Stability of Translating States for Self-propelled Swarms with Quadratic Potential
Abstract
The main result of this paper is proving the stability of translating states (flocking states) for the system of -coupled self-propelled agents governed by , . 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 , 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 identical self-propelled agents having all-to-all coupling
(1) |
Notations 1.1.
We denote by the position of the center of mass, and by 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 rather than we identify of a vector with the complex number The system (1) becomes
(2) |
Note that given constants and , if is a solution for (2), then the vectors and the complex conjugate vectors 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 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:
(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 denote the polar angle of the mean velocity, so


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 and and to recover the information about and via integration. The velocity-acceleration system for , obtained by differentiating (1) is:
(4) |
where the superscript 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 be a continuous scalar function, differentiable on Define as With the possible exception of the origin, the potential is differentiable with Consider the system with identical self propelling terms and rotationally symmetric potential:
(5) |
where are positive constants.
The system (1) is in the class of (5) with and Other well known classes of radial potentials are Morse potentials where is the difference between two exponential decays (see [8], [9]), and power law potentials where ( [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 , [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 in this paper (and in [5]). In fact (1) has very few configurations with symmetries:
Proposition 1.2.
If among the agents of the system (1) there exist agents (assumed to be the first ) whose motion is symmetrically distributed about the center of the swarm, i.e.
then either for all , or for all
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 into subsets where the agents in have fixed positions coinciding with the center at all times; each agent in oscillates on a segment centered at , satisfying the scalar Lienard equation (which has an attracting cycle); each agent in is asymptotic to a rotation
Definition 1.3.
A solution for (1) is called a rotating state if there exist angles with and a constant such that for all and all (a counterclockwise rotating state), or for all and all (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 are uniformly distributed in , 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 for (1) is called a flocking state if for all in . (Necessarily the common velocity is that of the center of mass.)
Note that for a flocking state If we subtract the equation for from that of we get that for all meaning that all particles have collapsed onto the center of mass, moving together according to
Unless for all (the trivial solution), the velocity of the flocking solutions maintains the initial direction of with speed approaching unit speed at a rate of (since solves the logistic equation A nontrivial flocking state has velocity approaching exponentially fast. Moreover, if denotes the absolutely convergent integral then the flocking state’s center of mass has nearly rectilinear motion: The limit holds since
Definition 1.5.
A solution for (1) is called a translating state if there exists such that for all (Necessarily, for a translating state there exists such that for all )
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 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 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 much slower than the conjectured exponential rate.
Theorem 1.6.
Consider the system (1) with initial conditions and For any there exists such that if and satisfy: for all in ; ; for all in , then for :
(6) |
Moreover,
(7) |
and there exists in with such that
(8) |
The proof of Theorem 1.6 is completed in three parts, corresponding to three dimension reductions of the system: form to in Section 2, to in Section 3, and to 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 dimensional invariant subspace for a system taking the form where is the Jacobian matrix and is a nonlinear function whose Maclaurin expansion has terms or order 3 or higher. Counted with multiplicity, the eigenvalues of consist of pairs of purely imaginary roots , and eigenvalues with negative real parts. The stability of the origin for systems whose Jacobian matrices have pairs of purely imaginary roots and 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 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 towards the origin. Our approach draws inspiration from a class of time-dependent systems with cubic nonlinearities instead:
(9) |
where is vector-valued from to , the coefficient functions are continuous and the remainder functions defined on satisfy the condition: there exists a function with such that for all In Section 3.3 we provide easily verifiable conditions on the coefficient functions that sufficiently ensure the asymptotic stability of the solution , with convergence rate of order 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 denote the column vectors with all- entries and all- entries in by
(10) |
Given positive integers denote the matrices of all ones and all zeros by and We may use just for a matrix if its dimensions are clear form context. Denote the identity matrix by Denote by the hyperplane orthogonal to in
(ii) Given (column) vectors and denote by the column vector
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
the generally accepted definition of asymptotic flocking requires two conditions:
-
•
Velocity alignment:
-
•
Forming a finite-diameter group:
Cucker-Smale models have constant mean velocity so the existence of a limiting value of is automatic; the problem reduces to an equivalent model with For the classical communications functions and with it is proven in [10] and [3] that there is unconditional flocking in the sense that all solutions exhibit asymptotic flocking with being exponentially small.
