Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
\xspaceaddexceptions

Q Q\ellroman_ℓ S

Co-evolutionary dynamics for two adaptively coupled Theta neurons

Felix Augustsson    Erik A. Martens erik.martens@math.lth.se aCentre for Mathematical Sciences, Lund University, Sölvegatan 18, 221 00 Lund, Sweden
(July 1, 2024)
Abstract

Natural and technological networks exhibit dynamics that can lead to complex cooperative behaviors, such as synchronization in coupled oscillators and rhythmic activity in neuronal networks. Understanding these collective dynamics is crucial for deciphering a range of phenomena from brain activity to power grid stability. Recent interest in co-evolutionary networks has highlighted the intricate interplay between dynamics on and of the network with mixed time scales. Here, we explore the collective behavior of excitable oscillators in a simple networks of two Theta neurons with adaptive coupling without self-interaction. Through a combination of bifurcation analysis and numerical simulations, we seek to understand how the level of adaptivity in the coupling strength, a𝑎aitalic_a, influences the dynamics. We first investigate the dynamics possible in the non-adaptive limit; our bifurcation analysis reveals stability regions of quiescence and spiking behaviors, where the spiking frequencies mode-lock in a variety of configurations. Second, as we increase the adaptivity a𝑎aitalic_a, we observe a widening of the associated Arnol’d tongues, which may overlap and give room for multi-stable configurations. For larger adaptivity, the mode-locked regions may further undergo a period-doubling cascade into chaos. Our findings contribute to the mathematical theory of adaptive networks and offer insights into the potential mechanisms underlying neuronal communication and synchronization.

adaptive networks, collective dynamics, learning, Theta neurons
preprint: AIP/123-QED

Networks appear in a variety of natural and technological systems connecting dynamic units, such as heart and nerve cells in the body, and generators and consumers in the power grid. The network coupling facilitates emergent collective behaviors, such as the synchronization of oscillatory units. Examples include neurons adhering to collective firing during epileptic seizure, and the frequency locking of alternating currents in the power grid. Maintaining (or losing) such collective behaviors is critical for the proper functioning of these systems. Much research effort has already been laid in studying systems with either dynamics on the network nodes, or of the network edges. Combining dynamics on and of the networks results in co-evolutionary or adaptive networks, leading to new exciting challenges in understanding and analyzing the resulting dynamics. To make progress, this relatively young research area in mathematics has focused on relatively simple phase oscillator models. Here, we aim to extend the understanding of such networks by considering excitable oscillators, namely Theta neurons with adaptive coupling. We find several dynamic phenomena intrinsic to adaptive coupling, including not only the widening of mode locking regions, where neurons spike with different frequency ratios; but also their overlapping in multi-stable regions, and a period-doubling cascade into chaos. Thus, the system offers a wide range of stable mode-locking configurations for neural spiking behavior.

I Introduction

Networks are ubiquitous in nature and technology, including a range of systems from small to large scales [1], such as metabolic cell circuits [2], vascular and leaf venation networks [3, 4, 5, 6], neural networks in the brain [7, 8], power grid networks [9, 10], the internet [11], and transport systems [12, 13]. While for some purposes, networks may be considered stationary, the inclusion of dynamic processes are of great significance in a variety of contexts. Dynamics may occur on the network nodes (dynamics on the network) or along the edges forming the network (dynamics of the network). Examples of dynamics on the nodes include neurons, heart cells, flashing fireflies, Josephson junctions, metronomes and pendulum clocks [14, 15]; dynamics of the network includes processes in social networks, synaptic plasticity in the brain, regulation of flow in vascular networks, and more [16, 17, 18, 19]. Dynamics on networks are known to lead to emergent cooperative or collective behaviors, such as swarming and flocking in fish and birds [20, 21], or the synchronization of frequencies or phases in networks of coupled oscillators [14, 22, 15] and collective firing of neurons [23, 24], which have been subject to a long series of studies [25, 26, 27, 28].

While these types of dynamics have been studied for a few decades, the combination of both in so-called co-evolutionary or adaptive networks has attracted increased interest more recently [29, 30]. The complex interactions that may occur between two high dimensional systems which could be coupled in a variety of ways and at mixed time scales [31] represents a challenge in terms of developing a mathematical theory. It thus makes sense to start studying this topic using simple models. Some effort has already been undertaken in attempting to understand the dynamics of simple phase oscillator models, such as Kuramoto model [32] where oscillators are rotating constantly and are coupled with first order harmonics, combined with coupling strengths that adapt in response to the phase difference between two individual nodes [33, 30, 34, 35, 36, 37].

In this study, we ask the question how excitable oscillators, such as neurons, behave collectively in an adaptive network. While such dynamics has been investigated computationally in a variety of settings [38], this topic is subject for developing further rigid mathematical analysis. To take first steps in the direction of answering this question, we choose a simple phase oscillator model, the Theta neuron model [39, 40, 41], combined with a simple adaptation rule for the coupling strength. The Theta neuron is not only a good candidate model because of its simplicity, but is also equivalent to the quadratic integrate-and-fire model, can be derived from higher dimensional neuronal models [39, 42], featuring exact dimensional reduction methods exist for populations of Theta neurons with uniform coupling, akin to neural masses, [23, 27, 43, 28, 44], and has been used to study a variety of applications in a neuroscientific contexts [27, 45, 43].

