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

Stability of Translating States for Self-propelled Swarms with Quadratic Potential

Irina Popovici Mathematics Department, United States Naval Academy, Annapolis, MD 21402, [email protected]
Abstract

The main result of this paper is proving the stability of translating states (flocking states) for the system of nn-coupled self-propelled agents governed by r¨k=(1|r˙k|2)r˙k1nj=1n(rkrj)\ddot{r}_{k}=(1-|\dot{r}_{k}|^{2})\dot{r}_{k}-\frac{1}{n}\sum_{j=1}^{n}(r_{k}-r_{j}), rk2r_{k}\in\mathbb{R}^{2}. A flocking state is a solution where all agents move with identical velocity. Numerical explorations have shown that for a large set of initial conditions, after some drift, the particles’ velocities align, and the distance between agents tends to zero. We prove that every solution starting near a translating state asymptotically approaches a translating state nearby, an asymptotic behavior exclusive to swarms in the plane. We quantify the rate of convergence for the directional drift, the mean field speed, and the oscillations in the direction normal to the motion. The latter decay at a rate of 1/t1/\sqrt{t}, mimicking the oscillations of some systems with almost periodic coefficients and cubic nonlinearities. We give sufficient conditions for that class of systems to have an asymptotically stable origin.

1 Introduction

The emergence of patterns in multi-agent dynamical systems has received a lot of attention since Turing featured it in the seminal paper [1]; interest was reignited by Kuramoto’s model of synchronization [2] and Cucker-Smale’s model of flocking [3]. Self organization has been observed in other natural swarms (school of fish, desert locusts, pacemaker cells , [23], [24], [25], [26]) and in artificial systems, sometimes as a desired effect (synchronization of power grids, lasers, [27], [28]), sometimes as detrimental (crowd-structure interactions in [29]).

In this paper we study the alignment in a swarm system with nn identical self-propelled agents having all-to-all coupling

r¨k=(1|r˙k|2)r˙k1nj=1n(rkrj), where rk(t)2\ddot{r}_{k}=(1-|\dot{r}_{k}|^{2})\dot{r}_{k}-\frac{1}{n}\sum_{j=1}^{n}(r_{k}-r_{j}),\;\mbox{ where }r_{k}(t)\in\mathbb{R}^{2} (1)
Notations 1.1.

We denote by R(t)=1nj=1nrj(t)\displaystyle R(t)=\frac{1}{n}\sum_{j=1}^{n}r_{j}(t) the position of the center of mass, and by V(t)=R˙=1nj=1nr˙j(t)V(t)=\dot{R}=\frac{1}{n}\sum_{j=1}^{n}\dot{r}_{j}(t) the velocity of the center of mass, a.k.a the mean velocity.

On occasion, it is convenient to think of the motion as taking place in the complex plane \mathbb{C} rather than 2;\mathbb{R}^{2}; we identify of a vector r=(x,y)r=(x,y) with the complex number 𝕣=x+iy.\mathbb{r}=x+iy. The system (1) becomes

𝕣¨k=(1|𝕣˙k|2)𝕣˙k(𝕣kR).\ddot{\mathbb{r}}_{k}=(1-|\dot{\mathbb{r}}_{k}|^{2})\dot{\mathbb{r}}_{k}-(\mathbb{r}_{k}-R). (2)

Note that given constants 𝕔0\mathbb{c}_{0}\in\mathbb{C} and θ0\theta_{0}\in\mathbb{R}, if (𝕣k(t))k=1..n\displaystyle\left(\mathbb{r}_{k}(t)\right)_{k=1..n} is a solution for (2), then the vectors eiθ0𝕣k(t)+𝕔0\displaystyle e^{i\theta_{0}}\mathbb{r}_{k}(t)+\mathbb{c}_{0} and the complex conjugate vectors 𝕣k(t)¯\overline{\mathbb{r}_{k}(t)} satisfy (2) as well. Thus the systems (1) and (2) are translation, rotation and reflection invariant.

It has been shown that the system’s coupling strength is enough to ensure that eventually all agents move in finite-diameter configurations ([4]), yet not too strong to lock the dynamics into a unique pattern of motion. We note that the self-propelling term (1|r˙k|2)r˙k(1-|\dot{r}_{k}|^{2})\dot{r}_{k} brings energy into the system when a particle’s speed is below one, and dissipates energy when the particle moves at high speed. This suggests that if agents are to align or synchronize, they would select patterns where all agents move at roughly unit speed. Numerical experiments show that for a large set of initial conditions the system reaches a limit state that has the center of mass either moving uniformly (at unit speed), or becoming stationary, the jargon for these configurations being flocking/translating and milling/rotating. (There are no generally accepted definitions for these terms; see definitions 1.3, 1.4 for how they are used in this paper). We want to emphasise that unlike the classical Cucker-Smale flocks [3] or Kuramoto oscillators [2], the center of mass for (1) does not have uniform motion:

V˙=1nj=1n(1|r˙j|2)r˙j.\dot{V}=\frac{1}{n}\sum_{j=1}^{n}(1-|\dot{r}_{j}|^{2})\dot{r}_{j}. (3)

Theoretical results in [4] and in this paper (Figures 1, 2 and Remark 4.7) show that as the center of mass moves toward its limit state it has a not-insignificant drift in its location or in its direction of motion. Let Θ(t)\Theta(t) denote the polar angle of the mean velocity, so V(t)=|V|eiΘ(t).\displaystyle V(t)=|V|e^{i\Theta(t)}.

Refer to caption
Figure 1: The position of two of the particles of a multi-agent swarm (1), shown as the solid black curves. Initially the center of mass has horizontal velocity and unit speed, as shown by the red vector. The flock’s eventual velocity VV_{\infty} has drifted away from the horizontal, as illustrated by the location of the center of mass (the blue square markers). The speed |V||V| approaches one at a rate of 1/t.1/t.
Refer to caption
Figure 2: The change in the direction of motion for a swarm with four agents. The oscillatory curve is the graph of m=dΘdtm=\frac{d\Theta}{dt} where Θ\Theta is the polar angle of the mean velocity V.V. The dotted line shows the mean value of mm over a time window of size 2π.2\pi. The flock’s eventual velocity VV_{\infty} has a convergent polar angle, with an overall counterclockwise drift equal to ΘΘ0=0m𝑑t>0.\Theta_{\infty}-\Theta_{0}=\int_{0}^{\infty}mdt>0.

The center of mass is very sensitive to perturbations, in the sense that small changes in initial conditions could result in large scale drift of the center of mass just on account of changing the direction of motion. In order to control the evolution of small perturbations it may be more expedient to focus on the velocities VV and vkv_{k} and to recover the information about RR and rkr_{k} via integration. The velocity-acceleration system for vk=r˙kv_{k}=\dot{r}_{k}, obtained by differentiating (1) is:

v¨k=(1|vk|2)v˙k2(vkTv˙k)vk(vkV)\ddot{v}_{k}=(1-|v_{k}|^{2})\dot{v}_{k}-2(v_{k}^{T}\dot{v}_{k})v_{k}-(v_{k}-V) (4)

where the superscript TT denotes transposition.

1.1 Special Configurations

Second order models of self-propelled agents with all-to-all gradient coupling have been extensively studied, notably by researchers in Prof. Bertozzi’s group ([6], for continuous models, and [7] for agent-based models). Let K:[0,)K:[0,\infty)\to\mathbb{R} be a continuous scalar function, differentiable on (0,).(0,\infty). Define U:dU:\mathbb{R}^{d}\to\mathbb{R} as U(x)=K(|x|).U(x)=K(|x|). With the possible exception of the origin, the potential UU is differentiable with U(x)=K(|x|)x|x|.\nabla U(x)=K^{\prime}(|x|)\frac{x}{|x|}. Consider the system with identical self propelling terms and rotationally symmetric potential:

r¨k=(αβ|r˙k|2)r˙k1njkU(xkxj)\ddot{r}_{k}=(\alpha-\beta|\dot{r}_{k}|^{2})\dot{r}_{k}-\frac{1}{n}\sum_{j\neq k}\nabla U(x_{k}-x_{j}) (5)

where α,β\alpha,\beta are positive constants.

The system (1) is in the class of (5) with K(r)=12r2K(r)=\frac{1}{2}r^{2} and α=β=1.\alpha=\beta=1. Other well known classes of radial potentials are Morse potentials where KK is the difference between two exponential decays (see [8], [9]), and power law potentials where K(r)=caracbrb,K(r)=c_{a}r^{a}-c_{b}r^{b}, ( [18], Section 8; [19]; [4], 3.3-3.5). Given the symmetry of (5) it is natural to expect that all emergent patterns feature rotational symmetries; in that spirit, with the exception of [4] and [5], most theoretical results address particles that are uniformly distributed around their center of mass and their similarly restricted perturbations, or have assumptions about the degeneracies of the system that for (1) only hold for n2n\leq 2, [13]. We emphasize that we make no uniform distribution assumption in this paper, which explains the sharp distinction between the rates of convergence to limit configuration: exponential rates obtained by authors working exclusively in the subset with rotational symmetries, versus 1t\frac{1}{\sqrt{t}} in this paper (and in [5]). In fact (1) has very few configurations with symmetries:

Proposition 1.2.

If among the nn agents of the system (1) there exist p3p\geq 3 agents (assumed to be the first pp) whose motion is symmetrically distributed about the center RR of the swarm, i.e.

𝕣k(t)=R(t)+e2πi(k1)p(𝕣1(t)R(t)) for k=1..p,\mathbb{r}_{k}(t)=R(t)+e^{2\pi i\frac{(k-1)}{p}}(\mathbb{r}_{1}(t)-R(t))\;\mbox{ for }\;k=1..p,

then either 𝕣1(t)=𝕣p(t)\mathbb{r}_{1}(t)=\dots\mathbb{r}_{p}(t) for all tt\in\mathbb{R}, or R(t)=R(0)R(t)=R(0) for all t.t\in\mathbb{R}.

Proof.

To improve the flow of the paper, the proof is presented in the Appendix. ∎

It was shown in [5] that the only possible configurations of (1) with a stationary center of mass come from partitioning the agents {1,n}\{1,\dots n\} into subsets ,,\mathcal{F},\mathcal{L},\mathcal{R} where the agents in \mathcal{F} have fixed positions coinciding with the center R(0)R(0) at all times; each agent in \mathcal{L} oscillates on a segment centered at R(0)R(0), satisfying the scalar Lienard equation a¨=(1a˙2)a˙a\ddot{a}=(1-\dot{a}^{2})\dot{a}-a (which has an attracting cycle); each agent in \mathcal{R} is asymptotic to a rotation e±iteiθ,k.e^{\pm it}e^{i\theta_{\infty,k}}.

Definition 1.3.

A solution (rk(t))k=1..n\displaystyle\left(r_{k}(t)\right)_{k=1..n} for (1) is called a rotating state if there exist angles θ0,k\theta_{0,k} with k=1neiθ0,k=0\displaystyle\sum_{k=1}^{n}e^{i\theta_{0,k}}=0 and a constant cc\in\mathbb{C} such that rk(t)=eiθ0,keit+cr_{k}(t)=e^{i\theta_{0,k}}e^{it}+c for all tt and all kk (a counterclockwise rotating state), or rk(t)=eiθ0,keit+cr_{k}(t)=e^{i\theta_{0,k}}e^{-it}+c for all tt and all kk (a clockwise rotating state).

Some authors, but not all, use the term mill for a rotating state; most authors define a mill as a rotating state whose angles θ0,k\theta_{0,k} are uniformly distributed in [0,2π][0,2\pi], a.k.a having rotational symmetry. All the rotating states of (1) are stable, per [5]. Configurations with rotational symmetries for this system and for other radial potentials have been extensively studied: see [16], [8], [9], [18] for some early theoretical and experimental results.

Definition 1.4.

A solution (rk(t))k=1..n\displaystyle\left(r_{k}(t)\right)_{k=1..n} for (1) is called a flocking state if r˙k1(t)=r˙k2(t)\dot{r}_{k_{1}}(t)=\dot{r}_{k_{2}}(t) for all k1,k2k_{1},k_{2} in 1..n1..n. (Necessarily the common velocity is that of the center of mass.)

Note that for a flocking state r¨k=V˙.\ddot{r}_{k}=\dot{V}. If we subtract the equation for r¨k1\ddot{r}_{k_{1}} from that of r¨k2\ddot{r}_{k_{2}} we get that rk1=rk2r_{k_{1}}=r_{k_{2}} for all k1,k2k_{1},k_{2} meaning that all particles have collapsed onto the center of mass, moving together according to

V˙=(1|V|2)V.\dot{V}=(1-|V|^{2})V.

Unless V(t)=0V(t)=0 for all tt (the trivial solution), the velocity of the flocking solutions maintains the initial direction of V(0)=|V(0)|eiΘ0V(0)=|V(0)|e^{i\Theta_{0}} with speed approaching unit speed at a rate of e2te^{-2t} (since |V|2|V|^{2} solves the logistic equation ddt(|V|2)=2(1|V|2)|V|2).\displaystyle\frac{d}{dt}(|V|^{2})=2(1-|V|^{2})|V|^{2}\;). A nontrivial flocking state has velocity approaching eiΘ0e^{i\Theta_{0}} exponentially fast. Moreover, if cc_{\infty} denotes the absolutely convergent integral c=0[V(τ)eiΘ0]𝑑τ\displaystyle c_{\infty}=\int_{0}^{\infty}\left[V(\tau)-e^{i\Theta_{0}}\right]d\tau then the flocking state’s center of mass has nearly rectilinear motion: limt[R(t)(c+teiΘ0)]=0.\lim_{t\to\infty}\left[R(t)-(c_{\infty}+te^{i\Theta_{0}})\right]=0. The limit holds since R(t)teiΘ0c=0t[V(τ)eiΘ0]𝑑τ0[V(τ)eiΘ0]𝑑τ=t[V(τ)eiΘ0]𝑑τ.R(t)-te^{i\Theta_{0}}-c_{\infty}=\int_{0}^{t}\left[V(\tau)-e^{i\Theta_{0}}\right]d\tau-\int_{0}^{\infty}\left[V(\tau)-e^{i\Theta_{0}}\right]d\tau=-\int_{t}^{\infty}\left[V(\tau)-e^{i\Theta_{0}}\right]d\tau.

Definition 1.5.

A solution (rk(t))k=1..n\displaystyle\left(r_{k}(t)\right)_{k=1..n} for (1) is called a translating state if there exists Θ0[0,2π]\Theta_{0}\in[0,2\pi] such that r˙k(t)=eiΘ0\dot{r}_{k}(t)=e^{i\Theta_{0}} for all k=1..n.k=1..n. (Necessarily, for a translating state there exists c0c_{0}\in\mathbb{C} such that rk(t)=c0+teiΘ0r_{k}(t)=c_{0}+te^{i\Theta_{0}} for all k.k.)

The main result of this paper is proving that the translating states are stable, in the sense that if the particles in (1) start close to each other, with velocities near some unit vector (cosΘ0,sinΘ0)(\cos\Theta_{0},\sin\Theta_{0}) then the particles remain close to each other, with velocities that are asymptotic to a common unit vector, whose direction does not drift far from Θ0,\Theta_{0}, as stated in Theorem 1.6. We also quantify the rate of convergence to the limiting configuration (Corollary 4.4.1), and we show that it is at a rate of 1t,\displaystyle\frac{1}{\sqrt{t}}, much slower than the conjectured exponential rate.

Theorem 1.6.

Consider the system (1) with initial conditions rk(0)=rk,0r_{k}(0)=r_{k,0} and r˙k(0)=r˙k,0.\dot{r}_{k}(0)=\dot{r}_{k,0}. For any ϵ>0\epsilon>0 there exists δ>0\delta>0 such that if rk,0r_{k,0} and r˙k,0\dot{r}_{k,0} satisfy: |rk1,0rk2,0|<δ|r_{k_{1},0}-r_{k_{2},0}|<\delta for all k1,k2k_{1},k_{2} in 1..n1..n ; 1δ<|V0|<1+δ1-\delta<|V_{0}|<1+\delta; |r˙k,0V0|<δ|\dot{r}_{k,0}-V_{0}|<\delta for all kk in 1..n1..n, then for t>0t>0 :

|rk1(t)rk2(t)|<ϵ for all k1,k2 in 1..n1ϵ<|V(t)|<1+ϵ|r˙k(t)V0|<ϵ for all k in 1..n\begin{array}[]{l}|r_{k_{1}}(t)-r_{k_{2}}(t)|<\epsilon\;\mbox{ for all }k_{1},k_{2}\;\mbox{ in }1..n\\ 1-\epsilon<|V(t)|<1+\epsilon\\ |\dot{r}_{k}(t)-V_{0}|<\epsilon\;\mbox{ for all }k\mbox{ in }1..n\end{array} (6)

Moreover,

limt|rk1(t)rk2(t)|=0 for all k1,k2 in 1..n\lim_{t\to\infty}|r_{k_{1}}(t)-r_{k_{2}}(t)|=0\;\mbox{ for all }k_{1},k_{2}\mbox{ in }1..n (7)

and there exists Θ\Theta_{\infty} in [0,2π][0,2\pi] with |V(0)(cosΘ,sinΘ)|<ϵ|V(0)-(\cos\Theta_{\infty},\sin\Theta_{\infty})|<\epsilon such that

limtr˙k(t)=limtV(t)=(cosΘ,sinΘ).\lim_{t\to\infty}\dot{r}_{k}(t)=\lim_{t\to\infty}V(t)=(\cos\Theta_{\infty},\sin\Theta_{\infty}). (8)

The proof of Theorem 1.6 is completed in three parts, corresponding to three dimension reductions of the system: form 4n4n to 4n34n-3 in Section 2, to 2n2n in Section 3, and to nn in Section 4, with the stability of the translating states for (1) being equivalent to that of the origin for the reduced systems.

The dynamical system produced in Section 2 evolves in a 4n34n-3 dimensional invariant subspace for a system taking the form w˙=Aw+G(w)\dot{w}=Aw+G(w) where AA is the Jacobian matrix and GG is a nonlinear function whose Maclaurin expansion has terms or order 3 or higher. Counted with multiplicity, the eigenvalues of AA consist of nn pairs of purely imaginary roots ±i\pm i, and 2n12n-1 eigenvalues with negative real parts. The stability of the origin for systems whose Jacobian matrices have qq pairs of purely imaginary roots (±iω1,,±iωq)(\pm i\omega_{1},\dots,\pm i\omega_{q}) and d2qd-2q roots with negative real part has been extensively investigated; the problem is commonly known as the critical case with purely imaginary roots. Most authors apply elaborate iterative transformations from normal form theory to obtain systems for which stability criteria such as Molchanov and Lyapunov’s second method can be employed. These transformations require that the frequencies ωk\omega_{k} satisfy non-resonance conditions which are violated in our setting,([33], [34]), or alternatively, they assume the a priory existence of some quadratic Lyapunov functions ( [35]), rendering their methods inapplicable here despite concluding similar decay rate of 1/t1/\sqrt{t} towards the origin. Our approach draws inspiration from a class of time-dependent systems with cubic nonlinearities instead:

x˙k=bk(t)xk3+ck(t)xk1nl=1ndl(t)xl2+Rk(x,t), for k=1n,\dot{x}_{k}=b_{k}(t)x_{k}^{3}+c_{k}(t)x_{k}\frac{1}{n}\sum\limits_{l=1}^{n}d_{l}(t)x_{l}^{2}+R_{k}(x,t),\;\mbox{ for }k=1\dots n, (9)

where x=(x1,xn)Tx=(x_{1},\dots x_{n})^{T} is vector-valued from \mathbb{R} to n\mathbb{R}^{n}, the coefficient functions bk,ck,dk:b_{k},c_{k},d_{k}:\mathbb{R}\to\mathbb{R} are continuous and the remainder functions RkR_{k} defined on n×\mathbb{R}^{n}\times\mathbb{R} satisfy the o3\textsl{o}_{3} condition: there exists a function ϵ:[0,)\epsilon:[0,\infty)\to\mathbb{R} with limr>0ϵ(r)r3=0\displaystyle\lim_{r->0}\frac{\epsilon(r)}{r^{3}}=0 such that |Rk(x,t)|ϵ(|x|)|R_{k}(x,t)|\leq\epsilon(|x|) for all (x,t)n×.(x,t)\in\mathbb{R}^{n}\times\mathbb{R}. In Section 3.3 we provide easily verifiable conditions on the coefficient functions that sufficiently ensure the asymptotic stability of the solution x=𝟘nx=\mathbb{0}_{n}, with convergence rate of order 1t.\frac{1}{\sqrt{t}}. We adapt the arguments from Section 3.3 to complete the proof of Theorem 1.6, in Section 4.1. We use the following to facilitate notations:

Notations 1.7.

(i) Given a positive integer dd denote the column vectors with all-11 entries and all-0 entries in d\mathbb{R}^{d} by

𝟙d=(1,1,1)T, and 𝟘d=(0,0,0)T.\mathbb{1}_{d}=(1,1,\dots 1)^{T},\;\mbox{ and }\mathbb{0}_{d}=(0,0,\dots 0)^{T}. (10)

Given positive integers d1,d2d_{1},d_{2} denote the d1×d2d_{1}\times d_{2} matrices of all ones and all zeros by 𝟙d1,d2=𝟙d1𝟙d2T,\displaystyle\mathbb{1}_{d_{1},d_{2}}=\mathbb{1}_{d_{1}}\mathbb{1}_{d_{2}}^{T}, and 𝕆d1,d2=𝟘d1𝟘d2T.\displaystyle\mathbb{O}_{d_{1},d_{2}}=\mathbb{0}_{d_{1}}\mathbb{0}_{d_{2}}^{T}. We may use just 𝕆\mathbb{O} for a matrix if its dimensions are clear form context. Denote the d×dd\times d identity matrix by Id.I_{d}. Denote by HdH_{d} the hyperplane orthogonal to 𝟙d\mathbb{1}_{d} in d.\mathbb{R}^{d}.

(ii) Given (column) vectors ad1a\in\mathbb{R}^{d_{1}} and bd2b\in\mathbb{R}^{d_{2}} denote by col(a,b)d1+d2\mbox{col}(a,b)\in\mathbb{R}^{d_{1}+d_{2}} the column vector (a1,ad1,b1,bd2)T.(a_{1},\dots a_{d_{1}},b_{1},\dots b_{d_{2}})^{T}.

We conclude this introductory subsection by examining how other authors interpret stability for limit patterns and comparing their settings to our results, as stated in Theorem 1.6, Remarks 4.5, and 4.7, since there is no universally accepted definition.

For Cucker-Smale (CS) models with communication function ψ\psi

r¨k=1nj=1nψ(|rjrk|)(r˙jr˙k)\ddot{r}_{k}=\frac{1}{n}\sum_{j=1}^{n}\psi(|r_{j}-r_{k}|)(\dot{r}_{j}-\dot{r}_{k})

the generally accepted definition of asymptotic flocking requires two conditions:

  • Velocity alignment: limtj=1n|r˙j(t)V(t)|2=0\lim_{t\to\infty}\sum\limits_{j=1}^{n}|\dot{r}_{j}(t)-V(t)|^{2}=0

  • Forming a finite-diameter group: sup0t<j=1n|rj(t)R(t)|2<.\sup_{0\leq t<\infty}\sum\limits_{j=1}^{n}|r_{j}(t)-R(t)|^{2}<\infty.

Cucker-Smale models have constant mean velocity V(t)=V(0)V(t)=V(0) so the existence of a limiting value of VV is automatic; the problem reduces to an equivalent model with V(t)=0.V(t)=0. For the classical communications functions ψ(s)=αsβ\psi(s)=\frac{\alpha}{s^{\beta}} and ψ(s)=α(1+s2)β/2\psi(s)=\frac{\alpha}{(1+s^{2})^{\beta/2}} with 0β10\leq\beta\leq 1 it is proven in [10] and [3] that there is unconditional flocking in the sense that all solutions exhibit asymptotic flocking with |r˙j(t)V(t)||\dot{r}_{j}(t)-V(t)| being exponentially small.

For Kuramoto oscillators (KO)

θ˙k=ωk+Knj=1nsin(θjθk|),θk(t)\dot{\theta}_{k}=\omega_{k}+\frac{K}{n}\sum_{j=1}^{n}\sin(\theta_{j}-\theta_{k}|),\;\;\theta_{k}(t)\in\mathbb{R}

the mean frequency ω=1nj=1nωj\omega=\frac{1}{n}\sum_{j=1}^{n}\omega_{j} is constant, ensuring that the mean phase 1nj=1nθj\frac{1}{n}\sum_{j=1}^{n}\theta_{j} is linear in time. For (KO) synchronization plays the role of (CS)’s velocity alignment, requiring that the angular frequencies θ˙k\dot{\theta}_{k} converge to the mean frequency. It was shown that under some mild conditions on the initial distribution of phase angles, even non-identical oscillators synchronize [31], and that the exponential rate of convergence proven in [30] for identical oscillators (i.e. when all ωk=ω\omega_{k}=\omega ) extends to more general systems. Not all systems synchronize, and some do at much slower rates. Moreover, [30] shows that identical oscillators with initial phase angles confined in an semicircle have phase angles that converge exponentially to a common θ.\theta_{\infty}.

