Skip to main content
Proceedings. Mathematical, Physical, and Engineering Sciences logoLink to Proceedings. Mathematical, Physical, and Engineering Sciences
. 2022 Mar 9;478(2259):20210817. doi: 10.1098/rspa.2021.0817

Collective states in a ring network of theta neurons

Oleh Omel’chenko 1,, Carlo R Laing 2
PMCID: PMC8908473  PMID: 35280327

Abstract

We consider a ring network of theta neurons with non-local homogeneous coupling. We analyse the corresponding continuum evolution equation, analytically describing all possible steady states and their stability. By considering a number of different parameter sets, we determine the typical bifurcation scenarios of the network, and put on a rigorous footing some previously observed numerical results.

Keywords: theta neurons, neural networks, bumps

1. Introduction

The collective or emergent behaviour of large networks of neurons is a topic of ongoing interest [1,2], with applications to the study of epilepsy [3,4], binocular rivalry [5], visual hallucinations [6,7] and working memory [810], among others. One type of network often considered is a ring, where neurons can be thought of as being arranged on a closed curve. This is natural if some property of a neuron is correlated with an angular variable such as heading in a head directional network [11], the direction in the plane to a visual stimulus that is to be remembered [12] or the orientation of a neuron’s receptive field [13].

A variety of model neurons have been considered when studying ring networks including leaky integrate-and-fire [9,10], quadratic integrate-and-fire (QIF) [1417], more realistic conductance-based models [5,18,19] and theta neurons [20,21]. Note that under a simple transformation, the quadratic integrate-and-fire neuron with infinite threshold and reset is exactly equivalent to a theta neuron. The theta neuron is the normal form for a saddle-node-on-an-invariant-circle (SNIC) bifurcation [22,23], and we will consider networks of theta neurons because they have remarkable mathematical properties, which allow us to perform many calculations exactly. Our results should be broadly applicable to ring networks of other Type-I neurons, i.e. neurons that start firing through a SNIC bifurcation.

Synaptic coupling between neurons is typically of longer range than just nearest-neighbour, so ring networks of neurons often have non-local coupling. This coupling is normally homogeneous, i.e. the strength of connections between two points on the ring depends on only the distance between these points [6].

There are several types of solutions of interest for a ring network of neurons. One is a ‘bump’ state, in which a spatially localized group of neurons is active, with all other neurons in the domain quiescent [10,24,25]. Because of the homogeneous coupling, such bumps can be positioned anywhere on the domain, and thus their position encodes a single angular variable. Such networks are often bistable, with the spatially uniform ‘all-off’ state being another attractor. It is this bistability that is of interest when modelling working memory [810,26].

In this paper, we consider a system of N synaptically coupled theta neurons as presented in [20]

dθjdt=1cosθj+(1+cosθj)(ηj+κIj),j=1,,N. 1.1

The excitability parameters ηj are chosen from a Lorentzian distribution with mean η0 and width γ0

g(η)=γπ1(ηη0)2+γ2.

Note that in the limit γ0 the distribution g(η) becomes a delta-distribution, therefore in this case we obtain system (1.1) for identical neurons. Each neuron receives a current input

Ij(t)=2πNk=1NKjkPn(θk(t)),

which is obtained as a weighted sum of pulses

Pn(θ)=an(1cosθ)n.

The positive integer n controls the width of the pulse and the constant

an=n!(2n1)!!=123n135(2n1)

is chosen according to the normalization condition

02πPn(θ)dθ=2π.

The pulse function Pn(θ) models the output pulse created when a neuron fires, i.e. θ increases through π. Note that it is possible to take the limit n, giving delta function coupling [27]. We only consider the case n=2 below, and based on our experience and the work of others [27] we expect that varying n will only move bifurcations in parameter space, not introduce new dynamics. The parameter κ is the overall coupling strength and the weights Kjk are defined according to the rule

Kjk=K(2π(jk)/N)

with some even non-constant 2π-periodic function K(x). Following [20], below we use a cosine coupling kernel

K(x)=12π(1+Acosx) 1.2

with AR such that A0. (Note that for A=0 equations (1.1) describe a population of mean-field (globally) coupled theta neurons as presented in [23].) For A>1, this form of connectivity is often referred to as ‘Mexican hat’, being positive for small x and negative for larger x, while for A<1 it is of inverted Mexican hat type, being negative for small x and positive for large x. While neurons are either excitatory or inhibitory, connectivity functions of mixed sign are often used to mimic the combined effects of excitatory and inhibitory populations [25].

From an application point of view, for every solution of equation (1.1) it is important to clarify which neurons are quiescent and which are firing. Mathematically, this question is equivalent to calculating the firing rate for neuron k

fk=12πdθkdtT, 1.3

where T denotes the time average. For identical neurons (i.e. γ=0) if all fk equal zero, then the corresponding solution of equation (1.1) is called the uniform rest state, see figure 1a. By contrast, if all fk are equal but positive, then the corresponding solution of equation (1.1) is called the uniform spiking state, see figure 1b. Note that the coincidence of all firing rates fk does not imply automatically the correlation of the spiking events in neighbouring neurons. In fact, the spikings can occur randomly or following a certain order such that the identical values of fk are obtained only after averaging over time.

Figure 1.

Figure 1.

(a) Uniform rest state and (b) uniform spiking state in equation (1.1). The top panels show snapshots of θk, the bottom panels show firing rates fk. Parameters: N=1024, γ=0, n=2, κ=1, A=3 and η0=0.2 (for both states). (Online version in colour.)

System (1.1) can also support spatially heterogeneous states, including modulated rest states, see figure 2a, and bump states, see figure 2b. Moreover, in system (1.1), one can also observe more complex non-stationary patterns, see figure 3, where more than one group of neurons may be active and their spiking events are modulated not only in space but also in time. Note that the patterns presented in figure 3 occur for parameter values for which all simpler collective states in system (1.1) are unstable. While interesting in their own right, in this paper, we study the existence and stability of only stationary patterns. The structure of the rest of the paper is as follows. In §2 we present the equations governing the continuum limit of (1.1), in §3 we discuss invariant sets of the continuum equations and in §4 we derive self-consistency equations whose solutions describe all stationary states. In §5 we show how to calculate the stability of any steady state and then in §§6 and 7, we use these results to analyse spatially uniform and non-uniform states, respectively. We conclude with a discussion in §8.

Figure 2.

Figure 2.

(a) Modulated rest state for A=3 and η0=0.33, and (b) bump state for A=5 and η0=2 in equation (1.1). The top panels show snapshots of θk, the bottom panels show firing rates fk. Other parameters: N=1024, γ=0, n=2 and κ=1. (Online version in colour.)

Figure 3.

Figure 3.

