting, stars (red online) are instead ob-tained after the control has been applied. Here, a1 = −3,a2 = −1, b1 = −1, b2 = 4, γ = 0.1 and d = 0.01. Thenetwork employed is identical to that used in drawing Fig. 1.

{xi = f(xi, yi) +

∑j ∆ijxj

yi = g(xi, yi) + d∑j ∆ijyj

(33)

with

f(xi, yi) = γ[b1xi − (x2i + a1y2i )xi] (34)

and

g(xi, yi) = γ[b2yi − (x2i + a2y2i )yi], (35)

where xi and yi are real and positive and i = 1, ...N , withN denoting the number of nodes. The parameters a1, a2,b1, b2, γ, d are also assumed to be real. As usual ∆ standsfor the discrete Laplacian operator. System (33) admitsan homogenous fixed point (x∗,y∗) whose componentsrespectively read:

x∗i =a1b2 − a2b1a1 − a2

(36)

y∗i =b1 − b2a1 − a2

. (37)

For a proper choice of the involved parameters, the ho-mogenous fixed point is stable (the detailed study of thestability of the homogenous solution is here omitted).Starting from this setting we insert a non homogenousperturbation in the form xi = x∗ + ui, yi = y∗ + vi andlinearize Eq. (33) around (x∗,y∗). The obtained linearsystem can be solved by expanding the perturbation on

Time20 40 60 80 100 120

Nod

e nu

mbe

r

20

40

60

80

100

2

2.5

3

3.5

4

Time20 40 60 80 100 120

Nod

e nu

mbe

r

20

40

60

80

1000.5

1

1.5

2

2.5

3

FIG. 11: Evolution of x (upper panel) and y (lower panel)versus time. The system is unstable and a non homogenousperturbation, inserted at time τ1 = 10, evolves in patchy dis-tribution. At τ2 = 60 the control is turned on (thus thenetwork of connections rewired) and the perturbation dampsaway. Parameters are set as in Fig. 10.

.

eigenvectors basis, once the underlying network has beenspecified. This enables one to isolate the region of pa-rameters for which the diffusion driven instability candevelop.

In Fig. 10, we report the dispersion relation λRe vs.

−Λ(α)Re for a proper selection of the model parameters.

The solid line stands for the continuum case, while thecircles (blue online) are obtained assuming a (symmetric)

network of couplings. Notice that λRe(Λ(α)Re = 0) < 0, as

expected because of the imposed stability of the homo-geneous solution with respect to homogeneous perturba-tion. Moreover, the circles populate the positive bumpof the dispersion relation, thus signaling the instability.The stars (red online) show the dispersion relation com-

Page 13: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

13

-$Re(,)

0 2 4 6 8 10

6R

e

-0.4

-0.3

-0.2

-0.1

0

0.1

FIG. 12: The dispersion relation λRe as a function of −Λ(α)Re .

The solid line stands for the continuous dispersion relation.The (blue online) circle are obtained for the system defined ona NW balanced newtork, with p = 0.27. The (red online) starsrepresent the dispersion relation obtained for the controlledmatrix of couplings. Here, a1 = −7, a2 = −1, b1 = −1,b2 = 1.5, γ = 0.1 and d = 0.01.

$Re(,)

-6 -4 -2 0

$Im(,

)

-5

0

5

FIG. 13: Eigenvalues in the complex plane (Λ(α)Re ,Λ

(α)Im). The

blue circles represent the eigenvalues of the initial Laplacian∆, while the red stars refer to the eigenvalues of ∆c. Theshaded area represents the instability region obtained by usingthe procedure discussed in Ref. [14]. The eigenvalues in thisregion correspond to the unstable modes, characterized byλRe > 0, in Fig. 12. Parameters are set as in Fig. 12.

puted from the controlled Laplacian. The symbols arenow characterized by λRe < 0: the stability is hence re-covered. Direct simulations of system (33) as displayedin Fig. 11 confirm the adequacy of the proposed controlscheme.