For a swarm of identical planar agents coupled with a power law potential as in (5), some authors ([32]) define flocking configurations as being solutions where all agents have rectilinear motion, with a common velocity: rk(t)=rk,0+v0tr_{k}(t)=r_{k,0}+v_{0}t for all k.k. Necessarily v0=αβeiθ0v_{0}=\sqrt{\frac{\alpha}{\beta}}e^{i\theta_{0}} for some θ0[0,2π).\theta_{0}\in[0,2\pi). Let r=(rk,0)k=1..nr_{\star}=(r_{k,0})_{k=1..n} in 2n\mathbb{R}^{2n} denote the initial locations of a flocking configuration. In [32] the authors define the stability of the flocking state evolving from rr_{\star} by referring to a manifold of configurations. They define the flock manifold (r)4n\mathcal{F}(r_{\star})\subset\mathbb{R}^{4n} as

(r)={(eiϕrk,0+tαβeiθ,αβeiθ)k=1..n, for all ϕ[0,2π),θ[0,2π),t>0}\mathcal{F}(r_{\star})=\{(e^{i\phi}r_{k,0}+t\sqrt{\frac{\alpha}{\beta}}e^{i\theta},\sqrt{\frac{\alpha}{\beta}}e^{i\theta})_{k=1..n},\mbox{ for all }\phi\in[0,2\pi),\theta\in[0,2\pi),t>0\}

and the ’local asymptotic stability’ of (r)\mathcal{F}(r_{\star}) as the existence of a δ\delta- neighborhood UU of (r)\mathcal{F}(r_{\star}) such that for initial conditions r0,r˙0r_{0},\dot{r}_{0} in U,U, the distance from the state variables r(t),r˙(t)r(t),\dot{r}(t) to the manifold (r)\mathcal{F}(r_{\star}) approaches zero as t.t\to\infty. Note that this is a rather weak condition on the behavior of V(t)V(t), as it only requires that |V|αβ|V|\to\sqrt{\frac{\alpha}{\beta}} without imposing any conditions on VV having a steady direction.

1.2 Obstacles on the way to traditional analysis

One of the difficulties in working with translating and flocking states is the fact that they are neither fixed points nor periodic cycles. Moreover, for (1) the perturbations of the translating states have centers of mass that do not move with constant velocity (or in a predetermined direction) as in Cucker-Smale where excising a linear-in-tt term out of the motion transforms the flocking states into fixed points.

In [4] the authors prove that there exit constants Cr,Cv,CaC_{r},C_{v},C_{a} that depend only on the number of particles, such that every solution to (1 ) satisfies |rk(t)R(t)|Cr,|r˙k(t)V(t)|Cv,|r_{k}(t)-R(t)|\leq C_{r},\;|\dot{r}_{k}(t)-V(t)|\leq C_{v}, and |r¨k(t)V˙(t)|Ca|\ddot{r}_{k}(t)-\dot{V}(t)|\leq C_{a} for large enough t.t. This suggests two possible workarounds for the unboundedness of flocking states: study (4) as a substitute for (1) or change the unknown functions of (1 ) from rkr_{k} and r˙k\dot{r}_{k} to rk(t)R(t)r_{k}(t)-R(t) and vk=r˙k.v_{k}=\dot{r}_{k}. Both approaches lead to dynamics evolving in a compact phase space, with the translating states of (1 ) corresponding to fixed points with all velocities vkv_{k} equal to the mean speed V0,V_{0}, a unit vector: |V0|=1.|V_{0}|=1.

This subsection briefly addresses some remaining obstacles in pursuing these model modifications. Although this subsection helps set the tone for the dimension reduction and the proof of Theorem 1.6 that begins in Section 2, the proof of the main results are independent of it.

Remark 1.8.

Any solution (rk(t))k=1..n\displaystyle\left(r_{k}(t)\right)_{k=1..n} for (1) produces velocity vectors vk=r˙kv_{k}=\dot{r}_{k} that solve (4). The converse is not true.

Proof.

(4) is obtained from differentiating (1). We provide a counterexample for the converse: let V0V_{0} be any nonzero vector in the plane with |V0|1.|V_{0}|\neq 1. Let vk(t)=V0v_{k}(t)=V_{0} for all tt and all k=1..n.k=1..n. The vectors vkv_{k} are a solution to (4) that can’t be realized as derivatives of solutions to (1), because if they did, they would be velocities for a flocking state whose asymptotic speed is neither zero nor one, an impossibility. ∎

Proposition 1.9.

Every fixed point of (4) with |V0|=1|V_{0}|=1 is unstable.

Note that the fixed points of (4) with |V0|=1|V_{0}|=1 correspond to the translating solutions of (1). We want to emphasise that the perturbations that destabilize (4) don’t correspond to derivatives of solutions to (1). The main theorem (1.6) in this paper makes this point.

Proof.

First note that a point (vk,0,v˙k,0)k=1..n\displaystyle\left(v_{k,0},\dot{v}_{k,0}\right)_{k=1..n} is a fixed point for (4) if and only if v˙k,0=0\dot{v}_{k,0}=0 for all kk and vk,0=V0v_{k,0}=V_{0} for all kk, where V0V_{0} is the initial mean velocity. (Necessarily for such a fixed point v˙k,0=0\dot{v}_{k,0}=0 for all kk; we get from (4) that 0=vk,0V00=v_{k,0}-V_{0} for all kk so all the velocities are equal.)

In order to prove that the fixed points of the system (4) with |V0|=1|V_{0}|=1 are unstable it is enough to do so when V0V_{0} is the unit vector in the xx- axis direction (due to the rotation invariance). Note that the linear subspace
{col(c1𝟙n,c2𝟙n,c3𝟙n,c4𝟙n),c1,c2,c3,c4}\{\mbox{col}(c_{1}\mathbb{1}_{n},c_{2}\mathbb{1}_{n},c_{3}\mathbb{1}_{n},c_{4}\mathbb{1}_{n}),\;c_{1},c_{2},c_{3},c_{4}\in\mathbb{R}\} of 4n\mathbb{R}^{4n} is invariant under the flow of (4), allowing us work within the class of solutions that satisfy vk(t)=V(t)v_{k}(t)=V(t) for all k=1..nk=1..n and all tt. Denote by vx,vy,axv_{x},v_{y},a_{x} and aya_{y} the horizontal and vertical components of VV and V˙\dot{V}, respectively. In this class (4) becomes:

vx˙=axvy˙=ayax˙=(13vx2vy2)ax2vxvyayay˙=(1vx23vy2)ay2vxvyax.\begin{array}[]{ll}\dot{v_{x}}=&a_{x}\\ \dot{v_{y}}=&a_{y}\\ \dot{a_{x}}=&(1-3v_{x}^{2}-v_{y}^{2})a_{x}-2v_{x}v_{y}a_{y}\\ \dot{a_{y}}=&(1-v_{x}^{2}-3v_{y}^{2})a_{y}-2v_{x}v_{y}a_{x}.\end{array} (11)

The system (11) has two conserved quantities:

ax(1vx2vy)2vx=c0x and ay(1vx2vy)2vy=c0ya_{x}-(1-v_{x}^{2}-v_{y})^{2}v_{x}=c_{0x}\;\mbox{ and }a_{y}-(1-v_{x}^{2}-v_{y})^{2}v_{y}=c_{0y}

where the constants of motion c0x=ax(0)(1vx(0)2vy(0))2vx(0)c_{0x}=a_{x}(0)-(1-v_{x}(0)^{2}-v_{y}(0))^{2}v_{x}(0) and c0y=ay(0)(1vx(0)2vy(0))2vy(0)c_{0y}=a_{y}(0)-(1-v_{x}(0)^{2}-v_{y}(0))^{2}v_{y}(0) are small if vx(0)v_{x}(0) is near one and vy(0)v_{y}(0) is near zero. Fix the perturbation parameter c0y=0c_{0y}=0 and allow c0=c0xc_{0}=c_{0x} to vary near zero; this reduces the study of (11) to that of the 3-dimensional system

vx˙=(1vx2vy2)vx+c0vy˙=(1vx2vy2)vyc0˙=0.\begin{array}[]{ll}\dot{v_{x}}=&(1-v_{x}^{2}-v_{y}^{2})v_{x}+c_{0}\\ \dot{v_{y}}=&(1-v_{x}^{2}-v_{y}^{2})v_{y}\\ \dot{c_{0}}=&0.\end{array} (12)

For c0=0c_{0}=0 all the points from the unit circle in the (vx,vy)(v_{x},v_{y}) plane are fixed points. For each c00c_{0}\neq 0 , small, the system (12) has three fixed points, one fixed point (v(c0),0,c0)(v_{\star}(c_{0}),0,c_{0}) near M(1,0,0)M(1,0,0) and two other fixed points near (0,0,0)(0,0,0) and (1,0,0).(-1,0,0). v(c0)v_{\star}(c_{0}) satisfies v(1v2)=c0.v_{\star}(1-v_{\star}^{2})=-c_{0}.

The system (12) can be seen as a one-parameter family of planar systems. Its stability and bifurcations are well known: a summary of the ensuing dynamics can be found in [Gu], page 70 (their parameter γ\gamma is c0-c_{0}). For c0<0c_{0}<0 small, the fixed point (v,0)(v_{\star},0) has a positive eigenvalue (λ=1v2\lambda=1-v_{\star}^{2}), and the flow of (12) has heteroclinic cycles with a saddle connection near (1,0)(1,0), a stable node near (1,0)(-1,0) and an unstable node near (0,0)(0,0)).

The locations of the aforementioned fixed points are shown Figure 3, with c0c_{0} on the vertical axis: the curve (v(c0),0,c0)(v_{\star}(c_{0}),0,c_{0}) through M(1,0,0)M(1,0,0) is illustrated in red with two \star markers; also in red are the almost-vertical curves of fixed points near the fixed points (0,0)(0,0) and (1,0)(-1,0) in the c0=0c_{0}=0 plane. The fixed points on the unit circle in the plane c0=0c_{0}=0 use ×\times markers. All the other curves in Figure 3 are trajectories within three representative planes (c0<0,c0=0,c_{0}<0,c_{0}=0, and c0>0c_{0}>0).

Refer to caption
Figure 3: The flow of (12) for three values of c0.c_{0}. The system (4) subsumes (12). The dynamics near the fixed point of (4) with V0=(1,0)V_{0}=(1,0) corresponds to the dynamics of (12) near M(1,0,0),M(1,0,0), which is unstable on account of having saddles and heteroclinic cycles near MM in the region c0<0.c_{0}<0.

The dynamics of (12) in the region c0<0c_{0}<0 near M(1,0,0)M(1,0,0) makes the fixed point (1,0,0,0)(1,0,0,0) of (11) unstable; thus the fixed point (𝟙n,𝟘n,𝟘n,𝟘n)(\mathbb{1}_{n},\mathbb{0}_{n},\mathbb{0}_{n},\mathbb{0}_{n}) of (4) is unstable.

The second natural approach for transforming the unbounded solutions of (1) and its translating states into bounded trajectories and fixed points is to work with σk=rkR=(σx,k,σy,k)\sigma_{k}=r_{k}-R=(\sigma_{x,k},\sigma_{y,k}) and r˙k=(vx,k,vy,k)\dot{r}_{k}=(v_{x,k},v_{y,k}) as the unknown functions. Note that since rnR=k=1n1σkr_{n}-R=-\sum_{k=1}^{n-1}\sigma_{k} we can eliminate it from our unknowns, and work within a 4n24n-2 dimensional phase space (if we were to keep the redundant rnRr_{n}-R as a state variable, we would have added two more zero eigenvalues to the system).

The system (1) becomes:

σx,k˙=vx,k1nj=1nvx,jσy,k˙=vy,k1nj=1nvy,jvx,k˙=(1vx,k2vy,k2)vx,kσx,kvy,k˙=(1vx,k2vy,k2)vy,kσy,kvx,n˙=(1vx,n2vy,n2)vx,n+k=1n1σx,kvy,n˙=(1vx,k2vy,k2)vy,n+k=1n1σy,k\begin{array}[]{ll}\dot{\sigma_{x,k}}=&v_{x,k}-\frac{1}{n}\sum_{j=1}^{n}v_{x,j}\\ \dot{\sigma_{y,k}}=&v_{y,k}-\frac{1}{n}\sum_{j=1}^{n}v_{y,j}\\ \hline\cr\\ \dot{v_{x,k}}=&(1-v_{x,k}^{2}-v_{y,k}^{2})v_{x,k}-\sigma_{x,k}\\ \dot{v_{y,k}}=&(1-v_{x,k}^{2}-v_{y,k}^{2})v_{y,k}-\sigma_{y,k}\\ \dot{v_{x,n}}=&(1-v_{x,n}^{2}-v_{y,n}^{2})v_{x,n}+\sum_{k=1}^{n-1}\sigma_{x,k}\\ \dot{v_{y,n}}=&(1-v_{x,k}^{2}-v_{y,k}^{2})v_{y,n}+\sum_{k=1}^{n-1}\sigma_{y,k}\end{array} (13)

where k=1..(n1).k=1..(n-1).

The translating states of (1) with V(t)=(cosθ0,sinθ0)V(t)=(\cos\theta_{0},\sin\theta_{0}) correspond to the fixed points Pθ0P_{\theta_{0}} with σx=σy=𝟘n1,\sigma_{x}=\sigma_{y}=\mathbb{0}_{n-1}, and vx=cosθ0𝟙n,vy=sinθ0𝟙n.v_{x}=\cos\theta_{0}\;\mathbb{1}_{n},\;v_{y}=\sin\theta_{0}\;\mathbb{1}_{n}.

Focus on the dynamics near P0=col(𝟘n1,𝟘n1,𝟙n1,𝟘n1,1,0).P_{0}=\mbox{col}(\mathbb{0}_{n-1},\mathbb{0}_{n-1},\mathbb{1}_{n-1},\mathbb{0}_{n-1},1,0). The other fixed points are obtained from P0P_{0} via rotation of angle θ0.\theta_{0}. The linearization of (13) at P0P_{0} yields the Jacobian matrix

J=[𝕆𝕆Bn1𝕆1n𝟙n1𝟘n1𝕆𝕆𝕆Bn1𝟘n11n𝟙n1In1𝕆2In1𝕆𝟘n1𝟘n1𝕆In1𝕆𝕆𝟘n1𝟘n1𝟙n1T𝟘n1T𝟘n1T𝟘n1T20𝟘n1T𝟙n1T𝟘n1T𝟘n1T00]J=\left[\begin{array}[]{ll|ccrr}\mathbb{O}&\mathbb{O}&B_{n-1}&\mathbb{O}&-\frac{1}{n}\mathbb{1}_{n-1}&\mathbb{0}_{n-1}\\ \mathbb{O}&\mathbb{O}&\mathbb{O}&B_{n-1}&\mathbb{0}_{n-1}&-\frac{1}{n}\mathbb{1}_{n-1}\\ \hline\cr\\ -\textrm{I}_{n-1}&\mathbb{O}&-2\textrm{I}_{n-1}&\mathbb{O}&\mathbb{0}_{n-1}&\mathbb{0}_{n-1}\\ \mathbb{O}&-\textrm{I}_{n-1}&\mathbb{O}&\mathbb{O}&\mathbb{0}_{n-1}&\mathbb{0}_{n-1}\\ \mathbb{1}_{n-1}^{T}&\mathbb{0}_{n-1}^{T}&\mathbb{0}_{n-1}^{T}&\mathbb{0}_{n-1}^{T}&-2&0\\ \mathbb{0}_{n-1}^{T}&\mathbb{1}_{n-1}^{T}&\mathbb{0}_{n-1}^{T}&\mathbb{0}_{n-1}^{T}&0&0\end{array}\right]

where Bn1B_{n-1} is the square matrix In11n𝟙n1,n1\textrm{I}_{n-1}-\frac{1}{n}\mathbb{1}_{n-1,n-1} and 𝕆=𝕆n1,n1.\mathbb{O}=\mathbb{O}_{n-1,n-1}. We get that

det(JλI4n2)=λ(1+λ2)n1(2+λ)(1+λ)2n2.\mbox{det}(J-\lambda I_{4n-2})=\lambda(1+\lambda^{2})^{n-1}(2+\lambda)(1+\lambda)^{2n-2}.

The eigenvector for λ=0\lambda=0 is col(𝟘n1,𝟘n1,𝟘n1,𝟙n1,0,1)\mbox{col}(\mathbb{0}_{n-1},\mathbb{0}_{n-1},\mathbb{0}_{n-1},\mathbb{1}_{n-1},0,1); it has the direction of the line tangent to the curve of fixed points {Pθ0,θ0[0,2π]}\{P_{\theta_{0}},\;\theta_{0}\in[0,2\pi]\} at P0.P_{0}. The eigenvalue λ=0\lambda=0 captures the rotation invariance of the model. Note that due to the ±i\pm i eigenvalues there are more degeneracies associated with (13) than the symmetries of the problem.

If in the system (13) all particles’ yy-components are initially zero, they stay zero forever, with the remaining 2n12n-1 dimensional system in σx,vx,vx,n1\sigma_{x},v_{x},v_{x,n-1} having the Jacobian JxJ_{x} at σx=𝟘n1,vx=𝟙n1,vx,n1=1\sigma_{x}=\mathbb{0}_{n-1},v_{x}=\mathbb{1}_{n-1},v_{x,n-1}=1 equal to

Jx=[𝕆Bn11n𝟙n1In12In1𝟘n1𝟙n1T𝟘n1T2]J_{x}=\left[\begin{array}[]{lll}\mathbb{O}&B_{n-1}&-\frac{1}{n}\mathbb{1}_{n-1}\\ -\textrm{I}_{n-1}&-2\textrm{I}_{n-1}&\mathbb{0}_{n-1}\\ \mathbb{1}_{n-1}^{T}&\mathbb{0}_{n-1}^{T}&-2\end{array}\right]

and characteristic polynomial (λ+2)(λ+1)2n2.(\lambda+2)(\lambda+1)^{2n-2}. In the absence of a yy- component, all trajectories approach the fixed point P0P_{0} exponentially fast, meaning that the entire stable manifold for (13) at P0P_{0} consists of the trajectories with no yy component. The flow on the 2n12n-1 dimensional central manifold captures the drift in the direction of motion, and the perturbations/ oscillations that are orthogonal to the direction of motion.

In order to understand the dynamics, one needs a suitable approximation of a central manifold map hh whose graph is a central manifold, typically done using Taylor polynomials (as in [11], pages 3-5). Unfortunately, Taylor approximations are ill fitted to capture the correct scales of the flow in the presence of non-isolated fixed points, which is the case for (13). A presentation of that inadequacy is in [5] 111[5] provides an alternate construction for the central manifold for systems with non-isolated fixed points having no purely imaginary eigenvalues. Their method does not apply in the context of (13). The predicament of having non-isolated fixed points and imaginary eigenvalues of high multiplicity requires we further refine the model until all the translating states are identified with a single fixed point. That is done in Section 2.

Remark 1.10.

If we allow the motion to take place in space (i.e replacing the assumption that rk2r_{k}\in\mathbb{R}^{2} from (1) with rk3r_{k}\in\mathbb{R}^{3}), there exist initial conditions spawning solutions that fail the asymptotic conditions (7) and (8) even though the initial conditions are arbitrarily close to those of flocking states.

For example, starting from the flocking solution rk(t)=(0,0,t)r_{k}(t)=(0,0,t) for all tt and kk, given ϵ>0\epsilon>0 small, consider the perturbed initial conditions rk,0=(ϵcosθk,ϵsinθk,0)r_{k,0}=(\epsilon\cos\theta_{k},\epsilon\sin\theta_{k},0) and r˙k,0=(ϵsinθk,ϵcosθk,1ϵ2)\dot{r}_{k,0}=(-\epsilon\sin\theta_{k},\epsilon\cos\theta_{k},\sqrt{1-\epsilon^{2}}) where θk\theta_{k} are angles such that k=1neiθk=0.\displaystyle\sum_{k=1}^{n}e^{i\theta k}=0. Note that for these perturbations R(0)=(0,0,0)R(0)=(0,0,0) and V(0)=(0,0,1ϵ2).V(0)=(0,0,\sqrt{1-\epsilon^{2}}). It is immediate to verify that the particles in the 3\mathbb{R}^{3} version of (1) move with unit speed on helix-shaped trajectories according to

rk(t)=(ϵcos(t+θk),ϵsin(t+θk),t1ϵ2).r_{k}(t)=(\epsilon\cos(t+\theta_{k}),\epsilon\sin(t+\theta_{k}),t\sqrt{1-\epsilon^{2}}).

The perturbed trajectories have R(t)=(0,0,t1ϵ2)R(t)=(0,0,t\sqrt{1-\epsilon^{2}}) and therefore |rk(t)R(t)|=ϵ|r_{k}(t)-R(t)|=\epsilon and |r˙k(t)V(t)|=ϵ|\dot{r}_{k}(t)-V(t)|=\epsilon for all t,t, failing the asymptotics (7) and (8).


2 Eliminating the Zero Eigenvalue(s)

One of the difficulties in working with the system (1) is the drift in the direction of motion, as illustrated in Figure 1. For configurations having nonzero mean velocity (V(t)0V(t)\neq 0 for t>0t>0), we decouple the direction of motion from the other variables of the system. We achieve this by using a frame that moves and rotates according to the center of mass, as illustrated in Figure 4. The new frame annihilates the directional drift and the rotation symmetry of the system, just as working with rkRr_{k}-R eliminated the translation invariance. Since the rotation and translation invariance were associated with the multiplicity-three zero eigenvalue for the linearization of (1), the eigenvalues of the ensuing reframed dynamics will be shown to be nonzero (either equal to ±i\pm i or with negative real parts).

Notations 2.1.

Within this section, if a=(a1,a2)a=(a_{1},a_{2}) and b=(b1,b2)b=(b_{1},b_{2}) are vectors in 2\mathbb{R}^{2} with b0,b\neq 0, define the vector aa^{\bot} and the scalars compba\mbox{comp}_{b}a and ortba\mbox{ort}_{b}a to be:

a=(a2,a1),compba=aTb|b|,ortba=aTb|b|.a^{\bot}=(-a_{2},a_{1}),\;\;\mbox{comp}_{b}a=\frac{a^{T}b}{|b|},\;\;\mbox{ort}_{b}a=\frac{a^{T}b^{\bot}}{|b|}. (14)

Note that (a)=a,|a|2=|a|2=(compba)2+(ortba)2\displaystyle(a^{\bot})^{\bot}=-a,\;|a|^{2}=|a^{\bot}|^{2}=(\mbox{comp}_{b}a)^{2}+(\mbox{ort}_{b}a)^{2} and compbb=|b|.\displaystyle\mbox{comp}_{b}b=|b|. Also:

a=(compba)b|b|+(ortba)b|b|.a=(\mbox{comp}_{b}a)\frac{b}{|b|}+(\mbox{ort}_{b}a)\frac{b^{\bot}}{|b|}. (15)
Proposition 2.2.

Let II be in interval and a,b:I2a,b:I\to\mathbb{R}^{2} be differentiable functions, b0.b\neq 0. Then compba\mbox{comp}_{b}a and ortba\mbox{ort}_{b}a are differentiable and

ddt(compba)=compba˙+1|b|(ortba)(ortbb˙)ddt(ortba)=ortba˙1|b|(compba)(ortbb˙).\begin{array}[]{cl}\frac{d}{dt}\left(\mbox{comp}_{b}a\right)&=\mbox{comp}_{b}\dot{a}+\frac{1}{|b|}(\mbox{ort}_{b}a)(\mbox{ort}_{b}\dot{b})\\ \frac{d}{dt}\left(\mbox{ort}_{b}a\right)&=\mbox{ort}_{b}\dot{a}-\frac{1}{|b|}(\mbox{comp}_{b}a)(\mbox{ort}_{b}\dot{b}).\end{array} (16)
Proof.