Non-stationary states in equation (1.1) for η0=2 (ac), for η0=0.9 (df) and for η0=0.3 (gi). The panels (a), (d), (g) show snapshots of θk, the panels (b), (e), (h) show firing rates fk averaged over 100 time units, and the panels (c), (f), (i) show space–time plots of θk. Other parameters: N=1024, γ=0, n=2, κ=1 and A=5.

2. Continuum limit

Using well-established techniques, namely the Ott/Antonsen ansatz [28,29], it was shown in [20] that in the continuum limit (N), the long-term coarse-grained dynamics of system (1.1) can be described by a mean-field equation

zt=(iη0γ)(1+z)2i(1z)22+κi(1+z)22KHn(z), 2.1

for a local order parameter z(x,t), which represents the value of eiθj(t) averaged over all indices j such that 2πj/Nx[0,2π]. Note that the symbol K in equation (2.1) denotes an integral operator

(Kφ)(x)=02πK(xy)φ(y)dy

and

Hn(z)=an[C0+q=1nCq(zq+z¯q)]

is a nonlinear function with

Cq=k=0nm=0kδk2m,q(1)kn!2k(nk)!m!(km)!.

While the complex quantity z may not seem to have an obvious biological interpretation, one can use the equivalence of theta and QIF neurons to determine relevant quantities. Defining W(1z¯)/(1+z¯) one can show [21,30] that the instantaneous firing rate of neurons at position x and time t is the flux through θ=π:

f(x,t)=1πReW=1|z|2π|1+z|2. 2.2

If z(x,t) does not depend on t, then formula (2.2) yields a continuum limit analogue of the averaged firing rates defined by equation (1.3). Similarly, if Vj=tan(θj/2) is the voltage of the jth QIF neuron, the mean voltage at position x and time t is given by ImW.

Remark 2.1. —

It follows from the above definition that Hn(1)=0 for every nN. Indeed, using the substitution m=kl, we obtain

Cq=k=0nm=0kδk2m,q(1)kn!2k(nk)!m!(km)!=k=0nl=0kδk+2l,q(1)kn!2k(nk)!(kl)!l!=Cq.

Therefore,

Hn(1)an=q=nnCq=q=nnk=0nm=0kδk2m,q(1)k2k(nk)(km)=0,

where we used the identity

q=nnδk2m,q=1valid for0mkn

and the two other identities

m=0k(km)=2kandk=0n(1)k(nk)=0

that follow from the properties of the binomial coefficients.

Remark 2.2. —

Note that the function Hn(z) is real even for a complex argument z. To stress this fact, we can rewrite this function in the form

Hn(z)=anC0+2ReDn(z)whereDn(z)=anq=1nCqzq.

Remark 2.3. —

Recall that for every continuous kernel K(x) the above operator K is a bounded compact operator on the space of continuous functions C([0,2π];C). Moreover, in the case of the cosine kernel (1.2), for every φ(x)C([0,2π];C), we have

(Kφ)(x)=φ+Acosxφcosy+Asinxφsiny,

where

φ=12π02πφ(y)dy.

Furthermore, if φ(x) is an even function (i.e. φ(x)=φ(x)), then

(Kφ)(x)=φ+Acosxφcosy. 2.3

Equation (2.1) can be rewritten in the form

zt=(iJ(x,t)γ)(1+z)2i(1z)22, 2.4

where J(x,t)=η0+κKHn(z(x,t)). Then, for every point x0[0,2π] it can be considered as an ODE with respect to z(x0,t) with a driving force J(x0,t). This fact and the special structure of equation (2.4) have important consequences for the dynamics of solutions to equation (2.1), which we discuss in the next section.

3. Invariant sets of equation (2.1)

Let us consider an ordinary differential equation

dudt=(iJ(t)γ)(1+u)2i(1u)22 3.1

with a complex-valued unknown function u(t). In the following, we suppose that J(t) is a real coefficient and γ0. Moreover, we denote by D={uC:|u|<1} the open unit disc in the complex plane, by D¯={uC:|u|1} the closure of D, and by S={uC:|u|=1} the boundary of D.

Proposition 3.1. —

The closed unit disc D¯ is an invariant set of equation (3.1). In other words, if u(0)D¯, then the corresponding solution u(t) of equation (3.1) lies in D¯ for all t>0.

Moreover, if γ>0, then the ω-limit set of any initial condition u(0)D¯ lies in the open unit disc D.

Furthermore, if γ=0, then the unit circle S and the open unit disc D are distinct invariant sets of equation (3.1).

Proof. —

The complex conjugate of equation (3.1) reads

du¯dt=(iJ(t)γ)(1+u¯)2+i(1u¯)22.

Then, simple calculations yield

d|u|2dt=udu¯dt+u¯dudt=(J(t)1)(1|u|2)Imuγ(2|u|2+(1+|u|2)Reu).

For |u|=1, this equation implies

d|u|2dt=2γ(1+Reu)0. 3.2

Hence, if |u(0)|1, then |u(t)| cannot grow above one, and therefore |u(t)|1 for all t>0.

On the other hand, if γ>0, then for every uS except u=1, the differential inequality (3.2) becomes a strict inequality

d|u|2dt<0.

Moreover, at u=1, the right-hand side of equation (3.1) equals 2i, and therefore u=1 is not an equilibrium of equation (3.1). From all these facts, it follows that any initial condition u(0)S is pushed inside the disc D for t>0. Moreover, if u(0)D, then u(t) also remains inside D for all t>0.

The last assertion of the proposition follows from a simple observation that

d|u|2dt=0,

for all |u|=1 and γ=0.

4. Self-consistency equation

We will start with an auxiliary proposition.

Proposition 4.1. —

For any cR, the equation

dudt=(icγ)(1+u)2i(1u)22 4.1

has exactly one stable equilibrium in the closed unit disc D¯. More precisely, if γ=0, then the equilibrium is neutrally stable, but if γ>0 then it is linearly stable.

Proof. —

All equilibria of equation (4.1) satisfy a quadratic equation

(1u1+u)2=c+iγ,

which can be rewritten in the equivalent form

1u1+u=ξ, 4.2

where ξ denotes one of the two distinct values of the square root of c+iγ. The rational function in the left-hand side of equation (4.2) is a particular Möbius transformation, which maps the closed unit circle D¯ on to the closed half-plane {uC:Reu>0}. Therefore, to get a solution uD¯ from equation (4.2), we need to choose ξ=c+iγ with Reξ0. In this case, equation (4.2) yields

u=1ξ1+ξ. 4.3

Let us linearize equation (4.1) around the above equilibrium u. In this case, the dynamics of small perturbations v obeys a linear equation

dvdt=[(icγ)(1+u)+i(1u)]v.

Therefore, the linear stability of equilibrium u requires

Re[(icγ)(1+u)+i(1u)]0.

Using (4.2) and (4.3), we convert the expression in square brackets as follows:

(icγ)(1+u)+i(1u)=i[(c+iγ)(1+u)+(1u)]=i(ξ2+ξ)(1+u)=2iξ.

