Despite their simplicity, networks of coupled phase oscillators can give rise to intriguing collective dynamical phenomena. However, the symmetries of globally and identically coupled identical units do not allow solutions where distinct oscillators are frequency-unlocked—a necessary condition for the emergence of chimeras. Thus, forced symmetry breaking is necessary to observe chimera-type solutions. Here, we consider the bifurcations that arise when full permutational symmetry is broken for the network to consist of coupled populations. We consider the smallest possible network composed of four phase oscillators and elucidate the phase space structure, (partial) integrability for some parameter values, and how the bifurcations away from full symmetry lead to frequency-unlocked weak chimera solutions. Since such solutions wind around a torus they must arise in a global bifurcation scenario. Moreover, periodic weak chimeras undergo a period-doubling cascade leading to chaos. The resulting chaotic dynamics with distinct frequencies do not rely on amplitude variation and arise in the smallest networks that support chaos.

Networks of coupled oscillators occur in a wide range of systems in biology, medicine, and technology. The proper functioning of such systems often relies on the emergence (or absence) of collective modes such as synchronization, where oscillators lock their frequencies and/or phases1,2—a phenomenon that can be studied in the prominent Kuramoto model of phase oscillators.3,4 Chimeras are symmetry-breaking solutions where part of the population synchronizes while the other oscillates incoherently,5 even if oscillators are identical. While this phenomenon has received much attention in recent years,6,7 the bifurcations of chimera solutions in finite networks have remained elusive. Previous studies8,9 concerning variants of the Kuramoto model, composed of two populations of identical oscillators with heterogeneous intra- and inter-population coupling/phase lags, found that bifurcations leading to chimera states emerge for phase lags near multiples of π2—these are parameters that can be related to resonance points in physical/mechanical oscillator models.10,11 Here, we analyze such a variant of the Kuramoto model with two populations with two phase oscillators near these parameter points. We present a detailed analysis of how phase space is organized and elucidate the emergence and bifurcations of chaotic weak chimeras.9,12,13

Coupled phase oscillator networks have been instrumental in understanding collective dynamic phenomena in real-world oscillatory systems.1,2 The state of each dynamical node is given by a single phase variable on the circle T:=R/2πZ. To understand the dynamics, in many cases, it makes sense to assume an idealized system where all nodes are identically connected to all other nodes (such as in the Kuramoto model3,4) and/or all nodes are identical (for example, in the classical master stability approach14). Both of these assumptions together lead to full permutational symmetry, which restricts the possible dynamics: While these systems can still exhibit intriguing collective dynamics,15,16 the symmetry restricts the oscillators’ frequencies: For networks of phase oscillators, all nodes have to be frequency synchronized since oscillators’ phases cannot pass each other.8 

Dynamics where different oscillators evolve at different (asymptotic) frequencies have attracted significant attention in recent years, a phenomenon commonly known as a chimera (see Refs. 7, 17, and 18 for reviews). As such solutions/invariant sets are impossible in fully symmetric phase oscillator networks, breaking system symmetries—also known as forced symmetry breaking19—is necessary for frequency-unlocked solutions to arise. A less symmetric system will have fewer invariant subspaces that restrict the dynamics and create the possibility for frequency-unlocked chimera-type solutions. Note that we focus here on first-order phase oscillators rather than systems with additional degrees of freedom in which chimera phenomena have been observed. These include oscillators with amplitude variation20–22 or second-order phase equations;23 due to the larger phase space, chimera-type dynamics24,25 can occur in these extended systems also for fully symmetric networks.

A classical way to break full permutational symmetry is to consider coupled populations of phase oscillators: While oscillators are still all-to-all coupled, distinct populations form when coupling within populations can differ from the coupling between populations. Such a coupling structure for L=MN phase oscillators of Kuramoto-type, where the coupling is mediated by a single harmonic, arises by introducing distinct coupling strengths Ks,Kn>0 and phase lags αs,αnT within and between populations. Specifically, the phase θσ,k of oscillator k{1,,N} of population σ{1,,M} evolves according to

(1)

In other words, coupling between oscillators is mediated by the coupling functions gs(ϕ)=Kssin(ϕαs), gn(ϕ)=Knsin(ϕαn) that determine the self-coupling within populations and coupling to other populations (its “neighbors”), respectively. While system (1) is not symmetric with respect to any permutation of oscillator indices, it still has a wreath product symmetry that permutes oscillators within populations or permutes the populations26 (see also Ref. 27). In particular, the populations in (1) are interchangeable in contrast to networks with distinct populations.28 

Splitting the oscillators into just M=2 populations can lead to frequency-unlocked chimera dynamics. Indeed, the classical work by Abrams et al.29 demonstrated the emergence of chimera solutions for two coupled populations with identical phase lags αs=αn but disparate coupling strength KnKs in the limit of N oscillators. More generic interactions with nonidentical phase lags αsαn in the limit of large networks may lead to chaotic collective dynamics.9,13 Chimera solutions not only arise in the mean-field limit but also for finite networks6 and general sinusoidal coupling is sufficient to obtain chaotic chimera dynamics.13 While it was already indicated in Ref. 13 that chaotic dynamics are possible even in the smallest networks of M=2 populations of N=2 oscillators, the bifurcations that lead to chaotic chimera dynamics remained elusive.

The main focus of this paper is to analyze small networks that consist of M=2 populations of N=2 oscillators with generic sinusoidal coupling of Kuramoto-type (1) [see Fig. 1(a)]. This is the smallest network of this type in which chaotic dynamics can occur. Even as the full permutational symmetry is broken, phase space is organized by special structures for specific parameter: We first analyze degenerate cases where the dynamical system is either a gradient system or Hamiltonian-like with conserved quantities. These special cases allow us to understand how frequency-unlocked dynamics arise in these small networks. Perturbing away from the special parameter values yields global bifurcation scenarios that give rise to frequency-unlocked limit cycle solutions—they correspond to weak chimeras. As such, these solutions have nonzero winding number, wrap around the torus, and have nontrivial homology. How (some of) these solutions bifurcate further is shown in the numerical bifurcation diagram in Fig. 1(b) in terms of the phase difference ψ1:=θ1,1θ1,2 (note that further branches co-exist): The dynamics eventually undergo a period-doubling cascade that leads to chaotic weak chimeras. Finally, we highlight that for certain parameter values, the system has simultaneously conservative and dissipative dynamics in different regions of phase space.

FIG. 1.

The road to chaotic weak chimeras for networks of M=2 populations of N=2 oscillators (1) with sinusoidal coupling. The network is sketched in panel (a) and has the symmetries of the square D4. Panel (b) shows a bifurcation diagram in the phase difference ψ1:=θ1,1θ1,2 of the first population obtained via quasi-continuation (see Sec. V C) for fixed relative coupling strengths KsKn=0.7 and phase lag αs=0.44. Vertical lines delineate αs-values for the trajectories shown later in Fig. 9, referring to the corresponding panel labels. Panel (c) shows the maximal Lyapunov exponent for varying (αs,αn) calculated by numerically integrating from a fixed initial condition for T=10000 time units. The dashed line indicates parameter values shown in (a).

FIG. 1.

The road to chaotic weak chimeras for networks of M=2 populations of N=2 oscillators (1) with sinusoidal coupling. The network is sketched in panel (a) and has the symmetries of the square D4. Panel (b) shows a bifurcation diagram in the phase difference ψ1:=θ1,1θ1,2 of the first population obtained via quasi-continuation (see Sec. V C) for fixed relative coupling strengths KsKn=0.7 and phase lag αs=0.44. Vertical lines delineate αs-values for the trajectories shown later in Fig. 9, referring to the corresponding panel labels. Panel (c) shows the maximal Lyapunov exponent for varying (αs,αn) calculated by numerically integrating from a fixed initial condition for T=10000 time units. The dashed line indicates parameter values shown in (a).

Close modal

The paper is organized as follows. In Sec. II, we collect basic information about the governing equation (1), identify invariant subspaces and equilibria, and analyze their linear stability. In Sec. III, we consider the case of pure sine coupling (odd coupling) for which the equations of motion are a gradient system. In Sec. IV, we consider the case of pure cosine coupling (even coupling) that leads to the emergence of conserved quantities; the phase space structure elucidates the emergence or frequency-unlocked solutions corresponding to weak chimeras. Subsequently, in Sec. V, we show how these weak chimeras bifurcate. Specifically, we detail the bifurcations in Fig. 1 that lead to the emergence of chaotic dynamics. In Sec. VI, we briefly comment on coexisting conservative–dissipative dynamics for combined pure sine/cosine coupling. We conclude with a discussion in Sec. VII; here, we also briefly consider the dynamics of small networks with M=3 populations.

Before we analyze coupled phase oscillator networks, we briefly recall some notions related to dynamical systems that are equivariant with respect to the action of a group of symmetries. Let F:XTX be a smooth vector field on X where TX denotes the tangent bundle. Suppose that a group Γ acts on X. The vector field F is Γ-equivariant if

for all γΓ where γ^ is the induced action on the tangent space. A Γ-equivariant vector field F defines a Γ-equivariant dynamical system,

(2)

on X.30,31 The group Γx:={γΓ|γx=x} denotes the stabilizer or isotropy subgroup of xX. While the isotropy subgroup describes the symmetries of a single point, the symmetries of a set XX are Σ(X):={γΓ|γXX} where γX={γx|xX}. If HΓ is a subgroup, then the set Fix(H):={xX|γx=xγH} is invariant under the flow induced by (2).

While the later analysis focuses on small networks, we first collect general properties of the general network (1) consisting of M populations of N phase oscillators. We consider the coupling of Kuramoto type with gs(ϕ)=Kssin(ϕαs) and gn(ϕ)=Knsin(ϕαn) throughout the paper but the symmetry analysis in this section is independent of this specific choice. Since the coupling is through phase differences only, the system is T-equivariant: There is a continuous T symmetry where γT acts on the phase space TL by a common phase shift θσ,kθσ,k+γ to all oscillators. For Kuramoto-type coupling, note that by rescaling time, we can set Ks+Kn=1 without loss of generality and parameterize the coupling strength through the disparity parameter A:=KsKn (see also Ref. 29).

1. Globally and identically coupled oscillators

If the coupling within and between populations is identical, that is, A=0 (Ks=Kn) and αs=αn=:α, then we have a globally coupled network of L=MN identical oscillators. Omitting the population index σ, write θ=(θ1,,θL) and the evolution of the phase θk of oscillator k{1,,L} is

(3)

System (3) is SL-equivariant, where SL denotes the symmetric group on L symbols, which acts on TL by permuting the oscillator indices. This implies that the sets Pmn:={θ|θm=θn} are dynamically invariant as fixed-point sets of the transposition that swaps oscillators m and n. Thus, the dynamics can be reduced to the canonical invariant region

that is bounded by Pnm (see Refs. 32 and 33 for details). Indeed, since the coupling function only contains a first harmonic, the canonical invariant region is foliated into dynamically invariant sheets on which the dynamics are effectively two-dimensional.34 

The symmetries and type of coupling imply that certain phase configurations are dynamically invariant. The phase configuration

(4)

where all oscillators are phase synchronized is dynamically invariant as the intersection of all Pmn. Phase synchrony is typically quantified by the Kuramoto order parameter

(5)

with i:=1: We have |Z|=1 if and only if θ0. Define the antiphase or incoherent phase configurations as

(6)

The set I is dynamically invariant for globally coupled networks and is a union of manifolds (see Ref. 33).