To simplify the analysis, we follow a similar paradigm as in previous work by one of the authors [36], and study the dynamics for two adaptively coupled Theta neurons; we parameterize the strength of adaptation with the parameter a𝑎aitalic_a, starting from the non-adaptive limit of coupled neurons, and ask the questions: what dynamics occur for increasing adaptivity, which can genuinely be attributed to adaptive dynamics? Does adaptivity strengthen or weaken the emergent behaviors such as collective spiking? To obtain some answers we set up the model in Sec. II and begin the study in Sec. III with a careful bifurcation analysis for the dynamics of non-adaptively coupled neurons, for non-identical (Sec. III.C) and identical excitability parameters (Sec. III.D). Understanding these limiting cases, we are ready to begin a numerical analysis for the system with adaptive coupling. We conclude the study with a discussion of the obtained results in Sec. IV.

II Model

II.1 Network of Theta neurons

We consider a model of N𝑁Nitalic_N interacting Theta neurons, where the phase θk(t)𝕋=S1/2πsubscript𝜃𝑘𝑡𝕋superscript𝑆1similar-to-or-equals2𝜋\theta_{k}(t)\in\mathbb{T}=S^{1}\simeq\mathbb{R}/2\pi\mathbb{Z}italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t ) ∈ blackboard_T = italic_S start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT ≃ blackboard_R / 2 italic_π blackboard_Z of the k𝑘kitalic_kth neuron, k[N]:={1,,N}𝑘delimited-[]𝑁assign1𝑁k\in[N]:=\{1,\ldots,N\}italic_k ∈ [ italic_N ] := { 1 , … , italic_N }, evolves according to

θ˙k=1cosθk+(1+cosθk)ιk,subscript˙𝜃𝑘1subscript𝜃𝑘1subscript𝜃𝑘subscript𝜄𝑘\dot{\theta}_{k}=1-\cos{\theta_{k}}+\left(1+\cos{\theta_{k}}\right)\iota_{k},over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 1 - roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + ( 1 + roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , (1)

with parameter ιk=ηk+Iksubscript𝜄𝑘subscript𝜂𝑘subscript𝐼𝑘\iota_{k}=\eta_{k}+I_{k}italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_I start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT where ηksubscript𝜂𝑘\eta_{k}italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is the excitability threshold and Iksubscript𝐼𝑘I_{k}italic_I start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT the synaptic input current. The Theta neuron (1) is the normal form of the saddle-node-on-invariant-circle (SNIC) bifurcation [41] and is a canonical type 1 neuron [39]. For ιk<0subscript𝜄𝑘0\iota_{k}<0italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT < 0, a stable and unstable fixed point (θ,θ+subscript𝜃subscript𝜃\theta_{-},\theta_{+}italic_θ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT) occur on the phase circle 𝕋𝕋\mathbb{T}blackboard_T; for ιk=0subscript𝜄𝑘0\iota_{k}=0italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 0, these fixed points coalesce in a saddle-node bifurcation; for ιk>0subscript𝜄𝑘0\iota_{k}>0italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT > 0, the flow on the circle results in a periodic motion (spiking).

If ιk<0subscript𝜄𝑘0\iota_{k}<0italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT < 0, the Theta neuron is said to be excitable or quiescent: in the absence of perturbations, the phase relaxes to the stable fixed point θsubscript𝜃\theta_{-}italic_θ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT on the phase circle 𝕋𝕋\mathbb{T}blackboard_T; however, a perturbation may lead to a single spike (at θ=π𝜃𝜋\theta=\piitalic_θ = italic_π) before returning to the stable fixed point. This may happen in at least two ways: (i) the phase is perturbed across the unstable fixed point (constituting a threshold), this is possible, considering that the Theta model derives from a higher dimensional model [39] so that the circle is embedded in a higher dimensional space; (ii) a very short-lived (time scale of a single cycle) increase in ιksubscript𝜄𝑘\iota_{k}italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT momentarily pushes the system across the bifurcation threshold ιk=0subscript𝜄𝑘0\iota_{k}=0italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 0. If ιk>0subscript𝜄𝑘0\iota_{k}>0italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT > 0, the neuron is excited and is said to be spiking or firing (periodically).

In a network of Theta neurons, the input current may result from a variety of interactions [27]. Here, we assume that the input current for neuron k𝑘kitalic_k is given by a sum of pulses emitted from adjacent neurons l𝑙litalic_l,

Ik=1d(k)l[N]κklPs(θl),subscript𝐼𝑘1𝑑𝑘subscript𝑙delimited-[]𝑁subscript𝜅𝑘𝑙subscript𝑃𝑠subscript𝜃𝑙I_{k}=\frac{1}{d(k)}\sum_{l\in[N]}\kappa_{kl}P_{s}(\theta_{l}),italic_I start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_d ( italic_k ) end_ARG ∑ start_POSTSUBSCRIPT italic_l ∈ [ italic_N ] end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) , (2)

with pulses [46, 27] given by

Ps(θ)=as(1cosθ)s,subscript𝑃𝑠𝜃subscript𝑎𝑠superscript1𝜃𝑠\displaystyle P_{s}(\theta)=a_{s}\left(1-\cos{\theta}\right)^{s},italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_θ ) = italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( 1 - roman_cos italic_θ ) start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT , (3)