Therefore, the equilibrium u is (neutrally) stable if and only if Imξ0.

In the appendix, it is shown that for any cR and any γ0 there exists exactly one value ξ=c+iγ such that Reξ0 and Imξ0. This ends the proof.

Let Uγ(c) be a function that determines the equilibrium uD¯ of equation (4.1) mentioned in proposition 4.1. More precisely,

Uγ(c)=1c+iγ1+c+iγ, 4.4

where the square root c+iγ is calculated according to proposition A.1 in the appendix.

Remark 4.2. —

In the degenerate case γ=0, function Uγ(c) simplifies as follows:

U0(c)={1c1+cforc0,1ic1+icforc<0.

Then, if we calculate the firing rates (2.2) corresponding to z=U0(c), we obtain f(x,t)=c/π for c0, and f(x,t)=0 for c<0.

Suppose that a(x) is an equilibrium of equation (2.1). Then a(x) is also an equilibrium of equation (2.4) with J(x,t)=w(x):=η0+κKHn(a). Using proposition 4.1, we find a(x)=Uγ(w(x)). Therefore, the last two formulae are consistent with each other iff w(x) satisfies the integral equation

w(x)=η0+κKHn(Uγ(w)). 4.5

The self-consistency equation (4.5) becomes particularly simple in the case of the cosine kernel K(x). Indeed, if in this case, we look for an even solution of equation (4.5), then due to remark 2.3 we have

w(x)=w^0+w^1cosx, 4.6

with some coefficients w^0,w^1R. Inserting expressions (2.3) and (4.6) into equation (4.5) and equating the x-independent terms and the terms proportional to cosx separately, we obtain a nonlinear two-dimensional system

w^0=η0+κHn(Uγ(w^0+w^1cosy)) 4.7

and

w^1=κAHn(Uγ(w^0+w^1cosy))cosy, 4.8

which can be solved approximately with a standard Newton’s method. If we find a pair (w^0,w^1)TR2 that satisfies (4.7) and (4.8), then formula (4.6) yields the corresponding solution of equation (4.5), while the formula a(x)=Uγ(w(x)) yields the corresponding equilibrium of equation (2.1). Note that similar self-consistency equations have been found in similar systems [3133] but those authors were studying chimera states.

Having seen how to find equilibria of equation (2.1) we now move on to determine their stability.

5. Stability of equilibria of equation (2.1)

Suppose that z=a(x) is a stationary solution of equation (2.1). Then, inserting the ansatz z=a(x)+v(x,t) into equation (2.1) and linearizing the resulting relation with respect to small perturbations v(x,t), we obtain a differential equation determining the linear stability of a(x)

vt=μ(x)v+κi(1+a(x))22K(Dn(a)v+Dn(a)¯v¯), 5.1

where

μ(x)=[i(η0+κKHn(a))γ](1+a(x))+i(1a(x))

and

Dn(z)=ddzDn(z)=anq=1nqCqzq1.

To investigate the decay of different spatial modes, we insert the ansatz

v(x,t)=v+(x)eλt+v¯(x)eλ¯t

into equation (5.1) and equate separately the terms at eλt and eλ¯t. Thus, we obtain a spectral problem

λ(v+v)=(μ(x)00μ(x)¯)(v+v)+κi2((1+a)2(1+a)2(1+a¯)2(1+a¯)2)(KDn(a)v+KDn(a)¯v). 5.2

Note that the right-hand side of equation (5.2) is a sum of two bounded operators on C([0,2π];C2). Moreover, the first term of this sum is a multiplication operator, whereas the second term is a compact operator (see remark 2.3). Operators with similar structure have been often considered in the context of studying the chimera states [34]. Using the results of these works, we can draw the following conclusions:

  • (i)

    The spectrum determined by equation (5.2) lies in a bounded region of the complex plane.

  • (ii)

    The spectrum consists of two parts: essential σess and discrete σdiscr spectra.

  • (iii)
    The essential spectrum σess is determined by the multiplication operator in the right-hand side of (5.2). It can be computed explicitly by
    σess={μ(x):x[0,2π]}{c.c.}.
  • (iv)

    The discrete spectrum σdiscr consists of a finite number of eigenvalues with finite multiplicity.

Proposition 5.1. —

Suppose that a(x)=Uγ(w(x)), where w(x) is a solution of the self-consistency equation (4.5). Then the essential spectrum corresponding to a(x) is given by

σess={2iw(x)+iγ:x[0,2π]}{c.c.}.

Proof. —

Replacing a(x) by Uγ(w(x)) in the definition of μ(x) and using equation (4.5), we obtain

μ(x)=i(w(x)+iγ)(1+Uγ(w(x)))+i(1Uγ(w(x))).

Then, formula (4.4) yields μ(x)=2iw(x)+iγ.

Remark 5.2. —

Recall that the branch of the complex square root in (4.4) is chosen so that Im(w(x)+iγ)0. Therefore, Re(2iw(x)+iγ)0 and hence the essential spectrum σess in proposition 5.1 is always neutrally stable. Moreover, for γ>0 the value of w(x)+iγ cannot be real, therefore in this case, the essential spectrum σess lies entirely in the open half-plane {λC:Reλ<0}.

The spectral problem (5.2), in general, is infinite-dimensional. However, in the case of cosine kernel (1.2), it can be reduced to a finite-dimensional form. Indeed, remark 2.3 implies that

(KDn(a)v+KDn(a)¯v)=k=13Vkϕk(x), 5.3

where VkC2, and ϕ1(x)=1, ϕ2(x)=cosx, ϕ3(x)=sinx are three linearly independent functions that span the range of the operator K. Inserting (5.3) into equation (5.2) and expressing (v+,v)T from the resulting equation, we obtain

(Dn(a)v+Dn(a)¯v)=k=13L(x,λ)Vkϕk(x), 5.4

where

L(x,λ)=(Dn(a)00Dn(a)¯)(λμ(x)00λμ(x)¯)1κi2((1+a)2(1+a)2(1+a¯)2(1+a¯)2)

is a (2×2)-matrix. Now, using remark 2.3 and the definition of functions ϕk(x) given after (5.3), we conclude

(KDn(a)v+KDn(a)¯v)=k=13(L(x,λ)ϕ1(x)ϕk(x)ϕ1(x)+j=23AL(x,λ)ϕj(x)ϕk(x)ϕj(x))Vk.

Because of the linear independence of ϕj(x), this formula agrees with (5.3) if and only if

Vj=k=13Bjk(λ)Vk, 5.5

where

B1k(λ)=L(x,λ)ϕ1(x)ϕk(x)

and

Bjk(λ)=AL(x,λ)ϕj(x)ϕk(x)forj=2,3.