For Kuramoto oscillators (KO)
the mean frequency is constant, ensuring that the mean phase is linear in time. For (KO) synchronization plays the role of (CS)’s velocity alignment, requiring that the angular frequencies 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 ) 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
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: for all Necessarily for some Let in denote the initial locations of a flocking configuration. In [32] the authors define the stability of the flocking state evolving from by referring to a manifold of configurations. They define the flock manifold as
and the ’local asymptotic stability’ of as the existence of a neighborhood of such that for initial conditions in the distance from the state variables to the manifold approaches zero as Note that this is a rather weak condition on the behavior of , as it only requires that without imposing any conditions on 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- term out of the motion transforms the flocking states into fixed points.
In [4] the authors prove that there exit constants that depend only on the number of particles, such that every solution to (1 ) satisfies and for large enough 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 and to and Both approaches lead to dynamics evolving in a compact phase space, with the translating states of (1 ) corresponding to fixed points with all velocities equal to the mean speed a unit vector:
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.
Proof.
(4) is obtained from differentiating (1). We provide a counterexample for the converse: let be any nonzero vector in the plane with Let for all and all The vectors 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 is unstable.
Note that the fixed points of (4) with 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 is a fixed point for (4) if and only if for all and for all , where is the initial mean velocity. (Necessarily for such a fixed point for all ; we get from (4) that for all so all the velocities are equal.)
In order to prove that the fixed points of the system (4) with are unstable it is enough to do so when is the unit vector in the axis direction (due to the rotation invariance). Note that the linear subspace
of is invariant under the flow of (4), allowing us work within the class of solutions
that satisfy
for all and all .
Denote by and the horizontal and vertical components of and , respectively. In this class (4) becomes:
(11) |
The system (11) has two conserved quantities:
where the constants of motion and are small if is near one and is near zero. Fix the perturbation parameter and allow to vary near zero; this reduces the study of (11) to that of the 3-dimensional system
(12) |
For all the points from the unit circle in the plane are fixed points. For each , small, the system (12) has three fixed points, one fixed point near and two other fixed points near and satisfies
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 is ). For small, the fixed point has a positive eigenvalue (), and the flow of (12) has heteroclinic cycles with a saddle connection near , a stable node near and an unstable node near ).
The locations of the aforementioned fixed points are shown Figure 3, with on the vertical axis: the curve through is illustrated in red with two markers; also in red are the almost-vertical curves of fixed points near the fixed points and in the plane. The fixed points on the unit circle in the plane use markers. All the other curves in Figure 3 are trajectories within three representative planes ( and ).

The dynamics of (12) in the region near makes the fixed point of (11) unstable; thus the fixed point 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 and as the unknown functions. Note that since we can eliminate it from our unknowns, and work within a dimensional phase space (if we were to keep the redundant as a state variable, we would have added two more zero eigenvalues to the system).
The system (1) becomes:
(13) |
where
The translating states of (1) with correspond to the fixed points with and
Focus on the dynamics near The other fixed points are obtained from via rotation of angle The linearization of (13) at yields the Jacobian matrix
where is the square matrix and We get that
The eigenvector for is ; it has the direction of the line tangent to the curve of fixed points at The eigenvalue captures the rotation invariance of the model. Note that due to the eigenvalues there are more degeneracies associated with (13) than the symmetries of the problem.
If in the system (13) all particles’ components are initially zero, they stay zero forever, with the remaining dimensional system in having the Jacobian at equal to
and characteristic polynomial In the absence of a component, all trajectories approach the fixed point exponentially fast, meaning that the entire stable manifold for (13) at consists of the trajectories with no component. The flow on the 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 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.
For example, starting from the flocking solution for all and , given small, consider the perturbed initial conditions and where are angles such that Note that for these perturbations and It is immediate to verify that the particles in the version of (1) move with unit speed on helix-shaped trajectories according to
The perturbed trajectories have and therefore and for all 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 ( for ), 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 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 or with negative real parts).
Notations 2.1.
Within this section, if and are vectors in with define the vector and the scalars and to be:
(14) |
Note that and Also:
(15) |
Proposition 2.2.
Let be in interval and be differentiable functions, Then and are differentiable and
(16) |
Notations 2.3.
For the system (1) with , define the scalar functions of time and to be
(19) |
The scalars 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. and are introduced to streamline notations – they can be expressed in terms of , as in (20) and (21).