where the shape parameter sN𝑠𝑁s\in Nitalic_s ∈ italic_N controls the pulse width, and the normalization constant as=2s(s!)2/(2s)!subscript𝑎𝑠superscript2𝑠superscript𝑠22𝑠a_{s}=2^{s}(s!)^{2}/(2s)!italic_a start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 2 start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ( italic_s ! ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 2 italic_s ) ! satisfies 02πPs(θ)\dlθ=2πsuperscriptsubscript02𝜋subscript𝑃𝑠𝜃\dl𝜃2𝜋\int_{0}^{2\pi}P_{s}(\theta)\,\dl\theta=2\pi∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_θ ) italic_θ = 2 italic_π. The pulse function Pssubscript𝑃𝑠P_{s}italic_P start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT will attain its maxima at θ=π𝜃𝜋\theta=\piitalic_θ = italic_π, which is referred to as a spike. Pulses received from neuron l𝑙litalic_l are weighted by an synaptic interaction or coupling strength κklsubscript𝜅𝑘𝑙\kappa_{kl}italic_κ start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT, normalized by the number d(k)𝑑𝑘d(k)italic_d ( italic_k ) of neurons incident to neuron k𝑘kitalic_k.

In this study, we assume that neurons be autaptic, i.e., neurons have no self-interactions with κll=0subscript𝜅𝑙𝑙0\kappa_{ll}=0italic_κ start_POSTSUBSCRIPT italic_l italic_l end_POSTSUBSCRIPT = 0 for l[N]𝑙delimited-[]𝑁l\in[N]italic_l ∈ [ italic_N ], and d(k)=n1𝑑𝑘𝑛1d(k)=n-1italic_d ( italic_k ) = italic_n - 1.

II.2 Co-evolutionary network model

Many models assume that the coupling strengths, κklsubscript𝜅𝑘𝑙\kappa_{kl}italic_κ start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT, be constant, in which case we say the coupling is non-adaptive. In this study, however, our aim is to study the effects of adaptive coupling strengths, i.e., coupling strengths that adapt according to the activity of neurons. We consider the case where the coupling adapts on a time scale ε1superscript𝜀1\varepsilon^{-1}italic_ε start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT according to the rule

ε1κ˙kl=𝒜(θk,θl)κkl,superscript𝜀1subscript˙𝜅𝑘𝑙𝒜subscript𝜃𝑘subscript𝜃𝑙subscript𝜅𝑘𝑙\varepsilon^{-1}\dot{\kappa}_{kl}=\mathcal{A}(\theta_{k},\theta_{l})-\kappa_{% kl},italic_ε start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over˙ start_ARG italic_κ end_ARG start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT = caligraphic_A ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) - italic_κ start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT , (4)

with an adaptation function 𝒜=𝒜(θk,θl)𝒜𝒜subscript𝜃𝑘subscript𝜃𝑙\mathcal{A}=\mathcal{A}(\theta_{k},\theta_{l})caligraphic_A = caligraphic_A ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ). Thus, the adaptation is implemented to be pairwise or local between two neurons k𝑘kitalic_k and l𝑙litalic_l, and is homogeneous across the network. Combining Eqs. (1) and (4) results in adaptive or co-evolutionary network dynamics [29, 30]. The second term in (4) guarantees that the coupling strength stays bounded as long as A𝐴Aitalic_A is bounded. In this study, we choose the adaptive function

𝒜(θk,θl)=b+acos(θkθl+β),𝒜subscript𝜃𝑘subscript𝜃𝑙𝑏𝑎subscript𝜃𝑘subscript𝜃𝑙𝛽\mathcal{A}(\theta_{k},\theta_{l})=b+a\cos(\theta_{k}-\theta_{l}+\beta),caligraphic_A ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) = italic_b + italic_a roman_cos ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT + italic_β ) , (5)

where b𝑏bitalic_b implements a baseline for the coupling strength, a𝑎aitalic_a is the adaptivity or adaptation strength, and β𝛽\betaitalic_β an adaptation shift.

Non-adaptive dynamics is retrieved when a=0𝑎0a=0italic_a = 0. In this case the coupling converges to the baseline value κklbsubscript𝜅𝑘𝑙𝑏\kappa_{kl}\to bitalic_κ start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT → italic_b as t𝑡\ t\to\inftyitalic_t → ∞. Varying the adaptation shift tunes the type of adaptation. For β=0,π𝛽0𝜋\beta=0,\piitalic_β = 0 , italic_π, the adaptation function is symmetric, 𝒜(θk,θl)=𝒜(θl,θk)𝒜subscript𝜃𝑘subscript𝜃𝑙𝒜subscript𝜃𝑙subscript𝜃𝑘\mathcal{A}(\theta_{k},\theta_{l})=\mathcal{A}(\theta_{l},\theta_{k})caligraphic_A ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ) = caligraphic_A ( italic_θ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ). For β=0𝛽0\beta=0italic_β = 0 and a>0𝑎0a>0italic_a > 0, the coupling for in-phase configurations, i.e., neurons with phase difference close to 0 (or π𝜋\piitalic_π), is amplified (or suppressed). For β=π𝛽𝜋\beta=\piitalic_β = italic_π and a>0𝑎0a>0italic_a > 0 (which is equivalent to β=0𝛽0\beta=0italic_β = 0 and a<0𝑎0a<0italic_a < 0), the coupling for anti-phase configurations, i.e., neurons with phase difference close to π𝜋\piitalic_π (or 0), is amplified (or suppressed). Other choices for β𝛽\betaitalic_β may break the symmetry of the system, resulting in more complicated dynamics [36].