(Note that in the above expressions the averaging over x is performed for each matrix element separately.) Finally, we see that system (5.5) has a non-zero solution (V1,V2,V3)T if and only if

det(I6B(λ))=0, 5.6

where

B(λ)=(B11(λ)B12(λ)B13(λ)B21(λ)B22(λ)B23(λ)B31(λ)B32(λ)B33(λ))

and I6 is the (6×6) unit matrix. Equation (5.6) is a characteristic equation that determines the discrete spectrum σdiscr associated with any equilibrium of equation (2.1) in the case of cosine coupling kernel (1.2).

In the next section, we consider the existence of spatially uniform steady states of equation (2.1) and use the results in this section to determine their stability.

6. Spatially uniform states

(a) . Existence of uniform states

It is easy to verify that for any cosine kernel (1.2) (in fact, even for any kernel that satisfies 02πK(x)dx=2π) equation (4.5) has a set of spatially uniform solutions

w(x)=pR,

where the constant p satisfies p=η0+κHn(Uγ(p)), or equivalently

η0=pκHn(Uγ(p)). 6.1

By construction, formula (6.1) yields a parametric representation of all uniform states of equation (2.1) with stable essential spectra. Indeed, for every (p,κ)R2, we can find the corresponding excitability parameter η0 using equation (6.1) and the corresponding equilibrium of equation (2.1) by using the formula Uγ(p). Importantly, it follows from remark 4.2 that for γ=0 the sign of the parameter p allows us to distinguish between the uniform spiking states (p>0) and the uniform rest states (p<0). Moreover, for small values of γ, the same criterion can be used to distinguish between the partially synchronized spiking states and the partially synchronized rest states introduced in [23].

Formula (6.1) gives a global representation of the branch of uniform states. However, in practice, it is more important to answer the other question: How many different uniform states can be found in equation (2.1) for given parameters (η0,κ)R2 and what are their types? This question can be answered geometrically, if we rewrite equation (6.1) in the form

pη0κ=Fn,γ(p), 6.2

where

Fn,γ(p)=Hn(Uγ(p))=anC0+2Re(Dn(Uγ(p))).

Note that the right-hand side of equation (6.2) does not depend on either η0 or κ. Moreover, for any fixed γ0 we have Uγ(p)1 as p±. This implies that Fn,γ(p)Hn(1) as p±, and therefore for any κR and any η0R, equation (6.2) has at least one real solution. In the next proposition, we show that under certain conditions equation (6.2) can also have three different roots.

Proposition 6.1. —

Let the following conditions be satisfied:

  • (a)

    Fn,0(p)+ for p+0,

  • (b)

    Fn,0(p)0 for p+,

  • (c)

    Fn,0(p)<0 for p>0,

  • (d)

    Fn,0(p)0 for p0 and for p,

  • (e)

    there exists pmin<0 such that Fn,0(p)<0 for p<pmin and Fn,0(p)>0 for pmin<p<0.

Then, equation (6.2) with γ=0 has the following properties:

  • (i)
    For every κ>0, there exists p0>0 such that Fn,0(p0)=1/κ. Moreover,
    • if η0>0, then equation (6.2) has one positive root,
    • if p0κFn,0(p0)<η0<0, then equation (6.2) has two distinct positive roots and one negative root,
    • if η0<p0κFn,0(p0), then equation (6.2) has one negative root.
  • (ii)

    If 1/Fn,0(pmin)<κ<0 and η0<0, then equation (6.2) has one negative root. On the other hand, if 1/Fn,0(pmin)<κ<0 and η0>0, then equation (6.2) has one positive root.

  • (iii)
    For every κ<1/Fn,0(pmin), there exist p1<p2<0 such that Fn,0(p1,2)=1/κ. Moreover,
    • if η0<p2κFn,0(p2) or p1κFn,0(p1)<η0<0, then equation (6.2) has one negative root,
    • if 0<η0<p1κFn,0(p1), then equation (6.2) has two distinct negative roots and one positive root,
    • if p2κFn,0(p2)<η0<min{0,p1κFn,0(p1)}, then equation (6.2) has three distinct negative roots,
    • if η0>max{0,p1κFn,0(p1)}, then equation (6.2) has one positive root.

Proof. —

Suppose κ>0. Then the existence of a unique value p0>0 such that Fn,0(p0)=1/κ follows from the assumptions (a)–(c). The remaining part of the assertion (i) follows from the geometric consideration of equation (6.2) explained in figure 4. Indeed, if the slope κ is fixed, then changing η0 we can encounter three qualitatively different situations, which are separated by two critical cases: the line (pη0)/κ is tangent to the graph of Fn,0(p) for p>0, or this line passes through the point (0,0).

Two other assertions (ii) and (iii) are proved by a similar geometric argument.

Figure 4.

Figure 4.

Geometric solution of equation (6.2) with γ=0 and n=2 for several values of the coupling strength κ. The thick black curve shows the graph of F2,0(p). Solid thin red lines show ‘extreme’ lines (pη0)/κ. Panels (i)–(iii) illustrate how the number and sign of the roots of equation (6.2) change for varying parameter κ. (Online version in colour.)

Remark 6.2. —

Assumptions (d) and (e) of proposition 6.1 imply that p1(κ) and p2(κ) are smooth monotonic functions for κ<1/Fn,0(pmin). Moreover,

p1(κ)κFn,0(p1(κ))p2(κ)κFn,0(p2(κ)). 6.3

Indeed, by definition we have κ=1/Fn,0(p1(κ)), therefore the Implicit Function Theorem implies that

dp1dκ=[Fn,0(p1(κ))]2Fn,0(p1(κ)).

Hence, using the chain rule, we obtain

ddκ(p1(κ)κFn,0(p1(κ)))=ddp(pFn,0(p)Fn,0(p))|p=p1(κ)dp1dκ=Fn,0(p1(κ)).

A similar formula can be also obtained for p2(κ), therefore

ddκ(p1(κ)κFn,0(p1(κ))p2(κ)+κFn,0(p2(κ)))=Fn,0(p2(κ))Fn,0(p1(κ))0,

where we used the fact that Fn,0(p) is a decreasing function for p<0 (see assumptions (d) and (e)). Since the relation (6.3) is satisfied for κ=1/Fn,0(pmin) (in this case p1(κ)=p2(κ)), it obviously will remain true for all smaller values of κ.

Remark 6.3. —

It is easy to verify that the assumptions of proposition 6.1 are satisfied for n=2. In particular,

H2(z)=1+2ReD2(z)andD2(z)=23z+16z2,

and

F2,0(p)=23Re[(U0(p)2)U0(p)]andF2,0(p)=23Re[U0(p)2+(U0(p)2)U0(p)].

The conclusion of proposition 6.1, in this case, is represented geometrically in figure 5.

Figure 5.

Figure 5.