The symmetries also constrain the frequencies of the oscillators. For a solution θ(t) of (3) with initial condition θ(0)=θ0, define the asymptotic average frequency24,35 as

(7)

Since the sets Pmn are dynamically invariant, the phase difference between oscillators is bounded. This implies that

(8)

for all m,n{1,,L}.

While no precise mathematical definition of a chimera exists (but attempts to classify them36), a distinguishing feature of a weak chimera (as defined in Ref. 33) is the separation of frequencies: A trajectory θ(t) with initial condition θ(0)=θ0 converges to a weak chimera if there are distinct m,n,n such that Ωm(θ0)Ωn(θ0)=Ωn(θ0). Thus, weak chimeras cannot exist in fully symmetric systems.

2. Breaking full symmetry: M populations of N oscillators

Breaking the symmetry is necessary to observe weak chimeras in a network of phase oscillators: Here, we now consider the dynamics of M populations of N oscillators each. Equation (1) that determines the evolution is SNSM=(SN)MSMSL equivariant, where denotes a semidirect product. Each copy of SN acts on TL by permuting the oscillator indices within a given population σ and SM acts by permuting the populations (i.e., the population indices σ); the symmetries within and between populations do not necessarily commute and, hence, the semi direct product.

The system has fewer symmetries than the fully symmetric system. For example—in the notation of Sec. II A1—the sets Pmn are only dynamically invariant if oscillators m,n belong to the same population. Note that the oscillators in each population are still fully symmetric. This implies that for (1) there are invariant sets Pσ,mn:={θ|θσ,m=θσ,n} for m,n{1,,N} and any σ{1,,M}.

The (global) Kuramoto order parameter Z is defined as in (5); it quantifies coherence of all oscillators in the network. Naturally, we also define phase synchrony and incoherence for each population: The (local) Kuramoto order parameter for population σ is

(9)

and we have Z=1Mσ=1MZσ.

The main focus of our analysis is the smallest multi-population phase oscillator networks that allow for chaotic dynamics. Consider networks (1) that consist of M=2 populations of N=2 oscillators; the network is sketched in Fig. 1(a). In this case, the governing equations (1) read

(10a)
(10b)
(10c)
(10d)

with ω=ωKs4sinαs. If the coupling is fully symmetric, i.e., Ks=Kn and αs=αn, then (10) is S4-equivariant. This means that up to the rotational symmetry, the canonical invariant region is a tetrahedron bounded by the dynamically invariant Pmn (as in Sec. II A 1) (see Refs. 33 and 37 for details). The phase configuration 0 with full phase synchrony corresponds to each of the four corners of the tetrahedron. The incoherent phase configurations I in C are the points with S2×S2 isotropy, that is, a line of points parameterized by (a,b,a+π,b+π)T4.

For a generic choice of coupling parameters, the dynamical system (10) is S2S2D4-equivariant, the symmetries of the square [cf. Fig. 1(a)]. Specifically, D4 is generated by the rotational symmetry

(11a)

[a counterclockwise rotation by an angle of π2 in Fig. 1(a)] and the mirror symmetry

(11b)

[a flip about the main diagonal in Fig. 1(a)]. General properties of D4-equivariant phase oscillator networks can be found in Ref. 32. The codimension-1 invariant subspaces are P1,12={θ1,1=θ1,2} and P2,12={θ2,1=θ2,2}. Since we only have two oscillators per population, we will write P1:=P1,12, P2:=P2,12 for simplicity.

Three values of the disparity parameter stand out where the network structure is qualitatively different. As already discussed, A=0 corresponds to Kn=Ks, which is the strength of interactions within and between populations is the same. If in addition also αn=αs, the oscillators have full permutational symmetry and we have the Kuramoto–Sakaguchi equations for identical oscillators. If A=1, we have Kn=0 and the populations are uncoupled; this means that the green links are absent in Fig. 1(a). If A=1, we have Ks=0 and the oscillators within populations are uncoupled. This configuration is equivalent to a ring of oscillators with nearest-neighbor coupling as illustrated in Fig. 1(a): If the purple links are absent, for a given oscillator, the two oscillators of the other population are the direct neighbors.

1. Reduced dynamics and symmetries

The phase-shift symmetry of (10) means that the system dynamics is effectively three-dimensional. We exploit this by rewriting the system in terms of phase differences ψ1:=θ1,1θ1,2,ψ2:=θ1,2θ2,1,ψ3:=θ2,1θ2,2. This yields the reduced system

(12a)
(12b)
(12c)

If the coupling is fully symmetric, i.e., A=0 (Ks=Kn) and αs=αn, then the phase space is organized by invariant planes Pnm, the canonical invariant region, and its images under the symmetry. In the reduced system (12), these invariant planes are given by ψ1=0, ψ3=0, ψ2=0, ψ1+ψ2=0, ψ2+ψ3=0, ψ1+ψ2+ψ3=0, which split T3 into six dynamically invariant regions, the canonical invariant regions and its images. The geometry is illustrated in Fig. 3(a) where one invariant region is bounded by coordinate planes and a light gray plane ψ3=2πψ1ψ2. The intersection of all planes corresponds to full phase synchrony. As defined above, the set of incoherent phase configurations are the points where the (global) Kuramoto order vanishes; in phase difference coordinates, these are given by

For arbitrary parameters αs, αn, A, the network (10) has dihedral symmetry S2S2D4. In phase differences (12), the generators of the symmetry action (11) are

(13)
(14)

that correspond to the rotation and reflection mentioned above. Solutions transform in a nontrivial way along variable ψ2 that describes the state of the two populations relative to one another under γr, γm. The system also has parameter symmetries

(15)
(16)
(17)

where (17) is time reversing for A<0.

2. Equilibria, invariant subspaces, and linear stability

The four oscillator network (12) supports phase-synchronized (or coherent) solutions. Specifically, there are fixed points

(18)
(19)

The first fixed point corresponds to the phase configuration where all oscillators are in phase (the point of full isotropy where all phases are equal) [cf. Fig. 2(a)]. The second corresponds to the configuration where both populations are in phase but antiphase with respect to each other [cf. Fig. 2(b)].

FIG. 2.

Phase configurations in a network of M=2 populations of N=2 oscillators: (a) full synchrony 0, (b) antiphase clusters π, (c) antiphase configurations (π,0,π) and (π,π,π), (d) set I0, (e) set I1 and I2, (f) set S0, (g) sets S1 and S2, (h) cluster with |Z1|=1 (ψ1=0) (includes flat chimera), (i) cluster with |Z2|=1 (ψ3=0) (includes flat chimera), (j) two-cluster configuration (isotropy S3×S1 in the globally coupled system), (k) the plane ψ2=0 (isotropy S2 in the globally coupled system), and (l) schematic representation of the weak chimera states when oscillators inside each cluster are phase locked (|θσ,1θσ,2|<2π for all times) and two groups are phase unlocked (rotate relative to one another).

FIG. 2.

Phase configurations in a network of M=2 populations of N=2 oscillators: (a) full synchrony 0, (b) antiphase clusters π, (c) antiphase configurations (π,0,π) and (π,π,π), (d) set I0, (e) set I1 and I2, (f) set S0, (g) sets S1 and S2, (h) cluster with |Z1|=1 (ψ1=0) (includes flat chimera), (i) cluster with |Z2|=1 (ψ3=0) (includes flat chimera), (j) two-cluster configuration (isotropy S3×S1 in the globally coupled system), (k) the plane ψ2=0 (isotropy S2 in the globally coupled system), and (l) schematic representation of the weak chimera states when oscillators inside each cluster are phase locked (|θσ,1θσ,2|<2π for all times) and two groups are phase unlocked (rotate relative to one another).

Close modal

Linear stability of the solutions 0 and π is determined by the eigenvalues

(20a)
(20b)
(20c)

respectively; the signs are for 0 and π, respectively.

The incoherent phase configurations in I can be parameterized as three lines

(21a)
(21b)
(21c)

such that I=I0I1I2. The set I0 is the phase configuration where each population is incoherent (that is, the phase difference of the two oscillators is π), and the sets I1 and I2 correspond to phase configurations where oscillators in distinct populations have a phase difference of π [cf. Figs. 2(c) and 2(d)].

The line I0 is a continuum of equilibria. Linearizing the equations, we obtain that linear stability is determined by the eigenvalues,

(22a)
(22b)
(22c)

Writing a=Ks4cosαs, b=Kn4cos(αn+φ), and c=Kn4cos(αnφ) we have that λ2,3=2(abc). This yields the linear stability properties of the equilibria: Phase configurations in I0 are linearly stable if a<0 and bc<a2 or if a<min{0,bc}. How the equilibria bifurcate depends on φ: The eigenvalues λ2,λ3 are a complex conjugated pair if φ(|αnπ2|π,|αnπ2|)(|αnπ2|,π|αnπ2|), which suggests a (degenerate) Hopf bifurcation for Kscosαs=0. For other φ, the eigenvalues are real. In the limiting case of global coupling (Ks=Kn:=1, αs=αn=α), the eigenvalues evaluate to

Now consider the incoherent phase configurations I1 and I2. These are invariant lines of (12) for any parameter values. The dynamics on each of these lines is given by

Thus, I1 and I2 are continua of nonisolated fixed points if αs=αn=±π2 as in the traditional Kuramoto–Sakaguchi equations. Linear stability of these fixed points for αs=αn=α=±π2 is determined by the nontrivial eigenvalues

This implies that the set of incoherent phase configurations I=I0I1I2 are surrounded by sets of periodic orbits when α=±π2 and KnKs>0 (|A|<1). If αs,αn±π2, the invariant lines persist and are surrounded by spiral trajectories.

Finally, we consider the set of two-cluster synchronized phase configurationsS, which are the points S2×S2 isotropy in the globally coupled system: These are the points where there are two clusters of two oscillators each. In phase difference coordinates on T3 for (12), we define

and have S=S0S1S2 [see Figs. 2(f) and 2(g)]. The set S0 consists of the phase configurations where each of the populations is phase synchronized—this means that S0=P1P2={ψ1=0}{ψ3=0} (cf. Fig. 3). The sets S1,S2 are phase configurations where two oscillators of distinct populations are phase synchronized. The set S is dynamically invariant for any values of αs, αn, and A; S0 is the fixed-point set of a subgroup of D4. Now, the dynamics on S0 are determined by

For αn±π2, there are exactly two fixed points on this line that corresponds to 0 and π, respectively. For αn=±π2, the set S0 is a continuum of fixed points. Linearization yields the eigenvalues

FIG. 3.