We turn then to considering Eq. (33) placed on a di-rected and balanced network, generated according to the

Time200 400 600 800 1000 1200

Nod

e nu

mbe

r

20

40

60

80

1001.3

1.35

1.4

1.45

1.5

Time200 400 600 800 1000 1200

Nod

e nu

mbe

r

20

40

60

80

100

0.5

0.6

0.7

0.8

0.9

FIG. 14: Evolution of x (upper panel) and y (lower panel)versus time, for the situations depicted in Figs. 12 and 13. Theasymmetry in the couplings makes the system unstable andthe non homogeneous perturbation inserted at time τ1 = 100evolves in a quasi wave distribution. At τ2 = 500 the controlis turned on (thus the network of connections rewired) andthe perturbation fades away.

generalized NW recipe discussed in the paper. In Fig. 12we report the dispersion relation before (blue circles) andafter (red stars) the supervised rewiring of the adjacencymatrix. As it can be clearly appreciated by visual in-spection, the spectrum of the modified Laplacian matrixyields a stable dispersion relation. The same conclusioncan be drawn by analyzing the data reported in Fig. 13:the (red online) stars are scattered outside the region de-puted to the instability, obtained by using the conditionsreported in Ref. [14]. Numerical simulations displayed inFig. 14 confirm the predicted scenario.

Page 14: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

14

[1] Mark C Cross and Pierre C Hohenberg. Pattern forma-tion outside of equilibrium. Reviews of modern physics,65(3):851, 1993.

[2] Rebecca B Hoyle. Pattern formation: an introduction tomethods. Cambridge University Press, 2006.

[3] Philip G Drazin and William Hill Reid. Hydrodynamicstability. Cambridge university press, 2004.

[4] JD Murray. Mathematical biology ii: Spatial models andbiochemical applications, volume ii, 2003.

[5] Albert Goldbeter et al. Biochemical oscillations andcellular rhythms. Biochemical Oscillations and CellularRhythms, by Albert Goldbeter, Foreword by MJ Berridge,Cambridge, UK: Cambridge University Press, 1997, 1,1997.

[6] John Wyller, Patrick Blomquist, and Gaute T Einevoll.Turing instability and pattern formation in a two-population neuronal network model. Physica D: Non-linear Phenomena, 225(1):75–93, 2007.

[7] Hiroya Nakao and Alexander S Mikhailov. Turing pat-terns in network-organized activator-inhibitor systems.Nature Physics, 6(7):544–550, 2010.

[8] Malbor Asllani, Tommaso Biancalani, Duccio Fanelli,and Alan J McKane. The linear noise approximation forreaction-diffusion systems on networks. The EuropeanPhysical Journal B, 86(11):1–10, 2013.

[9] Malbor Asllani, Francesca Di Patti, and Duccio Fanelli.Stochastic turing patterns on a network. Physical ReviewE, 86(4):046105, 2012.

[10] SH Strogatz. Nonlinear Dynamics and Chaos: With Ap-plications to Physics, Biology, Chemistry and Engineer-ing, volume 272. Westview Press, 2001.

[11] James D Murray. Mathematical biology i: an introduc-tion, vol. 17 of interdisciplinary applied mathematics,2002.

[12] Hans G Othmer and LE Scriven. Instability and dynamicpattern in cellular networks. Journal of theoretical biol-ogy, 32(3):507–537, 1971.

[13] Hans G Othmer and LE Scriven. Non-linear aspects ofdynamic pattern in cellular networks. Journal of theoret-ical biology, 43(1):83–112, 1974.

[14] Malbor Asllani, Joseph D Challenger, Francesco SaverioPavone, Leonardo Sacconi, and Duccio Fanelli. The the-ory of pattern formation on directed networks. Naturecommunications, 5, 2014.

[15] Christopher N Angstmann, Isaac C Donnelly, and Bruce IHenry. Pattern formation on networks with reactions: Acontinuous-time random-walk approach. Physical ReviewE, 87(3):032804, 2013.

[16] Malbor Asllani, Daniel M. Busiello, Timoteo Carletti,Duccio Fanelli, and Gwendoline Planchon. Turing pat-terns in multiplex networks. Phys. Rev. E, 90:042814,Oct 2014.

[17] Peter J Mucha, Thomas Richardson, Kevin Macon, Ma-son A Porter, and Jukka-Pekka Onnela. Communitystructure in time-dependent, multiscale, and multiplexnetworks. science, 328(5980):876–878, 2010.

[18] Jesus Gomez-Gardenes, Irene Reinares, Alex Arenas, andLuis Mario Florıa. Evolution of cooperation in multiplexnetworks. Scientific reports, 2, 2012.

[19] Ginestra Bianconi. Statistical mechanics of multiplexnetworks: Entropy and overlap. Physical Review E,

87(6):062806, 2013.[20] Richard G Morris and Marc Barthelemy. Transport

on coupled spatial networks. Physical review letters,109(12):128703, 2012.