Number of solutions of equation (6.2) with γ=0 and n=2 for different values of the parameters η0 and κ. The symbols ‘+’ and ‘’ indicate the number of solutions and their signs in different regions bounded by the solid red curves. (Online version in colour.)

(b) . Stability of uniform states

Let w(x)=p0R be a constant solution of equation (4.5). It corresponds to the equilibrium z=Uγ(p0) of equation (2.1), and therefore we can perform its stability analysis according to the approach described in §5. Using proposition 5.1, we find that the essential spectrum σess consists of two points μ0 and μ¯0 where μ0=2ip0+iγ.

To calculate the discrete spectrum σdiscr, we need to consider equation (5.6). Note that the matrix L(x,λ) of a uniform state does not depend on x. Moreover, simple calculations yield

B1k(λ)=δ1kL(,λ)

and

Bjk(λ)=AδjkL(,λ)/2forj=2,3,

where δjk is the Kronecker delta. Therefore, equation (5.6) factorizes into three independent equations

det(I2B11(λ))=0, 6.4
det(I2B22(λ))=0, 6.5
det(I2B33(λ))=0. 6.6

Note that equations (6.5) and (6.6) obviously coincide with each other and therefore determine identical eigenvalues.

Proposition 6.4. —

Equation (6.4) has two solutions

λ1,±=Re(μ0+2ζ0)±4|ζ0|2[Im(μ0+2ζ0)]2,

while equation (6.5) has two solutions

λ2,±=Re(μ0+Aζ0)±A2|ζ0|2[Im(μ0+Aζ0)]2,

where

ζ0=κi4Dn(a0)(1+a0)2witha0=Uγ(p0).

Proof. —

Let us consider equation (6.5). Using the definition of Bjk(λ), we rewrite it in the form

det[I2A(λμ000λμ¯0)1(ζ0ζ0ζ¯0ζ¯0)]=0.

For every λσess (that means λμ0 and λμ¯0), the latter equation is equivalent to

det[(λμ000λμ¯0)A(ζ0ζ0ζ¯0ζ¯0)]=(λμ0Aζ0)(λμ¯0Aζ¯0)A2|ζ0|2=λ22λRe(μ0+Aζ0)+|μ0+Aζ0|2A2|ζ0|2=0.

Solving it with respect to λ we obtain the formula for λ2,±. The other equation (6.4) can be considered by analogy.

Remark 6.5. —

Using (4.4), it is easy to demonstrate that

i4(1+a0)2=ip0+iγUγ(p0)=μ02Uγ(p0)

and

ζ0=μ02κDn(a0)Uγ(p0)=μ02κddpDn(Uγ(p))|p=p0.

Let us consider the equation

λ22λRe(μ0+2ζ0)+|μ0+2ζ0|24|ζ0|2=0

that determines λ1,± and check when it has at least one zero solution. This situation corresponds to a static bifurcation of the uniform state a0=Uγ(p0) and is equivalent to the condition

|μ0+2ζ0|24|ζ0|2=0.

Inserting here the expression of ζ0 from remark 6.5, we obtain

|μ0|2(12κRe[Dn(a0)Uγ(p0)])=0.

The case μ0=0 corresponds to the degenerate essential spectrum and therefore we discard it. Then the only possibility to satisfy the above equation is to have

12κRe[Dn(a0)Uγ(p0)]=1κFn,γ(p0)=0.

It is easy to verify that the latter condition is equivalent to the situation when the line in the left-hand side of equation (6.2) is tangent to the graph of Fn,γ(p). In other words, all static bifurcations described by equation (6.4) occur at the fold points of the graph of p versus η0 with fixed κ.

Proposition 6.6. —

In the case γ=0, the eigenvalues λ1,± determined by equation (6.4) are either both real or complex-conjugate and purely imaginary. Therefore equation (6.4) in this case can determine only static bifurcations, i.e. bifurcations characterized by real positive eigenvalues appearing from zero.

Proof. —

If p0>0, then μ0=2ip0 and the derivatives U0(p0) and Dn(U0(p0)) are both real. Hence, the value of ζ0 is purely imaginary and therefore

λ1,±=±4|ζ0|2[Im(μ0+2ζ0)]2.

On the other hand, if p0<0, then μ0=2p0 and therefore

λ1,±=μ0+Re(2ζ0)±4|ζ0|2[Im(2ζ0)]2=μ0+Re(2ζ0)±|Re(2ζ0)|.

This ends the proof.

Propositions 6.1 and 6.6 give a complete description of the uniform states in equation (2.1) with γ=0 and A=0. Indeed, if A=0, then equations (6.5) and (6.6) have no solutions, and therefore the stability of the uniform state is determined by the roots of equation (6.4) only. One can ask what happens to these states if we slightly increase the parameter γ in equation (2.1)? To answer this question, we can use the fact that the parametric representation (6.1) as well as the explicit formula of eigenvalues λ1,±, see proposition 6.4, in general smoothly depend on γ. Therefore, if max(Reλ1,,Reλ1,+)<0 or max(Reλ1,,Reλ1,+)>0, then the same inequality will remain true for small enough values of γ. However, a similar argument does not work in the case of complex-conjugate and purely imaginary eigenvalues λ1,±. From proposition 6.6, it follows that such a situation occurs if p0>0 and 4|ζ0|2[Im(μ0+2ζ0)]2<0. We consider this case in more detail below.

Remark 6.7. —

If γ=0 and p0>0, then μ0=2ip0 and ζ0 (see remark 6.5) are purely imaginary, therefore

4|ζ0|2[Im(μ0+2ζ0)]2=μ02+4μ0ζ0=μ02(12κddpDn(U0(p))|p=p0)=4p0(1κFn,0(p0)).

Hence, in this case, the inequality 4|ζ0|2[Im(μ0+2ζ0)]2<0 is equivalent to 1κFn,0(p0)>0.

Proposition 6.8. —

Let p0>0 and 1κFn,0(p0)>0. Then for sufficiently small γ>0, there exists a uniform state z=Uγ(p0) of equation (2.1) with η0=p0κFn,γ(p0). Moreover, the eigenvalues λ1,± are complex conjugate and satisfy

Reλ1,±<0,if 2>κ(Fn,0(p0)+2p0Fn,0(p0))

and

Reλ1,±>0,if 2<κ(Fn,0(p0)+2p0Fn,0(p0)).

See figure 6 for a graphical representation of this proposition.

Figure 6.

Figure 6.

Graphical representation of the assertion of proposition 6.8 for n=2. The solid red curve has a vertical asymptote η0=2/3 at the left end. (Online version in colour.)

Proof. —

The existence of the uniform state z=Uγ(p0) is obvious and we need to consider only the corresponding expression

λ1,±=Re(μ0+2ζ0)±4|ζ0|2[Im(μ0+2ζ0)]2, 6.7

taking into account that μ0 and ζ0 depend on γ. Due to the assumptions p0>0 and 1κFn,0(p0)>0 we can be sure that the square root in equation (6.7) remains purely imaginary for sufficiently small γ, therefore

Reλ1,±=Re(μ0+2ζ0).