The phase space of the coupled oscillator network (12) for (ψ1,ψ2,ψ3)T3 is organized by invariant manifolds. Panel (a) shows the invariant lines for identical Kuramoto–Sakaguchi coupling Ks=Kn=K (A=0) and αs=αn=α. The sets I0, I1, I2 are shown as red lines, the sets S0, S1, S2 with isotropy S2×S2 as green lines, and blue lines are points of isotropy S3×S1. Blue balls indicate synchronous state 0=(0,0,0) and its copies, and red balls show the fixed point π=(0,π,0) as well as the points (π,0,π), (π,π,π) with isotropy (S2)2×Z2 (and their copies). The last fixed points locate on intersections of I. Light gray plane together with coordinate planes ψ=0 forms one of the canonical invariant regions. Panel (b) shows the phase space T3 for the uncoupled groups Kn=0 (A=1) and αs=±π2. The planes ψ3=ψ1 and ψ3=ψ1 are filled with fixed points. These two planes split phase space into two regions that are filled with straight lines (phase-unlocked periodic orbits) oriented in two opposite directions. Panel (c) gives a schematic diagram of the relative position of invariant lines and one of the invariant cylinders L(C) [cf. (28)] for the Hamiltonian-like case αs=αn=±π2, A(0,1). Blue lines on the cylinder correspond to a heteroclinic (homoclinic on the torus) network. Orange line shows one of the continuous family phase uncoupled periodic orbits that correspond to chimeras with a nontrivial winding number. There are two chimera strips of periodic orbits bounded by heteroclinic chimeras that have an opposite orientation along variable ψ2. Panel (d) shows the projection of cylinders L(C) into the plane (ψ1,ψ3) for different values of parameter C[0,1] [typical level lines for the surface (28)].

FIG. 3.

The phase space of the coupled oscillator network (12) for (ψ1,ψ2,ψ3)T3 is organized by invariant manifolds. Panel (a) shows the invariant lines for identical Kuramoto–Sakaguchi coupling Ks=Kn=K (A=0) and αs=αn=α. The sets I0, I1, I2 are shown as red lines, the sets S0, S1, S2 with isotropy S2×S2 as green lines, and blue lines are points of isotropy S3×S1. Blue balls indicate synchronous state 0=(0,0,0) and its copies, and red balls show the fixed point π=(0,π,0) as well as the points (π,0,π), (π,π,π) with isotropy (S2)2×Z2 (and their copies). The last fixed points locate on intersections of I. Light gray plane together with coordinate planes ψ=0 forms one of the canonical invariant regions. Panel (b) shows the phase space T3 for the uncoupled groups Kn=0 (A=1) and αs=±π2. The planes ψ3=ψ1 and ψ3=ψ1 are filled with fixed points. These two planes split phase space into two regions that are filled with straight lines (phase-unlocked periodic orbits) oriented in two opposite directions. Panel (c) gives a schematic diagram of the relative position of invariant lines and one of the invariant cylinders L(C) [cf. (28)] for the Hamiltonian-like case αs=αn=±π2, A(0,1). Blue lines on the cylinder correspond to a heteroclinic (homoclinic on the torus) network. Orange line shows one of the continuous family phase uncoupled periodic orbits that correspond to chimeras with a nontrivial winding number. There are two chimera strips of periodic orbits bounded by heteroclinic chimeras that have an opposite orientation along variable ψ2. Panel (d) shows the projection of cylinders L(C) into the plane (ψ1,ψ3) for different values of parameter C[0,1] [typical level lines for the surface (28)].

Close modal

Note that the invariant planes P1={ψ1=0} and P2={ψ3=0} bound the phase differences within each population (see Fig. 2 for the corresponding phase configurations). Thus, Ω11,12=0 and Ω21,22=0. By contrast, the phase difference ψ2 between the two populations may not be bounded—this corresponds to weak chimeras. In Secs. IIIVI, we will elucidate the dynamical mechanisms that lead to such solutions.

If the coupling function is a pure sine function (αs=αn=0), system (10) is a gradient system.38 More precisely, with the potential

Eq. (10) can be written as

(23)

for σ,k{1,2}. The system has a parameter symmetry given by the action of γ(A) as defined in (17). This allows us to restrict the parameter range to A(1,1) and makes the bifurcation behavior of the system for the special parameter values A=±1 (corresponding to uncoupled populations and a ring topology, respectively) more transparent. Note that for these parameter values γ() is an additional Z2-symmetry.

We first analyze the equilibria and their stability for αs=αn=0. According to (20), the stability of the coherent equilibrium 0 is determined by the eigenvalues λ1=(A1)/2,λ2,3=1/2. Thus, 0 is an attractor for A<1 and a saddle equilibrium otherwise. Similarly, the stability of the equilibrium π is determined by λ1=(1A)/2,λ2,3=A/2. Thus, π is a source for A0, a saddle for A(0,1), and a sink for A1 (with neutral linear stability along I0 for A=1). Finally, the linearization of the vector field at the incoherent invariant line I0={(π,φ,π)|φT} has eigenvalues (22), which evaluate to

Therefore, different points I0 have different transverse stability for certain values of the parameter A. Specifically, for A<1, the equilibria (π,φ,π) are stable if

where φ=arccos(1+A|1A|). Decreasing A from zero, the equilibria on I0 are (transversely) stable around φ=±π2; the part of I0 with transverse stability increases as A is decreased. Thus, there is multistability of the equilibrium 0 and segments of the manifold I0 if A<1. Conversely, for A(1,0) equilibria on I0 are repellers if φ(φπ,φ)(φ,πφ). For A>0, the entire line I0 is repelling. In all other cases, the points of I0 are of saddle type.

The system has a number of equilibria apart from 0,π, and the continuum I0. First, the points (π,0,0), (0,0,π), (π,π,0), and (0,π,π)—two-cluster configurations of one and three oscillators [cf. Figs. 2(j) and 2(k)]—are equilibria for any A when αs=αn=0. Linear stability is determined by the eigenvalues

These equilibria are of saddle type since Kn>0 implies λ1λ2<0, Kn<0 implies λ1λ3<0, and Kn=0 implies λ2λ3<0.

Second, for A>0, the system has four symmetric equilibria (2ψ,ψ,0), (2ψ,ψ,0), (0,ψ,2ψ), and (0,ψ,2ψ) with ψ=arccos(A1A+1). These equilibria are of saddle type and move along a straight line as the parameter A is varied. Pairs of these saddle points (in each plane ψ1=0, ψ3=0) disappear in pitchfork bifurcations with π at A=0.

The system bifurcates at A=1, A=0, and A=1. In each of these cases, the dynamics have additional symmetries and there can be conserved quantities.

If A=0, the system is fully symmetric and the well-known Kuramoto model for identical oscillators. The system has a global attractor 0 and repellers I=I0I1I2 consisting entirely of equilibria. Other equilibria with phase difference coordinates 0 and π are saddles in this case. Breaking the full permutational symmetry for A0 leads to the disappearance of fixed points from invariant varieties I1 and I2 but these lines remain invariant.

For A=±1, the system has conserved quantities. If A=1, the oscillator populations are uncoupled and there is a conserved quantity. We have

Proposition 1
For A=1 and any αs,αn (thus also for αs=αn=0), the system has a first integral

The proof is a computation that is given in  Appendix A. Moreover, for the specific gradient cases, we have an additional constant of motion; thus, trajectories are completely determined.

Proposition 2
For A=1 and αs=0 the system has a first integral
(24)

If A=1, there is no coupling between distinct populations and we have a network of identical Kuramoto oscillators on a ring. This system also has a conserved quantity.

Proposition 3
For A=1 and αs=αn=0, the system has a first integral
(25)

Note that the constants of motion H(1,) and H(1,0) do not depend on the interpopulation phase difference ψ2. Thus, all solutions belong to cylinder-like surfaces. Figure 4 shows the projections of surfaces determined by H(1,)(ψ1,ψ3)=C and H(1,0)(ψ1,ψ3)=C for fixed C[1,1] onto the plane (ψ1,ψ3) for different values of parameter C. In addition, for A=1, all solutions of the system belong to the parallel planes H(1,0)=β, βT: Each regular trajectory starts from a point of manifold I0 and tends to a point of the set ψ1=ψ3=0 [see Fig. 4(c)]. The manifold S consists completely of fixed points in this case. For A=1, the whole manifold I consists of saddle equilibria (transverse to I). In addition, the system has two more invariant lines of degenerate saddle points given by ψ1=π, 2ψ2+ψ3=0 and ψ1+2ψ2=0, ψ3=π. Stable and unstable one-dimensional invariant manifolds of each saddle belong are contained in the same cylinder H(1,0)=C for fixed C (again, the saddles are neutral in the third direction). All trajectories that are not equilibria or separatrices connect the source π to the attractor 0 [see Fig. 4(d)]. In particular, these trajectories on each cylinder H(1,0)=C are bounded by one-dimensional invariant manifolds of two saddle points.

FIG. 4.

Level lines of the first integrals H(1,)(ψ1,ψ3)=C, H(1,0)(ψ1,ψ3)=C for different constants C (projection of system solutions on the plane (ψ1,ψ3)T2). (a) Level lines for the first integrals of the system (12) for A=1 and arbitrary αs; (b) the same for A=1 and αs=αn=0. (c) and (d) show schematic phase portraits on level surfaces H(1,)=C and H(1,0)=C in coordinates (ψ2,ϕ), where ϕ is a parameterization variable along one of the level lines. The red dots in (a) and (b), as well as the red line in (c) correspond to the invariant manifold I0.

FIG. 4.

Level lines of the first integrals H(1,)(ψ1,ψ3)=C, H(1,0)(ψ1,ψ3)=C for different constants C (projection of system solutions on the plane (ψ1,ψ3)T2). (a) Level lines for the first integrals of the system (12) for A=1 and arbitrary αs; (b) the same for A=1 and αs=αn=0. (c) and (d) show schematic phase portraits on level surfaces H(1,)=C and H(1,0)=C in coordinates (ψ2,ϕ), where ϕ is a parameterization variable along one of the level lines. The red dots in (a) and (b), as well as the red line in (c) correspond to the invariant manifold I0.

Close modal

To summarize, for the gradient case, we observe the following stability/bifurcation behavior. For A<0, the system has multistability of 0 and two parts of the manifold I0 described in Sec. III A. Moreover, the equilibrium π is a repeller. For A(1,0), the system has only one attractor 0 and repellers π and parts of I0. For A>0 the entire line I0 becomes a repeller. For A(0,1), the equilibrium 0 is the only attractor. Finally, if A>1, the equilibrium π is an attractor and 0 is repelling. One-dimensional manifolds of equilibria only exist at the bifurcation values of A; except for I0, these do not persist as the degeneracies are broken.