From (14) and the fact that the derivative and \bot commute,

ddt(compba)=(a˙)Tb|b|+aTddtb|b|=compba˙+aTddtb|b|ddt(ortba)=(a˙)Tb|b|+aT(ddtb|b|)=ortba˙+aT(ddtb|b|).\begin{array}[]{cl}\frac{d}{dt}\left(\mbox{comp}_{b}a\right)&=(\dot{a})^{T}\frac{b}{|b|}+a^{T}\frac{d}{dt}\frac{b}{|b|}=\mbox{comp}_{b}\dot{a}+a^{T}\frac{d}{dt}\frac{b}{|b|}\\ \frac{d}{dt}\left(\mbox{ort}_{b}a\right)&=(\dot{a})^{T}\frac{b^{\bot}}{|b|}+a^{T}\left(\frac{d}{dt}\frac{b}{|b|}\right)^{\bot}=\mbox{ort}_{b}\dot{a}+a^{T}\left(\frac{d}{dt}\frac{b}{|b|}\right)^{\bot}.\end{array} (17)

From ddtb|b|=1|b|b˙bTb˙|b|3b=1|b|((compbb˙)b|b|+(ortbb˙)b|b|)compbb˙b|b|2\displaystyle\frac{d}{dt}\frac{b}{|b|}=\frac{1}{|b|}\dot{b}-\frac{b^{T}\dot{b}}{|b|^{3}}b=\frac{1}{|b|}\left((\mbox{comp}_{b}\dot{b})\frac{b}{|b|}+(\mbox{ort}_{b}\dot{b})\frac{b^{\bot}}{|b|}\right)-\mbox{comp}_{b}\dot{b}\;\frac{b}{|b|^{2}}
we get

ddtb|b|=1|b|2(ortbb˙)b\frac{d}{dt}\frac{b}{|b|}=\frac{1}{|b|^{2}}(\mbox{ort}_{b}\dot{b})\;b^{\bot} (18)

Substituting (18) and (b)=b\displaystyle(b^{\bot})^{\bot}=-b into (17) yields (16).

Notations 2.3.

For the system (1) with V(t)0V(t)\neq 0, define the scalar functions of time xk,yk,uk,wkx_{k},y_{k},u_{k},w_{k} and m,sm,s to be

xk=compV(rkR) and yk=ortV(rkR)uk=compV(r˙k)wk=ortV(r˙k)s=compV(V˙)m=1|V|ortV(V˙).\begin{array}[]{lcl}x_{k}=\mbox{comp}_{V}(r_{k}-R)&\mbox{ and }&y_{k}=\mbox{ort}_{V}(r_{k}-R)\\ u_{k}=\mbox{comp}_{V}(\dot{r}_{k})&&w_{k}=\mbox{ort}_{V}(\dot{r}_{k})\\ s=\mbox{comp}_{V}(\dot{V})&&m=\frac{1}{|V|}\mbox{ort}_{V}(\dot{V}).\end{array} (19)

The scalars x1,wnx_{1},\dots w_{n} are the coordinates of the agents relative to a frame that moves along the center of mass, and rotates according to the direction of the mean velocity, as shown in Figure 4. ss and mm are introduced to streamline notations – they can be expressed in terms of x1,wnx_{1},\dots w_{n}, as in (20) and (21).

Refer to caption
Figure 4: The moving frame used to define x1,wn.x_{1},\dots w_{n}. The solid black curve represents the location of the center of mass, RR; tangent to it are the velocity vectors VV. The shorter curve illustrates the position of agent 1.1. The point AA shows center of mass at some time t;t; at that time the displacement vector of the first agent is r1(t)R(t),r_{1}(t)-R(t), shown as the red diagonal vector. The sides of the dotted rectangle represent x1(t)x_{1}(t) and y1(t)y_{1}(t).

Note that V˙=sV|V|+mV\displaystyle\dot{V}=s\frac{V}{|V|}+mV^{\bot} thus m(t)m(t) captures how fast the direction of VV is changing relative to the speed, where s(t)s(t) captures the changes in the speed of the center of mass. The geometrical meanings of mm and ss are apparent if we represent the velocity VV in polar coordinates, V(t)=|V|(cosΘ(t),sinΘ(t)).V(t)=|V|(\cos\Theta(t),\sin\Theta(t)). Then

V˙=d|V|dt(cosΘ(t),sinΘ(t))+dΘdt|V|(sinΘ(t)),cosΘ(t))=d|V|dtV|V|+dΘdtV.\dot{V}=\frac{d|V|}{dt}(\cos\Theta(t),\sin\Theta(t))+\frac{d\Theta}{dt}|V|(-\sin\Theta(t)),\cos\Theta(t))=\frac{d|V|}{dt}\frac{V}{|V|}+\frac{d\Theta}{dt}V^{\bot}.

On the other hand V=sV|V|+mV,V=s\frac{V}{|V|}+mV^{\bot}, therefore s=d|V|dt and m=dΘdt.\displaystyle s=\frac{d|V|}{dt}\;\mbox{ and }\;m=\frac{d\Theta}{dt}.

From (3) and (15) we get V˙=1nk=1n(1uk2wk2)(ukV|V|+wkV|V|).\displaystyle\dot{V}=\frac{1}{n}\sum_{k=1}^{n}(1-u_{k}^{2}-w_{k}^{2})(u_{k}\frac{V}{|V|}+w_{k}\frac{V^{\bot}}{|V|}). Collecting the tangential and normal components of the acceleration V˙\dot{V} as in (19) yields

s=1nk=1n(1uk2wk2)uk=d|V|dt and s=\frac{1}{n}\sum_{k=1}^{n}(1-u_{k}^{2}-w_{k}^{2})u_{k}=\frac{d|V|}{dt}\;\mbox{ and } (20)
m=1|V|1nk=1n(1uk2wk2)wk=dΘdt.m=\frac{1}{|V|}\frac{1}{n}\sum_{k=1}^{n}(1-u_{k}^{2}-w_{k}^{2})w_{k}=\frac{d\Theta}{dt}. (21)

Before reformulating (1) in terms of x1,wnx_{1},\dots w_{n}, we point out that xnx_{n} will not be used as state variable, since xn=k=1n1xk,\displaystyle x_{n}=-\sum_{k=1}^{n-1}x_{k}, which follows from k=1n(rkR)=0.\displaystyle\sum_{k=1}^{n}(r_{k}-R)=0. The latter and 1nknr˙k=V\displaystyle\frac{1}{n}\sum_{k}^{n}\dot{r}_{k}=V imply:

k=1nxk=k=1nyk=0, and k=1nwk=0,1nknuk=|V|.\sum_{k=1}^{n}x_{k}=\sum_{k=1}^{n}y_{k}=0,\mbox{ and }\sum_{k=1}^{n}w_{k}=0,\;\;\frac{1}{n}\sum_{k}^{n}u_{k}=|V|. (22)

Unlike xn,x_{n}, we keep wnw_{n} and yny_{n} as state variables because eliminating them would hinder downstream computations (by contributing off-pattern terms to the mean-field variables VV and V˙\dot{V} ).

Proposition 2.4.

The differential equations governing the (4n1)(4n-1) dynamical system with unknowns w1,wnw_{1},\dots w_{n}, y1,yny_{1},\dots y_{n}, x1,xn1x_{1},\dots x_{n-1} and u1,unu_{1},\dots u_{n} are

w˙k=(1uk2wk2)wkykmuk,k=1..ny˙k=wkmxk,k=1..n1y˙n=wnmxn=wnm(k=1n1xk),x˙k=uk1nk=1nuk+myk,k=1..n1u˙k=(1uk2wk2)ukxk+mwk,k=1..n1u˙n=(1un2wn2)un+l=1n1xl+mwn\begin{array}[]{ll}\dot{w}_{k}=(1-u_{k}^{2}-w_{k}^{2})w_{k}-y_{k}-mu_{k},&k=1..n\\ \dot{y}_{k}=w_{k}-mx_{k},&k=1..n-1\\ \dot{y}_{n}=w_{n}-mx_{n}=w_{n}-m(-\sum_{k=1}^{n-1}x_{k}),&\\ \dot{x}_{k}=u_{k}-\frac{1}{n}\sum_{k=1}^{n}u_{k}+my_{k},&k=1..n-1\\ \dot{u}_{k}=(1-u_{k}^{2}-w_{k}^{2})u_{k}-x_{k}+mw_{k},&k=1..n-1\\ \dot{u}_{n}=(1-u_{n}^{2}-w_{n}^{2})u_{n}+\sum_{l=1}^{n-1}x_{l}+mw_{n}&\end{array} (23)

where mm was defined in (21) as m=11nk=1nuk1nk=1n(1uk2wk2)wk.\displaystyle m=\frac{1}{\frac{1}{n}\sum_{k=1}^{n}u_{k}}\frac{1}{n}\sum_{k=1}^{n}(1-u_{k}^{2}-w_{k}^{2})w_{k}.

Proof.

Apply the results of (16) for b=V,\displaystyle b=V, with compV(V˙)=s,\mbox{comp}_{V}(\dot{V})=s, and ortV(V˙)=m|V|.\mbox{ort}_{V}(\dot{V})=m|V|. We get that for a differentiable function aa

ddt(compVa)=compVa˙+mortVaddt(ortVa)=ortVa˙mcompVa.\begin{array}[]{ll}\frac{d}{dt}\left(\mbox{comp}_{V}a\right)=&\mbox{comp}_{V}\dot{a}+m\;\mbox{ort}_{V}a\\ \frac{d}{dt}\left(\mbox{ort}_{V}a\right)=&\mbox{ort}_{V}\dot{a}-m\;\mbox{comp}_{V}a.\end{array} (24)

Use (24) for a=rkRa=r_{k}-R; recall that compV(rkR)=xk\displaystyle\mbox{comp}_{V}(r_{k}-R)=x_{k} and ortV(rkR)=yk.\displaystyle\mbox{ort}_{V}(r_{k}-R)=y_{k}. Moreover a˙=r˙kV\dot{a}=\dot{r}_{k}-V and thus a˙\dot{a} has components compVa˙=uk|V|\displaystyle\mbox{comp}_{V}\dot{a}=u_{k}-|V| and ortVa˙=wk.\displaystyle\mbox{ort}_{V}\dot{a}=w_{k}. We get

x˙k=uk|V|+myky˙k=wkmxk.\begin{array}[]{ll}\dot{x}_{k}=u_{k}-|V|+my_{k}\\ \dot{y}_{k}=w_{k}-mx_{k}.\end{array} (25)

Use (24) again for a=r˙k.a=\dot{r}_{k}. Note that a˙=(1uk2wk2)r˙k(rkR)\dot{a}=(1-u_{k}^{2}-w_{k}^{2})\dot{r}_{k}-(r_{k}-R) so it has components compVa˙=(1uk2wk2)ukxk\displaystyle\mbox{comp}_{V}\dot{a}=(1-u_{k}^{2}-w_{k}^{2})u_{k}-x_{k} and ortVa˙=(1uk2wk2)wkyk.\displaystyle\mbox{ort}_{V}\dot{a}=(1-u_{k}^{2}-w_{k}^{2})w_{k}-y_{k}. We get

u˙k=(1uk2wk2)ukxk+mwkk=1nw˙k=(1uk2wk2)wkykmukk=1n.\begin{array}[]{lll}\dot{u}_{k}=&(1-u_{k}^{2}-w_{k}^{2})u_{k}-x_{k}+mw_{k}&k=1\dots n\\ \dot{w}_{k}=&(1-u_{k}^{2}-w_{k}^{2})w_{k}-y_{k}-mu_{k}&k=1\dots n.\end{array} (26)

Remark 2.5.

The dynamical system (23)(\ref{State}) is reflection invariant, in the sense that if (w,y,x,u)(w,y,x,u) is a solution for (23)(\ref{State}), then (w,y,x,u)(-w,-y,x,u) is also a solution.

Let HH denote the hyperplane orthogonal to 𝟙n\mathbb{1}_{n} in n.\mathbb{R}^{n}. The subspace H×H×2n1H\times H\times\mathbb{R}^{2n-1} is invariant under (23).

The reflection invariance holds since the functions mxmx and mumu are odd functions of (w,y)(w,y) and even in (x,u)(x,u) and thus the vector field given from (23)(\ref{State}) has its (w˙,y˙)(\dot{w},\dot{y}) components given by odd function of (w,y)(w,y) and even in (x,u),(x,u), whereas the (x˙,u˙)(\dot{x},\dot{u}) components of FF are even functions of (w,y)(w,y) and (x,u).(x,u). The invariance of H×H×2n1H\times H\times\mathbb{R}^{2n-1} follows from 1nk=1nw˙k=1nk=1nyk\displaystyle\frac{1}{n}\sum_{k=1}^{n}\dot{w}_{k}=\frac{1}{n}\sum_{k=1}^{n}y_{k} and 1nk=1ny˙k=1nk=1nwk.\displaystyle\frac{1}{n}\sum_{k=1}^{n}\dot{y}_{k}=\frac{1}{n}\sum_{k=1}^{n}w_{k}.

Remark 2.6.

Any translating state rk(t)=(c0,1,c0,2)+teiΘ0r_{k}(t)=(c_{0,1},c_{0,2})+te^{i\Theta_{0}} for k=1..nk=1..n of (1) corresponds to the fixed point wk=0,yk=0w_{k}=0,y_{k}=0 and uk=1u_{k}=1 for k=1..nk=1..n and xk=0x_{k}=0 for k=1..n1k=1..n-1 of (23).

Proposition 2.7.

Let ZfZ_{f} denote the fixed point of (23) with wk=0,yk=0,xl=0w_{k}=0,y_{k}=0,x_{l}=0 and uk=1u_{k}=1 (for k=1..nk=1..n and l=1..n1l=1..n-1). Denote by JJ the matrix of the linearization of (23) at Zf.Z_{f}. Then

det(λIJ)=(λ2+1)n(λ+1)2(n1)(λ+2).\mbox{det}(\lambda I-J)=(\lambda^{2}+1)^{n}(\lambda+1)^{2(n-1)}(\lambda+2).

The center subspace EcE^{c} of JJ has dimension 2n2n and is spanned by the variables w,y.w,y. The stable subspace EsE^{s} has dimension 2n12n-1 and is spanned by the variables x,u.x,u. There is no unstable subspace: 4n1=EcEs=2n2n1.\displaystyle\mathbb{R}^{4n-1}=E^{c}\oplus E^{s}=\mathbb{R}^{2n}\oplus\mathbb{R}^{2n-1}.

Proof.

Note that at ZfZ_{f} the speed is |V|=1nk=1nuk=1.|V|=\frac{1}{n}\sum_{k=1}^{n}u_{k}=1. Near ZfZ_{f} all the terms (1uk2wk2)wk(1-u_{k}^{2}-w_{k}^{2})w_{k} are quadratic or smaller since both (1uk2wk2)(1-u_{k}^{2}-w_{k}^{2}) and wkw_{k} are zero at Zf,Z_{f}, therefore (1uk2wk2)wk(1-u_{k}^{2}-w_{k}^{2})w_{k} have zero gradients. From m=1|V|1nk=1n(1uk2wk2)wk\displaystyle m=\frac{1}{|V|}\frac{1}{n}\sum_{k=1}^{n}(1-u_{k}^{2}-w_{k}^{2})w_{k} we conclude that mm and the terms muk,mxk,mykmu_{k},mx_{k},my_{k} and mwkmw_{k} vanish and have zero gradient at ZfZ_{f} as well. The nonlinear term (ukuk3wk2uk)(u_{k}-u_{k}^{3}-w_{k}^{2}u_{k}) that is part of u˙k\dot{u}_{k} is linearly approximated by (2)(uk1)(-2)(u_{k}-1) at uk=1,wk=0.u_{k}=1,w_{k}=0. Finally, note that the pattern for u˙nxl\displaystyle\frac{\partial\dot{u}_{n}}{\partial x_{l}} is different than that for the partial derivatives of u˙k\dot{u}_{k} and k=1..n1k=1..n-1 so when using block notations for the Jacobian matrix, we have to separate away the last uu component from the rest.

The Jacobian matrix (23) at ZfZ_{f} is

J=[𝕆nIn𝕆𝕆0n,1In𝕆n𝕆𝕆0n,1𝕆𝕆𝕆𝕟𝟙Bn11n𝟙n1𝕆𝕆In12In10n1,101,n01,n𝟙n1T01,n12]J=\left[\begin{array}[]{ll|lll}\mathbb{O}_{n}&-I_{n}&\mathbb{O}&\mathbb{O}&0_{n,1}\\ I_{n}&\mathbb{O}_{n}&\mathbb{O}&\mathbb{O}&0_{n,1}\\ \hline\cr&&&&\\ \mathbb{O}&\mathbb{O}&\mathbb{O_{n-1}}&B_{n-1}&\frac{-1}{n}\mathbb{1}_{n-1}\\ \mathbb{O}&\mathbb{O}&-\textrm{I}_{n-1}&-2\textrm{I}_{n-1}&0_{n-1,1}\\ 0_{1,n}&0_{1,n}&\mathbb{1}_{n-1}^{T}&0_{1,n-1}&-2\end{array}\right] (27)

where Bn1B_{n-1} is the square matrix In11n𝟙n1,n1\textrm{I}_{n-1}-\frac{1}{n}\mathbb{1}_{n-1,n-1}; on the left of the separating line 𝕆\mathbb{O} means 𝕆n1,n\mathbb{O}_{n-1,n} and on the right of the separating line 𝕆\mathbb{O} means 𝕆n,n1\mathbb{O}_{n,n-1}. Note that JJ is already in a block diagonal form. The first block, of size 2n2n, corresponding to the state variables ww and yy, has characteristic polynomial (λ2+1)n.\displaystyle(\lambda^{2}+1)^{n}. The second block, of size (2n1)(2n-1), corresponding to the state variables xx and uu has characteristic polynomial (λ+1)2n2(λ+2).\displaystyle(\lambda+1)^{2n-2}(\lambda+2). One can check that by performing the following operations: first subtract the last column from each column in the middle block, then add all the rows of the middle block to the last row (the identity matrix II being In1\textrm{I}_{n-1}):

|λII+1n𝟙1n𝟙n1I(λ+2)I0n1,1𝟙n1T01,n1λ+2|=|λII1n𝟙n1I(λ+2)I0n1,1𝟙n1T(λ+2)𝟙n1Tλ+2|=\left|\begin{array}[]{lll}\lambda I&-I+\frac{1}{n}\mathbb{1}&\frac{1}{n}\mathbb{1}_{n-1}\\ I&(\lambda+2)I&0_{n-1,1}\\ -\mathbb{1}_{n-1}^{T}&0_{1,n-1}&\lambda+2\end{array}\right|=\left|\begin{array}[]{lll}\lambda I&-I&\frac{1}{n}\mathbb{1}_{n-1}\\ I&(\lambda+2)I&0_{n-1,1}\\ -\mathbb{1}_{n-1}^{T}&-(\lambda+2)\mathbb{1}_{n-1}^{T}&\lambda+2\end{array}\right|=
=|λII1n𝟙n1I(λ+2)I0n1,100λ+2|=(λ+2)(λ2+2λ+1)n1.=\left|\begin{array}[]{lll}\lambda I&-I&\frac{1}{n}\mathbb{1}_{n-1}\\ I&(\lambda+2)I&0_{n-1,1}\\ 0&0&\lambda+2\end{array}\right|=(\lambda+2)(\lambda^{2}+2\lambda+1)^{n-1}.

Notations 2.8.

Given an initial condition vector Z0Z_{0} in 4n1\mathbb{R}^{4n-1} denote by Z=ψt(Z0)Z=\psi_{t}(Z_{0}) the solution to (23). Given an initial condition vector X0X_{0} in 4n\mathbb{R}^{4n} consisting of rk(0)r_{k}(0) and vk(0)v_{k}(0) for (1), denote by ϕt(X0)\phi_{t}(X_{0}) the solution to (1).

Remark 2.9.

For the proof of the Main Theorem 1.6, it is enough to show that the fixed point Zf=col(𝟘n,𝟘n,𝟘n1,𝟙n)Z_{f}=col(\mathbb{0}_{n},\mathbb{0}_{n},\mathbb{0}_{n-1},\mathbb{1}_{n}) is an asymptotic fixed point for (23) on H×H×2n1H\times H\times\mathbb{R}^{2n-1} and that the integral 0|m|𝑑t\int_{0}^{\infty}|m|dt converges and it is small for solutions starting near ZfZ_{f}.

Proof.

The assumptions of Main Theorem 1.6 and the assertions (6) and (7) can be restated using ψt=col(w,y,x,u)\psi_{t}=col(w,y,x,u) alone, since they are norm-based and

|rk1rk2|2=|(rk1R)(rk2R)|2=(xk1xk2)2+(yk1yk2)2|r_{k_{1}}-r_{k_{2}}|^{2}=|(r_{k_{1}}-R)-(r_{k_{2}}-R)|^{2}=(x_{k_{1}}-x_{k_{2}})^{2}+(y_{k_{1}}-y_{k_{2}})^{2}

and

|V|=1nk=1nuk and |r˙kV|2=(uk|V|)2+wk2.|V|=\frac{1}{n}\sum_{k=1}^{n}u_{k}\;\mbox{ and }\;|\dot{r}_{k}-V|^{2}=(u_{k}-|V|)^{2}+w_{k}^{2}.

This turns the assertions (6) and (7) into statements about the asymptotic stability of the flow ψt\psi_{t} at Zf.Z_{f}.

For the assertion (8) we need to examine how, given a solution ψt(Z0)\psi_{t}(Z_{0}) in the (4n3)(4n-3) dimensional space H×H×2n1H\times H\times\mathbb{R}^{2n-1}, we can identify the trajectories ϕt\phi_{t} in 4n\mathbb{R}^{4n} that in the center-mass frame (19) lead back to ψt(Z0)\psi_{t}(Z_{0}) . The trajectories ϕt\phi_{t} can only be determined up to a 3-dimensional parameter c0c_{0}, interpreted as a location (c0,1,c0,2)(c_{0,1},c_{0,2}) in 2\mathbb{R}^{2} together with an angle c0,3=Θ0.c_{0,3}=\Theta_{0}.

For ψt=(x1,un)\psi_{t}=(x_{1},\dots u_{n}) in H×H×2n1H\times H\times\mathbb{R}^{2n-1}, define Θ(t)=Θ0+0tm𝑑τ\displaystyle\Theta(t)=\Theta_{0}+\int_{0}^{t}m\;d\tau where m=1|V|1nk=1n(1uk2wk2)wk\displaystyle m=\frac{1}{|V|}\frac{1}{n}\sum_{k=1}^{n}(1-u_{k}^{2}-w_{k}^{2})w_{k} (see (21)). Note that |Θ(t)Θ0|0|m|𝑑t.|\Theta(t)-\Theta_{0}|\leq\int_{0}^{\infty}|m|dt.

Rebuild VV using V=|V|(cosΘ(t),sinΘ(t))V=|V|(\cos\Theta(t),\sin\Theta(t)) and from it find
R(t)=(c0,1,c0,2)+0tV𝑑t.\displaystyle R(t)=(c_{0,1},c_{0,2})+\int_{0}^{t}Vdt. Note that Θ0\Theta_{0} captures the indeterminacy due to the rotational invariance of (1) and (c0,1,c0,2)(c_{0,1},c_{0,2}) captures the indeterminacy due to the translation invariance of (1). Finally, we can recover the directions of the vectors from ϕt\phi_{t} using

rk=R(t)+xkV|V|+ykV|V| and r˙k=ukV|V|+wkV|V|.r_{k}=R(t)+x_{k}\frac{V}{|V|}+y_{k}\frac{V^{\bot}}{|V|}\;\;\mbox{ and }\;\dot{r}_{k}=u_{k}\frac{V}{|V|}+w_{k}\frac{V^{\bot}}{|V|}. (28)

Once the convergence of limt|V(t)|=1\lim_{t\to\infty}|V(t)|=1 is established, assertion (8) reduces to proving the convergence of the angle Θ\Theta, which would follow from the convergence of 0|m|𝑑t.\int_{0}^{\infty}|m|dt.

Remark 2.10.

Let 𝒰\mathcal{U} be the open subset of 4n1\mathbb{R}^{4n-1} where 1nk=1nuk>0.\frac{1}{n}\sum_{k=1}^{n}u_{k}>0. Let t0,Z0t_{0},Z_{0} be such that ψt0(Z0)𝒰.\psi_{t_{0}}(Z_{0})\in\mathcal{U}. Then there exists a time interval (t0ϵ,t0+ϵ)(t_{0}-\epsilon,t_{0}+\epsilon) such that the function tψt(Z0)t\to\psi_{t}(Z_{0}) is analytic on that interval.

This holds since the vector field (23) is defined by algebraic operations on m,w,y,x,u.m,w,y,x,u. The condition that ψt0(Z0)𝒰\psi_{t_{0}}(Z_{0})\in\mathcal{U} ensures that the fraction 1|V|=nuk\displaystyle\frac{1}{|V|}=\frac{n}{\sum u_{k}} used to define mm in terms of w,y,x,uw,y,x,u (see (21)) is real analytic in a neighborhood of ψt0(Z0).\psi_{t_{0}}(Z_{0}).

2.1 The Polar Angles in the (wk,yk)(w_{k},y_{k}) Planes

In this subsection we are only interested in solutions ψt(Z0)\psi_{t}(Z_{0}) with t0t\geq 0 that lie in the set 𝒰\mathcal{U} where 1nk=1nuk>0.\frac{1}{n}\sum_{k=1}^{n}u_{k}>0. Consider the projection of ψt(Z0)\psi_{t}(Z_{0}) on the plane (wk,yk);(w_{k},y_{k}); denote by γk\gamma_{k} the curve with parametrization (wk(t),yk(t)).(w_{k}(t),y_{k}(t)).