On the other hand, using the formula μ0=2ip0+iγ and remark 6.5, we obtain

dμ0dγ|γ=0=1p0anddζ0dγ|γ=0=κ2p0ddpDn(U0(p))|p=p0+κp0d2dp2Dn(U0(p))|p=p0.

This implies that

Reλ1,±=2κ(Fn,0(p0)+2p0Fn,0(p0))2p0γ+O(γ2)

for γ0.

Proposition 6.8 has an important consequence for the stability of the uniform states of equation (2.1). Although for γ=0 such states can undergo only static bifurcations (see proposition 6.6), for γ>0 they can lose stability through a Hopf bifurcation too. In particular, for small γ, a Hopf bifurcation occurs in the vicinity of the curve described parametrically by the equations

κ=2(Fn,0(p)+2pFn,0(p))1andη0=pκFn,0(p)

with p>0 that satisfy the inequality 1κFn,0(p)>0. An example of such a curve for n=2 is shown in figure 6.

7. Spatially non-uniform states

In the previous section, we considered the constant equilibria of equation (2.1), which correspond to uniform rest states and uniform spiking states of coupled theta neurons (1.1). However, these are in general not the only possible solutions of equation (2.1). Below we show that for a cosine coupling (1.2) with A0 equation (2.1) also has a variety of non-constant equilibria representing more complex patterns in system (1.1). We focus on the case γ=0, n=2 and test three distinct values κ=1, 1 and 2 corresponding to three qualitatively different scenarios in figure 5. For each of these cases, we consider three distinct values of A, namely 0, 3 and 5.

In figure 7, for all nine combinations of κ and A values, we plot either p (for spatially uniform states) or w^0 (for non-uniform states) as a function of η0. The branch of uniform states is always independent of A, but the stability of different states on it can change for different A. If A0, then using proposition 6.4, we typically find two bifurcation points determined by the condition λ2,±=0. Then, using the self-consistency equations (4.7)–(4.8) we can compute a branch of non-constant equilibria of equation (2.1) that connects these two bifurcation points. Every solution (w^0,w^1) of equations (4.7)–(4.8) with w^10 corresponds to a non-constant even solution (4.6) of equation (4.5) and hence to a spatially modulated equilibrium a(x)=Uγ(w(x)) of equation (2.1). The stability of this equilibrium can be analysed using the results of §5. More precisely, the essential spectrum σess of the operator on the right-hand side of the linearized equation (5.1) is determined by proposition 5.1, while the discrete spectrum σdiscr of this operator can be calculated by numerically solving the characteristic equation (5.6). Note that since w(x) is an even function, the matrix L(x,λ) in the definition of B(λ) is an even function of x too. Therefore, Bjk(λ)=0 for j=1,2 and k=3 as well as for j=3 and k=1,2. Consequently, equation (5.6) factorizes into two independent equations

det[I4(B11(λ)B12(λ)B21(λ)B22(λ))]=0 7.1

and

det[I2B33(λ)]=0. 7.2

It is easy to verify that, by construction, equation (7.1) determines eigenvalues corresponding to perturbations that do not break the Z2-symmetry of the considered equilibrium, while equation (7.2) determines eigenvalues corresponding to symmetry-breaking perturbations. It turns out that all instabilities shown in figure 7 are due to equation (7.1) only. On the other hand, for every non-constant even equilibrium a(x) equation (7.2) yields a zero eigenvalue that is related to the translation invariance of equation (2.1).

Figure 7.

Figure 7.

Solution branches of equation (6.2) (for spatially uniform states) or of (4.7) and (4.8) (for non-uniform states) with γ=0 and n=2 for different pairs of the parameters κ and A. Solid and dotted curves show stable and unstable solutions. Black curves correspond to uniform states, while red curves correspond to spatially modulated states and bump states. The dots in panels (b), (f) and (h) show points corresponding to the solutions shown in figures 8, 9, 12 and 13. In panel (c), the red curve has two tiny stable parts close to the fold point on the lower branch and close to the right end of the curve. Panel (b) shows a situation analogous to that shown in fig. 9 in [21].

Figures 813 provide additional information about the solution branches in figure 7. Each row of these figures shows an equilibrium a(x) of equation (2.1), the corresponding firing rate f(x) calculated by (2.2), and the linear stability spectrum of a(x). The sampled equilibria are chosen so that they illustrate all qualitatively different spatial profiles a(x) and their linear stability properties. Note that we do not show any equilibria for figure 7i, because they look very similar to those shown in figure 12.

Figure 9.

Figure 9.

The equilibria a(x) of equation (2.1), the firing rates f(x) and the spectra of the corresponding linearized operator for the three red points in figure 7b. The purple lines and the black dots in the far right panels show essential and discrete spectra, respectively. The top panel shows an unstable bump solution, the middle shows a neutrally stable bump, while the bottom shows an unstable spatially modulated state. The solution in the middle panel is analogous to that shown in fig. 2 of [20].

Figure 8.

Figure 8.

The equilibria a(x) of equation (2.1), the firing rates f(x) and the spectra of the corresponding linearized operator for the three black points in figure 7b. The purple and the black dots in the far right panels show essential and discrete spectra, respectively. The top panels correspond to a stable ‘all-off’ state while the bottom panels correspond to a neutrally stable ‘all-on’ state.

Figure 13.

Figure 13.

The equilibria a(x) of equation (2.1), the firing rates f(x) and the spectra of the corresponding linearized operator for the four red points in figure 7h.

Figure 12.

Figure 12.

The equilibria a(x) of equation (2.1), the firing rates f(x) and the spectra of the corresponding linearized operator for the four red points in figure 7f. The results here are similar to those shown in figure 9. This is not unexpected, since for κ=1 and A<1, the effective coupling is of Mexican hat type with a positive value of κ.

Our results reveal several parameter combinations suitable for the observation of stable bump states. In particular, such states can be found for (κ,A)=(1,3), (κ,A)=(1,5) and (κ,A)=(2,5). Although in all these cases, the bump states have similar spatial profiles, their stability diagrams for positive and negative κ are a bit different. More precisely, for κ=1, the η0-interval of stable bumps is bounded by two fold points, while for κ=1 (as well as for κ=2) the stability interval of bumps is bounded by a fold point at the right end and by a Hopf bifurcation at the left end.

Apart from the bump states, equation (2.1) has only one more type of stable spatially non-uniform equilibria. These are modulated rest states shown in the first and third rows of both figures 10 and 11. Even though none of the neurons are firing here, their state is not spatially uniform.

Figure 10.

Figure 10.