Note that thus captures how fast the direction of is changing relative to the speed, where captures the changes in the speed of the center of mass. The geometrical meanings of and are apparent if we represent the velocity in polar coordinates, Then
On the other hand therefore
From (3) and (15) we get Collecting the tangential and normal components of the acceleration as in (19) yields
(20) |
(21) |
Before reformulating (1) in terms of , we point out that will not be used as state variable, since which follows from The latter and imply:
(22) |
Unlike we keep and as state variables because eliminating them would hinder downstream computations (by contributing off-pattern terms to the mean-field variables and ).
Proposition 2.4.
The differential equations governing the dynamical system with unknowns , , and are
(23) |
where was defined in (21) as
Proof.
Apply the results of (16) for with and We get that for a differentiable function
(24) |
Use (24) for ; recall that and Moreover and thus has components and We get
(25) |
Use (24) again for Note that so it has components and We get
(26) |
∎
Remark 2.5.
The dynamical system is reflection invariant, in the sense that if is a solution for , then is also a solution.
Let denote the hyperplane orthogonal to in The subspace is invariant under (23).
The reflection invariance holds since the functions and are odd functions of and even in and thus the vector field given from has its components given by odd function of and even in whereas the components of are even functions of and The invariance of follows from and
Remark 2.6.
Proposition 2.7.
Proof.
Note that at the speed is Near all the terms are quadratic or smaller since both and are zero at therefore have zero gradients. From we conclude that and the terms and vanish and have zero gradient at as well. The nonlinear term that is part of is linearly approximated by at Finally, note that the pattern for is different than that for the partial derivatives of and so when using block notations for the Jacobian matrix, we have to separate away the last component from the rest.
The Jacobian matrix (23) at is
(27) |
where is the square matrix ; on the left of the separating line means and on the right of the separating line means . Note that is already in a block diagonal form. The first block, of size , corresponding to the state variables and , has characteristic polynomial The second block, of size , corresponding to the state variables and has characteristic polynomial 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 being ):
∎
Notations 2.8.
Remark 2.9.
Proof.
The assumptions of Main Theorem 1.6 and the assertions (6) and (7) can be restated using alone, since they are norm-based and
and
This turns the assertions (6) and (7) into statements about the asymptotic stability of the flow at
For the assertion (8) we need to examine how, given a solution in the dimensional space , we can identify the trajectories in that in the center-mass frame (19) lead back to . The trajectories can only be determined up to a 3-dimensional parameter , interpreted as a location in together with an angle
For in , define where (see (21)). Note that
Rebuild using and from it find
Note that captures the indeterminacy due to the rotational invariance of (1) and captures the indeterminacy due to the translation invariance of (1).
Finally, we can recover the directions of the vectors from using
(28) |
Once the convergence of is established, assertion (8) reduces to proving the convergence of the angle , which would follow from the convergence of ∎
Remark 2.10.
Let be the open subset of where Let be such that Then there exists a time interval such that the function is analytic on that interval.
This holds since the vector field (23) is defined by algebraic operations on The condition that ensures that the fraction used to define in terms of (see (21)) is real analytic in a neighborhood of
2.1 The Polar Angles in the Planes
In this subsection we are only interested in solutions with that lie in the set where Consider the projection of on the plane denote by the curve with parametrization
The upper block of shows that each plane is an eigenspace for It is natural to recast the dynamics in each of these planes using polar coordinates, (’a’ is for amplitude), but we run into a technical difficulty: we would like to find continuous functions to represent as If for all then a smooth polar angle exists. However, the property is not invariant under the flow: the origin, is a fixed point, but points such as are not, meaning that could go through and in doing so it could enter then emerge at different polar angles, forcing a jump in (see Figure 5).

Proposition 2.11.
(i) If and are such that with for , then there exits a continuous function such that one of the following is true
(ii) Let be such that the solution to remains in for all with not identically zero. Then there exists a polar angle such that and such that the functions and are continuous. If is any continuously differentiable, periodic function with period , the function is continuous and piecewise differentiable.
Proof.
(i) We will use the analyticity of and and the injectivity of the tangent modulo ( implies ).