[21] Vincenzo Nicosia, Ginestra Bianconi, Vito Latora, andMarc Barthelemy. Growing multiplex networks. Physicalreview letters, 111(5):058701, 2013.

[22] Mikko Kivela, Alex Arenas, Marc Barthelemy, James PGleeson, Yamir Moreno, and Mason A Porter. Multilayernetworks. Journal of complex networks, 2(3):203–271,2014.

[23] Stefano Boccaletti, Ginestra Bianconi, Regino Criado,Charo I Del Genio, Jesus Gomez-Gardenes, Miguel Ro-mance, Irene Sendina-Nadal, Zhen Wang, and Massim-iliano Zanin. The structure and dynamics of multilayernetworks. Physics Reports, 544(1):1–122, 2014.

[24] Nikos E Kouvaris, Shigefumi Hata, and Albert Dıaz-Guilera. Pattern formation in multiplex networks. Sci-entific reports, 5, 2015.

[25] M Mimura and JD Murray. On a diffusive prey-predatormodel which exhibits patchiness. Journal of TheoreticalBiology, 75(3):249–262, 1978.

[26] John L Maron and Susan Harrison. Spatial pattern for-mation in an insect host-parasitoid system. Science,278(5343):1619–1621, 1997.

[27] Martin Baurmann, Thilo Gross, and Ulrike Feudel.Instabilities in spatially extended predator–prey sys-tems: spatio-temporal patterns in the neighborhood ofturing–hopf bifurcations. Journal of Theoretical Biology,245(2):220–229, 2007.

[28] Max Rietkerk and Johan Van de Koppel. Regular pat-tern formation in real ecosystems. Trends in ecology &evolution, 23(3):169–175, 2008.

[29] Hans Meinhardt and Alfred Gierer. Pattern formationby local self-activation and lateral inhibition. Bioessays,22(8):753–760, 2000.

[30] Matthew P Harris, Scott Williamson, John F Fallon,Hans Meinhardt, and Richard O Prum. Molecular ev-idence for an activator–inhibitor mechanism in develop-ment of embryonic feather branching. Proceedings of theNational Academy of Sciences of the United States ofAmerica, 102(33):11734–11739, 2005.

[31] Philip K Maini, Ruth E Baker, and Cheng-Ming Chuong.The turing model comes of molecular age. Science (NewYork, NY), 314(5804):1397, 2006.

[32] Stuart A Newman and Ramray Bhat. Activator-inhibitordynamics of vertebrate limb pattern formation. BirthDefects Research Part C: Embryo Today: Reviews,81(4):305–319, 2007.

[33] Takashi Miura and Kohei Shiota. Tgfβ2 acts as an acti-vator molecule in reaction-diffusion model and is involvedin cell sorting phenomenon in mouse limb micromass cul-ture. Developmental Dynamics, 217(3):241–249, 2000.

[34] Anatol M Zhabotinsky, Milos Dolnik, and Irving R Ep-stein. Pattern formation arising from wave instability ina simple reaction-diffusion system. The Journal of chem-ical physics, 103(23):10306–10314, 1995.

[35] Alan Mathison Turing. The chemical basis of morpho-genesis. Philosophical Transactions of the Royal Societyof London B: Biological Sciences, 237(641):37–72, 1952.

[36] Silvia Contemori, Francesca Di Patti, Duccio Fanelli,and Filippo Miele. Multiple-scale theory of topology-driven patterns on directed networks. Physical ReviewE, 93(3):032317, 2016.

Page 15: arXiv:1608.01572v1 [cond-mat.stat-mech] 4 Aug 2016 · PACS numbers: 89.75.-k 89.75.Kd 89.75.Fb 02.30.Yy 05.45.Xt INTRODUCTION Self-organized collective dynamics may emerge in sys-tems

15

[37] Julien Petit, Timoteo Carletti, Malbor Asllani, and Duc-cio Fanelli. Delay-induced turing-like waves for one-species reaction-diffusion model on a network. EPL (Eu-rophysics Letters), 111(5):58002, 2015.

[38] Arkady Pikovsky, Michael Rosenblum, and JurgenKurths. Synchronization: a universal concept in non-linear sciences, volume 12. Cambridge university press,2003.

[39] Renato E Mirollo and Steven H Strogatz. Synchroniza-tion of pulse-coupled biological oscillators. SIAM Journalon Applied Mathematics, 50(6):1645–1662, 1990.