II.3 Governing equations for N=2𝑁2N=2italic_N = 2 neurons with adaptation

For the sake of this study, to simplify the analysis we restrict our focus to N=2𝑁2N=2italic_N = 2 neurons with pulse shape parameter s=1𝑠1s=1italic_s = 1. Further, we choose β=0𝛽0\beta=0italic_β = 0 (which also covers β=π𝛽𝜋\beta=\piitalic_β = italic_π if the sign of a𝑎aitalic_a is flipped), which implies that the difference coupling Δ:=κ21κ120assignΔsubscript𝜅21subscript𝜅120\Delta:=\kappa_{21}-\kappa_{12}\to 0roman_Δ := italic_κ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT - italic_κ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT → 0 as t𝑡t\to\inftyitalic_t → ∞, since we then have Δ˙=Δ˙ΔΔ\dot{\Delta}=-\Deltaover˙ start_ARG roman_Δ end_ARG = - roman_Δ. Thus κ:=(κ12+κ21)/2assign𝜅subscript𝜅12subscript𝜅212\kappa:=(\kappa_{12}+\kappa_{21})/2italic_κ := ( italic_κ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT + italic_κ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ) / 2 is the only dynamic inter-neuron coupling strength in the system (if transients are ignored). Using the assumption of autaptic neurons, we then have Ik=κ(1cos(θk))subscript𝐼𝑘𝜅1subscript𝜃𝑘I_{k}=\kappa(1-\cos(\theta_{k}))italic_I start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_κ ( 1 - roman_cos ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ). The dynamics then obeys the governing equations

θ1˙˙subscript𝜃1\displaystyle\dot{\theta_{1}}over˙ start_ARG italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG =1cosθ1+(1+cosθ1)(η1+κ(1cosθ2)),absent1subscript𝜃11subscript𝜃1subscript𝜂1𝜅1subscript𝜃2\displaystyle=1-\cos\theta_{1}+\left(1+\cos\theta_{1}\right)\left(\eta_{1}+% \kappa\left(1-\cos\theta_{2}\right)\right),= 1 - roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ( 1 + roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ( italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_κ ( 1 - roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ) , (6a)
θ2˙˙subscript𝜃2\displaystyle\dot{\theta_{2}}over˙ start_ARG italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG =1cosθ2+(1+cosθ2)(η2+κ(1cosθ1)),absent1subscript𝜃21subscript𝜃2subscript𝜂2𝜅1subscript𝜃1\displaystyle=1-\cos\theta_{2}+\left(1+\cos\theta_{2}\right)\left(\eta_{2}+% \kappa\left(1-\cos\theta_{1}\right)\right),= 1 - roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + ( 1 + roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ( italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_κ ( 1 - roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ) , (6b)
κ˙˙𝜅\displaystyle\dot{\kappa}over˙ start_ARG italic_κ end_ARG =ε(b+acos(θ2θ1)κ).absent𝜀𝑏𝑎subscript𝜃2subscript𝜃1𝜅\displaystyle=\varepsilon\left(b+a\cdot\cos(\theta_{2}-\theta_{1})-\kappa% \right).= italic_ε ( italic_b + italic_a ⋅ roman_cos ( italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) - italic_κ ) . (6c)

We note that one expects that trajectories in the asymptotic time limit of (6) reside in the compact space given by

(θ1,θ2,κ)𝕋2×[b|a|,b+|a|],subscript𝜃1subscript𝜃2𝜅superscript𝕋2𝑏𝑎𝑏𝑎\left(\theta_{1},\theta_{2},\kappa\right)\in\mathbb{T}^{2}\times\left[b-\lvert a% \rvert,b+\lvert a\rvert\right],( italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_κ ) ∈ blackboard_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT × [ italic_b - | italic_a | , italic_b + | italic_a | ] , (7)

since the range of cos𝑐𝑜𝑠cositalic_c italic_o italic_s lies in the interval [1,1]11[-1,1][ - 1 , 1 ].

In what follows, we assume that the adaptation evolves with ε=0.01𝜀0.01\varepsilon=0.01italic_ε = 0.01 and has a zero base line (b=0𝑏0b=0italic_b = 0).

III Analysis for non-adaptive coupling (a=0)𝑎0(a=0)( italic_a = 0 )

III.1 Fixed point analysis and saddle-node bifurcations

The fixed point conditions for (6) are

00\displaystyle 0 =1cosθ1+(1+cosθ1)(η1+κ(1cosθ2)),absent1subscript𝜃11subscript𝜃1subscript𝜂1𝜅1subscript𝜃2\displaystyle=1-\cos\theta_{1}+\left(1+\cos\theta_{1}\right)\left(\eta_{1}+% \kappa\left(1-\cos\theta_{2}\right)\right),= 1 - roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ( 1 + roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ( italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_κ ( 1 - roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ) , (8a)
00\displaystyle 0 =1cosθ2+(1+cosθ2)(η2+κ(1cosθ1)).absent1subscript𝜃21subscript𝜃2subscript𝜂2𝜅1subscript𝜃1\displaystyle=1-\cos\theta_{2}+\left(1+\cos\theta_{2}\right)\left(\eta_{2}+% \kappa\left(1-\cos\theta_{1}\right)\right).= 1 - roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + ( 1 + roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ( italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_κ ( 1 - roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ) . (8b)