A polar angle is well defined and continuous on and on due to . We need to show that at the angle has sided limits that are equal or that differ by a multiple of
If one of or is identically zero, is along the coordinate axis; its polar angle is either or If neither nor are identically zero let be the first nonzero term in the Taylor series for and let be the first nonzero term for If then is tangent to the vector so the two sided limits of exist and are in the set If then is tangent to the axis and the polar angles have sided limits in the set If then is tangent to the axis and the polar angles have sided limits in the set so they differ by a multiple of Let and let be the Heaviside function. Then is continuous with
For (ii): there is nothing to prove if is nonzero. Assume that at some positive times. Due to the analyticity of the set of times when is finite or consists of times going to infinity. By (i), the polar angles corresponding to the intervals have well defined side limits that differ by or at meaning that the double angles coincide modulo This implies that and are continuous. The continuity properties for follow since is the sum of a Fourier series that converges in the sup norm. ∎
Notations 2.12.
(i) For the components of a solution to denote by and by the usual continuous polar coordinates if ; let be the piecewise smooth polar angle from Remark 2.11, part (ii) if is not the identically zero function, and let if for all Thus
(29) |
(ii) Denote by or simply the Lipschitz continuous function
(30) |
Note that for all smooth functions of period the functions 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 while working with analytic functions of Central manifolds and their flows are not necessarily let alone analytic.
3 The Dynamics on The Central Manifold
We shift coordinates in order to bring the fixed point at the origin of
Notations 3.1.
Let for and let (thus
Let denote the vector field in given by
(31) |
where , with the index being in for the top components, and in for the components below the line.
Note that we used Also: gives the rate of convergence of the mean field speed to one. We get
(32) |
In the coordinates the dynamical system (23) becomes
An alternate expression for is
We want to prove that the system (31) has the origin as an asymptotically stable fixed point. The Jacobian matrix for (31) at the origin is the same as that of (23) at the point given in (27), meaning that the central subspace for is associated with the eigenvalues , has dimension , and it consists of the points
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 from the origin. is assumed small enough so that the local central manifold is defined within the ball , with the central manifold map having continuous order derivatives.
We use the customary big-O notation to convey a function that has absolute value at most a positive multiple of . For example both and are since is the largest amplitude among
Notations 3.2.
For a positive we use to denote
Theorem 3.3.
The flow on the central manifold of (31) is governed by the dimensional system
(33) |
Proof.
Let denote the map whose graph is the central manifold, as in [11]. In our notations, the central manifold is described by Since the projections onto the components of the vector fields and are equal, the map can be constructed to be even, In particular the Taylor expansion of has no odd-power terms.
For a scalar function denote by the row vector of the partial derivatives with and denote by the row vector of the derivatives with
We seek to produce an approximation of such as Given that has no third degree terms in its expansion, we seek quadratic polynomials such that We separate the terms of (31) that are smaller than the needed precision. A priory and Their gradients are We get
(34) |
We conclude that
(35) |
Following the approximation technique from [11] Theorem 3, page 5, based on the vector field (35), we need quadratic polynomials with and with such that the differences given below are
For a more compact notation, denote by the linear differential operator
Moving the error terms to the left, we get
(36) |
We meet the requirement for the differences if the right hand side of (36) is identically zero. It is more expedient to use defined as
rather than . To achieve precision we need that for the following hold:
Note that by subtracting the last equation from the middle row equations we get
(37) |
Before proceeding to finding the quadratic polynomials we perform preliminary calculations to determine how acts on simple quadratic terms. We compute From these and
(38) |
we get
(39) |
Using vector notation we rewrite (39) as
(40) |
Let denote the matrix of the operator,
Motivated by the fact that the original system (1) had index-permutation invariance, and from the identities and we look for quadratic polynomials with similar features. We want to find constants such that for
Group the unknown coefficients into the row vectors , and We get that
From (40) we get
Use matrix notation to rewrite (37) as:
and
Use identification of coefficients in the latter equality: we require The solution is We conclude
By identification of coefficients we require that the 6-dimensional row vector solves
We get that and
Recall that We get that for
Substitute into the differential equations (35). Up to
thus To complete the approximation of note that the exact equation differs from by The term differs from by Overall,
∎
Remark 3.4.
On the central manifold, the mean speed is always under one, and
Proof: Use and the fact that the quadratic form is positive definite with minimum equal to
3.2 Amplitude Variation on The Central Manifold
We consider trajectories in the subset of the central manifold, for as long as they remain in the small ball where the central manifold is . We use the polar coordinate representation from (29) and Remark 2.12.