Weak chimeras as solutions where the two populations have distinct frequencies arise for phase-lag parameters close to ±π2 (cf. Refs. 8, 9, and 12). Here, we elucidate the mechanisms that lead to such solutions. We mainly consider the singular case αs=αn=α=±π2: then, the vector field is conservative (Hamiltonian-like). Indeed, if the coupling function is even (this is the case if αs=αn=±π2), system (12) is divergence-free, that is, if G denotes the right hand side of (12), then

for any values of Ks and Kn (see also Ref. 33). This means that the system does not have any attractors or repellers and there are one-parameter families of fixed points and two-parametric families of periodic orbits or families of homo/heteroclinic cycles. Moreover, the system has the time-reversing symmetry given by

(26)

with fixed point space Fix(γ(t))={(ψ1,ψ2,ψ3)|ψ1=ψ3}. System (12) with parameters A, αs=±π2, αs=π2 is equivalent to the same system with parameters A1, αs=αn=±π2 due to the parameter symmetry

Consider the case that the phase-lag parameters αn,αs are identical, that is, α:=αn=αs=±π2. For A=0 (Kn=Ks), the system is fully symmetric and the phase space is organized into the canonical invariant region and its symmetric copies (cf. Sec. II B). Specifically, the sets S1, S2 are dynamically invariant as points with isotropy S2×S2. For α=±π2, these sets remain invariant even for nonidentical coupling strength A0 (this is not true in general): They form continua of equilibria whose linear stability is determined by the eigenvalues,

(27a)
(27b)

The zero eigenvalue corresponds to the direction along S1 or S2, respectively, and the equilibria are degenerate saddles or degenerate centers depending on the sign KsKn (or 1A2).

By direct calculation, one can verify the existence of a first integral (see also the constants of motion in Refs. 6 and 34).

Proposition 4
For αn=αs=±π2, system (12) has the first integral

The existence of the conserved quantity implies that the phase space is organized by invariant sets

(28)

parameterized by C(0,1). These are cylindrical surfaces if lifted to R3 and two-dimensional tori in T3; in the following, we will simply refer to L(C) as invariant cylinders. Note that we can also parameterize the invariant cylinders by their diameter d=42arccos(C) (at ψ3=ψ1). One of these cylinders is shown in Fig. 3(c) and typical projections of such cylinders on the (ψ1,ψ3)-plane for different values of C are shown in Fig. 3(d). In the limiting case of C=1, we have L(C)=I0. If C=0, it corresponds to the “square” cylinder of invariant planes ψ1=0, ψ1=π, ψ3=0, and ψ3=π. For C(0,1), the set L(C) contains eight equilibria, which correspond to the intersection of L(C) with the sets I1, I2 and S1, S2: With the parameterization of I,S as above, the intersection is at φ=2arcsin(C) and we have

The intersections are shown in Fig. 3(c): Each invariant line I, =1,2, intersects the cylinder L(C) twice, at the point I and the symmetric point I=I (or I=2πI in R3) for given C(0,1). The invariant line S intersects the cylinder L(C) at points S and its symmetric counterpart S=S. In the limiting case C=0, the points I1, I1, I2, I2 correspond to π (and its symmetric copies) while S1, S1, S2, S2 correspond to the fully synchronized phase configuration 0.

The global dynamics are determined by the dynamics on the invariant cylinders. To obtain coordinates on the invariant cylinder (torus), write the variables ψ1, ψ3 in polar coordinates ρ, ϕ. Now, the two-dimensional dynamics on L(C) can be expressed in coordinates (ψ2,ϕ); these depend on the value of C and may bifurcate as C is varied. Note that the dynamics for any C is present since C parameterizes the family of invariant tori that foliates phase space rather than being a system parameter.39,40 First consider the special case C=0; in this case, L(C)=L(0) corresponds to the degenerate cylinder that consists of the invariant planes. The dynamics on L(0) depend on the coupling parameter A as shown in Figs. 5(a)5(c): For A(0,1), there are continua of heteroclinic trajectories from equilibria in S0 as shown in Fig. 5(a); these are bounded by homoclinic trajectories from 0 to itself (on the torus) and degenerate to the equilibrium π. Moreover, there are families of periodic orbits with a nontrivial winding number (shaded areas). For A=0, these families of periodic orbits disappear as the stable and unstable manifolds of 0 form a heteroclinic “web” on L(0) [see Fig. 5(b)]. Finally, for A(1,0), there are homoclinic/heteroclinic trajectories from (points in) S0 to itself. For fixed A(0,1), the bifurcation scenario is similar with 0 replaced by Sj,Sj, j=1,2 and π replaced by Ij,Ij, j=1,2 [see Figs. 5(d)5(f)]. Define the critical cylinder size as

(29)

(Here, we used the symmetry γ(A,t) to obtain the formula for A1.) For C(0,C), we have families of periodic orbits where the populations rotate relative to each other. For C=C, we have a heteroclinic web between Sj,Sj. Finally, for C(C,1), there are periodic frequency-locked solutions.

FIG. 5.

The phase portraits on the surface of the cylinder L(C) for αs=αn=±π2 vary for different parameters A[0,1) and C. We show schematic phase diagrams on L(C) in the variables (ψ2,ϕ), where ϕT1 is an angle that parameterizes the curve sin(ψ1/2)sin(ψ3/2)=C. Phase portraits are skewed along the horizontal variable (ψ2ψ2+2ϕ) for a better presentation of symmetries of the original system (10). The left row shows the dynamics on the boundary cylinder, which consists of invariant planes: A(0,1), C=0 in panel (a); A=0, C=0 in panel (b); and A(1,0), C=0 in Panel (c). Green lines indicate the intersections of invariant planes P1,P2. The right row sketches the dynamics on the cylinders inside the invariant region: A(0,1), 0<C<C=C(A)<1 in Panel (d); A(0,1), C=C(0,1) in panel (e); and A(1,1), 1>C>C>0 in panel (f). Red points indicate the intersection of I with L(C), green points indicate intersections S with L(C) [they correspond to π in (b)], and blue points are 0. Light blue and yellow regions correspond to one-parametric families of periodic chimeras that have opposite directions. Blue lines show homoclinic and heteroclinic connections, which bound the regions with phase-unlocked dynamics in (a) and (d). Phase-locked (non-chimera) periodic orbits are shown in magenta. Orange and light blue curves in (c) and (f) also correspond to phase-locked solutions in phase space T3 of (12).

FIG. 5.

The phase portraits on the surface of the cylinder L(C) for αs=αn=±π2 vary for different parameters A[0,1) and C. We show schematic phase diagrams on L(C) in the variables (ψ2,ϕ), where ϕT1 is an angle that parameterizes the curve sin(ψ1/2)sin(ψ3/2)=C. Phase portraits are skewed along the horizontal variable (ψ2ψ2+2ϕ) for a better presentation of symmetries of the original system (10). The left row shows the dynamics on the boundary cylinder, which consists of invariant planes: A(0,1), C=0 in panel (a); A=0, C=0 in panel (b); and A(1,0), C=0 in Panel (c). Green lines indicate the intersections of invariant planes P1,P2. The right row sketches the dynamics on the cylinders inside the invariant region: A(0,1), 0<C<C=C(A)<1 in Panel (d); A(0,1), C=C(0,1) in panel (e); and A(1,1), 1>C>C>0 in panel (f). Red points indicate the intersection of I with L(C), green points indicate intersections S with L(C) [they correspond to π in (b)], and blue points are 0. Light blue and yellow regions correspond to one-parametric families of periodic chimeras that have opposite directions. Blue lines show homoclinic and heteroclinic connections, which bound the regions with phase-unlocked dynamics in (a) and (d). Phase-locked (non-chimera) periodic orbits are shown in magenta. Orange and light blue curves in (c) and (f) also correspond to phase-locked solutions in phase space T3 of (12).

Close modal

If A=1 (Kn=0), the network consists of two uncoupled populations that evolve independently of one another. In this case, γ(A,t) acts as a symmetry of the system since it maps AA1. The system (12) reduces to

and its solutions form a continuum of lines

parameterized by the initial conditions (ψ1(0),ψ2(0),ψ3(0))=(ψ10,ψ20,ψ30). The dynamics are schematically shown in Fig. 3(b). For most initial conditions, the phase difference ψ2 between the populations is increasing or decreasing monotonously. Specifically, we have that each population is frequency synchronized and they rotate relative to one another with Ω2,jΩ1,k=cos(ψ3(0))cos(ψ1(0)), j,k=1,2. Thus, ψ2 increases monotonically if |ψ3|>|ψ1| for ψ1,ψ3[π,π] and ψ2 decreases monotonically if |ψ3|<|ψ1| for ψ1,ψ3[π,π]. The direction of the flow is shown in Fig. 3(b) in the cube [0,2π]3. On the planes ψ3=±ψ1, the dynamics are trivial (there are nonisolated fixed points). Note that the invariant lines I0 and S0 are the intersections of these planes.

For coupling parameters A (Kn) close to the singular case of uncoupled populations, the coupling between populations is weak. Frequency unlocked solutions persist but now deviate from straight lines. This gives the frequency-unlocked solutions on the invariant cylinders L(C) for C<C=C(A) and given A(0,1) described above. For |A|>1, the parameter symmetry γ(A,t) implies that system (12) has the same phase portraits but with the shift ψ2ψ2+π and a potential time reversal. The symmetry also means that the saddles S, S swap places with centers I, I. Local bifurcations occur at each fixed point on the invariant manifolds I0, S0 of the system (two eigenvalues of each point transform from pure imaginary to real symmetric or vice versa). More globally, all invariant manifolds of the fixed points S swap with those of I, =1,2. Figure 6 schematically shows the bifurcation that occurs in the narrow stripe of phase space (ψ2,ϕ)L(C) at the bifurcation point A=1: There is a heteroclinic cycle between the saddle S (S) and its copy with coordinate ψ2+2π, which bounds a family of periodic orbits around the center I (I) [see Fig. 6(a)]. At the bifurcation point A=1, the phase space is foliated by straight lines as discussed above [see Fig. 6(b)]. For A>1, a new heteroclinic loop appears between the saddle I (I) that bounds periodic orbits centered at S (S). Frequency-unlocked chimera solutions exist for all A>1.

FIG. 6.

There is a degenerate bifurcation on the cylinders L(C), C(0,1), [see (28)] at A=1 for αs=αn=±π2 on. Panel (a) gives a schematic for A<1 before the bifurcation; Panel (b) at the bifurcation A=1; and Panel (c) for A>1. Phase portraits are given for ψ2[0,4π] and ϕ[ϕ,ϕ] for ϕ small. Red and blue points (balls) indicate I and S, respectively, for =1,2, blue lines are heteroclinic trajectories, light blue and yellow regions indicate continuous families of phase-unlocked (chimera) solutions. The heteroclinic cycle connecting S is compressed to the line ϕ=0 and a new heteroclinic cycle appears connecting I. At the bifurcation point, the whole line ϕ=0 consists of fixed points as saddles S become centers and centers I become saddles. The points I and S undergo an analogous bifurcation elsewhere on L(C).