The upper block of JJ shows that each (wk,yk)(w_{k},y_{k}) plane is an eigenspace for ±i.\pm i. It is natural to recast the dynamics in each of these planes using polar coordinates, wk+iyk=akeiθkw_{k}+iy_{k}=a_{k}e^{i\theta_{k}}(’a’ is for amplitude), but we run into a technical difficulty: we would like to find continuous functions θk(t)\theta_{k}(t) to represent γk\gamma_{k} as γk=wk(t)2+yk(t)2eiθk(t).\displaystyle\gamma_{k}=\sqrt{w_{k}(t)^{2}+y_{k}(t)^{2}}e^{i\theta_{k}(t)}. If γk(t)0\gamma_{k}(t)\neq 0 for all t,t, then a smooth polar angle θk(t)\theta_{k}(t) exists. However, the property γk=0\gamma_{k}=0 is not invariant under the flow: the origin, a1==an=0a_{1}=\dots=a_{n}=0 is a fixed point, but points such as (a1,a2,0,an)(a_{1},a_{2},\dots 0,\dots a_{n}) are not, meaning that γk\gamma_{k} could go through (0,0)(0,0) and in doing so it could enter then emerge at different polar angles, forcing a jump in θk\theta_{k} (see Figure 5).

Refer to caption
Figure 5: For curves with smooth parametrization a continuous polar angle θ(t)\theta(t) may not exist if they go through the origin, as the illustrated curves do (when t=0t=0). Left: the cycloid (tsint,1cost)(t-\sin t,1-\cos t) has θ(0)=π/2=θ(0+)\theta(0^{-})=\pi/2=\theta(0^{+}) so its polar angle is continuous. Middle: the parabola (t,tt2)(t,t-t^{2}) has its polar angle jump from θ(0)=3π/4\theta(0^{-})=-3\pi/4 to θ(0+)=π/4.\theta(0^{+})=\pi/4. Right: the smooth curve e1/|t|(cos(1/t),sin(1/t))e^{-1/|t|}(\cos(1/t),\sin(1/t)), shown as a solid curve for t>0t>0 and dotted curve for t<0.t<0. Its polar angle has θ(0+)=,\theta(0^{+})=\infty, so it admits no continuous extension.
Proposition 2.11.

(i) If k,t0k,t_{0} and ϵ\epsilon are such that γk(t0)=0\gamma_{k}(t_{0})=0 with γk(t)0\gamma_{k}(t)\neq 0 for 0<|tt0|<ϵ0<|t-t_{0}|<\epsilon , then there exits a continuous function ϕk(t)\phi_{k}(t) such that one of the following is true

γk(t)=|γk(t)|eiϕk(t) for all t with  0<|tt0|<ϵ or\gamma_{k}(t)=|\gamma_{k}(t)|e^{i\phi_{k}(t)}\;\mbox{ for all }t\mbox{ with }\;0<|t-t_{0}|<\epsilon\;\;\;\;\;\;\;\mbox{ or}
γk(t)=|γk|eiϕk(t) for t<t0andγk(t)=|γk|eiϕk(t)+πi for t0<t\gamma_{k}(t)=|\gamma_{k}|e^{i\phi_{k}(t)}\;\mbox{ for }t<t_{0}\;\;{and}\;\;\gamma_{k}(t)=|\gamma_{k}|e^{i\phi_{k}(t)+\pi i}\;\mbox{ for }t_{0}<t

(ii) Let γk\gamma_{k} be such that the solution to (23)(\ref{State}) remains in 𝒰\mathcal{U} for all t>0,t>0, with γk\gamma_{k} not identically zero. Then there exists a polar angle θk(t)\theta_{k}(t) such that γk=|γk(t)|eiθk(t)\displaystyle\gamma_{k}=|\gamma_{k}(t)|e^{i\theta_{k}(t)} and such that the functions sin(2θk(t))\sin(2\theta_{k}(t)) and cos(2θk(t))\cos(2\theta_{k}(t)) are continuous. If PP is any continuously differentiable, periodic function with period π\pi, the function P(θk(t))P(\theta_{k}(t)) is continuous and piecewise differentiable.

Proof.

(i) We will use the analyticity of wkw_{k} and yky_{k} and the injectivity of the tangent modulo π\pi ( tanβ1=tanβ2\tan\beta_{1}=\tan\beta_{2} implies β1β2=0mod π\beta_{1}-\beta_{2}=0\;\mbox{mod }\pi).

A polar angle α\alpha is well defined and continuous on (t0ϵ,t0)(t_{0}-\epsilon,t_{0}) and on (t0,t0+ϵ),(t_{0},t_{0}+\epsilon), due to γk0\gamma_{k}\neq 0. We need to show that at t0t_{0} the angle has sided limits that are equal or that differ by a multiple of π.\pi.

If one of wkw_{k} or yky_{k} is identically zero, γk\gamma_{k} is along the coordinate axis; its polar angle is either {0,π}\{0,\pi\} or ±π/2.\pm\pi/2. If neither wkw_{k} nor yky_{k} are identically zero let c1(tt0)d1c_{1}(t-t_{0})^{d_{1}} be the first nonzero term in the Taylor series for wkw_{k} and let c2(tt0)d2c_{2}(t-t_{0})^{d_{2}} be the first nonzero term for yk.y_{k}. If d1=d2d_{1}=d_{2} then γk\gamma_{k} is tangent to the vector (c1,c2)(c_{1},c_{2}) so the two sided limits of α\alpha exist and are in the set arctan(c2/c1)+π.\arctan(c_{2}/c_{1})+\pi\mathbb{Z}. If d1<d2d_{1}<d_{2} then γk\gamma_{k} is tangent to the yky_{k} axis and the polar angles have sided limits in the set π/2+π.-\pi/2+\pi\mathbb{Z}. If d2<d1d_{2}<d_{1} then γk\gamma_{k} is tangent to the wkw_{k} axis and the polar angles have sided limits in the set π,\pi\mathbb{Z}, so they differ by a multiple of π.\pi. Let pπ=α(t0+)α(t0)p\pi=\alpha(t_{0}^{+})-\alpha(t_{0}^{-}) and let HH be the Heaviside function. Then ϕk(t)=α(t)pπH(tt0)\phi_{k}(t)=\alpha(t)-p\pi H(t-t_{0}) is continuous with γk0(t)=±|γk0(t)|eiϕk(t).\gamma_{k_{0}}(t)=\pm|\gamma_{k_{0}}(t)|e^{i\phi_{k}(t)}.

For (ii): there is nothing to prove if γk\gamma_{k} is nonzero. Assume that γk=0\gamma_{k}=0 at some positive times. Due to the analyticity of γk\gamma_{k} the set of times when γk=0\gamma_{k}=0 is finite or consists of times 0t1<t2<0\leq t_{1}<t_{2}<\dots going to infinity. By (i), the polar angles corresponding to the intervals (0,t1),(t1,t2),(0,t_{1}),\;(t_{1},t_{2}),\dots have well defined side limits that differ by 0π0\pi or π\pi at t1,t2,,t_{1},t_{2},\dots, meaning that the double angles coincide modulo 2π.2\pi. This implies that sin(2ϕk(t))\sin(2\phi_{k}(t)) and cos(2ϕk(t))\cos(2\phi_{k}(t)) are continuous. The continuity properties for PP follow since PP is the sum of a Fourier series that converges in the sup norm. ∎

Notations 2.12.

(i) For the components wk(t),yk(t)w_{k}(t),y_{k}(t) of a solution to (23)(\ref{State}) denote by ak(t)=wk2+yk2a_{k}(t)=\sqrt{w_{k}^{2}+y_{k}^{2}} and by θk(t)\theta_{k}(t) the usual continuous polar coordinates if (wk(t),yk(t))(0,0)(w_{k}(t),y_{k}(t))\neq(0,0); let θk(t)\theta_{k}(t) be the piecewise smooth polar angle from Remark 2.11, part (ii) if (wk(t),yk(t))(w_{k}(t),y_{k}(t)) is not the identically zero function, and let ak=0,θk(t)=ta_{k}=0,\theta_{k}(t)=t if (wk(t),yk(t))=(0,0)(w_{k}(t),y_{k}(t))=(0,0) for all t.t. Thus

wk=akcosθk,yk=aksinθk.w_{k}=a_{k}\cos\theta_{k},\;y_{k}=a_{k}\sin\theta_{k}. (29)

(ii) Denote by aMaxa_{Max} or simply aMa_{M} the Lipschitz continuous function

aM(t)=max{ak(t),k=1,..n}.a_{M}(t)=\max\{a_{k}(t),k=1,..n\}. (30)

Note that for all smooth functions PP of period π\pi the functions P(θk(t))P(\theta_{k}(t)) are continuous and piecewise smooth.

We conclude this section by noting that although we will only need the polar coordinates representation for the trajectories confined to the central manifold, we had to introduce θk\theta_{k} while working with analytic functions of t.t. Central manifolds and their flows are not necessarily C,C^{\infty}, let alone analytic.

3 The Dynamics on The Central Manifold

We shift coordinates in order to bring the fixed point ZfZ_{f} at the origin of 4n1:\mathbb{R}^{4n-1}:

Notations 3.1.

Let zk=uk1z_{k}=u_{k}-1 for k=1..nk=1..n and let z¯=1nk=1nzk\bar{z}=\frac{1}{n}\sum\limits_{k=1}^{n}z_{k} (thus z¯=|V|1).\bar{z}=|V|-1).
Let F=F(w,y,x,z)F=F(w,y,x,z) denote the vector field in 4n1\mathbb{R}^{4n-1} given by

F=[(2zkzk2wk2)wkykm(1+zk)wkmxkzk1nk=1nzl+myk(2zkzk2wk2)(1+zk)xk+mwk(2znzn2wn2)(1+zn)+l=1n1xl+mwn].F=\left[\begin{array}[]{l}(-2z_{k}-z_{k}^{2}-w_{k}^{2})w_{k}-y_{k}-m(1+z_{k})\\ w_{k}-mx_{k}\\ \hline\cr z_{k}-\frac{1}{n}\sum\limits_{k=1}^{n}z_{l}+my_{k}\\ (-2z_{k}-z_{k}^{2}-w_{k}^{2})(1+z_{k})-x_{k}+mw_{k}\\ (-2z_{n}-z_{n}^{2}-w_{n}^{2})(1+z_{n})+\sum\limits_{l=1}^{n-1}x_{l}+mw_{n}\end{array}\right]. (31)

where m=1|V|1nl=1n(1ul2wl2)wl=11+1nl=1nzl1nl=1n(2zkzk2wk2)wl\displaystyle m=\frac{1}{|V|}\frac{1}{n}\sum\limits_{l=1}^{n}(1-u_{l}^{2}-w_{l}^{2})w_{l}=\frac{1}{1+\frac{1}{n}\sum\limits_{l=1}^{n}z_{l}}\frac{1}{n}\sum\limits_{l=1}^{n}(-2z_{k}-z_{k}^{2}-w_{k}^{2})w_{l} , with the index kk being in 1..n1..n for the top components, and kk in 1..n11..n-1 for the components below the line.

Note that we used (1uk2wk2)=(2zkzk2wk2).(1-u_{k}^{2}-w_{k}^{2})=(-2z_{k}-z_{k}^{2}-w_{k}^{2}). Also: z¯=|V|1\overline{z}=|V|-1 gives the rate of convergence of the mean field speed to one. We get

m=11+z¯1nk=1n(2zkzk2wk2)wk.m=\frac{1}{1+\overline{z}}\frac{1}{n}\sum_{k=1}^{n}(-2z_{k}-z_{k}^{2}-w_{k}^{2})w_{k}. (32)

In the (w,y,x,z)(w,y,x,z) coordinates the 4n14n-1 dynamical system (23) becomes

col(w˙,y˙,x˙,z˙)=F(w,y,x,z).col(\dot{w},\dot{y},\dot{x},\dot{z})=F(w,y,x,z).

An alternate expression for x˙k\dot{x}_{k} is x˙k=zkz¯+myk.\displaystyle\dot{x}_{k}=z_{k}-\overline{z}+my_{k}.

We want to prove that the system (31) has the origin as an asymptotically stable fixed point. The Jacobian matrix DFDF for (31) at the origin is the same as that of (23) at the point Zf,Z_{f}, given in (27), meaning that the central subspace for DF(0)DF(0) is associated with the eigenvalues ±i\pm i, has dimension 2n2n, and it consists of the points col(w,y,0,0).col(w,y,0,0).

3.1 Approximation of the Map of The Central Manifold

This subsection uses the standard PDE technique from [11] to find a Taylor approximation for the central manifold map of (31) and to obtain differential equations for the flow on the central manifold.

All the calculations from this subsection and beyond describe the solutions of (31) for as long as they stay within a small distance δ\delta from the origin. δ\delta is assumed small enough so that the local central manifold is defined within the ball BδB_{\delta}, with the central manifold map hh having continuous 5th5^{th} order derivatives.

We use the customary big-O notation 𝒪(|ξ|q)\mathcal{O}(|\xi|^{q}) to convey a function that has absolute value at most a positive multiple of |ξ|q|\xi|^{q}. For example both |w||w| and |y||y| are 𝒪(aM),\mathcal{O}(a_{M}), since aMa_{M} is the largest amplitude among (wk,yk).(w_{k},y_{k}).

Notations 3.2.

For a positive qq we use 𝒪q\mathcal{O}_{q} to denote 𝒪(aMq).\mathcal{O}(a_{M}^{q}).

Theorem 3.3.

The flow on the central manifold WcW^{c} of (31) is governed by the 2n2n dimensional system

w˙k=ykm+wksk+wk𝒪4+m𝒪2,y˙k=wk+m𝒪2 for k=1..n, where sk=(1725wk21225wkyk+825yk2)+(43200σww+1100σwy+57200σyy)σww=1nk=1nwk2,σwy=1nk=1nwkyk,σyy=1nk=1nyk2 and m=1nk=1nskwk+𝒪5=𝒪3.\begin{array}[]{l}\dot{w}_{k}=-y_{k}-m+w_{k}s_{k}+w_{k}\mathcal{O}_{4}+m\mathcal{O}_{2},\\ \dot{y}_{k}=w_{k}+m\mathcal{O}_{2}\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\mbox{ for }k=1..n,\;\mbox{ where }\\ \\ s_{k}=-\left(\frac{17}{25}w_{k}^{2}-\frac{12}{25}w_{k}y_{k}+\frac{8}{25}y_{k}^{2}\right)+\left(\frac{43}{200}\sigma_{ww}+\frac{1}{100}\sigma_{wy}+\frac{57}{200}\sigma_{yy}\right)\\ \\ \sigma_{ww}=\frac{1}{n}\sum\limits_{k=1}^{n}w_{k}^{2},\;\;\;\sigma_{wy}=\frac{1}{n}\sum\limits_{k=1}^{n}w_{k}y_{k},\;\;\sigma_{yy}=\frac{1}{n}\sum\limits_{k=1}^{n}y_{k}^{2}\;\;\;\mbox{ and }\\ m=\frac{1}{n}\sum\limits_{k=1}^{n}s_{k}w_{k}+\mathcal{O}_{5}=\mathcal{O}_{3}.\end{array} (33)
Proof.

Let hh denote the map whose graph is the central manifold, as in [11]. In our notations, the central manifold is described by col(x,z)=h(w,y).\displaystyle col(x,z)=h(w,y). Since the projections onto the (x,z)(x,z) components of the vector fields F(w,y,x,z)F(-w,-y,x,z) and F(w,y,x,z)F(w,y,x,z) are equal, the map hh can be constructed to be even, h(w,y)=h(w,y).h(-w,-y)=h(w,y). In particular the Taylor expansion of hh has no odd-power terms.

For a scalar function ϕ=ϕ(w,y)\phi=\phi(w,y) denote by DwϕD_{w}\phi the 1×n1\times n row vector of the partial derivatives with w1,wn,w_{1},\dots w_{n}, and denote by DyϕD_{y}\phi the row vector of the derivatives with y1,yn.y_{1},\dots y_{n}.

We seek to produce an approximation [XZ]\displaystyle\left[\begin{array}[]{l}X\\ Z\end{array}\right] of hh such as [XZ]h=𝒪4.\displaystyle\left[\begin{array}[]{l}X\\ Z\end{array}\right]-h=\mathcal{O}_{4}. Given that hh has no third degree terms in its expansion, we seek quadratic polynomials such that [XZ]h=𝒪4.\displaystyle\left[\begin{array}[]{l}X\\ Z\end{array}\right]-h=\mathcal{O}_{4}. We separate the terms of (31) that are smaller than the needed precision. A priory xk=𝒪2x_{k}=\mathcal{O}_{2} and zk=𝒪2.z_{k}=\mathcal{O}_{2}. Their gradients are 𝒪1.\mathcal{O}_{1}. We get

(2zkzk2wk2)=(1uk2wk2)=2zkwk2+𝒪4=𝒪21|V|=11+1nk=1nzn=1+𝒪2m=1|V|1nk=1n(1uk2wk2)wk=11+𝒪21nk=1n𝒪2𝒪1=𝒪3(2zkzk2wk2)(1+zk)=(2zkzk2wk2)+𝒪4\begin{array}[]{l}(-2z_{k}-z_{k}^{2}-w_{k}^{2})=(1-u_{k}^{2}-w_{k}^{2})=-2z_{k}-w_{k}^{2}+\mathcal{O}_{4}=\mathcal{O}_{2}\\ \\ \frac{1}{|V|}=\frac{1}{1+\frac{1}{n}\sum\limits_{k=1}^{n}z_{n}}=1+\mathcal{O}_{2}\\ \\ m=\frac{1}{|V|}\frac{1}{n}\sum\limits_{k=1}^{n}(1-u_{k}^{2}-w_{k}^{2})w_{k}=\frac{1}{1+\mathcal{O}_{2}}\frac{1}{n}\sum\limits_{k=1}^{n}\mathcal{O}_{2}\mathcal{O}_{1}=\mathcal{O}_{3}\\ \\ (-2z_{k}-z_{k}^{2}-w_{k}^{2})(1+z_{k})=(-2z_{k}-z_{k}^{2}-w_{k}^{2})+\mathcal{O}_{4}\end{array} (34)

We conclude that

w˙k=yk2zkwkwk3m+𝒪4=yk+𝒪3y˙k=wkmxk=wk+m𝒪2x˙k=zk1nl=1nzl+𝒪4z˙k=2zkxkwk2+𝒪4z˙n=2zn+l=1n1xl(k=1n1wl)2+𝒪4\begin{array}[]{l}\dot{w}_{k}=-y_{k}-2z_{k}w_{k}-w_{k}^{3}-m+\mathcal{O}_{4}=-y_{k}+\mathcal{O}_{3}\\ \dot{y}_{k}=w_{k}-mx_{k}=w_{k}+m\mathcal{O}_{2}\\ \\ \dot{x}_{k}=z_{k}-\frac{1}{n}\sum\limits_{l=1}^{n}z_{l}+\mathcal{O}_{4}\\ \dot{z}_{k}=-2z_{k}-x_{k}-w_{k}^{2}+\mathcal{O}_{4}\\ \dot{z}_{n}=-2z_{n}+\sum\limits_{l=1}^{n-1}x_{l}-\left(\sum\limits_{k=1}^{n-1}w_{l}\right)^{2}+\mathcal{O}_{4}\end{array} (35)

Following the approximation technique from [11] Theorem 3, page 5, based on the vector field (35), we need quadratic polynomials XkX_{k} with k=1..n1k=1..n-1 and ZkZ_{k} with k=1..nk=1..n such that the differences δX,δZ\delta^{X},\delta^{Z} given below are 𝒪4.\mathcal{O}_{4}.

δkX=[DwXkDyXk][y+𝒪3w+𝒪3](Zk1nl=1nZl+𝒪4)δkZ=[DwZkDyZk][y+𝒪3w+𝒪3](2ZkXkwk2+𝒪4)δnZ=[DwZnDyZn][y+𝒪3w+𝒪3](2Zn+l=1n1Xl(k=1n1wl)2+𝒪4)\begin{array}[]{l}\delta^{X}_{k}=[D_{w}X_{k}\;D_{y}X_{k}]\left[\begin{array}[]{r}-y+\mathcal{O}_{3}\\ w+\mathcal{O}_{3}\end{array}\right]-\left(Z_{k}-\frac{1}{n}\sum\limits_{l=1}^{n}Z_{l}+\mathcal{O}_{4}\right)\\ \delta^{Z}_{k}=[D_{w}Z_{k}\;D_{y}Z_{k}]\left[\begin{array}[]{r}-y+\mathcal{O}_{3}\\ w+\mathcal{O}_{3}\end{array}\right]-\left(-2Z_{k}-X_{k}-w_{k}^{2}+\mathcal{O}_{4}\right)\\ \delta^{Z}_{n}=[D_{w}Z_{n}\;D_{y}Z_{n}]\left[\begin{array}[]{r}-y+\mathcal{O}_{3}\\ w+\mathcal{O}_{3}\end{array}\right]-\left(-2Z_{n}+\sum\limits_{l=1}^{n-1}X_{l}-(\sum\limits_{k=1}^{n-1}w_{l})^{2}+\mathcal{O}_{4}\right)\end{array}

For a more compact notation, denote by LL the linear differential operator
Lf=(Dwf)y+(Dyf)w.Lf=-(D_{w}f)y+(D_{y}f)w. Moving the error terms to the left, we get

δkX+𝒪4=LXk(Zk1nl=1nZl)δkZ+𝒪4=LZk(2ZkXkwk2)δnZ+𝒪4=LZn(2Zn+l=1n1Xl(k=1n1wl)2)\begin{array}[]{l}\delta^{X}_{k}+\mathcal{O}_{4}=LX_{k}-\left(Z_{k}-\frac{1}{n}\sum\limits_{l=1}^{n}Z_{l}\right)\\ \delta^{Z}_{k}+\mathcal{O}_{4}=LZ_{k}-\left(-2Z_{k}-X_{k}-w_{k}^{2}\right)\\ \delta^{Z}_{n}+\mathcal{O}_{4}=LZ_{n}-\left(-2Z_{n}+\sum\limits_{l=1}^{n-1}X_{l}-(\sum\limits_{k=1}^{n-1}w_{l})^{2}\right)\end{array} (36)

We meet the 𝒪4\mathcal{O}_{4} requirement for the differences δX,δZ\delta^{X},\delta^{Z} if the right hand side of (36) is identically zero. It is more expedient to use Z¯\overline{Z} defined as

Z¯=1nl=1nZl\overline{Z}=\frac{1}{n}\sum_{l=1}^{n}Z_{l}

rather than ZnZ_{n}. To achieve 𝒪4\mathcal{O}_{4} precision we need that for k=1..n1k=1..n-1 the following hold:

LXk(ZkZ¯)=0LZk+2Zk+Xk=wk2LZ¯+2Z¯=1nk=1nwk2\begin{array}[]{l}LX_{k}-(Z_{k}-\overline{Z})=0\\ LZ_{k}+2Z_{k}+X_{k}=-w_{k}^{2}\\ L\overline{Z}+2\overline{Z}=-\frac{1}{n}\sum\limits_{k=1}^{n}w_{k}^{2}\end{array}

Note that by subtracting the last equation from the middle row equations we get

LXk(ZkZ¯)=0L(ZkZ¯)+2(ZkZ¯)+Xk=(wk21nk=1nwk2)LZ¯+2Z¯=1nk=1nwk2\begin{array}[]{l}LX_{k}-(Z_{k}-\overline{Z})=0\\ L(Z_{k}-\overline{Z})+2(Z_{k}-\overline{Z})+X_{k}=-(w_{k}^{2}-\frac{1}{n}\sum\limits_{k=1}^{n}w_{k}^{2})\\ L\overline{Z}+2\overline{Z}=-\frac{1}{n}\sum\limits_{k=1}^{n}w_{k}^{2}\end{array} (37)

Before proceeding to finding the quadratic polynomials Xk,Zk,Z¯,X_{k},Z_{k},\bar{Z}, we perform preliminary calculations to determine how LL acts on simple quadratic terms. We compute Lwk2=2wkyk,Lwkyk=wk2yk2,Lyk2=2wkyk.Lw_{k}^{2}=-2w_{k}y_{k},\;\;Lw_{k}y_{k}=w_{k}^{2}-y_{k}^{2},\;\;Ly_{k}^{2}=2w_{k}y_{k}. From these and

σww=1nk=1nwk2,σwy=1nk=1nwkyk,,σyy=1nk=1nyk2.\sigma_{ww}=\frac{1}{n}\sum_{k=1}^{n}w_{k}^{2},\;\sigma_{wy}=\frac{1}{n}\sum_{k=1}^{n}w_{k}y_{k},,\;\;\sigma_{yy}=\frac{1}{n}\sum_{k=1}^{n}y_{k}^{2}. (38)

we get

Lσww=2σwy and L(wk2σww)=2(wkykσwy)Lσwy=σwwσyyL(wkykσwy)=(wk2σww)(yk2σyy)Lσyy=2σwyL(yk2σyy)=2(wkykσwy)\begin{array}[]{cllll}L\sigma_{ww}&=-2\sigma_{wy}&\mbox{ and }&L(w_{k}^{2}-\sigma_{ww})&=-2(w_{k}y_{k}-\sigma_{wy})\\ L\sigma_{wy}&=\sigma_{ww}-\sigma_{yy}&&L(w_{k}y_{k}-\sigma_{wy})&=(w_{k}^{2}-\sigma_{ww})-(y_{k}^{2}-\sigma_{yy})\\ L\sigma_{yy}&=2\sigma_{wy}&&L(y_{k}^{2}-\sigma_{yy})&=2(w_{k}y_{k}-\sigma_{wy})\\ \end{array} (39)

Using vector notation we rewrite (39) as

L[σwwσwyσyy]=[020101020][σwwσwyσyy]L\left[\begin{array}[]{l}\sigma_{ww}\\ \sigma_{wy}\\ \sigma_{yy}\\ \end{array}\right]=\left[\begin{array}[]{ccc}0&-2&0\\ 1&0&-1\\ 0&2&0\end{array}\right]\left[\begin{array}[]{l}\sigma_{ww}\\ \sigma_{wy}\\ \sigma_{yy}\\ \end{array}\right]
L[wk2σwwwkykσwyyk2σyy]=[020101020][wk2σwwwkykσwyyk2σyy].L\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right]=\left[\begin{array}[]{ccc}0&-2&0\\ 1&0&-1\\ 0&2&0\end{array}\right]\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right]. (40)

Let MLM_{L} denote the matrix of the operator, ML=[020101020].\displaystyle M_{L}=\left[\begin{array}[]{ccc}0&-2&0\\ 1&0&-1\\ 0&2&0\end{array}\right].