The equilibria a(x) of equation (2.1), the firing rates f(x) and the spectra of the corresponding linearized operator for several points in figure 7c. Parameters: η0=0.14, 0.2, 3.32 and 3.25 (upper branch).

Figure 11.

Figure 11.

The equilibria a(x) of equation (2.1), the firing rates f(x) and the spectra of the corresponding linearized operator for several points in figure 7e. Parameters: η0=0.14, 0.2, 0.33 and 0.35 (lower branch).

Finally, we emphasize that our consideration reveals also a wide parameter range, see figure 7c,e,f,i, where all equilibria of equation (2.1) are unstable. In this case, any simple collective dynamics of neurons is not possible, therefore more complex non-stationary patterns can emerge in system (1.1), see figure 3.

8. Discussion

In summary, we analysed the dynamics of a ring network of theta neurons, homogeneously coupled with a cosine kernel. We took the continuum limit that allows us to describe the network’s asymptotic dynamics by a complex-valued integrodifferential equation (2.1). We showed how to determine the existence and stability of stationary solutions, both spatially uniform and non-uniform. We have given a number of examples demonstrating the validity of our approach, mostly concentrating on the case of identical neurons. These results put on a rigorous basis some of the numerical observations in [20,21].

We now discuss how our results relate to those of others. Byrne et al. [17] considered either one or two populations of theta neurons with synaptic coupling using reversal potentials and characteristic timescales, and exponentially decaying coupling kernels. They considered domains with either periodic or Neumann boundary conditions and looked for Turing bifurcations of the spatially uniform state. They also performed numerical bifurcation analysis of time-dependent solutions such as wavetrains and travelling fronts. Esnaola-Acebes et al. [14] also considered two ring networks of excitatory and inhibitory QIF neurons with even coupling functions involving more than just the first cosine term. They chose parameters so that there was a single spatially uniform state and performed a Turing bifurcation analysis of this state. This analysis was used to explain the decaying oscillatory transient spatial modes seen when the spatially uniform state was perturbed. They also numerically followed a bump solution and found a scenario of the form shown in figure 7e.

Byrne et al. [16] considered a network of QIF neurons on a periodic domain, with synaptic dynamics and an inverted Mexican hat connectivity. They also included gap junctional coupling. By analysing the continuum equations they found Turing–Hopf bifurcations of the spatially uniform state that result in the appearance of travelling and standing waves. Schmidt & Avitabile [15] studied ring networks of QIF neurons with Mexican hat connectivity and also determined the conditions for a Turing bifurcation.

Most of these works presented largely numerical studies of the fully nonlinear and often time-dependent solutions that such networks can support whereas we have concentrated on stationary solutions, exploiting the form of the coupling function to explicitly calculate these and analytically determine their stability.

Our approach could easily be generalized to other forms of coupling such as synaptic coupling using reversal potentials [17] or gap junctional coupling [16,35], and different coupling kernels, as long as the form of the equations allows the use of the Ott/Antonsen ansatz.

Supplementary Material

Appendix A

Proposition A.1. —

For every cR and γ0, there exists exactly one square root value ξ=c+iγ in the quadrant Reξ0, Imξ0, which is given by

ξ=12(c2+γ2+c+ic2+γ2c).

Proof. —

A complex number c+iγ can be written in the polar form

c+iγ=c2+γ2eiϕ

where

cosϕ=cc2+γ2andsinϕ=γc2+γ2.

Since γ0, we have ϕ[0,π]. Therefore,

ξ=c2+γ24eiϕ/2 A\,1

is the value of the square root c+iγ that lies in the quadrant Reξ0, Imξ0 (the other value of the square root has the opposite sign and therefore satisfies Reξ0 and Imξ0). Using standard trigonometric identities and the fact that ϕ[0,π], we find

cosϕ2=1+cosϕ2=12(1+cc2+γ2)andsinϕ2=1+cosϕ2=12(1cc2+γ2).

Inserting these expressions into formula (A 1), we obtain the assertion of the proposition.

Contributor Information

Oleh Omel’chenko, Email: omelchenko@uni-potsdam.de.

Carlo R. Laing, Email: c.r.laing@massey.ac.nz.

Data accessibility

The codes used for producing the data for figures are available in electronic supplementary material [36].

Authors' contributions

O.O.: formal analysis, methodology, software, validation, visualization, writing—original draft, writing—review and editing; C.R.L.: conceptualization, writing—original draft, writing—review and editing.

Both authors gave final approval for publication and agreed to be held accountable for the work performed therein.

Competing interests

We declare we have no competing interests.

Funding

The work of O.O. was supported by Deutsche Forschungsgemeinschaft, grant no. OM 99/2-1.