FIG. 6.

There is a degenerate bifurcation on the cylinders L(C), C(0,1), [see (28)] at A=1 for αs=αn=±π2 on. Panel (a) gives a schematic for A<1 before the bifurcation; Panel (b) at the bifurcation A=1; and Panel (c) for A>1. Phase portraits are given for ψ2[0,4π] and ϕ[ϕ,ϕ] for ϕ small. Red and blue points (balls) indicate I and S, respectively, for =1,2, blue lines are heteroclinic trajectories, light blue and yellow regions indicate continuous families of phase-unlocked (chimera) solutions. The heteroclinic cycle connecting S is compressed to the line ϕ=0 and a new heteroclinic cycle appears connecting I. At the bifurcation point, the whole line ϕ=0 consists of fixed points as saddles S become centers and centers I become saddles. The points I and S undergo an analogous bifurcation elsewhere on L(C).

Close modal

We can now describe the fate of the frequency-unlocked chimera solutions in more detail. At the bifurcation point A=1, the whole phase space is filled with frequency-unlocked chimera solutions except on the invariant planes ψ3=±ψ1. Indeed, there are two “chimera tubes” with opposite directions of trajectories. As the parameter A deviates from one, narrow cylinders with phase-locked periodic orbits appear. The heteroclinic trajectories form the boundaries between the region of frequency locking and a region of frequency-unlocked chimera solutions. Tracing these trajectories along the invariant cylinders L(C) that foliate phase space for C(0,C), we obtain a tube of phase-unlocked chimera solutions (cf. Fig. 7). Since the surface bounding the tube resembles a snake, we will refer to it as a serpentine chimera region. Figure 7(c) shows boundary surface of the tube schematically (“skin of the chimera-snake”) that consists of piecewise smooth surfaces. As A0, the tube degenerates to a single heteroclinic connection with a nontrivial winding number. Given the parameter symmetry γ(A,t), we can conclude that there are phase-unlocked chimera solutions for any A>0.

FIG. 7.

Phase-unlocked dynamics between populations exist with a serpentine chimera region for (ψ1,ψ2,ψ3)[0,2π]×[0,4π]×[π,π]. Panel (a) shows the boundary surface (grey shading) in T3 of the region. It consists of a one-parametric family of homoclinic trajectories between the points S, S (green points and lines), two heteroclinic trajectories of points S:=S(C)2 and S:=S(C) (blue lines)—on the cylinder L(C)—and two homoclinic cycles of points 0 (magenta lines). The surface consists of four symmetric smooth pentagonal parts, some of which are related by symmetry. Red lines indicate I0; green lines indicate S1, S2; the magenta line is a homoclinic orbit of 0 within P2={ψ3=0}; the orange semicircle is a projection of L(C) onto ψ2=0. Panel (b) shows a projection of the region onto ψ2=0. Panel (c) shows the dynamics on the surface of the region flattened out; the magenta/purple lines are related by symmetry.

FIG. 7.

Phase-unlocked dynamics between populations exist with a serpentine chimera region for (ψ1,ψ2,ψ3)[0,2π]×[0,4π]×[π,π]. Panel (a) shows the boundary surface (grey shading) in T3 of the region. It consists of a one-parametric family of homoclinic trajectories between the points S, S (green points and lines), two heteroclinic trajectories of points S:=S(C)2 and S:=S(C) (blue lines)—on the cylinder L(C)—and two homoclinic cycles of points 0 (magenta lines). The surface consists of four symmetric smooth pentagonal parts, some of which are related by symmetry. Red lines indicate I0; green lines indicate S1, S2; the magenta line is a homoclinic orbit of 0 within P2={ψ3=0}; the orange semicircle is a projection of L(C) onto ψ2=0. Panel (b) shows a projection of the region onto ψ2=0. Panel (c) shows the dynamics on the surface of the region flattened out; the magenta/purple lines are related by symmetry.

Close modal

Finally, consider the case A=1 when the oscillators within each population are uncoupled (Ks=0)—they still remain coupled indirectly through the interaction with the other population. In this case, γ(A,t) acts as a time-reversing symmetry since it maps AA1. A direct calculation shows:

Proposition 5
For A=1 and αn=±π2, system (12) has the conserved quantity
Remark 1

Note that H(1,π2)=H(1,0) in Proposition 2.

Thus, for A=1, there are two conserved quantities, H(,π2) and H(1,π2), and, thus, solutions lie on the intersection of the invariant cylinders L(C) with the plane H(1)=C(1) for C(0,1), C(1)T, given by

Similar to the case A=1, local and global bifurcations occur when A=1. The time reversing symmetry organizes the solutions on the invariant cylinders (28): Invariant lines of saddle fixed points S mutually change the structure of their fixed points with the lines of center fixed points I, =1,2. Global bifurcation with phase-locked periodic orbits (lines in phase space (ψ2,ϕ)R2) for A=1 is of the same type that the global bifurcation with the appearance of the family of straight lines in the described case A=1. In the case A=1, global bifurcations occur on the square K(0,0)T3 (with vertices at the points 0, π and two their copies) instead of S0 as for A=1.

In the conservative case, a limit cycle solution with a nontrivial winding number along the phase difference ψ2 (the phase difference between the two populations) arises in phase space. How do these solutions bifurcate as parameters are varied and the additional structures are broken? In the following, we will refer to any limit cycle solution with a nontrivial winding number as a weak chimera.

Before we turn to chimeras, we consider the overall organization of phase space by invariant subsets and qualitatively describe what dynamics are possible. The conservative dynamics for αs=±αn=±π2 correspond to a bifurcation point. Recall that for these parameter values, the sets I1, I2 and S1, S2 are continua of equilibria; in particular, the linearization at each equilibrium has a zero eigenvalue with an eigenvector in the direction of the corresponding set. Away from the bifurcation point, there are only two equilibria for each set: These are {π,(π,0,π)}I1, {π,(π,π,π)}I2, {0,(π,0,π)}S1, and {0,(π,π,π)}S2. Before the bifurcation point, one of the two equilibria is repelling along the set while the other one is attracting. After the bifurcation, the situation is reversed. Hence, the bifurcation corresponds to reversing the flow along the four sets in a degenerate bifurcation. The stability transverse to the invariant set determines the dynamics of trajectories nearby: If there are complex conjugate eigenvalues, nearby trajectories spiral around the set as the conservative structure is broken.

Away from the bifurcation point αs,αn=±π2, the invariant cylinders [Fig. 3(c)] break up, which allows for more intricate dynamics. For small deviations from the bifurcation parameter αs=±π2ε, there is a slow drift transverse to the cylinders L(C) (which are not invariant anymore). Since the families of heteroclinic orbits that separate frequency-locked and frequency-unlocked solutions also disconnect, it is possible for trajectories to come close to the continuum of equilibria I0 as well as the degenerate cylinder L(0) that consists of the invariant planes ψ1=0 and ψ3=0.

Taken together, typical trajectories can exhibit dynamics that explore a large region of phase space beyond the bifurcation point. Trajectories can move from the vicinity of the points π (0), moving along the invariant line I1 or I2 (S1 or S2) in a spiraling fashion to approach I0, before then drifting along the level sets L(C) into the region where the phase difference between populations increases and approaching a frequency-unlocked solution on the invariant set L(0). In the following, we will consider the bifurcations of these asymptotic solutions.

We first consider the weak chimeras on the invariant planes P1,P2 defined by ψ1=0 or ψ3=0 (these correspond to the limiting cylinder L(0) in Sec. IV B); we will refer to these solutions as flat chimeras. These correspond to in-phase synchronization of one population while the other population is approximately in anti-phase.8,12 The relative phase difference between the two populations increases without bound as Ω1,20. Without loss of generality, we assume that ψ3=0, that is, the second population is phase synchronized. Dynamics (12) on the invariant subspace restrict to a two-dimensional system; therefore, no chaotic flat chimeras are possible. Recall that families of flat chimeras arise for αs=±π2, αn=±π2 as described above [cf. Fig. 5(a)]. This family of flat chimeras is usually bounded by homoclinic or heteroclinic cycles [Fig. 5(b)]. These flat chimeras are neurally stable transverse to the invariant plane ψ3=0 due to the (partial) integrability of the system.

Figure 8 elucidates the bifurcation structure for parameter values where the conservative structure is broken. Note that because of the symmetries, the system can have only an even number of fixed points on the plane ψ3=0. The system has 2, 4, or 6 fixed points depending on parameter values. Saddle-node and saddle-connection bifurcations must occur in pairs simultaneously. Specifically, the panels in Figs. 8(a)8(p) show all possible phase portraits for the dynamics of (ψ2,ψ1)T2; these are arranged to show bifurcation transitions from (a) to (b), from (b) to (c), and so on. The system exhibits the following bifurcations on the plane: A pitchfork bifurcation of 0 or π transverse to the invariant line ψ1=0 [Figs. 8(c)8(d), 8(k)8(l), 8(l)8(m), and 8(o)8(p)]; two simultaneous saddle-node bifurcations on the flat chimera limit cycle [Figs. 8(a)8(c) and 8(i)8(k)]; two simultaneous saddle-node bifurcations away from the chimera [Figs. 8(o)8(n)]; a subcritical pitchfork bifurcation of limit cycles [Figs. 8(h)8(g)]; a supercritical pitchfork bifurcation of limit cycles [Figs. 8(g)8(h)]; two simultaneous saddle-connection bifurcations [Figs. 8(f)8(g) and 8(m)8(n)].

FIG. 8.

Phase portraits on the invariant surface P2={ψ3=0} for (ψ2,ψ1)[0,2π]2 show the bifurcation behavior for increasing parameter αs[0,2π] keeping A=0.7, αn=0.44 fixed. Specifically, we have panel (a) αs=0; (b) αs=0.041143; (c) αs=0.1; (d) αs=1.378599; (e) αs=1.4; (f) αs=1.410375; (g) αs=1.546389; (h) αs=1.55; (i) αs=1.64; (j) αs=1.731144; (k) αs=3.125269; (l) αs=3.16; (m) αs=3.5; (n) αs=4.520192; (o) αs=4.687983; (p) αs=4.69; (q) αs=4.8; (r) αs=4.9. Fixed points are colored according to type: source (red), sink (blue), saddle (green), saddle-node (two-color). Stable limit cycles are shown in blue, unstable in red, homo/heteroclinic in magenta. All cycles are non-homologic to zero on the two-dimensional torus and they correspond to weak chimeras.

FIG. 8.