Motivated by the fact that the original system (1) had index-permutation invariance, and from the identities l=1nxl=0\sum\limits_{l=1}^{n}x_{l}=0 and l=1n(zlz¯)=0,\sum\limits_{l=1}^{n}(z_{l}-\overline{z})=0, we look for quadratic polynomials Xk,(ZkZ¯),Z¯X_{k},(Z_{k}-\overline{Z}),\overline{Z} with similar features. We want to find constants c1,c9c_{1},\dots c_{9} such that for k=1..n1k=1..n-1

Xk=c1(wk2σww)+c2(wkykσwy)+c3(yk2σyy)ZkZ¯=c4(wk2σww)+c5(wkykσwy)+c6(yk2σyy)Z¯=c7σww+c8σwy+c9σyy\begin{array}[]{cl}X_{k}&=c_{1}(w_{k}^{2}-\sigma_{ww})+c_{2}(w_{k}y_{k}-\sigma_{wy})+c_{3}(y_{k}^{2}-\sigma_{yy})\\ &\\ Z_{k}-\overline{Z}&=c_{4}(w_{k}^{2}-\sigma_{ww})+c_{5}(w_{k}y_{k}-\sigma_{wy})+c_{6}(y_{k}^{2}-\sigma_{yy})\\ &\\ \overline{Z}&=c_{7}\sigma_{ww}+c_{8}\sigma_{wy}+c_{9}\sigma_{yy}\end{array}

Group the unknown coefficients into the row vectors cX=[c1,c2,c3]c_{X}=[c_{1},c_{2},c_{3}], cZ=[c4,c5,c6]c_{Z}=[c_{4},c_{5},c_{6}] and c¯=[c7,c8,c9].\overline{c}=[c_{7},c_{8},c_{9}]. We get that

Xk=cX[wk2σwwwkykσwyyk2σyy],ZkZ¯=cZ[wk2σwwwkykσwyyk2σyy],Z¯=c¯[σwwσwyσyy].X_{k}=c_{X}\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right],\;\;Z_{k}-\overline{Z}=c_{Z}\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right],\;\;\overline{Z}=\overline{c}\left[\begin{array}[]{l}\sigma_{ww}\\ \sigma_{wy}\\ \sigma_{yy}\\ \end{array}\right].

From (40) we get

LXk=cXML[wk2σwwwkykσwyyk2σyy],L(ZkZ¯)=cZML[wk2σwwwkykσwyyk2σyy], and L\;X_{k}=c_{X}M_{L}\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right],\;\;L(Z_{k}-\overline{Z})=c_{Z}M_{L}\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right],\mbox{ and }
LZ¯=c¯ML[σwwσwyσyy].L\overline{Z}=\overline{c}M_{L}\left[\begin{array}[]{l}\sigma_{ww}\\ \sigma_{wy}\\ \sigma_{yy}\\ \end{array}\right].

Use matrix notation to rewrite (37) as:

(cXMLcZ)[wk2σwwwkykσwyyk2σyy]=0,(c_{X}\;M_{L}-c_{Z})\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right]=0,
(cZML+2cZ+cX)[wk2σwwwkykσwyyk2σyy]=[100][wk2σwwwkykσwyyk2σyy](c_{Z}M_{L}+2c_{Z}+c_{X})\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right]=\left[\begin{array}[]{lll}-1&0&0\end{array}\right]\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right]

and

(c¯ML+2c¯)[σwwσwyσyy]=[100][σwwσwyσyy](\overline{c}\;M_{L}+2\overline{c})\left[\begin{array}[]{l}\sigma_{ww}\\ \sigma_{wy}\\ \sigma_{yy}\\ \end{array}\right]=\left[\begin{array}[]{lll}-1&0&0\end{array}\right]\left[\begin{array}[]{l}\sigma_{ww}\\ \sigma_{wy}\\ \sigma_{yy}\\ \end{array}\right]

Use identification of coefficients in the latter equality: we require c¯(ML+2I3)=[1 0 0].\displaystyle\overline{c}\;(M_{L}+2I_{3})=[-1\;0\;0]. The solution is c¯=[38,14,18.]\overline{c}=[-\frac{3}{8},-\frac{1}{4},-\frac{1}{8}.] We conclude

Z¯=38σww28σwy18σyy.\overline{Z}=-\frac{3}{8}\sigma_{ww}-\frac{2}{8}\sigma_{wy}-\frac{1}{8}\sigma_{yy}.

By identification of coefficients we require that the 6-dimensional row vector [cX,cZ][c_{X},c_{Z}] solves

[cXcZ][MLI3I3ML+2I3]=[0 0 01 0 0][c_{X}\;c_{Z}]\left[\begin{array}[]{l|c }M_{L}&I_{3}\\ \hline\cr\\ -I_{3}&M_{L}+2I_{3}\end{array}\right]=[0\;0\;0\;\;-1\;0\;0]

We get that cX=[1125,425,1425]\displaystyle c_{X}=[-\frac{11}{25},\;-\frac{4}{25},\;-\frac{14}{25}] and cz=[425,625,425].\displaystyle c_{z}=[-\frac{4}{25},\;-\frac{6}{25},\;\frac{4}{25}].

Recall that Zk=cZ[wk2σwwwkykσwyyk2σyy]+c¯[σwwσwyσyy].Z_{k}=c_{Z}\left[\begin{array}[]{c}w_{k}^{2}-\sigma_{ww}\\ w_{k}y_{k}-\sigma_{wy}\\ y_{k}^{2}-\sigma_{yy}\\ \end{array}\right]+\overline{c}\left[\begin{array}[]{l}\sigma_{ww}\\ \sigma_{wy}\\ \sigma_{yy}\\ \end{array}\right]. We get that for k=1..nk=1..n

Zk=425wk2625wkyk+425yk2+(43200σww1100σwy57200σyy).Z_{k}=-\frac{4}{25}w_{k}^{2}-\frac{6}{25}w_{k}y_{k}+\frac{4}{25}y_{k}^{2}+\left(-\frac{43}{200}\sigma_{ww}-\frac{1}{100}\sigma_{wy}-\frac{57}{200}\sigma_{yy}\right).

Substitute h=col[X,Z]+𝒪4\displaystyle h=col[X,Z]+\mathcal{O}_{4} into the differential equations (35). Up to 𝒪4,\mathcal{O}_{4},

2zkwk2(1725wk21225wkyk+825yk2)+(43100σww+2100σwy+57100σyy),-2z_{k}-w_{k}^{2}\simeq-\left(\frac{17}{25}w_{k}^{2}-\frac{12}{25}w_{k}y_{k}+\frac{8}{25}y_{k}^{2}\right)+\left(\frac{43}{100}\sigma_{ww}+\frac{2}{100}\sigma_{wy}+\frac{57}{100}\sigma_{yy}\right),

thus 2zkwk2=sk+𝒪4.-2z_{k}-w_{k}^{2}=s_{k}+\mathcal{O}_{4}. To complete the approximation of w˙k\dot{w}_{k} note that the exact equation w˙k=(2zkzk2wk2)wkykm(1+zk)\dot{w}_{k}=(-2z_{k}-z_{k}^{2}-w_{k}^{2})w_{k}-y_{k}-m(1+z_{k}) differs from (2zkwk2)wkykm(-2z_{k}-w_{k}^{2})w_{k}-y_{k}-m by zk2wkmzk=wk𝒪4+m𝒪2.z_{k}^{2}w_{k}-mz_{k}=w_{k}\mathcal{O}_{4}+m\mathcal{O}_{2}. The term (2zkwk2)wk(-2z_{k}-w_{k}^{2})w_{k} differs from skwks_{k}w_{k} by 𝒪4wk.\mathcal{O}_{4}w_{k}. Overall,

w˙k=ykm+wksk+m𝒪2+wk𝒪4\dot{w}_{k}=-y_{k}-m+w_{k}s_{k}+m\mathcal{O}_{2}+w_{k}\mathcal{O}_{4}

and y˙k=wk+m𝒪2\dot{y}_{k}=w_{k}+m\mathcal{O}_{2} per (35). Substitute 2zkwk2=sk+𝒪4,zk2=𝒪4-2z_{k}-w_{k}^{2}=s_{k}+\mathcal{O}_{4},\;z_{k}^{2}=\mathcal{O}_{4} and z¯=𝒪2\bar{z}=\mathcal{O}_{2} in m=11+z¯1nk=1n(2zkzk2wk2)wk\displaystyle m=\frac{1}{1+\overline{z}}\frac{1}{n}\sum_{k=1}^{n}(-2z_{k}-z_{k}^{2}-w_{k}^{2})w_{k} from (32) to complete (33).

Remark 3.4.

On the central manifold, the mean speed |V||V| is always under one, and

1|V|C1nl=1n(wl2+yl2)+𝒪4.1-|V|\geq C\frac{1}{n}\sum_{l=1}^{n}(w_{l}^{2}+y_{l}^{2})+\mathcal{O}_{4}.

Proof: Use 1|V|=Z¯+𝒪4=1nl=1n(38wl2+28wlyl+18yl2)\displaystyle 1-|V|=-\overline{Z}+\mathcal{O}_{4}=\frac{1}{n}\sum_{l=1}^{n}(\frac{3}{8}w_{l}^{2}+\frac{2}{8}w_{l}y_{l}+\frac{1}{8}y_{l}^{2}) and the fact that the quadratic form 38a2+28ab+18b2\displaystyle\frac{3}{8}a^{2}+\frac{2}{8}ab+\frac{1}{8}b^{2} is positive definite with minimum equal to C=228(a2+b2).C=\frac{2-\sqrt{2}}{8}(a^{2}+b^{2}).

3.2 Amplitude Variation on The Central Manifold

We consider trajectories (w,y)(w,y) in the subset H×HH\times H of the central manifold, for as long as they remain in the small ball BδB_{\delta} where the central manifold is C5C^{5}. We use the polar coordinate representation from (29) and Remark 2.12.

Let A,EA,E be periodic functions of period π\pi defined as

A(θ)=cos2θ[1725cos2θ1225sinθcosθ+825sin2θ] and E(θ)=43100cos2θ+2100sinθcosθ+57100sin2θ.\begin{array}[]{l}A(\theta)=-\cos^{2}\theta\left[\frac{17}{25}\cos^{2}\theta-\frac{12}{25}\sin\theta\cos\theta+\frac{8}{25}\sin^{2}\theta\right]\;\;\;\mbox{ and }\\ \\ E(\theta)=\frac{43}{100}\cos^{2}\theta+\frac{2}{100}\sin\theta\cos\theta+\frac{57}{100}\sin^{2}\theta.\end{array} (41)

We get from the choice of θk\theta_{k} in (wk,yk)=(akcosθk,aksinθk)(w_{k},y_{k})=(a_{k}\cos\theta_{k},a_{k}\sin\theta_{k}) that the functions A(θk(t))A(\theta_{k}(t)) and E(θk(t))E(\theta_{k}(t)) are continuous for all tt, and piecewise C5.C^{5}. Moreover, non-differentiability of A(θk)A(\theta_{k}) and E(θk)E(\theta_{k}) can only happen for times toot_{oo} and indexes kook_{oo} with wkoo(too)=ykoo(too)=0w_{k_{oo}}(t_{oo})=y_{k_{oo}}(t_{oo})=0. We abbreviate this excepted set of times (this restriction) by oo.\mathcal{E}{oo}.

Proposition 3.5.

Using polar coordinates, the differential equations for wk,ykw_{k},y_{k} become:

a˙k=mcosθk+ak3A(θk)+akcos2θk1nl=1nal2E(θl)+𝒪4θ˙k=1+mak𝒪0+𝒪2, except for oo.\begin{array}[]{l}\dot{a}_{k}=-m\cos\theta_{k}+a_{k}^{3}A(\theta_{k})+a_{k}\cos^{2}\theta_{k}\frac{1}{n}\sum\limits_{l=1}^{n}a_{l}^{2}E(\theta_{l})+\mathcal{O}_{4}\\ \\ \dot{\theta}_{k}=1+\frac{m}{a_{k}}\mathcal{O}_{0}+\mathcal{O}_{2},\;\;\;\;\mbox{ except for }\;\mathcal{E}{oo}.\end{array} (42)

The condition k=1nwk=0\displaystyle\sum\limits_{k=1}^{n}w_{k}=0 rewrites as k=1nakcosθk=0.\displaystyle\sum\limits_{k=1}^{n}a_{k}\cos\theta_{k}=0.

Proof.

Use aka˙k=wkw˙k+yky˙k=mwk+wk2sk+wk𝒪4a_{k}\dot{a}_{k}=w_{k}\dot{w}_{k}+y_{k}\dot{y}_{k}=-mw_{k}+w_{k}^{2}s_{k}+w_{k}\mathcal{O}_{4} per (33). Note that wkak𝒪4\frac{w_{k}}{a_{k}}\mathcal{O}_{4} is 𝒪4.\mathcal{O}_{4}. The periodic functions A,EA,E allow us to rewrite wk2skw_{k}^{2}s_{k} as

wk2sk=ak4A(θk)+ak2cos2θk1nl=1nal2E(θl),w_{k}^{2}s_{k}=a_{k}^{4}A(\theta_{k})+a_{k}^{2}\cos^{2}\theta_{k}\frac{1}{n}\sum_{l=1}^{n}a_{l}^{2}E(\theta_{l}),

and to conclude a˙k=mcosθk+ak3A(θk)+akcos2θk1nl=1nal2E(θl)+𝒪4.\dot{a}_{k}=-m\cos\theta_{k}+a_{k}^{3}A(\theta_{k})+a_{k}\cos^{2}\theta_{k}\frac{1}{n}\sum_{l=1}^{n}a_{l}^{2}E(\theta_{l})+\mathcal{O}_{4}. Approximate θk˙\dot{\theta_{k}} from w˙k=(yk+skwkm)+wk𝒪4+m𝒪2\displaystyle\dot{w}_{k}=(-y_{k}+s_{k}w_{k}-m)+w_{k}\mathcal{O}_{4}+m\mathcal{O}_{2} and y˙k=wk+m𝒪2\dot{y}_{k}=w_{k}+m\mathcal{O}_{2}:

θ˙k=y˙kwkw˙kykak2=1+1ak2(mwk𝒪2skwkyk+mykwkyk𝒪4myk𝒪2)\dot{\theta}_{k}=\frac{\dot{y}_{k}w_{k}-\dot{w}_{k}y_{k}}{a_{k}^{2}}=1+\frac{1}{a_{k}^{2}}(mw_{k}\mathcal{O}_{2}-s_{k}w_{k}y_{k}+my_{k}-w_{k}y_{k}\mathcal{O}_{4}-my_{k}\mathcal{O}_{2})
=1+mak(cosθk𝒪2+sinθk+sinθk𝒪2)sksinθkcosθk+sinθkcosθk𝒪4.=1+\frac{m}{a_{k}}(\cos\theta_{k}\mathcal{O}_{2}+\sin\theta_{k}+\sin\theta_{k}\mathcal{O}_{2})-s_{k}\sin\theta_{k}\cos\theta_{k}+\sin\theta_{k}\cos\theta_{k}\mathcal{O}_{4}.

The factor following mak\frac{m}{a_{k}} has bounded terms, thus is 𝒪0\mathcal{O}_{0}, and θ˙k=1+mak𝒪0+𝒪2.\dot{\theta}_{k}=1+\frac{m}{a_{k}}\mathcal{O}_{0}+\mathcal{O}_{2}.

Per (42), the amplitude aka_{k} is forced down by the negative term ak3A(θk)a_{k}^{3}A(\theta_{k}) and amplified by the positive coupling terms cos2(θk)E(θl).\cos^{2}(\theta_{k})E(\theta_{l}). To understand which of the two has the stronger effect, we examine their average impact over a cycle. Note that A0A\leq 0 and has average μ(A)=1π0πA(θ)𝑑θ=59200=0.295.\mu(A)=\frac{1}{\pi}\int_{0}^{\pi}A(\theta)d\theta=-\frac{59}{200}=-0.295. EE is positive and has average μ(E)=0.5.\mu(E)=0.5. Estimating the impact of agent ll on the amplitude aka_{k} requires we investigate the products between cos2(θ)\cos^{2}(\theta) and all possible shifts E(θ+τ),E(\theta+\tau), as shown in Figure 6. Considering all potential translations of functions, a.k.a. their hulls, is a common practice in dynamical systems involving almost periodic functions that we would like to avoid. Subsection 3.3 provides an alternate approach.

Refer to caption
Figure 6: Left: A plot of the periodic functions A0A\leq 0 and E>0E>0, with their respective means, namely μ(A)=0.295\mu(A)=-0.295 and μ(E)=0.5.\mu(E)=0.5. Right: the functions y=cos2(θ)E(θ+π/2)\displaystyle y=\cos^{2}(\theta)E(\theta+\pi/2) of mean 0.26750.2675 and y=A(θ)y=A(\theta) of mean 0.295.-0.295.

3.3 The Stability of the Zero Solution for Non-autonomous Systems with Cubic Nonlinearities.

This subsection places the proof of stability for (42) from Section 4 in a broader context. However, the proof itself is independent of the content of this subsections, except for the calculations for the mean values of functions (50), Remark 3.8 and Proposition 3.9.

We consider the systems (9) where the unknown x=(x1,xn)Tx=(x_{1},\dots x_{n})^{T} is vector-valued from \mathbb{R} to n\mathbb{R}^{n}, the coefficient functions bk,ck,dk:b_{k},c_{k},d_{k}:\mathbb{R}\to\mathbb{R} are continuous and the remainder functions RkR_{k} defined on n×\mathbb{R}^{n}\times\mathbb{R} satisfy the o3\mbox{\it{o}}_{3} condition:

x˙k=bk(t)xk3+ck(t)xk1nl=1ndl(t)xl2+Rk(x,t), for k=1n.\dot{x}_{k}=b_{k}(t)x_{k}^{3}+c_{k}(t)x_{k}\frac{1}{n}\sum\limits_{l=1}^{n}d_{l}(t)x_{l}^{2}+R_{k}(x,t),\;\mbox{ for }k=1\dots n.

As explained in the introductory section, such systems are related to the dynamics of autonomous systems near fixed points whose Jacobian have purely imaginary roots ω=(±iω1,,±iωn).\omega=(\pm i\omega_{1},\dots,\pm i\omega_{n}). In most practical settings the coefficient functions are combinations of trigonometric functions of periods 2π/ωk,2\pi/\omega_{k}, so they are almost periodic.222 If 𝒯\mathcal{T} denotes the set of all real-valued trigonometric polynomials (all linear combination of cos(λkt)\cos(\lambda_{k}t) and sin(λkt)\sin(\lambda_{k}t), with real λ1,λ2,\lambda_{1},\lambda_{2},\dots ), then a real-valued function ff is called almost periodic if there exists a sequence of trigonometric polynomials Tk𝒯T_{k}\in\mathcal{T} such that TkT_{k} converges to ff in the sup norm on .\mathbb{R}. Let 𝒜\mathcal{A}_{\mathbb{R}} denote the set of real-valued almost periodic functions. 𝒜\mathcal{A}_{\mathbb{R}} is closed under function multiplication and it is a closed linear subspace of the space Cb()C_{b}(\mathbb{R}) of continuous and bounded functions ; in fact 𝒜\mathcal{A}_{\mathbb{R}} is the smallest subalgebra of Cb()C_{b}(\mathbb{R}) containing all the continuous periodic functions ([36]).

Notations 3.6.

Define the mean μ(f)\mu(f) of a function f:[0,)f:[0,\infty)\to\mathbb{R} to be

μ(f)=limP1P0Pf(t)𝑑t, whenever the limit exists.\mu(f)=\lim_{P\to\infty}\frac{1}{P}\int_{0}^{P}f(t)dt,\;\mbox{ whenever the limit exists.} (43)

Let \mathcal{BIM} denote the subspace of bounded, continuous functions that have a mean, and have bounded off-mean antiderivative:

={fCb([0,))|fhas mean ,(fμ(f))𝑑tCb([0,))}.\mathcal{BIM}=\{f\in C_{b}([0,\infty))\;|f\;\mbox{has mean },\int\left(f-\mu(f)\right)dt\in C_{b}([0,\infty))\}. (44)

If ff is a continuous, periodic function with period TT then μ(f)\mu(f) is the customary average value: μ(f)=1T0Tf(t)𝑑t.\mu(f)=\frac{1}{T}\int_{0}^{T}f(t)dt. Moreover, the antiderivative (f(t)μ(f))𝑑t\int(f(t)-\mu(f))dt is periodic, thus f.f\in\mathcal{BIM}. Any linear combination of periodic functions belongs to .\mathcal{BIM}. 333Almost periodic functions have a mean. There exist zero-mean almost periodic functions with unbounded antiderivatives, so the set of almost periodic functions is not a subset of .\mathcal{BIM}.

We can now state the stability result for (9); in keeping with earlier conventions, for a solution vector x=(x1,xn)Tx=(x_{1},\dots x_{n})^{T} use xM=xM(t)x_{M}=x_{M}(t) to denote the largest absolute value among its components: xM(t)=maxk|xk(t)|.x_{M}(t)=\max_{k}|x_{k}(t)|.

Proposition 3.7.

Consider the system (9), with continuous coefficient functions bk,ck,dkb_{k},c_{k},d_{k} and reminders Rk=o3.R_{k}=\mbox{\it{o}}_{3}. Additionally, assume that bkb_{k} and (ck+dk)2(c_{k}+d_{k})^{2} are in .\mathcal{BIM}.

If μ(bk)+1nl=1nμ(14(cl+dl)2)<0\mu(b_{k})+\frac{1}{n}\sum\limits_{l=1}^{n}\mu\left(\frac{1}{4}(c_{l}+d_{l})^{2}\right)<0 for all kk then the x=𝟘nx=\mathbb{0}_{n} solution is asymptotically stable, with xM(t)CxM(0)/tx_{M}(t)\leq Cx_{M}(0)/\sqrt{t} for large t.t.

Proof.

Within this proof, use 𝒪q\mathcal{O}_{q} for 𝒪((maxk|xk|)q).\mathcal{O}\left(\;(\max_{k}|x_{k}|)^{q}\;\right). Work in a small ball BδB_{\delta} centered at the origin in n\mathbb{R}^{n}.

The functions bkb_{k} are in ,\mathcal{BIM}, so the antiderivatives 0t(bk(τ)μ(bk))𝑑τ\int_{0}^{t}(b_{k}(\tau)-\mu(b_{k}))d\tau are bounded, (44); by adding large enough constants, we obtain antiderivatives that are positive and bounded. Let BkB_{k} denote an antiderivative of bkμ(bk)b_{k}-\mu(b_{k}) that has positive (and bounded) range. Let μk=14μ((ck(t)+dk(t))2)\mu_{k}=\frac{1}{4}\mu\left((c_{k}(t)+d_{k}(t))^{2}\right) and let CkC_{k} denote an antiderivative of 14(ck+dk)2μk\frac{1}{4}(c_{k}+d_{k})^{2}-\mu_{k} with positive (and bounded) range.

Define the Lyapunov functions WkW_{k} and WW as

Wk(t)=xk2(t)2xk2(t)(Bk(t))+1nl=1nCl(t))+1, and W=1nk=1nWk.W_{k}(t)=\frac{x_{k}^{2}(t)}{2x_{k}^{2}(t)\left(B_{k}(t))+\frac{1}{n}\sum\limits_{l=1}^{n}C_{l}(t)\right)+1},\;\mbox{ and }\;W=\frac{1}{n}\sum\limits_{k=1}^{n}W_{k}. (45)

For as long as x(t)x(t) is in the ball BδB_{\delta} , the denominators of WkW_{k} are close to 1,1, so WkW_{k} is approximately equal to xk2.x_{k}^{2}. In particular WW and xM2x_{M}^{2} are only within a factor of nn from each other.