By the change of variables, cosθ1=2x1subscript𝜃12𝑥1\cos\theta_{1}=2x-1roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 2 italic_x - 1, cosθ2=2y1subscript𝜃22𝑦1\cos\theta_{2}=2y-1roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 2 italic_y - 1 and νk=ηk+2κ1subscript𝜈𝑘subscript𝜂𝑘2𝜅1\nu_{k}=\eta_{k}+2\kappa-1italic_ν start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + 2 italic_κ - 1 (where 0x10𝑥10\leq x\leq 10 ≤ italic_x ≤ 1 and 0y10𝑦10\leq y\leq 10 ≤ italic_y ≤ 1), the fixed point equations in Eq. 8 can be expressed as

00\displaystyle 0 =2κxy+ν1x+1,absent2𝜅𝑥𝑦subscript𝜈1𝑥1\displaystyle=-2\kappa xy+\nu_{1}x+1,= - 2 italic_κ italic_x italic_y + italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x + 1 , (9a)
00\displaystyle 0 =2κxy+ν2y+1.absent2𝜅𝑥𝑦subscript𝜈2𝑦1\displaystyle=-2\kappa xy+\nu_{2}y+1.= - 2 italic_κ italic_x italic_y + italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_y + 1 . (9b)

However, from Eq. 9 we see that solutions with x=0𝑥0x=0italic_x = 0 or y=0𝑦0y=0italic_y = 0 are not valid solutions, so for any fixed point x,y(0,1]𝑥𝑦01x,y\in(0,1]italic_x , italic_y ∈ ( 0 , 1 ].

We wish to eliminate one variable. Subtracting (9a) from (9b) implies that

ν1x=ν2y.subscript𝜈1𝑥subscript𝜈2𝑦\nu_{1}x=\nu_{2}y.italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x = italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_y . (10)

Thus, solutions can only exist if sgnν1=sgnν2sgnsubscript𝜈1sgnsubscript𝜈2\operatorname{sgn}\nu_{1}=\operatorname{sgn}\nu_{2}roman_sgn italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_sgn italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. This condition can be satisfied in two ways:

(i) If ν1=ν2=0subscript𝜈1subscript𝜈20\nu_{1}=\nu_{2}=0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0, Eqs. (9) reduce to the condition

2κxy=1.2𝜅𝑥𝑦12\kappa xy=1.2 italic_κ italic_x italic_y = 1 . (11)

This condition defines a one-dimensional manifold of fixed points as long as κ>12𝜅12\kappa>\frac{1}{2}italic_κ > divide start_ARG 1 end_ARG start_ARG 2 end_ARG, and represents the (zero-dimensional) point x=y=1𝑥𝑦1x=y=1italic_x = italic_y = 1 when κ=12𝜅12\kappa=\frac{1}{2}italic_κ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG. On the other hand, if κ<12𝜅12\kappa<\frac{1}{2}italic_κ < divide start_ARG 1 end_ARG start_ARG 2 end_ARG there is no fixed point. Note that ν1=ν2=0subscript𝜈1subscript𝜈20\nu_{1}=\nu_{2}=0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 implies η1=η2=12κsubscript𝜂1subscript𝜂212𝜅\eta_{1}=\eta_{2}=1-2\kappaitalic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 - 2 italic_κ, i.e., neurons are identical. This curve is seen in Fig. 1 (dashed line).

(ii) If ν1,ν20subscript𝜈1subscript𝜈20\nu_{1},\nu_{2}\neq 0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≠ 0, the condition sgnν1=sgnν2sgnsubscript𝜈1sgnsubscript𝜈2\operatorname{sgn}\nu_{1}=\operatorname{sgn}\nu_{2}roman_sgn italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_sgn italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT means that either ν1,ν2>0subscript𝜈1subscript𝜈20\nu_{1},\nu_{2}>0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0 or ν1,ν2<0subscript𝜈1subscript𝜈20\nu_{1},\nu_{2}<0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < 0. We may without loss of generality assume that |ν1||ν2|subscript𝜈1subscript𝜈2\lvert\nu_{1}\rvert\leq\lvert\nu_{2}\rvert| italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | ≤ | italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT |. From (10) it follows that

y=ν1ν2x,𝑦subscript𝜈1subscript𝜈2𝑥y=\frac{\nu_{1}}{\nu_{2}}x,italic_y = divide start_ARG italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG italic_x , (12)

implying that y(0,1]𝑦01y\in(0,1]italic_y ∈ ( 0 , 1 ] if x(0,1]𝑥01x\in(0,1]italic_x ∈ ( 0 , 1 ]. Considering Eqs. (9), it is therefore sufficient in the case of non-zero, same sign ν1,ν2subscript𝜈1subscript𝜈2\nu_{1},\nu_{2}italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT to solve the quadratic equation,