Phase portraits on the invariant surface P2={ψ3=0} for (ψ2,ψ1)[0,2π]2 show the bifurcation behavior for increasing parameter αs[0,2π] keeping A=0.7, αn=0.44 fixed. Specifically, we have panel (a) αs=0; (b) αs=0.041143; (c) αs=0.1; (d) αs=1.378599; (e) αs=1.4; (f) αs=1.410375; (g) αs=1.546389; (h) αs=1.55; (i) αs=1.64; (j) αs=1.731144; (k) αs=3.125269; (l) αs=3.16; (m) αs=3.5; (n) αs=4.520192; (o) αs=4.687983; (p) αs=4.69; (q) αs=4.8; (r) αs=4.9. Fixed points are colored according to type: source (red), sink (blue), saddle (green), saddle-node (two-color). Stable limit cycles are shown in blue, unstable in red, homo/heteroclinic in magenta. All cycles are non-homologic to zero on the two-dimensional torus and they correspond to weak chimeras.

Close modal

We highlight the bifurcations that relate to the creation and destruction of stable flat chimeras. First, there are two simultaneous saddle-node bifurcations on an invariant circle that lead to a limit cycle solution with nonzero winding number, a flat chimera [Figs. 8(a)8(c) and 8(k)8(m)]. Second, there are pitchfork bifurcations of limit cycles, both sub- [Figs. 8(h) and 8(i)] and supercritical [Figs. 8(p) and 8(q)]; these bifurcations can stabilize flat chimeras (within the invariant subspace). Third, there are simultaneous saddle-connection bifurcations of saddle equilibria that lead to the emergence of two symmetric limit cycles with unbounded phase difference ψ2 between populations; the resulting limit cycles can be stable [Figs. 8(f) and 8(g)] or unstable [Figs. 8(m) and 8(n)].

Note that the stability of flat chimeras in the full system (12) depends on the transverse stability with respect to the third direction.

Two populations of two phase oscillators with Kuramoto–Sakaguchi coupling support chaotic weak chimeras; the bifurcation scenario, obtained numerically,41 is indicated in Fig. 1. In the following, we describe the bifurcations that lead to the emergence of chaotic weak chimeras in more detail. Since |D4|=8, every point with trivial isotropy has eight images under the symmetry action. Recall that Σ(A) are the symmetries that preserve the set A. If Aη is a family of limit sets indexed by a parameter η and Σ(Aη) changes as η is varied, we have a symmetry-increasing bifurcation.42 

We describe the bifurcations as a single parameter is varied; for concreteness, we fix A=0.7 and αn=0.44 and vary αs. For these parameters, there exists a flat chimera (limit cycle) on P1P2 that consists of points with nontrivial isotropy as the plane is invariant under a reflection: Consider the stable limit cycle shown in Figs. 8(k) and 8(n) within Pσ. A numerical calculation shows that it is also stable transverse to Pσ. This limit cycle loses (transverse) stability in a pitchfork bifurcation of limit cycles which leads to two stable limit cycles with trivial isotropy that are mapped to each other through the reflectional symmetry. [Note that other stable limit cycles on Pσ, such as those shown in Fig. 8(h), which are unstable transversely and, therefore, lead to saddle limit cycles in a pitchfork bifurcation.] Since the bifurcations happen simultaneously on P1 and P2, there is a total of four such solutions. The projections of four chimera states into the (ψ1,ψ3)-plane are shown in Fig. 9(a). Note that each of the resulting limit cycles has a setwise reflectional symmetry. This symmetry is broken in a supercritical pitchfork bifurcation of limit cycles at αs1.616, which leads to a total of eight nonsymmetric weak chimera limit cycle solutions [cf. Figs. 9(a) and 9(b)].

FIG. 9.

In the full three-dimensional phase space, the chimera attractors bifurcate and eventually become chaotic; here projections of the trajectories onto (ψ1,ψ3)[0,2π]2 are shown where the D4 symmetry acts in the obvious way. Parameters are A=0.7 and αn=0.44 in all panels and αs as in Fig. 1. Panel (a) shows self-symmetric eight-shape stable limit cycles for αs=1.607; Panel (b) shows pairs of stable limit cycles for αs=1.63 that emerge from eight-shaped ones in a symmetry-breaking bifurcation; Panel (c)–(f) show limit cycles after period-doubling bifurcations with αs=1.635 in (c), αs=1.638 in (d); αs=1.6385 in (e) and αs=1.639 in (f); Panel (g) shows eight symmetric chaotic attractors for αs=1.6415; Panel (h) shows four symmetric chaotic attractors for αs=1.64166 after a symmetry-increasing bifurcation. Line thickness was reduced in (f)–(h) for visualization. Line styles/colors are used to distinguish solutions that are related by symmetry. The red dot indicates I0.

FIG. 9.

In the full three-dimensional phase space, the chimera attractors bifurcate and eventually become chaotic; here projections of the trajectories onto (ψ1,ψ3)[0,2π]2 are shown where the D4 symmetry acts in the obvious way. Parameters are A=0.7 and αn=0.44 in all panels and αs as in Fig. 1. Panel (a) shows self-symmetric eight-shape stable limit cycles for αs=1.607; Panel (b) shows pairs of stable limit cycles for αs=1.63 that emerge from eight-shaped ones in a symmetry-breaking bifurcation; Panel (c)–(f) show limit cycles after period-doubling bifurcations with αs=1.635 in (c), αs=1.638 in (d); αs=1.6385 in (e) and αs=1.639 in (f); Panel (g) shows eight symmetric chaotic attractors for αs=1.6415; Panel (h) shows four symmetric chaotic attractors for αs=1.64166 after a symmetry-increasing bifurcation. Line thickness was reduced in (f)–(h) for visualization. Line styles/colors are used to distinguish solutions that are related by symmetry. The red dot indicates I0.

Close modal

Each stable limit cycle loses stability in a period-doubling bifurcation [see Figs. 9(b) and 9(c)]. The resulting limit cycle bounds a Möbius strip that wraps around the torus in the ψ2 direction; the original limit cycle is of saddle type after the bifurcation and lies at the center of the strip. Note that the width of the Möbius strip varies along variable ψ2 and for different parameter values. This indicates the heterogeneity of the attraction strength along the periodic solution that will have a further impact on the formation of chaos. The first period-doubling bifurcation is followed by a chain of period-doubling bifurcations: The nth period-doubling bifurcation leads to a 2n-limit cycle as shown in Figs. 9(c)9(f). At each period-doubling bifurcation, the geometry of the Möbius strips becomes more elaborate, allowing trajectories to follow different directions as the trajectory wraps around the torus in ψ2 direction.

A chaotic attractor with nontrivial winding number (i.e., a chaotic weak chimera) emerges as a result of a period-doubling cascade as shown in Fig. 9(g). We estimate the fractal dimension of the chaotic attractor to be slightly larger than two. Roughly speaking, as trajectories wind around ψ2 on the attractor, they can either take a direct path, closer to the original limit cycle, or make an “excursion” in the ψ1 or ψ3 direction. The Poincaré section shown in Fig. 10(a) shows the finer structure of the attractor. The attractor undergoes a symmetry-increasing bifurcation as two symmetrically related attractors merge [Figs. 9(g) and 9(h)]; see also Ref. 24 and note that a similar effect has also been observed in oscillators with amplitude variations.22,25 The enlargement of the chaotic attractor can also be seen in the Poincaré section in Fig. 10(b); note that the section does not necessarily respect the attractor symmetry.

FIG. 10.

Poincaré sections show the finer structure of the chaotic attractors that arise in (12). Panel (a) shows the attractor in Fig. 9(g). Panel (b) shows the attractor in Fig. 9(h).

FIG. 10.

Poincaré sections show the finer structure of the chaotic attractors that arise in (12). Panel (a) shows the attractor in Fig. 9(g). Panel (b) shows the attractor in Fig. 9(h).

Close modal

As the parameter αs is increased beyond αs=1.64166, there is multistability between the equilibrium 0 and the four stable symmetric chaotic weak chimeras. Moreover, there are two saddle equilibria S on each of the invariant planes P1,P2, as shown in Fig. 8(h) or Fig. 8(m). The stable manifold of each saddle is two-dimensional, intersects the appropriate invariant plane transversely, and separates the basin of attraction of 0 and the chaotic attractors. Finally, the chaotic attractor is destroyed through a homoclinic tangency43–46 and trajectories eventually converge to 0. The transients in the vicinity of the former chaotic attractor can be very long before the stable equilibrium is approached.

To summarize, we have given an overview of the solutions of (12) that correspond to weak chimeras for two coupled populations of phase oscillators (10). All of these solutions share the property that the phase difference between the populations (ψ2 in the reduced dynamics) is unbounded as time evolves. As they wrap around the torus, the solutions are nonhomological to zero and must arise in global bifurcations as described above.

  1. Limit cycle solutions on the invariant planes P1={ψ1=0} or P2={ψ3=0} [Figs. 8(k)8(n)], referred to as flat chimeras. The situation corresponds to phase synchronization of one of the populations with local order parameter |Z1|=1 (or |Z2|=1) and |Z2|=|Z2(t)|(0,1) [or |Z1|=|Z1(t)|(0,1)].

  2. A one-parameter family of neutral periodic orbits on the invariant plane for αs=±π2, αn=0, αn=±π2, αn=π [Fig. 11(a)].

  3. The limit cycles with setwise symmetry [Fig. 9(a)].

  4. The limit cycles without symmetry [Figs. 9(b)9(f)].

  5. A two-parameter family of neutrally stable periodic orbits (parameters as in item 2) [cf. Figs. 3(c), 5(a), and 5(d)].

  6. The eight nonsymmetric chaotic attractors [Figs. 9(g), 10(a), and 10(c)].

  7. The four symmetric chaotic attractors that appear as the result of the symmetry-increasing bifurcation [Figs. 9(h), 10(b), and 10(d)].

Note that trajectories close to homoclinic/heteroclinic orbits that are nonhomological to zero [see Fig. 8(f)] can also show (transient) frequency-unlocked dynamics between the populations.

FIG. 11.

Coexisting conservative and dissipative arise for M=2 populations of N=2 phase oscillators (10). Here, schematic phase portraits on the cylindric surfaces H(1,0)=C are shown for variables (ψ2,ϕ)T2, where ϕ is a parameterization of the level line of the first integral. Panels (a)–(c) show the bifurcation transition that leads to the disappearance of the conservative region (shaded in yellow) filled with periodic (chimera) trajectories. Panel (d) shows dynamics for C=0 when level surface consists of two half-planes; the intersection of these planes is I0 (green line). Fixed points are colored according to type: source (red), sink (blue), saddle (green), saddle-node (two-color). Purple lines are heteroclinic orbits.

FIG. 11.