Before differentiating WkW_{k} we take note of the following:

Remark 3.8.

For f,gf,g positive functions, with g,g˙g,\dot{g} bounded, and ff small,

ddtf2fg+1=f˙2f2g˙(2fg+1)2=(f˙2f2g˙)(1+𝒪(|f|)).\frac{d}{dt}\frac{f}{2fg+1}=\frac{\dot{f}-2f^{2}\;\dot{g}}{(2fg+1)^{2}}=(\dot{f}-2f^{2}\;\dot{g})(1+\mathcal{O}(|f|)).

To estimate W˙k\dot{W}_{k} apply the latter with f=xk2f=x_{k}^{2} and g=Bk+1nl=1nCl.g=B_{k}+\frac{1}{n}\sum\limits_{l=1}^{n}C_{l}. We get

12W˙k=[xkx˙kxk4(B˙k+1nl=1nC˙l)](1+𝒪2)=\frac{1}{2}\dot{W}_{k}=\left[x_{k}\dot{x}_{k}-x_{k}^{4}\left(\dot{B}_{k}+\frac{1}{n}\sum\limits_{l=1}^{n}\dot{C}_{l}\right)\right](1+\mathcal{O}_{2})=
bkxk4+ckxk2[1nl=1ndlxl2]+xk4(bk+μ(bk)1nl=1n(14(cl+dl)2μl))+𝒪4𝒪2b_{k}x_{k}^{4}+c_{k}x_{k}^{2}\left[\frac{1}{n}\sum\limits_{l=1}^{n}d_{l}x_{l}^{2}\right]+x_{k}^{4}\left(-b_{k}+\mu(b_{k})-\frac{1}{n}\sum\limits_{l=1}^{n}(\frac{1}{4}(c_{l}+d_{l})^{2}-\mu_{l})\right)+\mathcal{O}_{4}\mathcal{O}_{2}
=xk4(μ(bk)+1nl=1nμl)+ckxk2[1nl=1ndlxl2]xk4[1nl=1n14(cl+dl)2]+𝒪6=x_{k}^{4}\left(\mu(b_{k})+\frac{1}{n}\sum\limits_{l=1}^{n}\mu_{l}\right)+c_{k}x_{k}^{2}\left[\frac{1}{n}\sum\limits_{l=1}^{n}d_{l}x_{l}^{2}\right]-x_{k}^{4}\left[\frac{1}{n}\sum\limits_{l=1}^{n}\frac{1}{4}(c_{l}+d_{l})^{2}\right]+\mathcal{O}_{6} (46)

since the terms x˙kxk\dot{x}_{k}x_{k} are 𝒪4\mathcal{O}_{4} and B˙k,C˙l\dot{B}_{k},\dot{C}_{l} are bounded.

Use the constant KK^{\prime} for maxk{μ(bk)+1nl=1nμl}=K.\max_{k}\{\mu(b_{k})+\frac{1}{n}\sum\limits_{l=1}^{n}\mu_{l}\}=-K^{\prime}. By our assumptions K>0K^{\prime}>0 and xk4(μ(bk)+1nl=1nμl)Kxk4.\displaystyle x_{k}^{4}\left(\mu(b_{k})+\frac{1}{n}\sum\limits_{l=1}^{n}\mu_{l}\right)\leq-K^{\prime}x_{k}^{4}. Substituting into (46), and averaging over kk in 1n1\dots n gives:

12W˙K[1nk=1nxk4]+𝒪6+\frac{1}{2}\dot{W}\leq-K^{\prime}\left[\frac{1}{n}\sum\limits_{k=1}^{n}x_{k}^{4}\right]+\mathcal{O}_{6}+
+[1nk=1nckxk2][1nl=1ndlxl2][1nk=1nxk4][1nl=1n(12cl+12dl)2].+\left[\frac{1}{n}\sum\limits_{k=1}^{n}c_{k}x_{k}^{2}\right]\left[\frac{1}{n}\sum\limits_{l=1}^{n}d_{l}x_{l}^{2}\right]-\left[\frac{1}{n}\sum\limits_{k=1}^{n}x_{k}^{4}\right]\left[\frac{1}{n}\sum\limits_{l=1}^{n}(\frac{1}{2}c_{l}+\frac{1}{2}d_{l})^{2}\right].

Because W2W^{2} and 1nk=1nxk4\displaystyle\frac{1}{n}\sum\limits_{k=1}^{n}x_{k}^{4} are comparable, and larger than 𝒪6,\mathcal{O}_{6}, there exists K>0K>0 such that

W˙KW2+2[1nk=1nckxk2][1nl=1ndlxl2]2[1nk=1nxk4][1nl=1n(12cl+12dl)2].\dot{W}\leq-KW^{2}+2\left[\frac{1}{n}\sum\limits_{k=1}^{n}c_{k}x_{k}^{2}\right]\left[\frac{1}{n}\sum\limits_{l=1}^{n}d_{l}x_{l}^{2}\right]-2\left[\frac{1}{n}\sum\limits_{k=1}^{n}x_{k}^{4}\right]\left[\frac{1}{n}\sum\limits_{l=1}^{n}(\frac{1}{2}c_{l}+\frac{1}{2}d_{l})^{2}\right]. (47)

We will need the result of the following Proposition

Proposition 3.9.

Let dd be a positive integer, and let pk,qk,rk,k=1dp_{k},q_{k},r_{k},\;k=1\dots d be scalars in [0,).[0,\infty). Then

[k=1dpkqk][k=1dpkrk][k=1dpk2][k=1d(12qk+12rk)2].\left[\sum\limits_{k=1}^{d}p_{k}q_{k}\right]\left[\sum\limits_{k=1}^{d}p_{k}r_{k}\right]\leq\left[\sum\limits_{k=1}^{d}p_{k}^{2}\right]\left[\sum\limits_{k=1}^{d}\left(\frac{1}{2}q_{k}+\frac{1}{2}r_{k}\right)^{2}\right].
Proof.

Let a=k=1dpkqk\displaystyle a=\sum\limits_{k=1}^{d}p_{k}q_{k} and let b=k=1dpkrk.\displaystyle b=\sum\limits_{k=1}^{d}p_{k}r_{k}. Use the inequality ab(12a+12b)2.\displaystyle ab\leq(\frac{1}{2}a+\frac{1}{2}b)^{2}. From 12a+12b=k=1dpk(12qk+12rk).\displaystyle\frac{1}{2}a+\frac{1}{2}b=\sum\limits_{k=1}^{d}p_{k}(\frac{1}{2}q_{k}+\frac{1}{2}r_{k}). we get

[k=1dpkqk][k=1dpkrk][k=1dpk(12qk+12rk)]2.\left[\sum\limits_{k=1}^{d}p_{k}q_{k}\right]\left[\sum\limits_{k=1}^{d}p_{k}r_{k}\right]\leq\left[\sum\limits_{k=1}^{d}p_{k}(\frac{1}{2}q_{k}+\frac{1}{2}r_{k})\right]^{2}. (48)

For the vectors vv and vv^{\prime} in d\mathbb{R}^{d} with components vk=pkv_{k}=p_{k} and vk=12qk+12rkv^{\prime}_{k}=\frac{1}{2}q_{k}+\frac{1}{2}r_{k} use Cauchy Schwartz inequality to bound their dot product, v,v\langle v,v^{\prime}\rangle by the product of their Euclidean norms |v||v|.|v|\;|v^{\prime}|. We get

0k=1dpk(12qk+12rk)k=1dpk2k=1d(12qk+12rk)2.0\leq\sum\limits_{k=1}^{d}p_{k}(\frac{1}{2}q_{k}+\frac{1}{2}r_{k})\leq\sqrt{\sum\limits_{k=1}^{d}p_{k}^{2}}\;\sqrt{\leq\sum\limits_{k=1}^{d}(\frac{1}{2}q_{k}+\frac{1}{2}r_{k})^{2}}.

or equivalently

[k=1dpk(12qk+12rk)]2[k=1dpk2][k=1d(12qk+12rk)2].\left[\sum\limits_{k=1}^{d}p_{k}(\frac{1}{2}q_{k}+\frac{1}{2}r_{k})\right]^{2}\leq\left[\sum\limits_{k=1}^{d}p_{k}^{2}\right]\left[\sum\limits_{k=1}^{d}\left(\frac{1}{2}q_{k}+\frac{1}{2}r_{k}\right)^{2}\right]. (49)

Combine (48) and (49) to reach the conclusion of Proposition 3.9. ∎

Use Proposition 3.9 with d=nd=n and non-negative scalars pk=xk2,qk=ckp_{k}=x_{k}^{2},\;q_{k}=c_{k} and rk=dk)r_{k}=d_{k}). Conclude that the net contribution of the last two terms of (47) is negative and therefore

W˙KWk2.\dot{W}\leq-KW_{k}^{2}.

The solution to the differential equation a˙=Ka2\dot{a}=-Ka^{2} is a(t)=a(0)1+a(0)Kt.\displaystyle a(t)=\frac{a(0)}{1+a(0)Kt}. We get that W(t)W(0)1+W(0)Kt.\displaystyle W(t)\leq\frac{W(0)}{1+W(0)Kt}. From xM(t)2W(t)x_{M}(t)^{2}\approx W(t) we conclude

xM(t)K′′W(0)1+tW(0)K.x_{M}(t)\leq\frac{K^{\prime\prime}\sqrt{W(0)}}{\sqrt{1+tW(0)K}}.

This proves the asymptotic stability of the origin, and the stated decay rate. ∎

Note that the functions A(t),cos2tA(t),\cos^{2}t and E(t)E(t) defined in (41), used in setting up the polar coordinates system (42) are periodic with common period π.\pi. For these coefficient functions the averages mentioned in Proposition 3.7 can be calculated as:

μ(A)=59200=0.295,μ(14(cos2t+E(t))2)=4371600=.273125,μ(A)+μ(14(cos2t+E(t))2)=7320.\begin{array}[]{l}\mu(A)=-\frac{59}{200}=-0.295,\\ \mu(\frac{1}{4}(\cos^{2}t+E(t))^{2})=\frac{437}{1600}=.273125,\\ \mu(A)+\mu(\frac{1}{4}(\cos^{2}t+E(t))^{2})=-\frac{7}{320}.\end{array} (50)

4 Proving the Stability of the Translating States and of the Convergence Rates

In this section we complete the proof of stability of the translating state of (1) by working with (42), the polar coordinates reformulation of the dynamics on the subspace H×HH\times H of the central manifold system (33). We also prove that the direction of motion for the center of mass of the swarm (1) has a convergent angle Θ(t)\Theta(t) and that all agents’ velocities align with the mean velocity V.V. We give the rates of convergence to the limit configuration for all the relevant physical coordinates of (1).

4.1 Proving the Stability of the Origin

This subsection re-purposes the arguments of Subsection 3.3 to prove that the origin of (42) is asymptotically stable, with convergence rate of 1t,\frac{1}{\sqrt{t}}, as stated in Theorem 4.1.

Due to agents having a common oscillation frequency, one anticipates that the phase angles θk\theta_{k} will synchronize, that a˙k\dot{a}_{k} will become interchangeable with dak/dθkda_{k}/d\theta_{k} and the coefficient functions A(θk)A(\theta_{k}) interchangeable with the time-dependent (rather than phase-dependent) trigonometric polynomials Ak(t)A_{k}(t). However, although the differences θ˙k1\dot{\theta}_{k}-1 are small for some agents k,k, that may not hold true for the particles that are much closer to the origin than those further away. For particles closest to the origin it is possible for θ˙k1mak𝒪0\dot{\theta}_{k}-1\approx\frac{m}{a_{k}}\mathcal{O}_{0} to be very large even though m=𝒪3,m=\mathcal{O}_{3}, meaning that the phase of such particles can be wildly out of sync from the particles with larger amplitudes. Figure 7 illustrates that their amplitudes can have peculiar behaviour as well: while amplitudes of agents generally decrease over time, agents with much smaller magnitudes may undergo prolonged periods of amplitude growth.

Refer to caption
Figure 7: Left: the amplitude a3(t)a_{3}(t) of the agent closest to the origin; the dotted curve represents the average of a3a_{3} over a window of length 2π.2\pi. In average its oscillations increase to catch up with the larger amplitudes of the swarm. Right: a re-scaled illustration of the smallest amplitude a3a_{3} and largest amplitude a1(t)a_{1}(t). In average the larger oscillations decrease.
Theorem 4.1.

Let AA and EE be the periodic functions introduced in (41). For any ϵ>0\epsilon>0 there exists δ>0\delta>0 such that any solution (ak(t),θk(t))(a_{k}(t),\theta_{k}(t)) of

a˙k=mcosθk+ak3A(θk)+akcos2θk1nl=1nal2E(θl)+𝒪4θ˙k=1+mak𝒪0+𝒪2, except at oo\begin{array}[]{l}\dot{a}_{k}=-m\cos\theta_{k}+a_{k}^{3}A(\theta_{k})+a_{k}\cos^{2}\theta_{k}\frac{1}{n}\sum\limits_{l=1}^{n}a_{l}^{2}E(\theta_{l})+\mathcal{O}_{4}\\ \\ \dot{\theta}_{k}=1+\frac{m}{a_{k}}\mathcal{O}_{0}+\mathcal{O}_{2},\;\;\;\;\mbox{ except at }\;\mathcal{E}{oo}\end{array}

with k=1nakcosθk=0\displaystyle\sum\limits_{k=1}^{n}a_{k}\cos\theta_{k}=0 and m=𝒪3m=\mathcal{O}_{3} satisfies: if ak(0)<δa_{k}(0)<\delta for k=1..nk=1..n, then ak(t)<ϵa_{k}(t)<\epsilon for all t>0t>0 and k=1..n.k=1..n.

Moreover, there exist c1,c2c_{1},c_{2} that c1/tmaxkak(t)maxkak(0)c2)/t\displaystyle c_{1}/\sqrt{t}\leq\frac{\max_{k}a_{k}(t)}{\max_{k}a_{k}(0)}\leq c_{2})/\sqrt{t} for large tt.

Proof.

We consider orbits that start with amplitude vector a=(a1,an)a=(a_{1},\dots a_{n}) in a small neighborhood of the fixed point 𝟘nn\mathbb{0}_{n}\in\mathbb{R}^{n}. We will use a Lyapunov function akin to the one in Section 3.3, modified to segregate the agents that oscillate much closer to the origin from the rest.

Per (50) the functions A(θ)+59200\displaystyle A(\theta)+\frac{59}{200} and (12cos2θ+12E(θ))24371600\displaystyle\left(\frac{1}{2}\cos^{2}\theta+\frac{1}{2}E(\theta)\right)^{2}-\frac{437}{1600} have periodic, thus bounded, antiderivatives, with period π.\pi. Let C(θ)C(\theta) be an antiderivative of the latter, and B(θ)B(\theta) be an antiderivative of A(θ)+59200\displaystyle A(\theta)+\frac{59}{200} , both with strictly positive range. 444 The actual formulas for the antiderivatives C(θ)C(\theta) and B(θ)B(\theta) are unimportant in our calculations; however, one could use: B(θ)=A(θ)+59200dθ=17100sin(2θ)9800sin(4θ)325cos4(θ)+1\displaystyle B(\theta)=\int A(\theta)+\frac{59}{200}d\theta=-\frac{17}{100}\sin(2\theta)-\frac{9}{800}\sin(4\theta)-\frac{3}{25}\cos^{4}(\theta)+1 and C(θ)=1+23140000sin(4θ)1400cos(2θ)43160000cos(4θ).\displaystyle C(\theta)=1+\frac{231}{40000}\sin(4\theta)-\frac{1}{400}\cos(2\theta)-\frac{43}{160000}\cos(4\theta).

We use the following influence-diminishing function to weigh the impact of agents. Let Tnorm:[0,1]T_{norm}:\mathbb{R}\to[0,1] be the Lipschitz continuous function given by Tnorm(r)=1T_{norm}(r)=1 for |r|1|r|\geq 1 and Tnorm(r)=r2T_{norm}(r)=r^{2} for rr in [1,1].[-1,1]. Its graph is illustrated in Figure 8.

Refer to caption
Figure 8: The function TnormT_{norm}.

Given a point a=(a1,an)a=(a_{1},\dots a_{n}) in n\mathbb{R}^{n} denote by Tk(a)T_{k}(a) or simply by TkT_{k} the value of T(akaM2.75).\displaystyle T\left(\frac{a_{k}}{a_{M}^{2.75}}\right). Recall that aMa_{M} denotes the largest magnitude. The power 2.752.75 was selected to be just below the order of mm; m=𝒪3.m=\mathcal{O}_{3}. We have:

Tk(a)={ak2aM5.5 if ak<aM2.751 if akaM2.75T_{k}(a)=\begin{cases}\frac{a_{k}^{2}}{a_{M}^{5.5}}&\text{ if }a_{k}<a_{M}^{2.75}\\ 1&\text{ if }a_{k}\geq a_{M}^{2.75}\end{cases} (51)

For notational convenience let L=L(t)L=L(t) and S=S(t)S=S(t) denote subsets of {1,n}\{1,\dots n\} that correspond to agents whose magnitude at that instant is relatively large (LL) or relatively small (SS):

L={k|ak(t)aM2.75(t)},S={k|ak(t)<aM2.75(t)}.L=\{k|a_{k}(t)\geq a_{M}^{2.75}(t)\},\;\;S=\{k|a_{k}(t)<a_{M}^{2.75}(t)\}.

Note that if at some time kLk\in L then Tk=1,T_{k}=1, and if kSk\in S then Tk=ak2aM5.5.T_{k}=\frac{a_{k}^{2}}{a_{M}^{5.5}}.

Lemma 4.2.

If a=(a1,an)a=(a_{1},\dots a_{n}) is a solution to the system in Theorem 4.1, i.e. system (42) , then

ddtTk(a(t))=𝒪0.25 for all k=1..n\frac{d}{dt}T_{k}(a(t))=\mathcal{O}_{0.25}\;\;\mbox{ for all }k=1..n
Proof.

Fix kk in 1..n.1..n. The continuous function TkT_{k} is almost everywhere differentiable and has a weak derivative given almost everywhere by

T˙k={ddtak2aM5.5 if aw<aM2.750 if awaM2.75\dot{T}_{k}=\begin{cases}\frac{d}{dt}\frac{a_{k}^{2}}{a_{M}^{5.5}}&\text{ if }a_{w}<a_{M}^{2.75}\\ 0&\text{ if }a_{w}\geq a_{M}^{2.75}\end{cases}

If kLk\in L then T˙k=0.\dot{T}_{k}=0. For kSk\in S use

dTkdt=2aka˙kaM5.55.5ak2a˙MaM6.5.\frac{dT_{k}}{dt}=2\frac{a_{k}\dot{a}_{k}}{a_{M}^{5.5}}-5.5\frac{a_{k}^{2}\dot{a}_{M}}{a_{M}^{6.5}}.

Note that for al indexes kk the lowest terms in a˙k\dot{a}_{k} are of order three, so

ak|a˙k|aM5.5=akaM2.75|a˙k|aM3aM0.25𝒪0.25 since kS.\frac{a_{k}|\dot{a}_{k}|}{a_{M}^{5.5}}=\frac{a_{k}}{a_{M}^{2.75}}\frac{|\dot{a}_{k}|}{a_{M}^{3}}a_{M}^{0.25}\leq\mathcal{O}_{0.25}\;\;\mbox{ since }k\in S.

For the second term of T˙k\dot{T}_{k} use a˙M=𝒪3\dot{a}_{M}=\mathcal{O}_{3} and ak2aM3.5=(akaM2.75)2aM2\displaystyle\frac{a_{k}^{2}}{a_{M}^{3.5}}=\left(\frac{a_{k}}{a_{M}^{2.75}}\right)^{2}a_{M}^{2} to conclude that for kSk\in S we have ak2a˙MaM6.5=𝒪2.\displaystyle\frac{a_{k}^{2}\dot{a}_{M}}{a_{M}^{6.5}}=\mathcal{O}_{2}. Overall the sum of the two terms is the larger of the parts, i.e. 𝒪0.25.\mathcal{O}_{0.25}.

Lemma 4.3.

If a1,ana_{1},\dots a_{n} and θ1,θn\theta_{1},\dots\theta_{n} are solutions to the system in Theorem 4.1, i.e. system (42), then

(dθk(t)dt1)Tk(a)=𝒪0.25 for all k=1..n.\left(\frac{d\theta_{k}(t)}{dt}-1\right)T_{k}(a)=\mathcal{O}_{0.25}\;\;\;\mbox{ for all }\;k=1..n.
Proof.

The functions TkT_{k} are between zero and one, and θ˙k1=mak𝒪0+𝒪2\displaystyle\dot{\theta}_{k}-1=\frac{m}{a_{k}}\mathcal{O}_{0}+\mathcal{O}_{2} so we only need to prove that makTk=𝒪0.25.\displaystyle\frac{m}{a_{k}}T_{k}=\mathcal{O}_{0.25}.

If kk is an index from LL then Tk=1T_{k}=1 and mak=maM3aM2.75akaM0.25.\displaystyle\frac{m}{a_{k}}=\frac{m}{a_{M}^{3}}\frac{a_{M}^{2.75}}{a_{k}}\;a_{M}^{0.25}. The first fraction of the product is bounded, the second is less or equal to one since kk is an index in LL, leaving mak=𝒪0.25.\displaystyle\frac{m}{a_{k}}=\mathcal{O}_{0.25}.

If kk is an index from S,S, then Tk=ak2aM5.5\displaystyle T_{k}=\frac{a_{k}^{2}}{a_{M}^{5.5}} and thus

makTk=makak2aM5.5=maM3akaM2.75aM0.25=𝒪0.25.\frac{m}{a_{k}}T_{k}=\frac{m}{a_{k}}\frac{a_{k}^{2}}{a_{M}^{5.5}}=\frac{m}{a_{M}^{3}}\frac{a_{k}}{a_{M}^{2.75}}\;a_{M}^{0.25}=\mathcal{O}_{0.25}.

We used the fact that maM3=𝒪0\frac{m}{a_{M}^{3}}=\mathcal{O}_{0} and that when kSk\in S the ratio akaM2.75\frac{a_{k}}{a_{M}^{2.75}} is under one. ∎

Lemma 4.4.

ak2(1Tk(a))aM5.5\displaystyle a_{k}^{2}(1-T_{k}(a))\leq a_{M}^{5.5} and ak2(1Tk(a))aM5.5\displaystyle a_{k}^{2}(1-\sqrt{T_{k}(a)})\leq a_{M}^{5.5}for all k.k.

Proof.

If kk is in LL then Tk(a)=1T_{k}(a)=1 and the inequalities hold. If kk is in S,S, then ak2aM5.5a_{k}^{2}\leq a_{M}^{5.5} and both 1Tk1-T_{k} and 1Tk1-\sqrt{T_{k}} are in [0,1][0,1].

Returning to the proof of Theorem 4.1 :

Define the function Wk=Wk(a1(t),θn(t))W_{k}=W_{k}(a_{1}(t),\dots\theta_{n}(t)) as

Wk(t)=ak2(t)2ak2(t)(B(θk(t))+1nl=1nTl(a)C(θl))+1W_{k}(t)=\frac{a_{k}^{2}(t)}{2a_{k}^{2}(t)\left(B(\theta_{k}(t))+\frac{1}{n}\sum\limits_{l=1}^{n}T_{l}(a)C(\theta_{l})\right)+1} (52)

where BB and CC are the positive, bounded antiderivatives of 59/200+A,59/200+A, and of
(12cos2θ+12E(θ))2437/1600.\displaystyle(\frac{1}{2}\cos^{2}\theta+\frac{1}{2}E(\theta))^{2}-437/1600.

Define WW as W=1nl=1nWl.\displaystyle W=\frac{1}{n}\sum\limits_{l=1}^{n}W_{l}. Note that along trajectories other than the origin, the functions Wk,WW_{k},W are continuous functions of time. While aa is in small neighborhood of the origin, WkW_{k} are the composition between the Lipschitz continuous and piece-wise C5C^{5} functions 1/(1+x)1/(1+x), ak2(t),B(θk(t)),C(θk(t))a_{k}^{2}(t),B(\theta_{k}(t)),C(\theta_{k}(t)) and Tl,T_{l}, therefore the functions tWk(t)t\to W_{k}(t) are differentiable almost everywhere, with weak derivatives that can be computed with the usual chain rule for almost all tt. The denominators of WkW_{k} are very close to 1,1, so each WkW_{k} is comparable in value to ak2a_{k}^{2}. In fact WW and aM2a_{M}^{2} are only within a factor of nn from each other.

We claim that WW satisfies the differential inequality W˙KW2\displaystyle\dot{W}\leq-KW^{2} for some positive constant K.K. We start by estimating just W˙k.\dot{W}_{k}.