0=2κν1ν2x2+ν1x+1,02𝜅subscript𝜈1subscript𝜈2superscript𝑥2subscript𝜈1𝑥10=-2\kappa\frac{\nu_{1}}{\nu_{2}}x^{2}+\nu_{1}x+1,0 = - 2 italic_κ divide start_ARG italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_x + 1 , (13)

for x(0,1]𝑥01x\in(0,1]italic_x ∈ ( 0 , 1 ]. Saddle-node bifurcations occur in two scenarios. The first is when x=12(1+cosθ1)=1𝑥121subscript𝜃11x=\frac{1}{2}(1+\cos{\theta_{1}})=1italic_x = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 + roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = 1 (or y=1𝑦1y=1italic_y = 1 in the case of |ν2||ν1|subscript𝜈2subscript𝜈1\lvert\nu_{2}\rvert\leq\lvert\nu_{1}\rvert| italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | ≤ | italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT |). In the original parameters, this condition is equivalent to

00\displaystyle 0 =η1η2+2η2κη1, for η20formulae-sequenceabsentsubscript𝜂1subscript𝜂22subscript𝜂2𝜅subscript𝜂1 for subscript𝜂20\displaystyle=\eta_{1}\eta_{2}+2\eta_{2}\kappa-\eta_{1},\text{ for }\eta_{2}\leq 0= italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 2 italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_κ - italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , for italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ 0 (14a)
00\displaystyle 0 =η1η2+2η1κη2, for η10formulae-sequenceabsentsubscript𝜂1subscript𝜂22subscript𝜂1𝜅subscript𝜂2 for subscript𝜂10\displaystyle=\eta_{1}\eta_{2}+2\eta_{1}\kappa-\eta_{2},\text{ for }\eta_{1}\leq 0= italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 2 italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_κ - italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , for italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ 0 (14b)

The second is when the discriminant of the quadratic polynomial in Eq. (13) is 0 while the solution to the equation x(0,1]𝑥01x\in(0,1]italic_x ∈ ( 0 , 1 ] (or y(0,1]𝑦01y\in(0,1]italic_y ∈ ( 0 , 1 ] in the case of |ν2||ν1|subscript𝜈2subscript𝜈1\lvert\nu_{2}\rvert\leq\lvert\nu_{1}\rvert| italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | ≤ | italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT |), i.e.

(η1+2κ1)(η2+2κ1)+8κ=0, for 12|κ|η11+2|κ| and 12|κ|η21+2|κ|.formulae-sequenceformulae-sequencesubscript𝜂12𝜅1subscript𝜂22𝜅18𝜅0 for 12𝜅subscript𝜂112𝜅 and 12𝜅subscript𝜂212𝜅\begin{split}(\eta_{1}+2\kappa-1)(\eta_{2}+2\kappa-1)+8\kappa=0&,\\ \text{ for }\quad 1-2\lvert\kappa\rvert\leq\eta_{1}\leq 1+2\lvert\kappa\rvert&% \\ \text{ and }\quad 1-2\lvert\kappa\rvert\leq\eta_{2}\leq 1+2\lvert\kappa\rvert&% .\end{split}start_ROW start_CELL ( italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 italic_κ - 1 ) ( italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 2 italic_κ - 1 ) + 8 italic_κ = 0 end_CELL start_CELL , end_CELL end_ROW start_ROW start_CELL for 1 - 2 | italic_κ | ≤ italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ 1 + 2 | italic_κ | end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL and 1 - 2 | italic_κ | ≤ italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ 1 + 2 | italic_κ | end_CELL start_CELL . end_CELL end_ROW (15)

The resulting saddle-node bifurcation curves SN1 (Eq. (14)) and SN2 (Eq. (15)) are shown as black solid curves in Figs. 1 and 5.

It is furthermore possible to determine the number of fixed points that appear in the different regions outlined by the fold bifurcations. A derivation is given in Appendix A, and the results are summarized in Table 1 and 2.

In the following two sections, we specialize these results to discuss the cases of identical and non-identical excitabilities separately, see Fig. 1 and 5.

III.2 Dynamics for identical neurons (η1=η2subscript𝜂1subscript𝜂2\eta_{1}=\eta_{2}italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT)

We consider first the situation when the two neurons are identical, i.e., η:=η1=η2assign𝜂subscript𝜂1subscript𝜂2\eta:=\eta_{1}=\eta_{2}italic_η := italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and uncoupled, so that κ=b=0𝜅𝑏0\kappa=b=0italic_κ = italic_b = 0 in the governing equations. We then have ιk=ηsubscript𝜄𝑘𝜂\iota_{k}=\etaitalic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_η for k=1,2𝑘12k=1,2italic_k = 1 , 2, and each neuron effectively behaves like a single neuron. Recall that we already explained in Sec. III.1 following (8), such a neuron exhibits a SNIC bifurcation at the threshold defined by ιk=η=0subscript𝜄𝑘𝜂0\iota_{k}=\eta=0italic_ι start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_η = 0. An uncoupled neuron may thus only display two different types of dynamic states. For η0𝜂0\eta\leq 0italic_η ≤ 0 the phase θksubscript𝜃𝑘\theta_{k}italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ends up in a fixed point (quiescence) which we denote Q. For η>0𝜂0\eta>0italic_η > 0, the phase will periodically pass θ=π𝜃𝜋\theta=-\piitalic_θ = - italic_π (spiking) which we denote S. Since both neurons are identically parametrized, two uncoupled neurons are either both quiescent (Q Q) or both spiking (S S), as shown in Fig. 1.