Coexisting conservative and dissipative arise for M=2 populations of N=2 phase oscillators (10). Here, schematic phase portraits on the cylindric surfaces H(1,0)=C are shown for variables (ψ2,ϕ)T2, where ϕ is a parameterization of the level line of the first integral. Panels (a)–(c) show the bifurcation transition that leads to the disappearance of the conservative region (shaded in yellow) filled with periodic (chimera) trajectories. Panel (d) shows dynamics for C=0 when level surface consists of two half-planes; the intersection of these planes is I0 (green line). Fixed points are colored according to type: source (red), sink (blue), saddle (green), saddle-node (two-color). Purple lines are heteroclinic orbits.

Close modal

Finally, we remark that the system also has an interesting and unusual dynamics when αs=±π2, αn{0,π}—this corresponds to the function of the coupling function gs that determines the coupling with populations being even and the interaction function gn between populations being odd. In this case, the system has the first integral given in Proposition 3; for simplicity, we use the same notation here.

Proposition 6

For αs=π2, αn=0 and arbitrary A the system has the first integral H(1,0) as defined in (25).

Thus, the dynamics of the system evolve on the level sets H(1,0)(ψ1,ψ3)=C for fixed C[1,1] [see Fig. 4(b)]. In the cases C=±1, we have a planar system on the invariant planes P1={ψ3=0} and P2={ψ1=0}. As |C| is decreased, the level sets deform continuously and limit to dynamics on the half-planes ψ3=ψ1 for ψ1[0,π] and ψ3=2πψ1 for ψ1[π,2π] (or their symmetric counterparts) for C=0.

This system has coexistence of conservative (Hamiltonian-like) and dissipative dynamics in phase space for αs=π2, αn=0, and A[A,1/A] with A0.38146. Conservative–dissipative dynamics are often related to the presence of time-reversing symmetries; such dynamics has been described, for example, in a three-dimensional laser system,47 globally coupled superconducting Josephson junction arrays,48 chains of locally coupled phase oscillators,49 and circulant networks of phase oscillators with skew-symmetric coupling.50 Here, the system has the conservative region filled with a two-parameter family of neutral periodic orbits (weak chimeras) and this region is bounded with the surface of one-parameter family of heteroclinic cycles forming a “heteroclinic tube.” Outside the Hamiltonian-like region, the dynamics is dissipative with attractor 0 and repeller π when A<1 or attractor π and repeller 0 when A>1. The boundary surface of the three-dimensional conservative region has a structure similar to that shown in Fig. 7(a). The system has two one-parameter families of saddle points (which are curved lines compared to the case αs=αn=±π2). There are heteroclinic (homoclinic in T3) cycles that consist of these saddles and their stable and unstable invariant manifolds on the same level surface (each saddle is neutral in transversal towards the level surface). Note that while for fixed αs=π2 the dynamics for αn=π2 (see above) and αn=0 both take place on neutrally stable cylindrical surfaces, these surfaces have essentially different shapes [compare Figs. 3(d) and 4(b)]: Invariant lines that consist of neutral saddles are no longer straight lines (as I1 and I2 for αn=π2) but they deform when changing the parameter A.

The theory of bifurcations without parameters39,40 can also be used to study the dynamics of the current case. We can consider two-dimensional dynamics on surfaces H(1,0)=C using C as a parameter. This allows us to analyze conservative–dissipative dynamics on an individual level set and then extrapolate to the entire phase space T3 as C parameterizes the level sets and attractors and repellers are located on the common boundary of all level sets ψ1=ψ3=0. Figure 11 shows schematic phase portraits on surfaces H(1,0)=C for different A and C. The conservative region on the cylindric surface is the biggest for C=±1 for any A=1 and it shrinks with decreasing |C| to the heteroclinic cycle [transition from (a) to (c) in Fig. 11]. Figure 11(d) corresponds to the limit case C=0 with degenerate (green) line I0 of saddles in the entire phase space T3. There are two conservative regions close to the straight lines (0,ψ2,π) and (π,ψ2,0) and also one common dissipative region in T3.

For A=1, the three-dimensional conservative region corresponds to the whole phase space T3. Finally, the conservative region is destroyed in a saddle-connection bifurcation when the connections of stable and unstable invariant manifolds of each saddle are simultaneously broken. This happens either when one of the equalities αs=±π2, αn=0 is violated or when the parameter A leaves the interval [A,1/A]: If A reaches the bifurcation value A or 1/A [Fig. 8(c)], the conservative region collapses onto the one-dimensional heteroclinic cycle (between two saddles that belong to ψ1=0 or ψ3=0).

Here, we considered the properties of phase oscillator networks that consist of M populations of N phase oscillators. While we analyzed networks for M=N=2 with sinusoidal coupling in details, some of the observations hold also for larger networks with more general coupling. We, therefore, briefly discuss a few general properties of (1). First, system (1) has dihedral symmetry DMN for any M and N. Second, (1) is a gradient system for any odd coupling functions gs, gn for all M and N. For example, if gs(ϕ)=gn(ϕ)=g(ϕ) satisfies g(ϕ)=g(ϕ) then the system has the potential

where h(ϕ) is an even function such that h(ϕ)=g(ϕ). The gradient system (23) is a special case. Third, in the case of an even coupling function gs(ϕ)=gn(ϕ)=g(ϕ) with g(ϕ)=g(ϕ), system (1) is divergence-free (this generalizes results in Ref. 33 for Ks=Kn). Other properties will be discussed elsewhere.

Our results also shed light on the phase space structure in higher dimensions, for example, M=3 populations of N=2 oscillators. In this case, the system has a continuous set of neutral chimera solutions for coupling gs(x)=gn(x)=±cos(x) (or αs=αn=±π2). These solutions can be periodic, quasi-periodic or chaotic (similar to ABC flows,51,52 chaos that fills an entire torus). Figure 12 shows an example of such chaotic dynamics for M=3 populations of N=2 oscillators. This suggests that the situation with (N2)-parametric neutral chimeras also occurs in the case of arbitrary M, N for αs=αn=±π2.

FIG. 12.

Example of a neutral chaotic weak chimera in the six oscillator system for M=3, N=2 in the conservative case αs=αn=π2, A=12. (a) Time series of phase difference between two oscillators of the first group. (b) and (c) Projections of trajectories from T5 into phase planes of the phase differences. Phase differences are locked if two oscillators belong to the same group, and such differences are phase-unlocked if the oscillators belong to different groups (as θ1,2θ2,1).

FIG. 12.

Example of a neutral chaotic weak chimera in the six oscillator system for M=3, N=2 in the conservative case αs=αn=π2, A=12. (a) Time series of phase difference between two oscillators of the first group. (b) and (c) Projections of trajectories from T5 into phase planes of the phase differences. Phase differences are locked if two oscillators belong to the same group, and such differences are phase-unlocked if the oscillators belong to different groups (as θ1,2θ2,1).

Close modal

The system in (1) with M=2 populations and N>3 oscillators has been the subject of a number of studies17 with small and large finite oscillator systems,6,13,53 including the continuum limit9,13,29,54 of infinitely many oscillators, some based on low-dimensional descriptions obtained via the Watanabe–Strogatz or Ott–Antonsen reductions.34,55,56 For future work, it would be interesting to investigate how our findings extend to the dynamic behavior of such larger system sizes, particularly regarding invariant subspaces and how they are affected by the symmetry-breaking mechanism leading to weak chimera states. Finally, our results are, of course, also relevant for more general oscillator systems that allow for amplitude variation. If oscillators are weakly coupled so that the system possesses an attracting torus such that (1) provides an adequate phase reduction, then one would expect some persistence of the bifurcation scenarios. As nonchaotic weak chimeras have been found in real-world experiments with electrochemical oscillators,12 it would be interesting to see the bifurcations outlined here in such systems.

Previous studies found that bifurcation curves leading to chimera states are organized around points in parameter space where |α|=π2;8,9,12 here, we explained the symmetry-breaking bifurcations generating weak chimeras near these points. It is interesting to note that experiments using mechanical oscillators10,11,17 indicate that chimera states emergence in a scenario near resonance—this conjecture is confirmed as one can rigorously show11 that resonance in such systems is related to phase lags being αs,αn=±π2.

O.B. acknowledges support of the National Research Foundation of Ukraine (Project No. 2020.02/0089).

The authors have no conflicts to disclose.

Oleksandr Burylko: Conceptualization (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal). Erik A. Martens: Conceptualization (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal). Christian Bick: Conceptualization (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article.

The proofs of Propositions 1–6 are very similar and rely on direct computation with trigonometric identities. We need to show that the respective constant of motion H does not change over time, that is, dHdt=0 along solutions. For completeness, we include the proof of Proposition 1.