From Remark 3.8 we get that

W˙k=[2aka˙k2ak4(dBdθ|θkθk˙+1nl=1ndTldtC(θl)+1nl=1nTldCdθ|θlθl˙)](1+𝒪2)\dot{W}_{k}=\left[2a_{k}\dot{a}_{k}-2a_{k}^{4}\left(\frac{dB}{d\theta}\rvert_{\theta_{k}}\dot{\theta_{k}}+\frac{1}{n}\sum\limits_{l=1}^{n}\frac{dT_{l}}{dt}C(\theta_{l})+\frac{1}{n}\sum\limits_{l=1}^{n}T_{l}\frac{dC}{d\theta}\rvert_{\theta_{l}}\dot{\theta_{l}}\right)\right](1+\mathcal{O}_{2})

Recall that aka˙ka_{k}\dot{a}_{k} and ak4θk˙a_{k}^{4}\dot{\theta_{k}} are 𝒪4\mathcal{O}_{4} and that dBdθ,dCdθ,dTldt\frac{dB}{d\theta},\frac{dC}{d\theta},\frac{dT_{l}}{dt} and TlT_{l} are bounded, implying that the above bracket is 𝒪4.\mathcal{O}_{4}. We get that

12W˙k=[aka˙kak4(dBdθ|θkθk˙+1nl=1ndTldtC(θl)+1nl=1nTldCdθ|θlθl˙)]+𝒪6.\frac{1}{2}\dot{W}_{k}=\left[a_{k}\dot{a}_{k}-a_{k}^{4}\left(\frac{dB}{d\theta}\rvert_{\theta_{k}}\dot{\theta_{k}}+\frac{1}{n}\sum\limits_{l=1}^{n}\frac{dT_{l}}{dt}C(\theta_{l})+\frac{1}{n}\sum\limits_{l=1}^{n}T_{l}\frac{dC}{d\theta}\rvert_{\theta_{l}}\dot{\theta_{l}}\right)\right]+\mathcal{O}_{6}. (53)

Estimate the four terms of 12W˙k.\frac{1}{2}\dot{W}_{k}.

For the first term, use (42) to rewrite aka˙ka_{k}\dot{a}_{k} within an 𝒪5\mathcal{O}_{5} approximation:

aka˙k=makcosθk+ak4A(θk)+ak2cos2θk1nl=1nal2E(θl)+ak𝒪4a_{k}\dot{a}_{k}=-ma_{k}\cos\theta_{k}+a_{k}^{4}A(\theta_{k})+a_{k}^{2}\cos^{2}\theta_{k}\frac{1}{n}\sum\limits_{l=1}^{n}a_{l}^{2}E(\theta_{l})+a_{k}\mathcal{O}_{4}

For third term, ak41nl=1ndTldtC(θl),a_{k}^{4}\frac{1}{n}\sum\limits_{l=1}^{n}\frac{dT_{l}}{dt}C(\theta_{l}), use Lemma 4.2 to conclude that it is 𝒪4.25.\mathcal{O}_{4.25}.

For the second and fourth terms use (42) and Lemma 4.3, respectively:

ak4dBdθ|θk(θk˙1)=dBdθ|θk(mak3𝒪0+ak4𝒪2)=𝒪6 andak4TldCdθ|θl(θl˙1)=ak4dCdθ|θk𝒪0.25=𝒪4.25\begin{array}[]{l}a_{k}^{4}\frac{dB}{d\theta}\rvert_{\theta_{k}}(\dot{\theta_{k}}-1)=\frac{dB}{d\theta}\rvert_{\theta_{k}}(m\;a_{k}^{3}\mathcal{O}_{0}+a_{k}^{4}\mathcal{O}_{2})=\mathcal{O}_{6}\;\;\mbox{ and}\\ \\ a_{k}^{4}T_{l}\frac{dC}{d\theta}\rvert_{\theta_{l}}(\dot{\theta_{l}}-1)=a_{k}^{4}\frac{dC}{d\theta}\rvert_{\theta_{k}}\mathcal{O}_{0.25}=\mathcal{O}_{4.25}\end{array}

Thus in (53), replacing θ˙k,θ˙l\dot{\theta}_{k},\;\dot{\theta}_{l} by 11 and T˙\dot{T} by 0 keeps the approximation within 𝒪4.25\mathcal{O}_{4.25}:

12W˙k=[makcosθk+ak4A(θk)+ak2cos2θk1nl=1nal2E(θl)]+ak4(dBdθ|θk1nl=1nTldCdθ|θl)+𝒪4.25\begin{array}[]{l}\frac{1}{2}\dot{W}_{k}=\left[-ma_{k}\cos\theta_{k}+a_{k}^{4}A(\theta_{k})+a_{k}^{2}\cos^{2}\theta_{k}\frac{1}{n}\sum\limits_{l=1}^{n}a_{l}^{2}E(\theta_{l})\right]+\\ a_{k}^{4}\left(-\frac{dB}{d\theta}\rvert_{\theta_{k}}-\frac{1}{n}\sum\limits_{l=1}^{n}T_{l}\frac{dC}{d\theta}\rvert_{\theta_{l}}\right)+\mathcal{O}_{4.25}\end{array} (54)

From Lemma 4.4 we get ak2cos2θk1nl=1nal2E(θl)\displaystyle a_{k}^{2}\cos^{2}\theta_{k}\frac{1}{n}\sum\limits_{l=1}^{n}a_{l}^{2}E(\theta_{l}) and ak2Tkcos2θk1nl=1nal2TlE(θl)\displaystyle a_{k}^{2}\sqrt{T_{k}}\cos^{2}\theta_{k}\frac{1}{n}\sum\limits_{l=1}^{n}a_{l}^{2}\sqrt{T_{l}}E(\theta_{l}) are within 𝒪5.5\mathcal{O}_{5.5} from each other.

Substitute dBdθ=A+59200\frac{dB}{d\theta}=A+\frac{59}{200} and dCdθ=4371600(12cos2θ+12E(θ))2-\frac{dC}{d\theta}=\frac{437}{1600}-\left(\frac{1}{2}\cos^{2}\theta+\frac{1}{2}E(\theta)\right)^{2} into (54); the terms ak4A(θk)a_{k}^{4}A(\theta_{k}) cancel and we get:

12W˙k=[makcosθk59200ak4+ak2Tkcos2θk1nl=1nal2TlE(θl)]+4371600ak41nl=1nTlak4(1nl=1nTl(12cos2θl+12E(θl))2)+𝒪4.25\begin{array}[]{l}\frac{1}{2}\dot{W}_{k}=\left[-ma_{k}\cos\theta_{k}-\frac{59}{200}a_{k}^{4}+a_{k}^{2}\sqrt{T_{k}}\cos^{2}\theta_{k}\frac{1}{n}\sum\limits_{l=1}^{n}a_{l}^{2}\sqrt{T_{l}}E(\theta_{l})\right]+\\ \frac{437}{1600}a_{k}^{4}\frac{1}{n}\sum\limits_{l=1}^{n}T_{l}-a_{k}^{4}\left(\frac{1}{n}\sum\limits_{l=1}^{n}T_{l}\left(\frac{1}{2}\cos^{2}\theta_{l}+\frac{1}{2}E(\theta_{l})\right)^{2}\right)+\mathcal{O}_{4.25}\end{array} (55)

Since 0Tl10\leq T_{l}\leq 1, 1nl=1nTl1\frac{1}{n}\sum\limits_{l=1}^{n}T_{l}\leq 1 and 59200+43716001nl=1nTl7320.\displaystyle-\frac{59}{200}+\frac{437}{1600}\frac{1}{n}\sum\limits_{l=1}^{n}T_{l}\leq-\frac{7}{320}. Add the equations (55) corresponding to k=1..n,k=1..n, and divide by n.n. Use the fact that l=1nakcosθk=0\sum\limits_{l=1}^{n}a_{k}\cos\theta_{k}=0.

12W˙73201nk=1nak4+[1nk=1nak2Tkcos2θk][1nl=1nal2TlE(θl)]+[1nk=1nak4](1nl=1n(12Tlcos2θl+12TlE(θl))2)+𝒪4.25\begin{array}[]{l}\frac{1}{2}\dot{W}\leq-\frac{7}{320}\frac{1}{n}\sum\limits_{k=1}^{n}a_{k}^{4}+\left[\frac{1}{n}\sum\limits_{k=1}^{n}a_{k}^{2}\sqrt{T_{k}}\cos^{2}\theta_{k}\right]\left[\frac{1}{n}\sum\limits_{l=1}^{n}a_{l}^{2}\sqrt{T_{l}}E(\theta_{l})\right]+\\ -\left[\frac{1}{n}\sum\limits_{k=1}^{n}a_{k}^{4}\right]\left(\frac{1}{n}\sum\limits_{l=1}^{n}\left(\frac{1}{2}\sqrt{T_{l}}\cos^{2}\theta_{l}+\frac{1}{2}\sqrt{T_{l}}E(\theta_{l})\right)^{2}\right)+\mathcal{O}_{4.25}\end{array} (56)

Use Proposition 3.9 with d=nd=n and pk=ak2,qk=Tkcos2θkp_{k}=a_{k}^{2},\;q_{k}=\sqrt{T_{k}}\cos^{2}\theta_{k} and rk=TkE(θk)r_{k}=\sqrt{T_{k}}E(\theta_{k}) to conclude that the net contribution of the two products of sums in (56) is negative, therefore

W˙71601nk=1nak4++𝒪4.25.\dot{W}\leq-\frac{7}{160}\frac{1}{n}\sum\limits_{k=1}^{n}a_{k}^{4}++\mathcal{O}_{4.25}.

The function WW and aM2a_{M}^{2} are comparable in scale, therefore there exists a constant K>0K>0 that depends only on nn such that

W˙KW2,\dot{W}\leq-KW^{2},

meaning that for all t>0t>0

W(t)W(0)1+W(0)Kt and maxkak(t)maxkak(0)c21+W(0)Kt.W(t)\leq\frac{W(0)}{1+W(0)Kt}\;\;\mbox{ and }\frac{\max_{k}a_{k}(t)}{\max_{k}a_{k}(0)}\leq\frac{c_{2}}{\sqrt{1+W(0)Kt}}.

To get the lower bound of C/tC/\sqrt{t} for aM(t),a_{M}(t), or equivalently the lower bound of 1/t1/t for W(t)W(t), return to (55). Let K1K_{1} the maximum value of the periodic function (12cos2θ+12E(θ))2.\left(\frac{1}{2}\cos^{2}\theta+\frac{1}{2}E(\theta)\right)^{2}. We have

W˙kmakcosθk59200ak4+0K1ak4+𝒪4.25\dot{W}_{k}\geq-ma_{k}\cos\theta_{k}-\frac{59}{200}a_{k}^{4}+0-K_{1}a_{k}^{4}+\mathcal{O}_{4.25}

and on account that k=1nmakcosθk=0\sum_{k=1}^{n}ma_{k}\cos\theta_{k}=0 conclude

W˙(59200+K1)ak4+𝒪4.25CW2\dot{W}\geq-(\frac{59}{200}+K_{1})a_{k}^{4}+\mathcal{O}_{4.25}\geq-CW^{2}

ensuring that WW has a lower bound comparable to 1/t.1/t.

Corollary 4.4.1.

For initial conditions Z0H×H×2n1Z_{0}\in H\times H\times\mathbb{R}^{2n-1} near the fixed point ZfZ_{f}, Z0Z_{0} not on the stable manifold Ws(Zf),W^{s}(Z_{f}), there exists C>0C>0 depending on Z0Z_{0} such that

maxk(|rk(t)R(t)|+|r˙k(t)V(t)|)Ct for large t.\max_{k}\left(|r_{k}(t)-R(t)|+|\dot{r}_{k}(t)-V(t)|\right)\geq\frac{C}{\sqrt{t}}\;\;\mbox{ for large }t. (57)
Proof.

Let Z0H×H×2n1Z_{0}\in H\times H\times\mathbb{R}^{2n-1} be near the fixed point Zf,Z_{f}, with Z0Ws(Zf).Z_{0}\notin W^{s}(Z_{f}). Theorem 4.1 implies that the flow of (23) within H×H×2n1\in H\times H\times\mathbb{R}^{2n-1} is asymptotically stable near Zf,Z_{f}, therefore there exists an initial condition Z0,cZ_{0,c} in the central manifold, such that the two solutions ψt(Z0)\psi_{t}(Z_{0}) and ψt(Z0,c)\psi_{t}(Z_{0,c}) are exponentially close for tt in [0,).[0,\infty). Let rc,k,Rc,Vc,wc,k,uc,kr_{c,k},R_{c},V_{c},w_{c,k},\dots u_{c,k} denote the position vectors, velocities and their components for the orbit of Z0,c.Z_{0,c}. Since Z0WS(Zf),Z_{0}\notin W^{S}(Z_{f}), the projection onto the (w,y)(w,y) subspace of the point Z0,cZ_{0,c} is not the origin, so ac,M(0)=maxkwc,k2(0)+yc,k2(0)0.a_{c,M}(0)=\max_{k}\sqrt{w_{c,k}^{2}(0)+y_{c,k}^{2}(0)}\neq 0.

For the solution originating at Z0Z_{0} we have

|rk(t)R(t)|2+|r˙kV|2=xk2+yk2+(uk|V|)2+wk2wk2+yk2=ak2,|r_{k}(t)-R(t)|^{2}+|\dot{r}_{k}-V|^{2}=x_{k}^{2}+y_{k}^{2}+(u_{k}-|V|)^{2}+w_{k}^{2}\geq w_{k}^{2}+y_{k}^{2}=a_{k}^{2},

thus (|rk(t)R(t)|+|r˙kV|)ak.\left(|r_{k}(t)-R(t)|+|\dot{r}_{k}-V|\right)\geq a_{k}. Due to the exponential proximity of the two orbits, there exist positive CC and η\eta such that

|ak(t)ac,k(t)|Ceηt for t[0,).|a_{k}(t)-a_{c,k}(t)|\leq Ce^{-\eta t}\;\mbox{ for }t\in[0,\infty).

We get

maxk(|rk(t)R(t)|+|r˙kV|)maxkak(t)maxkac,k(t)Ceηt,\max_{k}\left(|r_{k}(t)-R(t)|+|\dot{r}_{k}-V|\right)\geq\max_{k}a_{k}(t)\geq\max_{k}a_{c,k}(t)-Ce^{-\eta t},

with maxkac,k(t)ac,M(0)t,\max_{k}a_{c,k}(t)\geq\frac{a_{c,M}(0)}{\sqrt{t}}, per Theorem 4.1. This proves (57). ∎

Theorem 4.1 implies that the flow on the central manifold of (31) is asymptotically stable in H×HH\times H. It also implies the asymptotic stability of the origin for the system (31) with decay rate of order 1/t1/\sqrt{t}, and the asymptotic stability of Zf=col(𝟘n,𝟘n,𝟘n1,𝟙n)Z_{f}=col(\mathbb{0}_{n},\mathbb{0}_{n},\mathbb{0}_{n-1},\mathbb{1}_{n}) for (23) in H×H×2n1.H\times H\times\mathbb{R}^{2n-1}. To complete the proof of Theorem 1.6 we are left to show that the direction of VV is convergent and close to that of V(0),V(0), for which it is sufficient to prove that 0|m|𝑑t\int_{0}^{\infty}|m|dt converges, per Remark (2.9). The solutions of the system (23) stemming from initial conditions near ZfZ_{f} evolve exponentially close to a solution on the center manifold. It is therefore enough to limit the investigation of 0|m|𝑑t\int_{0}^{\infty}|m|\;dt to trajectories from the central manifold.

4.2 Limit Behavior of the Mean Velocity V and of the Physical Coordinates of the Swarm

In this subsection we recast the results obtained for the amplitudes aka_{k} in terms of the physical coordinates of the swarm (1): the positions rkr_{k} and R,R, and the velocities r˙k\dot{r}_{k} and VV. We assume that the initial conditions for (23) are from a small neighborhood \mathcal{B} of ZfZ_{f}, within its basin of attraction.

We begin by addressing the limit behavior of the direction of motion, i.e. the polar angle Θ\Theta satisfying V(t)=|V(t)|eiΘ(t).\displaystyle V(t)=|V(t)|e^{i\Theta(t)}. Recall that the change in Θ(t)\Theta(t) is given by dΘdt=m,\displaystyle\frac{d\Theta}{dt}=m, according to (32) and (21).

Remark 4.5.

There exists Θ\Theta_{\infty} depending on Z0Z_{0} and c>0c>0 such that

|Θ(t)Θ|ct for large t.|\Theta(t)-\Theta_{\infty}|\leq\frac{c}{\sqrt{t}}\;\;\mbox{ for large }t.

Moreover, limtV(t)=(cosΘ,sinΘ).\lim_{t\to\infty}V(t)=(\cos\Theta_{\infty},\sin\Theta_{\infty}).

Proof: Due to the exponentially close proximity of arbitrary orbits to orbits from the central manifold, it is enough to prove the stated assertion for the central manifold flow.

Recall that Θ(t)=Θ0+0tm(s)𝑑s.\Theta(t)=\Theta_{0}+\int_{0}^{t}m(s)ds. For Z0Z_{0} on the central manifold m(t)=𝒪3.m(t)=\mathcal{O}_{3}. We get that for large tt the absolute value of mm is below a multiple of aM(0)3t3/2a_{M}(0)^{3}t^{-3/2} which has a convergent integral at infinity. This proves the existence of limtΘ(t),\displaystyle\lim_{t\to\infty}\Theta(t), denoted Θ.\Theta_{\infty}. Note that |ΘΘ(t)||\Theta_{\infty}-\Theta(t)| is bounded above by tcaM(0)3s3/2𝑑s\displaystyle\int_{t}^{\infty}\frac{ca_{M}(0)^{3}}{s^{3/2}}ds which is of order aM(0)3t1/2.a_{M}(0)^{3}t^{-1/2}.

Let V=(cosΘ,sinΘ).\displaystyle V_{\infty}=(\cos\Theta_{\infty},\sin\Theta_{\infty}). From V(t)=|V(t)|(cosΘ(t),sinΘ(t))V(t)=|V(t)|(\cos\Theta(t),\sin\Theta(t)) with |V|1|V|\to 1 and Θ(t)Θ\Theta(t)\to\Theta_{\infty} we conclude V(t)V(t) converges to (cosΘ,sinΘ).(\cos\Theta_{\infty},\sin\Theta_{\infty}).

Remark 4.6.

If ψt(Z0)\psi_{t}(Z_{0}) is a solution of (23) not in WsW^{s} then there exists c>0c>0 such that the mean speed satisfies 1|V|ct1-|V|\geq\frac{c}{t} for large t.t. Moreover, there exist initial conditions Z0Z_{0} for which the center of mass R(t)R(t) does not stay within bounded distance from rectilinear motion with velocity V,V_{\infty}, in the sense that for any T>0T>0

|R(t+T)R(T)tV| when t.|R(t+T)-R(T)-tV_{\infty}|\to\infty\;\mbox{ when }t\to\infty.

The center of mass falls behind exceedingly large distances compared to where rectilinear motion would have located it.

Proof.

If ψt(Z0)\psi_{t}(Z_{0}) is not in WsW^{s}, then it is exponentially close to a nontrivial solution from the central manifold, ψt(Z0c).\psi_{t}(Z_{0}^{c}). It suffices to prove the bound for Z0c.Z_{0}^{c}. Note that for Z0c0Z_{0}^{c}\neq 0 we have

1|V|=z¯=1nl=1n(38wl2+28wlyl+18yl2)2281nl=1nak2ct.1-|V|=-\overline{z}=\frac{1}{n}\sum_{l=1}^{n}(\frac{3}{8}w_{l}^{2}+\frac{2}{8}w_{l}y_{l}+\frac{1}{8}y_{l}^{2})\geq\frac{2-\sqrt{2}}{8}\frac{1}{n}\sum_{l=1}^{n}a_{k}^{2}\geq\frac{c}{t}.

To prove the existence of configurations with large gap between the center of mass RR and the rectilinear motion, consider agents with mirror symmetry with respect to xx-axis (i.e. m=0m=0). Due to the unidirectional motion, we have that V(t)=(|V|,0)V(t)=(|V|,0) with V=(1,0)V_{\infty}=(1,0) and we get that RR moves along the xx- axis: R(t)=(|R(t)|,0).R(t)=(|R(t)|,0). Identifying RR with its xx- component we get

tV(R(t+T)R(T))=Tt+T(1|V(s)|)𝑑sTT+tcs𝑑sclnt+TTtV_{\infty}-(R(t+T)-R(T))=\int_{T}^{t+T}(1-|V(s)|)ds\geq\int_{T}^{T+t}\frac{c}{s}ds\geq c\ln\frac{t+T}{T}\to\infty

as tt\to\infty. ∎

Remark 4.7.

For n3n\geq 3 the estimates of 𝒪(t3/2)\mathcal{O}(t^{-3/2}) for mm and for the drift Θ\Theta are sharp: there exist initial conditions Z0Z_{0} for which the solution ψt(Z0)\psi_{t}(Z_{0}) of (23) satisfies: there exist c>0c>0 and times t,t′′t^{\prime},t^{\prime\prime} going to infinity such that m(t)c/(t)3/2m(t^{\prime})\geq c/(t^{\prime})^{3/2} and such that the variation in the directions of motion Θ(t)\Theta(t) satisfies

|Θ(t′′)Θ(t)|C/(t)3/2|\Theta(t^{\prime\prime})-\Theta(t^{\prime})|\geq C/(t^{\prime})^{3/2}
Proof.

In order to prove that the estimate m=𝒪3m=\mathcal{O}_{3} is sharp, consider a configuration with n1n-1 of the particles having identical initial conditions, thus identical motion for all tt: wk=w(t),yk=y(t)w_{k}=w(t),y_{k}=y(t) for k=1..n1k=1..n-1. Necessarily, the last particle has wn=(n1)w,yn=(n1)y.w_{n}=-(n-1)w,\;y_{n}=-(n-1)y. Let a1a_{1} and θ\theta denote the common amplitude and polar angle for the first n1n-1 particles.

From Theorem 4.1 there exists c3>0c_{3}>0 such that a1(t)c3/t.a_{1}(t)\geq c_{3}/\sqrt{t}. For this configuration on account that wk=0\sum w_{k}=0 we get

m=1nk=1n(1725wk21225wkyk+825yk2)wk+𝒪5=(n1)(n2)(1725w21225wy+825y2)w+𝒪5c1tt(1725cos2θ1225cosθsinθ+825sin2θ)cosθ+𝒪5\begin{array}[]{l}m=-\frac{1}{n}\sum\limits_{k=1}^{n}\left(\frac{17}{25}w_{k}^{2}-\frac{12}{25}w_{k}y_{k}+\frac{8}{25}y_{k}^{2}\right)w_{k}+\mathcal{O}_{5}=\\ (n-1)(n-2)\left(\frac{17}{25}w^{2}-\frac{12}{25}wy+\frac{8}{25}y^{2}\right)w+\mathcal{O}_{5}\geq\\ c\frac{1}{t\sqrt{t}}\left(\frac{17}{25}\cos^{2}\theta-\frac{12}{25}\cos\theta\sin\theta+\frac{8}{25}\sin^{2}\theta\right)\cos\theta+\mathcal{O}_{5}\end{array} (58)

whenever the trigonometric function is positive.

Refer to caption
Figure 9: The graph of y=(1725cos2θ1225cosθsinθ+825sin2θ)cosθy=\left(\frac{17}{25}\cos^{2}\theta-\frac{12}{25}\cos\theta\sin\theta+\frac{8}{25}\sin^{2}\theta\right)\cos\theta, shown on [2π,2π];[-2\pi,2\pi]; the shaded part has θ[π/3,π/3]\theta\in[-\pi/3,\pi/3] and y>0.10.y>0.10.

Recall that dθdt=1+𝒪2=1+𝒪(1/t),\frac{d\theta}{dt}=1+\mathcal{O}_{2}=1+\mathcal{O}(1/t), meaning that the polar angle θ(t)\theta(t) is strictly increasing (to infinity), is Lipschitz (with constant close to one for intervals near infinity), and θ(t)\theta(t) is comparable in size with t.t. Let NN be large. Let t=t(N)t^{\prime}=t^{\prime}(N) and t′′=t′′(N)t^{\prime\prime}=t^{\prime\prime}(N) be time values when θ(t)=2Nππ/3\theta(t^{\prime})=2N\pi-\pi/3 and θ(t′′)=2Nπ+π/3.\theta(t^{\prime\prime})=2N\pi+\pi/3. We have that both tt^{\prime} and t′′t^{\prime\prime} are comparable to NN, and that for tt in [t,t′′][t^{\prime},t^{\prime\prime}] the amplitude a1(t)a_{1}(t) is comparable to 1/N1/\sqrt{N}. For such times, modulo 2π2\pi the polar angle θ\theta reduces to an angle in [π/3,π/3][-\pi/3,\pi/3] where the trigonometric function (1725cos2θ1225cosθsinθ+825sin2θ)cosθ\left(\frac{17}{25}\cos^{2}\theta-\frac{12}{25}\cos\theta\sin\theta+\frac{8}{25}\sin^{2}\theta\right)\cos\theta is above 0.100.10, see Figure 9. These estimates, and (58) prove

m(t)0.10ct3/2 for all tin the interval [t,t′′].m(t)\geq\frac{0.10c}{t^{3/2}}\;\mbox{ for all }t\;\mbox{in the interval }[t^{\prime},t^{\prime\prime}].