\begin{overpic}[width=433.62pt]{figures/nonadaptive-sym/stability-diagram} \put(96.0,4.0){\Large$\eta_{\sigma}$} \put(4.0,96.0){\Large$\kappa$} \put(51.5,70.0){SN\textsubscript{1} } \put(80.0,15.0){SN\textsubscript{2} } \put(51.5,37.0){Cusp} \put(46.5,44.5){a} \put(51.0,44.5){b} \put(55.5,44.5){c} \put(46.0,20.5){d} \put(51.5,20.5){e} \put(58.5,23.0){f} \put(69.0,23.0){g} \put(21.8,59.5){h} \put(21.8,64.0){i} \put(21.8,70.5){j} \put(22.0,35.0){\Large Q Q } \put(70.0,65.0){\Large S S } \put(22.0,80.0){\large Q\textsubscript{$\ell$} Q\textsubscript{$\ell$}/S S } \put(33.0,60.5){\rotatebox{-15.0}{Q Q/S S}} \put(52.0,10.0){Q Q/Q\textsubscript{$\ell$} Q\textsubscript{$\ell$} } \end{overpic}
Figure 1: Stability diagram for identical neurons in the non-adaptive system Eq. 6 with a=0𝑎0a=0italic_a = 0 and η1=η2=ηsubscript𝜂1subscript𝜂2𝜂\eta_{1}=\eta_{2}=\etaitalic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_η. Saddle-node bifurcations are shown as solid lines, the continuum of fixed points as a dashed line and global bifurcations as dash-dotted lines. Regions are colored based on their attractors: Q Q is blue, Q\ellroman_ℓ Q\ellroman_ℓ is yellow and S S is orange. Phase portraits for the points LABEL:sym-kappa-close0-I-LABEL:sym-kappa-grow-III are shown in Figs. 2, 3 and 4

When the two neurons are coupled, κ0𝜅0\kappa\neq 0italic_κ ≠ 0, the resulting dynamics and possible bifurcations can become more complicated. In general, three types of attractors can be observed for identical neurons. Both states/attractors from the uncoupled case carry over to the coupled case, i.e., both nodes are quiescent (Q Q) or spiking periodically (S S); however, a new dynamic scenario becomes possible, where the neurons do not spike but oscillate in a small-amplitude periodic orbit (Q\ellroman_ℓ Q\ellroman_ℓ), corresponding to a libration (spiking corresponds to a rotation).

We first focus on negative coupling strengths. The same bifurcation scenario as for the uncoupled case (κ=0𝜅0\kappa=0italic_κ = 0) occurs for the range of 12<κ012𝜅0-\frac{1}{2}<\kappa\leq 0- divide start_ARG 1 end_ARG start_ARG 2 end_ARG < italic_κ ≤ 0:

\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% close0/phase-plot-eta1-m0p1-eta2-m0p1-kappa-m0p1} \put(5.0,86.0){\ref{sym-kappa-close0-I})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% close0/phase-plot-eta1-0p-eta2-0p-kappa-m0p1} \put(5.0,86.0){\ref{sym-kappa-close0-II})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% close0/phase-plot-eta1-0p1-eta2-0p1-kappa-m0p1} \put(5.0,86.0){\ref{sym-kappa-close0-III})} \end{overpic}
Figure 2: Phase portrait for a=0𝑎0a=0italic_a = 0, κ=0.1𝜅0.1\kappa=-0.1italic_κ = - 0.1 with η=0.1,0,0.1𝜂0.100.1\eta=-0.1,0,0.1italic_η = - 0.1 , 0 , 0.1 and identical excitabilities, corresponding to the points LABEL:sym-kappa-close0-I-LABEL:sym-kappa-close0-III shown in Fig. 1, respectively. (Nullclines are shown as gray dashed lines, stable/unstable/saddle fixed points are shown as filled/empty/half-filled circles.)

When η<0𝜂0\eta<0italic_η < 0 there is a stable node, unstable node and two saddles (Q Q, Fig. 2 LABEL:sym-kappa-close0-I). At η=0𝜂0\eta=0italic_η = 0 the 4 fixed points coalesce in the origin in a saddle-node bifurcation (SN1, Fig. 2 LABEL:sym-kappa-close0-II). For η>0𝜂0\eta>0italic_η > 0, there are only spiking orbits (S S, Fig. 2 LABEL:sym-kappa-close0-III). Note that these spiking states are not limit cycles, but rather are foliating the entire phase space.

As we decrease κ𝜅\kappaitalic_κ further below 1212\frac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG, additional bifurcations occur. To explain the transitions for κ<12𝜅12\kappa<\frac{1}{2}italic_κ < divide start_ARG 1 end_ARG start_ARG 2 end_ARG we increase η𝜂\etaitalic_η from negative to positive values.