Let be periodic functions of period defined as
(41) |
We get from the choice of in that the functions and are continuous for all , and piecewise Moreover, non-differentiability of and can only happen for times and indexes with . We abbreviate this excepted set of times (this restriction) by
Proposition 3.5.
Using polar coordinates, the differential equations for become:
(42) |
The condition rewrites as
Proof.
Use per (33). Note that is The periodic functions allow us to rewrite as
and to conclude Approximate from and :
The factor following has bounded terms, thus is , and
∎
Per (42), the amplitude is forced down by the negative term and amplified by the positive coupling terms To understand which of the two has the stronger effect, we examine their average impact over a cycle. Note that and has average is positive and has average Estimating the impact of agent on the amplitude requires we investigate the products between and all possible shifts 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.

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 is vector-valued from to , the coefficient functions are continuous and the remainder functions defined on satisfy the condition:
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 In most practical settings the coefficient functions are combinations of trigonometric functions of periods so they are almost periodic.222 If denotes the set of all real-valued trigonometric polynomials (all linear combination of and , with real ), then a real-valued function is called almost periodic if there exists a sequence of trigonometric polynomials such that converges to in the sup norm on Let denote the set of real-valued almost periodic functions. is closed under function multiplication and it is a closed linear subspace of the space of continuous and bounded functions ; in fact is the smallest subalgebra of containing all the continuous periodic functions ([36]).
Notations 3.6.
Define the mean of a function to be
(43) |
Let denote the subspace of bounded, continuous functions that have a mean, and have bounded off-mean antiderivative:
(44) |
If is a continuous, periodic function with period then is the customary average value: Moreover, the antiderivative is periodic, thus Any linear combination of periodic functions belongs to 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
We can now state the stability result for (9); in keeping with earlier conventions, for a solution vector use to denote the largest absolute value among its components:
Proposition 3.7.
Consider the system (9), with continuous coefficient functions and reminders Additionally, assume that and are in
If for all then the solution is asymptotically stable, with for large
Proof.
Within this proof, use for Work in a small ball centered at the origin in .
The functions are in so the antiderivatives are bounded, (44); by adding large enough constants, we obtain antiderivatives that are positive and bounded. Let denote an antiderivative of that has positive (and bounded) range. Let and let denote an antiderivative of with positive (and bounded) range.
Define the Lyapunov functions and as
(45) |
For as long as is in the ball , the denominators of are close to so is approximately equal to In particular and are only within a factor of from each other.
Before differentiating we take note of the following:
Remark 3.8.
For positive functions, with bounded, and small,
To estimate apply the latter with and We get
(46) |
since the terms are and are bounded.
Use the constant for By our assumptions and Substituting into (46), and averaging over in gives:
Because and are comparable, and larger than there exists such that
(47) |
We will need the result of the following Proposition
Proposition 3.9.
Let be a positive integer, and let be scalars in Then
Proof.
Use Proposition 3.9 with and non-negative scalars and . Conclude that the net contribution of the last two terms of (47) is negative and therefore
The solution to the differential equation is We get that From we conclude
This proves the asymptotic stability of the origin, and the stated decay rate. ∎
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 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 and that all agents’ velocities align with the mean velocity 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 as stated in Theorem 4.1.
Due to agents having a common oscillation frequency, one anticipates that the phase angles will synchronize, that will become interchangeable with and the coefficient functions interchangeable with the time-dependent (rather than phase-dependent) trigonometric polynomials . However, although the differences are small for some agents 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 to be very large even though 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.

Theorem 4.1.
Let and be the periodic functions introduced in (41). For any there exists such that any solution of
with and satisfies: if for , then for all and
Moreover, there exist that for large .
Proof.
We consider orbits that start with amplitude vector in a small neighborhood of the fixed point . 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 and have periodic, thus bounded, antiderivatives, with period Let be an antiderivative of the latter, and be an antiderivative of , both with strictly positive range. 444 The actual formulas for the antiderivatives and are unimportant in our calculations; however, one could use: and
We use the following influence-diminishing function to weigh the impact of agents. Let be the Lipschitz continuous function given by for and for in Its graph is illustrated in Figure 8.