[40] Florian Dorfler and Francesco Bullo. Synchronizationand transient stability in power networks and nonuni-form kuramoto oscillators. SIAM Journal on Control andOptimization, 50(3):1616–1642, 2012.

[41] Florian Dorfler, Michael Chertkov, and Francesco Bullo.Synchronization in complex oscillator networks andsmart grids. Proceedings of the National Academy of Sci-ences, 110(6):2005–2010, 2013.

[42] Daniel Jung and Stefan Kettemann. Long-range responsein ac electricity grids. Phys. Rev. E, 94:012307, Jul 2016.

[43] Yerali Carolina Gandica Lopez, Renaud Lambiotte, Tim-oteo Carletti, F. Sampaio Dos Aidos, and J. Carvalho.Circadian Patterns on Wikipedia Edits, volume 644 ofStudies in Computational Intelligence, pages 293–300.Springer, part vii edition, 3 2016.

[44] Reza Olfati-Saber, J Alex Fax, and Richard M Murray.Consensus and cooperation in networked multi-agent sys-tems. Proceedings of the IEEE, 95(1):215–233, 2007.

[45] Shigefumi Hata, Hiroya Nakao, and Alexander SMikhailov. Global feedback control of turing patternsin network-organized activator-inhibitor systems. EPL(Europhysics Letters), 98(6):64004, 2012.

[46] Malbor Asllani, Timoteo Carletti, and Duccio Fanelli.Tune the topology to create or destroy patterns. arXivpreprint arXiv:1604.07046, 2016.

[47] Per Sebastian Skardal and Alex Arenas. On control-ling networks of limit-cycle oscillators. arXiv preprintarXiv:1603.00842, 2016.

[48] Igor S Aranson and Lorenz Kramer. The world of thecomplex ginzburg-landau equation. Reviews of ModernPhysics, 74(1):99, 2002.

[49] Hiroya Nakao. Complex ginzburg-landau equation onnetworks and its non-uniform dynamics. The Euro-pean Physical Journal Special Topics, 223(12):2411–2421,2014.

[50] Tat Leung Yee, Alan Cheng Hou Tsang, Boris Malomed,and Kwok Wing Chow. Exact solutions for domain walls

in coupled complex ginzburg–landau equations. Journalof the Physical Society of Japan, 80(6):064001, 2011.

[51] T Brooke Benjamin and JE Feir. The disintegration ofwave trains on deep water part 1. theory. Journal of FluidMechanics, 27(03):417–430, 1967.

[52] JT Stuart and RC DiPrima. The eckhaus and benjamin-feir resonance mechanisms. In Proceedings of the RoyalSociety of London A: Mathematical, Physical and Engi-neering Sciences, volume 362, pages 27–41. The RoyalSociety, 1978.

[53] Duncan J Watts and Steven H Strogatz. Collective dy-namics of small-worldnetworks. nature, 393(6684):440–442, 1998.

[54] Francesca Di Patti, Duccio Fanelli, Filippo Miele, andTimoteo Carletti. Benjamin-feir instabilities on directednetworks. arXiv preprint arXiv:1607.06301, 2016.

[55] Howard E Bell. Gershgorin’s theorem and the zerosof polynomials. The American Mathematical Monthly,72(3):292–295, 1965.

[56] Mark EJ Newman and Duncan J Watts. Scaling andpercolation in the small-world network model. PhysicalReview E, 60(6):7332, 1999.

[57] A diagonalizable and connected Laplacian matrix is in-stead a minimal requirement to be satisfied by our analyt-ical treatment both in the symmetric and in the directedcase.

[58] The eigenvectors of a symmetric matrix generate an or-thonormal basis.

[59] The network is balanced if and only if∑j ∆ij = 0. Indeed

we have:∑j ∆ij =

∑j Aij − kouti = kini − kouti = 0.

So∑j ∆ij = 0 if and only if kini = kouti . Note that in

the above we assumed the Laplacian defined as ∆ij =Aij − kouti δij .

[60] Remember that the mathematical proofs provided in theprevious section hold in general, assuming a directed(balanced and diagonalizable) Laplacian

[61] The control can be implemented in different ways. Theeigenvalues can be moved for instance horizontally, byacting on their real component, vertically by modify-ing their imaginary part, or diagonally, by resorting to alinear combination of the two aforementioned strategies.Real eigenvalues lay on the horizontal axis: when fallingin the region of instability, they are transferred into thestable domain by sliding them horizontally, namely byimposing a real correction, the imaginary part proving,in this respect, useless.