For this configuration, the drift of the direction of motion, Θ(t′′)Θ(t)\displaystyle\Theta(t^{\prime\prime})-\Theta(t^{\prime}) satisfies: Θ(t′′)Θ(t)=tt′′m(t)𝑑ttt′′0.10ctt𝑑t.\displaystyle\Theta(t^{\prime\prime})-\Theta(t^{\prime})=\int_{t^{\prime}}^{t^{\prime\prime}}m(t)dt\geq\int_{t^{\prime}}^{t^{\prime\prime}}\frac{0.10c}{t\sqrt{t}}dt. Next we show that the length of the interval [t,t′′][t^{\prime},t^{\prime\prime}] is bounded below and above by absolute constants: the functions of tθ(t)t\to\theta(t) and its inverse θt\theta\to t have Lipschitz constants close to one, so there exist positive constants c4c_{4} and c5c_{5} such that

2π3=|θ(t′′)θ(t)|c4|t′′t|c4c5|θ(t′′)θ(t)|=c4c52π3.\frac{2\pi}{3}=|\theta(t^{\prime\prime})-\theta(t^{\prime})|\leq c_{4}|t^{\prime\prime}-t^{\prime}|\leq c_{4}c_{5}|\theta(t^{\prime\prime})-\theta(t^{\prime})|=c_{4}c_{5}\frac{2\pi}{3}.

We conclude that tt′′0.10ctt𝑑t\displaystyle\int_{t^{\prime}}^{t^{\prime\prime}}\frac{0.10c}{t\sqrt{t}}dt is comparable to (t)3/2(t^{\prime})^{-3/2} and that the drift Θ(t′′)Θ(t)\Theta(t^{\prime\prime})-\Theta(t^{\prime}) exceeds (t)3/2.(t^{\prime})^{-3/2}.

5 Appendix

Proposition 1.2 If among the nn agents of the system (1) there exist p3p\geq 3 agents (assumed to be the first pp) whose motion is symmetrically distributed about the center RR of the swarm, i.e.

rk(t)=R(t)+e2πi(k1)p(r1(t)R(t)) for k=1..p,r_{k}(t)=R(t)+e^{2\pi i\frac{(k-1)}{p}}(r_{1}(t)-R(t))\;\mbox{ for }\;k=1..p, (59)

then either r1(t)==rp(t)r_{1}(t)=\dots=r_{p}(t) for all tt\in\mathbb{R}, or R(t)=R(0)R(t)=R(0) for all t.t\in\mathbb{R}.

Proof.

Within this proof, we identify all the planar vectors as points in \mathbb{C} so that we can multiply, divide, and take complex conjugate.

Assume that there exists p3p\geq 3 such that (59) holds while the center of mass is not stationary, i.e. there exists a time interval I1=(t1,t2)I_{1}=(t_{1},t_{2}) with V(t)0V(t)\neq 0 for tI1.t\in I_{1}. We want to show that there exists an open time interval such that all pp particles’ locations equal R(t).R(t). (If such an interval exists, it means the velocities for the pp particles match V(t)V(t) in that interval, so the pp particles have identical initial conditions; by the uniqueness of solutions to (1) we get that the particles satisfy r1(t)==rp(t)r_{1}(t)=\dots=r_{p}(t) for all tt\in\mathbb{R}.)

To simplify notations, within this proof 555Outside of this Appendix, we used θk\theta_{k} and mm to denote real-valued functions θk=θk(t)\theta_{k}=\theta_{k}(t) and m(t)m(t) with a different meaning. In this proof θk\theta_{k} are constants, and m(t)m(t) is complex-valued. let θk=2π(k1)p\theta_{k}=2\pi\frac{(k-1)}{p} for k=1..pk=1..p and let σ(t),m(t)\sigma(t),m(t) and mk(t)m_{k}(t) be complex valued function defined by:

σ(t)=r1(t)R(t),m(t)=σ˙V, and mk(t)=eiθk+m(t)\sigma(t)=r_{1}(t)-R(t),\;\;m(t)=\frac{\dot{\sigma}}{V},\;\mbox{ and }m_{k}(t)=e^{-i\theta_{k}}+m(t) (60)

We get that for k=1..p,k=1..p,

rk=R+eiθkσ and r˙k=V+eiθkVm.r_{k}=R+e^{i\theta_{k}}\sigma\;\mbox{ and }\;\dot{r}_{k}=V+e^{i\theta_{k}}Vm. (61)

To prove the appendix, we need to show that if p3,p\geq 3, then there exists a time interval I2I_{2} where σ(t)=0\sigma(t)=0 for tI2.t\in I_{2}.

Using (61), the first pp equations of (2) become

V˙+eiθkV˙m+eiθkVm˙=(1|V|2|1+eiθkm|2)V(1+eiθkm)eiθkσ\dot{V}+e^{i\theta_{k}}\dot{V}m+e^{i\theta_{k}}V\dot{m}=(1-|V|^{2}|1+e^{i\theta_{k}}m|^{2})V(1+e^{i\theta_{k}}m)-e^{i\theta_{k}}\sigma
V˙eiθk(eiθk+m(t))+eiθkVm˙=(1|V|2|eiθk+m|2)Veiθk(eiθk+m)eiθkσ\dot{V}e^{i\theta_{k}}(e^{-i\theta_{k}}+m(t))+e^{i\theta_{k}}V\dot{m}=(1-|V|^{2}|e^{-i\theta_{k}}+m|^{2})Ve^{i\theta_{k}}(e^{-i\theta_{k}}+m)-e^{i\theta_{k}}\sigma

Divide by eiθke^{i\theta_{k}}; recall that eiθk+m(t)=mk.e^{-i\theta_{k}}+m(t)=m_{k}. We get

V˙mk+Vm˙=(1|V|2|mk|2)Vmkσ.\dot{V}m_{k}+V\dot{m}=(1-|V|^{2}|m_{k}|^{2})Vm_{k}-\sigma.

Divide by V|V|2V|V|^{2} and collect the mkm_{k} terms to get that for k=1..pk=1..p and tI1t\in I_{1}

Vm˙+σV|V|2=mk(V˙V|V|2+1|V|2|mk|2)\frac{V\dot{m}+\sigma}{V|V|^{2}}=m_{k}\left(-\frac{\dot{V}}{V|V|^{2}}+\frac{1}{|V|^{2}}-|m_{k}|^{2}\right) (62)

Employ the notations: α(t)=1|V|2V˙V|V|2\alpha(t)=\frac{1}{|V|^{2}}-\frac{\dot{V}}{V|V|^{2}} and β(t)=Vm˙+σV|V|2\beta(t)=\frac{V\dot{m}+\sigma}{V|V|^{2}} to further simplify (62) to:

β(t)=mk(t)(α(t)|mk(t)|2), for k=1,p.\beta(t)=m_{k}(t)\left(\alpha(t)-|m_{k}(t)|^{2}\right),\;\mbox{ for }k=1,\dots p. (63)
Lemma 5.1.

If α,β,m\alpha,\beta,m\in\mathbb{C} are such that

β=(m+eiθk)(α|m+eiθk|2) for θk=2π(k1)p,k=1,p,\beta=(m+e^{-i\theta_{k}})\left(\alpha-|m+e^{-i\theta_{k}}|^{2}\right)\;\mbox{ for }\theta_{k}=\frac{2\pi(k-1)}{p},k=1,\dots p, (64)

then α\alpha is the positive real number α=2|m|2+1\alpha=2|m|^{2}+1 and β=m(|m|21).\beta=m(|m|^{2}-1).

Proof.

Note that

|m+eiθk|2=|m|2+1+meiθk+m¯eiθk|m+e^{-i\theta_{k}}|^{2}=|m|^{2}+1+me^{i\theta_{k}}+\overline{m}e^{-i\theta_{k}} (65)

and that the pp roots of unity, eiθk,\displaystyle e^{i\theta_{k}}, satisfy:

k=1peiθk=0,k=1peiθk=0,k=1pei2θk=0 and k=1pei2θk=0.\sum_{k=1}^{p}e^{i\theta_{k}}=0,\;\sum_{k=1}^{p}e^{-i\theta_{k}}=0,\;\sum_{k=1}^{p}e^{i2\theta_{k}}=0\;\mbox{ and }\sum_{k=1}^{p}e^{-i2\theta_{k}}=0. (66)

Averaging over the equations (65 ), from (66) we get

1pk=1p|m+eiθk|2=|m|2+11pk=1peiθk|m+eiθk|2=m1pk=1peiθk|m+eiθk|2=m¯.\begin{array}[]{l}\frac{1}{p}\sum_{k=1}^{p}|m+e^{-i\theta_{k}}|^{2}=|m|^{2}+1\\ \\ \frac{1}{p}\sum_{k=1}^{p}e^{-i\theta_{k}}|m+e^{-i\theta_{k}}|^{2}=m\\ \\ \frac{1}{p}\sum_{k=1}^{p}e^{i\theta_{k}}|m+e^{-i\theta_{k}}|^{2}=\overline{m}.\end{array} (67)

Rewrite the equations of (64) as

β=(m+eiθk)αm|m+eiθk|2eiθk|m+eiθk|2\beta=(m+e^{-i\theta_{k}})\alpha-m\;|m+e^{-i\theta_{k}}|^{2}-e^{-i\theta_{k}}|m+e^{-i\theta_{k}}|^{2}

and average over k=1..p.k=1..p. We get

β=(m+0)αm(|m|2+1)m=m(α|m|22).\beta=(m+0)\alpha-m(|m|^{2}+1)-m=m(\alpha-|m|^{2}-2). (68)

Substituting β=mαm(|m|2+2)\beta=m\alpha-m(|m|^{2}+2) in the left side of (64) yields

mαm(|m|2+2)=mα+eiθkαm|m+eiθk|2eiθk|m+eiθk|2 and m\alpha-m(|m|^{2}+2)=m\alpha+e^{-i\theta_{k}}\alpha-m\;|m+e^{-i\theta_{k}}|^{2}-e^{-i\theta_{k}}|m+e^{-i\theta_{k}}|^{2}\;\mbox{ and }
eiθkα=m(|m|2+2)+m|m+eiθk|2+eiθk|m+eiθk|2.e^{-i\theta_{k}}\alpha=-m(|m|^{2}+2)+m\;|m+e^{-i\theta_{k}}|^{2}+e^{-i\theta_{k}}|m+e^{-i\theta_{k}}|^{2}.

Solve for α\alpha:

α=m(|m|2+2)eiθk+m|m+eiθk|2eiθk+|m+eiθk|2\alpha=-m(|m|^{2}+2)e^{i\theta_{k}}+m\;|m+e^{-i\theta_{k}}|^{2}e^{i\theta_{k}}+|m+e^{-i\theta_{k}}|^{2}

Averaging over k=1..pk=1..p, based on (66) we obtain

α=m(|m|2+2)0+mm¯+|m|2+1=2|m|2+1.\alpha=-m(|m|^{2}+2)0+m\overline{m}+|m|^{2}+1=2|m|^{2}+1.

Substituting α=2|m|2+1\alpha=2|m|^{2}+1 into (68) gives β=m(|m|21),\beta=m(|m|^{2}-1), concluding the proof of the lemma. ∎

Rewriting (63) using Lemma 5.1, we get that mkm_{k} are solutions for the equations

m(|m|21)=mk(2|m|2+1|mk(t)|2), for k=1,pm(|m|^{2}-1)=m_{k}\left(2|m|^{2}+1-|m_{k}(t)|^{2}\right),\;\mbox{ for }k=1,\dots p (69)

implying that if the time is fixed, the equations z(2|m|2+1|z|2)=m(|m|21)\displaystyle z(2|m|^{2}+1-|z|^{2})=m(|m|^{2}-1) have at least pp distinct complex solutions zz (since mk=eiθk+m(t)m_{k}=e^{-i\theta_{k}}+m(t) are all distinct solutions).

Case 1: There exists a time when β=m(|m|21)\beta=\displaystyle m(|m|^{2}-1) is nonzero. We claim that in this case p2.p\leq 2.

At a time when β0\beta\neq 0, necessarily (2|m|2+1|mk|2)0\;(2|m|^{2}+1-|m_{k}|^{2})\neq 0 . Express β\beta using polar coordinates as β=beiψ\displaystyle\beta=be^{i\psi} where b>0.b>0. Since the solutions zz to z(2|m|2+1|z|2)=beiψz(2|m|^{2}+1-|z|^{2})=be^{i\psi} satisfy zeiψ=b2|m|2+1|z|2ze^{-i\psi}=\frac{b}{2|m|^{2}+1-|z|^{2}}\in\mathbb{R} , they are collinear (on the line through the origin, making an angle ψ\psi or ψ+π\psi+\pi with the xx- axis).We conclude that all the points m+eiθkm+e^{-i\theta_{k}} are collinear, which is impossible for the pthp^{th} roots of unity with p3.p\geq 3.

Case 2: β=m(|m|21)\beta=m(|m|^{2}-1) is the zero function on the interval I1.I_{1}. We claim that this implies that m(t)=0m(t)=0 for all tI1t\in I_{1} or p2.p\leq 2.

Assume not: then p3p\geq 3 and there exists a subinterval I2I_{2} of I1I_{1} with m(t)0,β(t)=0m(t)\neq 0,\beta(t)=0 for tt in I2.I_{2}. Since β=m(|m|21),\beta=m(|m|^{2}-1), we get that |m(t)|=1,α(t)=2|m|2+1=3|m(t)|=1,\alpha(t)=2|m|^{2}+1=3 and therefore the equations (69) become: at times tI2t\in I_{2}

0=(m+eiθk)(3|m+eiθk|2)0=(m+e^{-i\theta_{k}})(3-|m+e^{-i\theta_{k}}|^{2}) (70)

For a fixed t,t, the equation 3|m+eiθk|2=03-|m+e^{-i\theta_{k}}|^{2}=0 can be satisfied by at most two roots of unity, because otherwise we would have three or more roots satisfy |m¯+eiθk|=3,|\overline{m}+e^{i\theta_{k}}|=\sqrt{3}, implying that the circle circumscribing those three roots of unity is centered at m¯,-\overline{m}, and has radius 3\sqrt{3} rather than being the unit circle, a contradiction. Satisfying (70) when p3p\geq 3 requires that the factor (m+eiθk)(m+e^{-i\theta_{k}}) has a zero of its own, implying that the continuous function m(t)m(t) takes values in the discrete set {eiθk,k=1..p}.\{-e^{-i\theta_{k}},k=1..p\}. Thus there exists an index, denoted q,q, such that m(t)=eiθqm(t)=-e^{-i\theta_{q}} for all tI2.t\in I_{2}. We will show this together with (60) imply that V=0,V=0, contradicting the starting assumption of V0.V\neq 0.

Based on m=eiθqm=-e^{-i\theta_{q}} and (61) we get that on I2I_{2}

r˙q=V+eiθqmV=V+eiθq(1)eiθqV=0\dot{r}_{q}=V+e^{i\theta_{q}}mV=V+e^{i\theta_{q}}(-1)e^{-i\theta_{q}}V=0

meaning that the location of the qthq^{th} agent remains unchanged in time. The equation of motion states r¨q=(1|r˙q|2)r˙q(rqR),\displaystyle\ddot{r}_{q}=(1-|\dot{r}_{q}|^{2})\dot{r}_{q}-(r_{q}-R), thus R=rqR=r_{q} is constant, so R˙=V=0,\dot{R}=V=0, contradicting V0V\neq 0 on I2.I_{2}.

Our two cases show that if p3p\geq 3 then necessarily m(t)=0m(t)=0 on the interval I1.I_{1}. We get that σ˙(t)=0\dot{\sigma}(t)=0 and rk(t)=R(t)+eiθkσ0r_{k}(t)=R(t)+e^{i\theta_{k}}\sigma_{0} for all tt in I1.I_{1}. Substituting r˙k=V\dot{r}_{k}=V and r¨k=V˙\ddot{r}_{k}=\dot{V} in (1) yields that for multiple angles θk\theta_{k} the equalities V˙=(1|V|2)Veiθkσ0\dot{V}=(1-|V|^{2})V-e^{i\theta_{k}}\sigma_{0} hold, which require that σ0=σ(t)=0,\sigma_{0}=\sigma(t)=0, and therefore all pp particles’ positions coincide on I1.I_{1}.

References

  • [1] A. M. Turing. The chemical basis of morphogensis. Phil. Trans. Roy. Soc. B, vol. 237, 37–72, 1952.
  • [2] Yoshiki Kuramoto. Self-entrainment of a population of coupled non-linear oscillators. Lecture Notes in Physics, International Symposium on Mathematical Problems in Theoretical Physics. H. Araki (ed.) Vol. 39. Springer-Verlag, New York. p. 420. (1975).
  • [3] Felipe Cucker and Steve Smale. Emergent behavior in flocks IEEE Trans. Automat. Control, vol 52, no. 5, 852–862, 2007.
  • [4] C. Medynets and I. Popovici. On spatial cohesiveness of second-order self-propelled swarming systems. SIAM Journal on Applied Mathematics, Vol. 83, Iss. 6 2169–2188, 2023.
  • [5] Carl Kolon, Constantine Medynets and Irina Popovici. On the stability of Rotating States in Second-Order Self-Propelled Multi-Particle Systems preprint arXiv:2105.11419, 2021.
  • [6] Chad Topaz and Andrea L. Bertozzi, Swarming Patterns in a Two-Dimensional Kinematic Model for Biological Groups , SIAM Journal on Applied Mathematics Volume 65, Number 1, 152–174, 2004.
  • [7] B. Q. Nguyen, Y-L Chuang, D. Tung, C. Hsieh, Z. Jin, L. Shi, D. Marthaler, A. L. Bertozzi, R. M. Murray. Virtual attractive-repulsive potentials for cooperative control of second order dynamic vehicles on the Caltech MVWT, Proceedings of the American Control Conference, Portland 2005, pp. 1084-1089, https://www.math.ucla.edu/ bertozzi/papers/potential.pdf.
  • [8] Maria Dorsogna, Yao-Li Chuang, Andrea Bertozzi, and L Chayes. Self-propelled particles with soft-core interactions: Patterns, stability, and collapse. Physical review letters, 96:104302, 04 2006.
  • [9] G. Albi, D. Balagué, J. A. Carrillo, and J. von Brecht. Stability analysis of flock and mill rings for second order models in swarming. SIAM Journal on Applied Mathematics, 74(3):794–818, 2014.
  • [10] Seung-Yeal Ha and Jian-Guo Liu. A simple proof of the Cucker-Smale flocking dynamics and mean-field limit. Communications in Mathematical Sciences, vol. 7, no. 2, 297 – 325, 2009.
  • [11] Jack Carr. Applications of centre manifold theory, volume 35 of Applied Mathematical Sciences. Springer-Verlag, New York-Berlin, 1981.
  • [12] Vanderbauwhede, A. Centre Manifolds, Normal Forms and Elementary Bifurcations, In: Kirchgraber, U., Walther, H.O. (eds) Dynamics Reported. Dynamics Reported, vol 2. Vieweg+Teubner Verlag, Wiesbaden, 1989.
  • [13] J A Carrillo, Y Huang, and S Martin. Nonlinear stability of flock solutions in second-order swarming models. Nonlinear Analysis: Real World Applications, 17, 332–343, 2014.
  • [14] Alain Haraux and Mohamed Ali Jendoubi. The convergence problem for dissipative autonomous systems. SpringerBriefs in Mathematics. Springer, Cham; BCAM Basque Center for Applied Mathematics, Bilbao, 2015. Classical methods and recent advances, BCAM SpringerBriefs.
  • [15] Theodore Kolokolnikov, Hui Sun, David Uminsky, and Andrea L. Bertozzi. Stability of ring patterns arising from two-dimensional particle interactions. Phys. Rev. E, 84:015203, Jul 2011.
  • [16] Yao li Chuang, Maria R. D’Orsogna, Daniel Marthaler, Andrea L. Bertozzi, and Lincoln S. Chayes. State transitions and the continuum limit for a 2d interacting, self-propelled particle system. Physica D: Nonlinear Phenomena, 232(1):33–47, 2007.
  • [17] John Guckenheimer and Philip Holmes. Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. Springer New York, NY, 1983.
  • [18] A.L. Bertozzi, T. Kolokolnikov, H. Sun, D. Uminsky and J. von Brecht, J. Ring patterns and their bifurcations in a nonlocal model of biological swarms. Communications in Mathematical Sciences, Volume 13 (4). Pages: 955-985 , 2015.
  • [19] J. Carrillo, M. D’orsogna, and V. Panferov, Double milling in self-propelled swarms from kinetic theory, Kinetic and Related Models, 2 , 363–-378, 2009.
  • [20] Kevin P. O’Keeffe, Hyunsuk Hong, and Steven H. Strogatz. Oscillators that sync and swarm. Nature Communications, 8(1):1504, 2017.
  • [21] Arkady Pikovsky, Michael Rosenblum, and Jürgen Kurths. Synchronization: a universal concept in nonlinear sciences, volume 12 of Camb. Nonlinear Sci. Ser. Cambridge: Cambridge University Press, reprint of the 2001 hardback edition edition, 2003.
  • [22] Klementyna Szwaykowska, Ira B. Schwartz, Luis Mier y Teran Romero, Christoffer R. Heckman, Dan Mox, and M. Ani Hsieh. Collective motion patterns of swarms with delay coupling: Theory and experiment. Physical Review E, 93(3), mar 2016.
  • [23] Ballerini M, Cabibbo N, Candelier R, Cavagna A, Cisbani E, Giardina I, Lecomte V, Orlandi A, Parisi G, Procaccini A, Viale M, Zdravkovic V. Interaction ruling animal collective behavior depends on topological rather than metric distance: evidence from a field study. Proc Natl Acad Sci U S A., vol 105(4):1232–7, 2008.
  • [24] Scott Camazine, Jean-Louis Deneubourg, Nigel R. Franks, James Sneyd, Guy Theraula, and Eric Bonabeau Self-Organization in Biological Systems Princeton Studies in Complexity, 2003.
  • [25] J. Buhl, David J. T. Sumpter, Iain D. Couzin, Joseph J. Hale, Emma Despland, Esther R. Miller and Stephen J. Simpson, From Disorder to Order in Marching Locusts Science, vol. 312, 1402 –1406, 2006.
  • [26] C. Liu, D.R. Weaver, S.H. Strogatz, and S.M. Reppert. Cellular Construction of a Circadian Clock: Period Determination in the Suprachiasmatic Nuclei Cell 91, 855–860, 1997.
  • [27] Hellmann, F., Schultz, P., Jaros, P. et al. Network-induced multistability through lossy coupling and exotic solitary states. Nat Commun 11, 592, 2020.
  • [28] Herbert Winful and S. S. Wang. Stability of phase locking in coupled semiconductor laser arrays. Applied Physics Letters, 53(20): 1894–1896, 1988.
  • [29] S. H. Strogatz, D. M. Abrams, A. McRobie, B. Eckhardt, E. Ott. Theoretical Mechanics: Crowd synchrony on the millennium bridge. Nature 438 (3), pp. 43–44, 2005.
  • [30] Seung-Yeal Ha, Taeyoung Ha, and Jong-Ho Kim, On the complete synchronization of the Kuramoto phase model, Physica D: Nonlinear Phenomena, Volume 239, Issue 17, 1692–1700, 2010.
  • [31] N. Chopra and M. W. Spong, On Synchronization of Kuramoto Oscillators, Proceedings of the 44th IEEE Conference on Decision and Control, Seville, Spain, 3916–3922, 2005.
  • [32] J.A. Carrillo, D. Kalise, F. Rossi, E. Trelat, Controlling Swarms toward Flocks and Mills, SIAM Journal on Control and Optimization , Vol. 60, Iss.3, 1863–1891, 2022.
  • [33] Grushkovskaya, V., Zuyev, A. Asymptotic behavior of solutions of a nonlinear system in the critical case of q pairs of purely imaginary eigenvalues, Nonlinear Analysis 80(2013), 156–- 178.
  • [34] P.S. Krasilnikov, Algebraic criteria for asymptotic stability at 1- 1 resonance, Journal of Applied Mathematics and Mechanics, Volume 57, Issue 4, 1993, 583–590.
  • [35] V. Grushkovskaya, On the influence of resonances on the asymptotic behavior of trajectories of nonlinear systems in critical cases, Nonlinear Dynamics, 86 (2016), 587–603.
  • [36] C. Corduneanu with N. Gheorghiu and V. Barbu, Almost periodic functions, Chapter I, Tracts in Mathematics, Translated from the Romanian ed. by G. Bernstein and E. Tomer, 1961.