Given a point in denote by or simply by the value of Recall that denotes the largest magnitude. The power was selected to be just below the order of ; We have:
(51) |
For notational convenience let and denote subsets of that correspond to agents whose magnitude at that instant is relatively large () or relatively small ():
Note that if at some time then and if then
Proof.
Fix in The continuous function is almost everywhere differentiable and has a weak derivative given almost everywhere by
If then For use
Note that for al indexes the lowest terms in are of order three, so
For the second term of use and to conclude that for we have Overall the sum of the two terms is the larger of the parts, i.e. ∎
Proof.
The functions are between zero and one, and so we only need to prove that
If is an index from then and The first fraction of the product is bounded, the second is less or equal to one since is an index in , leaving
If is an index from then and thus
We used the fact that and that when the ratio is under one. ∎
Lemma 4.4.
and for all
Proof.
If is in then and the inequalities hold. If is in then and both and are in .
∎
Returning to the proof of Theorem 4.1 :
Define the function as
(52) |
where and are the positive, bounded antiderivatives of and of
Define as Note that along trajectories other than the origin, the functions are continuous functions of time. While is in small neighborhood of the origin, are the composition between the Lipschitz continuous and piece-wise functions , and therefore the functions are differentiable almost everywhere, with weak derivatives that can be computed with the usual chain rule for almost all . The denominators of are very close to so each is comparable in value to . In fact and are only within a factor of from each other.
We claim that satisfies the differential inequality for some positive constant We start by estimating just
From Remark 3.8 we get that
Recall that and are and that and are bounded, implying that the above bracket is We get that
(53) |
Estimate the four terms of
For the first term, use (42) to rewrite within an approximation:
For third term, use Lemma 4.2 to conclude that it is
For the second and fourth terms use (42) and Lemma 4.3, respectively:
Thus in (53), replacing by and by keeps the approximation within :
(54) |
From Lemma 4.4 we get and are within from each other.
Substitute and into (54); the terms cancel and we get:
(55) |
Since , and Add the equations (55) corresponding to and divide by Use the fact that .
(56) |
Use Proposition 3.9 with and and to conclude that the net contribution of the two products of sums in (56) is negative, therefore
The function and are comparable in scale, therefore there exists a constant that depends only on such that
meaning that for all
To get the lower bound of for or equivalently the lower bound of for , return to (55). Let the maximum value of the periodic function We have
and on account that conclude
ensuring that has a lower bound comparable to ∎
Corollary 4.4.1.
For initial conditions near the fixed point , not on the stable manifold there exists depending on such that
(57) |
Proof.
Let be near the fixed point with Theorem 4.1 implies that the flow of (23) within is asymptotically stable near therefore there exists an initial condition in the central manifold, such that the two solutions and are exponentially close for in Let denote the position vectors, velocities and their components for the orbit of Since the projection onto the subspace of the point is not the origin, so
Theorem 4.1 implies that the flow on the central manifold of (31) is asymptotically stable in . It also implies the asymptotic stability of the origin for the system (31) with decay rate of order , and the asymptotic stability of for (23) in To complete the proof of Theorem 1.6 we are left to show that the direction of is convergent and close to that of for which it is sufficient to prove that converges, per Remark (2.9). The solutions of the system (23) stemming from initial conditions near evolve exponentially close to a solution on the center manifold. It is therefore enough to limit the investigation of 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 in terms of the physical coordinates of the swarm (1): the positions and and the velocities and . We assume that the initial conditions for (23) are from a small neighborhood of , within its basin of attraction.
We begin by addressing the limit behavior of the direction of motion, i.e. the polar angle satisfying Recall that the change in is given by according to (32) and (21).
Remark 4.5.
There exists depending on and such that
Moreover,
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 For on the central manifold We get that for large the absolute value of is below a multiple of which has a convergent integral at infinity. This proves the existence of denoted Note that is bounded above by which is of order
Let From with and we conclude converges to
Remark 4.6.
If is a solution of (23) not in then there exists such that the mean speed satisfies for large Moreover, there exist initial conditions for which the center of mass does not stay within bounded distance from rectilinear motion with velocity in the sense that for any
The center of mass falls behind exceedingly large distances compared to where rectilinear motion would have located it.
Proof.