References

  • 1.Horn D, Opher I. 1997. Solitary waves of integrate-and-fire neural fields. Neural Comput. 9, 1677-1690. ( 10.1162/neco.1997.9.8.1677) [DOI] [PubMed] [Google Scholar]
  • 2.Ermentrout B. 2006. Gap junctions destroy persistent states in excitatory networks. Phys. Rev. E 74, 031918. ( 10.1103/PhysRevE.74.031918) [DOI] [PubMed] [Google Scholar]
  • 3.Jiruska P, de Curtis M, Jefferys JGR, Schevon CA, Schiff SJ, Schindler K. 2013. Synchronization and desynchronization in epilepsy: controversies and hypotheses. J. Physiol. 591, 787-797. ( 10.1113/jphysiol.2012.239590) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Netoff TI, Schiff SJ. 2002. Decreased neuronal synchronization during experimental seizures. J. Neurosci. 22, 7297-7307. ( 10.1523/JNEUROSCI.22-16-07297.2002) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Laing CR, Chow CC. 2002. A spiking neuron model for binocular rivalry. J. Comput. Neurosci. 12, 39-53. ( 10.1023/A:1014942129705) [DOI] [PubMed] [Google Scholar]
  • 6.Ermentrout B. 1998. Neural networks as spatio-temporal pattern-forming systems. Rep. Prog. Phys. 61, 353-430. ( 10.1088/0034-4885/61/4/002) [DOI] [Google Scholar]
  • 7.Bressloff PC. 2012. Spatiotemporal dynamics of continuum neural fields. J. Phys. A: Math. Theor. 45, 033001. ( 10.1088/1751-8113/45/3/033001) [DOI] [Google Scholar]
  • 8.Wang XJ. 2001. Synaptic reverberation underlying mnemonic persistent activity. Trends Neurosci. 24, 455-463. ( 10.1016/S0166-2236(00)01868-3) [DOI] [PubMed] [Google Scholar]
  • 9.Compte A, Brunel N, Goldman-Rakic PS, Wang XJ. 2000. Synaptic mechanisms and network dynamics underlying spatial working memory in a cortical network model. Cereb. Cortex 10, 910-923. ( 10.1093/cercor/10.9.910) [DOI] [PubMed] [Google Scholar]
  • 10.Laing C, Chow C. 2001. Stationary bumps in networks of spiking neurons. Neural Comput. 13, 1473-1494. ( 10.1162/089976601750264974) [DOI] [PubMed] [Google Scholar]
  • 11.Zhang K. 1996. Representation of spatial orientation by the intrinsic dynamics of the head-direction cell ensemble: a theory. J. Neurosci. 16, 2112-2126. ( 10.1523/JNEUROSCI.16-06-02112.1996) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Funahashi S, Bruce CJ, Goldman-Rakic PS. 1989. Mnemonic coding of visual space in the monkey’s dorsolateral prefrontal cortex. J. Neurophysiol. 61, 331-349. ( 10.1152/jn.1989.61.2.331) [DOI] [PubMed] [Google Scholar]
  • 13.Hubel DH, Wiesel TN. 1977. Ferrier lecture-Functional architecture of macaque monkey visual cortex. Proc. R. Soc. Lond. B 198, 1-59. ( 10.1098/rspb.1977.0085) [DOI] [PubMed] [Google Scholar]
  • 14.Esnaola-Acebes JM, Roxin A, Avitabile D, Montbrió E. 2017. Synchrony-induced modes of oscillation of a neural field model. Phys. Rev. E 96, 052407. ( 10.1103/PhysRevE.96.052407) [DOI] [PubMed] [Google Scholar]
  • 15.Schmidt H, Avitabile D. 2020. Bumps and oscillons in networks of spiking neurons. Chaos 30, 033133. ( 10.1063/1.5135579) [DOI] [PubMed] [Google Scholar]
  • 16.Byrne A, Ross J, Nicks R, Coombes S. 2022. Mean-field models for EEG/MEG: from oscillations to waves. Brain Topogr. 35, 36-53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Byrne A, Avitabile D, Coombes S. 2019. Next-generation neural field model: the evolution of synchrony within patterns and waves. Phys. Rev. E 99, 012313. ( 10.1103/PhysRevE.99.012313) [DOI] [PubMed] [Google Scholar]
  • 18.Gutkin BS, Laing CR, Colby CL, Chow CC, Ermentrout GB. 2001. Turning on and off with excitation: the role of spike-timing asynchrony and synchrony in sustained neural activity. J. Comput. Neurosci. 11, 121-134. ( 10.1023/A:1012837415096) [DOI] [PubMed] [Google Scholar]
  • 19.Tegnér J, Compte A, Wang XJ. 2002. The dynamical stability of reverberatory neural circuits. Biol. Cybern. 87, 471-481. ( 10.1007/s00422-002-0363-9) [DOI] [PubMed] [Google Scholar]
  • 20.Laing CR. 2014. Derivation of a neural field model from a network of theta neurons. Phys. Rev. E 90, 010901. ( 10.1103/PhysRevE.90.010901) [DOI] [PubMed] [Google Scholar]
  • 21.Laing CR. 2015. Exact neural fields incorporating gap junctions. SIAM J. Appl. Dyn. Syst. 14, 1899-1929. ( 10.1137/15M1011287) [DOI] [Google Scholar]
  • 22.Ermentrout GB, Kopell N. 1986. Parabolic bursting in an excitable system coupled with a slow oscillation. SIAM J. Appl. Math. 46, 233-253. ( 10.1137/0146017) [DOI] [Google Scholar]
  • 23.Luke TB, Barreto E, So P. 2013. Complete classification of the macroscopic behavior of a heterogeneous network of theta neurons. Neural Comput. 25, 3207-3234. ( 10.1162/NECO_a_00525) [DOI] [PubMed] [Google Scholar]
  • 24.Amari SI. 1977. Dynamics of pattern formation in lateral-inhibition type neural fields. Biol. Cybern. 27, 77-87. ( 10.1007/BF00337259) [DOI] [PubMed] [Google Scholar]
  • 25.Pinto DJ, Ermentrout GB. 2001. Spatially structured activity in synaptically coupled neuronal networks: II. Lateral inhibition and standing pulses. SIAM J. Appl. Math. 62, 226-243. ( 10.1137/S0036139900346465) [DOI] [Google Scholar]
  • 26.Laing CR, Troy WC, Gutkin B, Ermentrout GB. 2002. Multiple bumps in a neuronal model of working memory. SIAM J. Appl. Math. 63, 62-97. ( 10.1137/S0036139901389495) [DOI] [Google Scholar]
  • 27.Pazó D, Montbrió E. 2014. Low-dimensional dynamics of populations of pulse-coupled oscillators. Phys. Rev. X 4, 011009. ( 10.1103/PhysRevX.4.011009) [DOI] [Google Scholar]
  • 28.Ott E, Antonsen TM. 2008. Low dimensional behavior of large systems of globally coupled oscillators. Chaos 18, 037113. ( 10.1063/1.2930766) [DOI] [PubMed] [Google Scholar]
  • 29.Ott E, Antonsen TM. 2009. Long time evolution of phase oscillator systems. Chaos 19, 023117. ( 10.1063/1.3136851) [DOI] [PubMed] [Google Scholar]
  • 30.Montbrió E, Pazó D, Roxin A. 2015. Macroscopic description for networks of spiking neurons. Phys. Rev. X 5, 021028. ( 10.1103/PhysRevX.5.021028) [DOI] [Google Scholar]
  • 31.Abrams D, Strogatz S. 2006. Chimera states in a ring of nonlocally coupled oscillators. Int. J. Bifurcat. Chaos 16, 21-37. ( 10.1142/S0218127406014551) [DOI] [Google Scholar]
  • 32.Laing CR. 2009. The dynamics of chimera states in heterogeneous Kuramoto networks. Physica D 238, 1569-1588. ( 10.1016/j.physd.2009.04.012) [DOI] [Google Scholar]
  • 33.Omel’chenko OE, Knobloch E. 2019. Chimerapedia: coherence-incoherence patterns in one, two and three dimensions. New J. Phys. 21, 093034. ( 10.1088/1367-2630/ab3f6b) [DOI] [Google Scholar]
  • 34.Omel’chenko OE. 2018. The mathematics behind chimera states. Nonlinearity 31, R121-R164. ( 10.1088/1361-6544/aaaa07) [DOI] [Google Scholar]
  • 35.Laing CR. 2021. Effects of degree distributions in random networks of Type-I neurons. Phys. Rev. E 103, 052305. ( 10.1103/PhysRevE.103.052305) [DOI] [PubMed] [Google Scholar]
  • 36.Omel’chenko O, Laing CR. 2022. Collective states in a ring network of theta neurons. Figshare.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Availability Statement

The codes used for producing the data for figures are available in electronic supplementary material [36].


Articles from Proceedings. Mathematical, Physical, and Engineering Sciences are provided here courtesy of The Royal Society

RESOURCES