Proof of Proposition 1

Proof
Proof of Proposition 1
We need to show that for A=1 and any αs,αn the quantity
(A1)
is constant along trajectories.
Recall that for A=1, we have Kn=0 and Ks=1 so the equations of motion (12) reduce to
(A2a)
(A2b)
(A2c)
Taking the derivative of (A1), substituting the equations of motion, and using trigonometric identities yield
which proves the assertion.
1.
S. H.
Strogatz
,
Sync: The Emerging Science of Spontaneous Order
(
Penguin
,
2004
).
2.
A.
Pikovsky
,
M.
Rosenblum
, and
J.
Kurths
,
Synchronization: A Universal Concept in Nonlinear Sciences
(
Cambridge University Press
,
2003
).
3.
Y.
Kuramoto
, Chemical Oscillations, Waves, and Turbulence, Springer Series in Synergetics Vol. 19 (Springer, Berlin, 1984).
4.
J.
Acebrón
,
L.
Bonilla
,
C.
Pérez Vicente
,
F.
Ritort
, and
R.
Spigler
, “
The Kuramoto model: A simple paradigm for synchronization phenomena
,”
Rev. Mod. Phys.
77
(
1
),
137
185
(
2005
).
5.
Y.
Kuramoto
and
D.
Battogtokh
, “
Coexistence of coherence and incoherence in nonlocally coupled phase oscillators
,”
Nonlinear Phenom. Complex Syst.
4
,
380
385
(
2002
).
6.
M. J.
Panaggio
,
D. M.
Abrams
,
P.
Ashwin
, and
C. R.
Laing
, “
Chimera states in networks of phase oscillators: The case of two small populations
,”
Phys. Rev. E
93
(
1
),
012218
(
2016
).
7.
O. E.
Omel’chenko
, “
The mathematics behind chimera states
,”
Nonlinearity
31
(
5
),
R121
R164
(
2018
).
8.
P.
Ashwin
and
O.
Burylko
, “
Weak chimeras in minimal networks of coupled phase oscillators
,”
Chaos
25
,
013106
(
2015
).
9.
E. A.
Martens
,
C.
Bick
, and
M. J.
Panaggio
, “
Chimera states in two populations with heterogeneous phase-lag
,”
Chaos
26
(
9
),
094819
(
2016
).
10.
E. A.
Martens
,
S.
Thutupalli
,
A.
Fourrière
, and
O.
Hallatschek
, “
Chimera states in mechanical oscillator networks
,”
Proc. Natl. Acad. Sci. U.S.A.
110
(
26
),
10563
10567
(
2013
).
11.
E. A.
Martens
,
M.
Panaggio
, and
S.
Thutupalli
, “On the origin of mechanical chimera states” (unpublished).
12.
C.
Bick
,
M.
Sebek
, and
I. Z.
Kiss
, “
Robust weak chimeras in oscillator networks with delayed linear and quadratic interactions
,”
Phys. Rev. Lett.
119
(
16
),
168301
(
2017
).
13.
C.
Bick
,
M. J.
Panaggio
, and
E. A.
Martens
, “
Chaos in Kuramoto oscillator networks
,”
Chaos
28
(
7
),
071102
(
2018
).
14.
L. M.
Pecora
and
T. L.
Carroll
, “
Master stability functions for synchronized coupled systems
,”
Phys. Rev. Lett.
80
(
10
),
2109
2112
(
1998
).
15.
C.
Bick
,
M.
Timme
,
D.
Paulikat
,
D.
Rathlev
, and
P.
Ashwin
, “
Chaos in symmetric phase oscillator networks
,”
Phys. Rev. Lett.
107
(
24
),
244101
(
2011
).
16.
A.
Pikovsky
and
M.
Rosenblum
, “
Dynamics of globally coupled oscillators: Progress and perspectives
,”
Chaos
25
(
9
),
097616
(
2015
).
17.
M. J.
Panaggio
and
D. M.
Abrams
, “
Chimera states: Coexistence of coherence and incoherence in networks of coupled oscillators
,”
Nonlinearity
28
(
3
),
R67
R87
(
2015
).
18.
E.
Schöll
, “
Synchronization patterns and chimera states in complex networks: Interplay of topology and dynamics
,”
Eur. Phys. J. Spec. Top.
225
(
6
),
891
919
(
2016
).
19.
F.
Guyard
and
R.
Lauterbach
, “Forced symmetry breaking: Theory and applications,” in Pattern Formation in Continuous and Coupled Systems. The IMA Volumes in Mathematics and its Applications, edited by M. Golubitsky, D. Luss, and S. H. Strogatz (Springer, New York, 1999), Chap. 10, pp. 121–135.
20.
G. C.
Sethia
and
A.
Sen
, “
Chimera states: The existence criteria revisited
,”
Phys. Rev. Lett.
112
(
14
),
144101
(
2014
).
21.
J. D.
Hart
,
K.
Bansal
,
T. E.
Murphy
, and
R.
Roy
, “
Experimental observation of chimera and cluster states in a minimal globally coupled network
,”
Chaos
26
(
9
),
094801
(
2016
).
22.
S. W.
Haugland
and
K.
Krischer
, “
Connecting minimal chimeras and fully asymmetric chaotic attractors through equivariant pitchfork bifurcations
,”
Phys. Rev. E
103
,
L060201
(
2021
).
23.
Y. L.
Maistrenko
,
S.
Brezetsky
,
P.
Jaros
,
R.
Levchenko
, and
T.
Kapitaniak
, “
Smallest chimera states
,”
Phys. Rev. E
95
(
1
),
010203
(
2017
).
24.
C.
Bick
, “
Isotropy of angular frequencies and weak chimeras with broken symmetry
,”
J. Nonlinear Sci.
27
(
2
),
605
626
(
2017
).
25.
F. P.
Kemeth
,
S. W.
Haugland
, and
K.
Krischer
, “
Symmetries of chimera states
,”
Phys. Rev. Lett.
120
(
21
),
214101
(
2018
).
26.
Note that, by definition, the perturbed system retains the wreath product symmetry; thus, the type of symmetry breaking is different from adding heterogeneity [
P.
Ashwin
,
O.
Burylko
,
Y. L.
Maistrenko
, and
O. V.
Popovych
Extreme sensitivity to detuning for globally coupled phase oscillators,
Phys. Rev. Lett.
96
,
054102
(
2006
)], which, generically, does not preserve any symmetries.
27.
B.
Dionne
,
M.
Golubitsky
, and
I.
Stewart
, “
Coupled cells with internal symmetry: I. Wreath products
,”
Nonlinearity
9
(
2
),
559
574
(
1996
).
28.
H.
Hong
and
S. H.
Strogatz
, “
Kuramoto model of coupled oscillators with positive and negative coupling parameters: An example of conformist and contrarian oscillators
,”
Phys. Rev. Lett.
106
(
5
),
054102
(
2011
).
29.
D. M.
Abrams
,
R. E.
Mirollo
,
S. H.
Strogatz
, and
D. A.
Wiley
, “
Solvable model for chimera states of coupled oscillators
,”
Phys. Rev. Lett.
101
(
8
),
084103
(
2008
).
30.
M.
Golubitsky
and
I.
Stewart
, The Symmetry Perspective, Progress in Mathematics Vol. 200 (Birkhäuser Verlag, Basel, 2002).
31.
M. J.
Field
, Dynamics and Symmetry, ICP Advanced Texts in Mathematics Vol. 3 (Imperial College Press, 2007).
32.
P.
Ashwin
and
J. W.
Swift
, “
The dynamics of n weakly coupled identical oscillators
,”
J. Nonlinear Sci.
2
(
1
),
69
108
(
1992
).
33.
P.
Ashwin
,
C.
Bick
, and
O.
Burylko
, “
Identical phase oscillator networks: Bifurcations, symmetry and reversibility for generalized coupling
,”
Front. Appl. Math. Statist.
2
(
7
),
1
16
(
2016
).
34.
S.
Watanabe
and
S. H.
Strogatz
, “
Constants of motion for superconducting Josephson arrays
,”
Physica D
74
(
3–4
),
197
253
(
1994
).
35.
C.
Bick
and
P.
Ashwin
, “
Chaotic weak chimeras and their persistence in coupled populations of phase oscillators
,”
Nonlinearity
29
(
5
),
1468
1486
(
2016
).
36.
F. P.
Kemeth
,
S. W.
Haugland
,
L.
Schmidt
,
I. G.
Kevrekidis
, and
K.
Krischer
, “
A classification scheme for chimera states
,”
Chaos
26
,
094815
(
2016
).
37.
O.
Burylko
and
A.
Pikovsky
, “
Desynchronization transitions in nonlinearly coupled phase oscillators
,”
Physica D
240
(
17
),
1352
1361
(
2011
).
38.
Note that the dynamics αs=αn=π are identical up to a sign change.
39.
S.
Liebscher
, “
Dynamics near manifolds of equilibria of codimension one and bifurcation without parameters
,”
Electron. J. Differ. Equ.
2011
(
63
),
1
12
(
2011
).
40.
B.
Fiedler
,
S.
Liebscher
, and
J. C.
Alexander
, “
Generic Hopf bifurcation from lines of equilibria without parameters: I. Theory
,”
J. Differ. Equ.
167
,
16
35
(
2000
).
41.
The bifurcation diagram was obtained via quasi-continuation, i.e., the dynamical equations were numerically integrated over T=10000 time units for each parameter value, whereas end solutions for each parameter value were used as initial condition for the next parameter value. The initial condition (ψ1,ψ2,ψ3)=(4.7204,1.6028,3.1790) was used as starting point at αs=1.6 for quasi-continuation with endvalues αs=1.56 and αs=1.6415 in backward and forward directions, respectively.
42.
P.
Chossat
and
M.
Golubitsky
, “
Symmetry-increasing bifurcation of chaotic attractors
,”
Physica D
32
(
3
),
423
436
(
1988
).
43.
V.
Lukyanov
and
L.
Shilnikov
, “
On some bifurcations of dynamical systems with homoclinic structures
,”
Sov. Math. Dokl.
19
(
6
),
1314
1318
(
1978
).
44.
S. V.
Gonchenko
,
L. P.
Shil'nikov
, and
D. V.
Turaev
, “
Quasiattractors and homoclinic tangencies
,”
Comput. Math. Appl.
34
(
2–4
),
195
227
(
1997
).
45.
C.
Bonatti
,
L. J.
Díaz
, and
M.
Viana
, Dynamics Beyond Uniform Hyperbolicity, Encyclopaedia of Mathematical Sciences (Mathematical Physics) Vol. 102 (Springer Verlag, 2004).
46.
A.
Delshams
,
M.
Gonchenko
, and
S.
Gonchenko
, “
On dynamics and bifurcations of area-preserving maps with homoclinic tangencies
,”
Nonlinearity
28
(
9
),
3027
3071
(
2015
).
47.
A.
Politi
,
G. L.
Oppo
, and
R.
Badii
, “
Coexistence of conservative and dissipative behavior in reversible dynamical systems
,”
Phys. Rev. A
33
,
4055
4060
(
1986
).
48.
K. Y.
Tsang
,
R. E.
Mirollo
,
S. H.
Strogatz
, and
K.
Wiesenfeld
, “
Dynamics of a globally coupled oscillator array
,”
Physica D
48
(
1
),
102
112
(
1991
).
49.
D.
Topaj
and
A.
Pikovsky
, “
Reversibility vs. synchronization in oscillator lattices
,”
Physica D
170
(
2
),
118
130
(
2002
).
50.
O.
Burylko
,
A.
Mielke
,
M.
Wolfrum
, and
S.
Yanchuk
, “
Coexistence of Hamiltonian-like and dissipative dynamics in rings of coupled phase oscillators with skew-symmetric coupling
,”
SIAM J. Appl. Dyn. Syst.
17
(
3
),
2076
2105
(
2018
).
51.
V. I.
Arnold
, Sur la topologie des écoulements stationnaires des fluides parfaits, (Springer, 1965), Chap. Vladimir I. Arnold—Collected Works, Vol. 2, pp. 15–18.
52.
T.
Dombre
,
U.
Frisch
,
J. M.
Greene
,
M.
Hénon
,
A.
Mehr
, and
A. M.
Soward
, “
Chaotic streamlines in the ABC flows
,”
J. Fluid Mech.
167
,
353
391
(
1986
).
53.
E.
Montbrió
,
J.
Kurths
, and
B.
Blasius
, “
Synchronization of two interacting populations of oscillators
,”
Phys. Rev. E
70
(
5
),
056125
(
2004
).
54.
E. A.
Martens
,
M. J.
Panaggio
, and
D. M.
Abrams
, “
Basins of attraction for chimera states
,”
New J. Phys.
18
(
2
),
022002
(
2016
).
55.
E.
Ott
and
T. M.
Antonsen
, “
Low dimensional behavior of large systems of globally coupled oscillators
,”
Chaos
18
(
3
),
037113
(
2008
).
56.
C.
Bick
,
M.
Goodfellow
,
C. R.
Laing
, and
E. A.
Martens
, “
Understanding the dynamics of biological and neural oscillator networks through exact mean-field reductions: A review
,”
J. Math. Neurosci.
10
(
1
),
9
(
2020
).