If is not in , then it is exponentially close to a nontrivial solution from the central manifold, It suffices to prove the bound for Note that for we have
To prove the existence of configurations with large gap between the center of mass and the rectilinear motion, consider agents with mirror symmetry with respect to axis (i.e. ). Due to the unidirectional motion, we have that with and we get that moves along the axis: Identifying with its component we get
as . ∎
Remark 4.7.
For the estimates of for and for the drift are sharp: there exist initial conditions for which the solution of (23) satisfies: there exist and times going to infinity such that and such that the variation in the directions of motion satisfies
Proof.
In order to prove that the estimate is sharp, consider a configuration with of the particles having identical initial conditions, thus identical motion for all : for . Necessarily, the last particle has Let and denote the common amplitude and polar angle for the first particles.
From Theorem 4.1 there exists such that For this configuration on account that we get
(58) |
whenever the trigonometric function is positive.

Recall that meaning that the polar angle is strictly increasing (to infinity), is Lipschitz (with constant close to one for intervals near infinity), and is comparable in size with Let be large. Let and be time values when and We have that both and are comparable to , and that for in the amplitude is comparable to . For such times, modulo the polar angle reduces to an angle in where the trigonometric function is above , see Figure 9. These estimates, and (58) prove
For this configuration, the drift of the direction of motion, satisfies: Next we show that the length of the interval is bounded below and above by absolute constants: the functions of and its inverse have Lipschitz constants close to one, so there exist positive constants and such that
We conclude that is comparable to and that the drift exceeds ∎
5 Appendix
Proposition 1.2 If among the agents of the system (1) there exist agents (assumed to be the first ) whose motion is symmetrically distributed about the center of the swarm, i.e.
(59) |
then either for all , or for all
Proof.
Within this proof, we identify all the planar vectors as points in so that we can multiply, divide, and take complex conjugate.
Assume that there exists such that (59) holds while the center of mass is not stationary, i.e. there exists a time interval with for We want to show that there exists an open time interval such that all particles’ locations equal (If such an interval exists, it means the velocities for the particles match in that interval, so the particles have identical initial conditions; by the uniqueness of solutions to (1) we get that the particles satisfy for all .)
To simplify notations, within this proof 555Outside of this Appendix, we used and to denote real-valued functions and with a different meaning. In this proof are constants, and is complex-valued. let for and let and be complex valued function defined by:
(60) |
We get that for
(61) |
To prove the appendix, we need to show that if then there exists a time interval where for
Using (61), the first equations of (2) become
Divide by ; recall that We get
Divide by and collect the terms to get that for and
(62) |
Employ the notations: and to further simplify (62) to:
(63) |
Lemma 5.1.
If are such that
(64) |
then is the positive real number and
Proof.
Note that
(65) |
and that the roots of unity, satisfy:
(66) |
Averaging over the equations (65 ), from (66) we get
(67) |
Rewrite the equations of (64) as
and average over We get
(68) |
Substituting in the left side of (64) yields
Solve for :
Averaging over , based on (66) we obtain
Substituting into (68) gives concluding the proof of the lemma. ∎
Rewriting (63) using Lemma 5.1, we get that are solutions for the equations
(69) |
implying that if the time is fixed, the equations have at least distinct complex solutions (since are all distinct solutions).
Case 1: There exists a time when is nonzero. We claim that in this case
At a time when , necessarily . Express using polar coordinates as where Since the solutions to satisfy , they are collinear (on the line through the origin, making an angle or with the axis).We conclude that all the points are collinear, which is impossible for the roots of unity with
Case 2: is the zero function on the interval We claim that this implies that for all or
Assume not: then and there exists a subinterval of with for in Since we get that and therefore the equations (69) become: at times
(70) |
For a fixed the equation can be satisfied by at most two roots of unity, because otherwise we would have three or more roots satisfy implying that the circle circumscribing those three roots of unity is centered at and has radius rather than being the unit circle, a contradiction. Satisfying (70) when requires that the factor has a zero of its own, implying that the continuous function takes values in the discrete set Thus there exists an index, denoted such that for all We will show this together with (60) imply that contradicting the starting assumption of
Based on and (61) we get that on
meaning that the location of the agent remains unchanged in time. The equation of motion states thus is constant, so contradicting on
Our two cases show that if then necessarily on the interval We get that and for all in Substituting and in (1) yields that for multiple angles the equalities hold, which require that and therefore all particles’ positions coincide on
∎
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.