\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% neg/phase-plot-eta1-m0p1-eta2-m0p1-kappa-m1p2} \put(5.0,86.0){\ref{sym-kappa-neg-I})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% neg/phase-plot-eta1-0p05-eta2-0p05-kappa-m1p2-color} \put(5.0,86.0){\ref{sym-kappa-neg-II})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% neg/phase-plot-eta1-0p15-eta2-0p15-kappa-m1p2-color} \put(5.0,86.0){\ref{sym-kappa-neg-III})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% neg/phase-plot-eta1-0p4-eta2-0p4-kappa-m1p2} \put(5.0,86.0){\ref{sym-kappa-neg-IV})} \end{overpic}
Figure 3: Phase portraits for the non-adaptive system with a=0𝑎0a=0italic_a = 0, κ=1.2𝜅1.2\kappa=-1.2italic_κ = - 1.2 and η=0.1,0.05,0.15,0.4𝜂0.10.050.150.4\eta=-0.1,0.05,0.15,0.4italic_η = - 0.1 , 0.05 , 0.15 , 0.4 and identical excitabilities, corresponding to the points LABEL:sym-kappa-neg-I-LABEL:sym-kappa-neg-IV in Fig. 1, respectively. (Nullclines are shown as gray dashed lines, stable/unstable/saddle/center fixed points are shown as filled/empty/half-filled/center-dot circles.)

For η<0𝜂0\eta<0italic_η < 0 we have two saddles and a stable and an unstable node (Q Q, Fig. 3 LABEL:sym-kappa-neg-I). As we augment η𝜂\etaitalic_η and traverse SN1 from left to right, four additional fixed points, two saddles and two center points, are born at the origin (Fig. 3 LABEL:sym-kappa-neg-II). We denote these as the inner fixed points, as opposed to the outer fixed points present in LABEL:sym-kappa-neg-I. The center points are surrounded by a foliation of librations (Q\ellroman_ℓ Q\ellroman_ℓ) bounded by heteroclinic curves connecting the two saddles (green orbits in Fig. 3 LABEL:sym-kappa-neg-II). Thus, this region displays bi-stability between quiescent (Q Q) and librating (Q\ellroman_ℓ Q\ellroman_ℓ) states. As we increase η𝜂\etaitalic_η further, the heteroclinic connections change their nature as the dash-dotted curve is traversed111This boundary is numerically determined by considering orbits starting near the outer saddle and determining whether the resulting orbit remains oscillatory (libration) or converges to the stable node: instead of connecting the saddles to each other, they now have become homoclinic curves connecting each saddle to itself (blue and red orbits in Fig. 3 LABEL:sym-kappa-neg-III). Note that the system still provides the same type of bi-stability (Q Q/Q\ellroman_ℓ Q\ellroman_ℓ). Finally, as we further increase η𝜂\etaitalic_η, we traverse SN2 and all eight fixed points coalesce pairwise on the diagonals |θ1|=|θ2|subscript𝜃1subscript𝜃2|\theta_{1}|=|\theta_{2}|| italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | = | italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | in a saddle-node bifurcation, leaving behind only periodic spiking orbits (S S, Fig. 3 LABEL:sym-kappa-neg-IV).

For positive coupling (κ>0𝜅0\kappa>0italic_κ > 0), we first focus on the bifurcations in the region of negative excitability (η<0𝜂0\eta<0italic_η < 0).

\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% grow/phase-plot-eta1-m0p7-eta2-m0p7-kappa-0p5} \put(5.0,86.0){\ref{sym-kappa-grow-I})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% grow/phase-plot-eta1-m0p7-eta2-m0p7-kappa-0p7-color} \put(5.0,86.0){\ref{sym-kappa-grow-II})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-sym/phase-plots/kappa-% grow/phase-plot-eta1-m0p7-eta2-m0p7-kappa-1p-color} \put(5.0,86.0){\ref{sym-kappa-grow-III})} \end{overpic}
Figure 4: Phase portraits for the non-adaptive system with a=0𝑎0a=0italic_a = 0, η=0.7𝜂0.7\eta=-0.7italic_η = - 0.7 and κ=0.5,0.7,1𝜅0.50.71\kappa=0.5,0.7,1italic_κ = 0.5 , 0.7 , 1 and identical excitabilities, corresponding to the parameter points LABEL:sym-kappa-grow-I-LABEL:sym-kappa-grow-III in Fig. 1, respectively. Two homoclinic orbits are marked in blue and red, and two heteroclinic orbits are marked in green. (Nullclines are shown as gray dashed lines, stable/unstable/saddle/center fixed points are shown as filled/empty/half-filled/center-dot circles.)

For sufficiently low κ𝜅\kappaitalic_κ, the system only displays four fixed points (stable/unstable node, two saddles) so that the only stable state corresponds to quiescence (Q Q, Fig. 4 LABEL:sym-kappa-grow-I). As we increase κ𝜅\kappaitalic_κ, the topology of the phase space changes substantially in a global bifurcation222The location of this bifurcation was determined numerically by sampling many initial conditions and determining whether trajectories only converge to a quiescent fixed point (stable node Q Q), or whether some trajectories correspond to rotations (S S) in phase space. (dash-dotted line in Fig. 1). The stable and unstable manifolds of the two saddles merge to form a homoclinic connection for each saddle separately (these homoclinic orbits are highlighted in red and blue in Fig. 4 LABEL:sym-kappa-grow-II). Between these homoclinic orbits lie a foliation of periodic spiking orbits. Thus this region displays bi-stability between quiescent (Q Q) and spiking (S S) states. Further increasing κ𝜅\kappaitalic_κ, we cross the boundary defined by ν=η+2κ1=0𝜈𝜂2𝜅10\nu=\eta+2\kappa-1=0italic_ν = italic_η + 2 italic_κ - 1 = 0 (dashed line in Fig. 1), where the two fixed point conditions in Eq. 8 become identical (i.e., nullclines are identical) and produce a continuum of fixed points. Above this global bifurcation, on one hand, the two saddles transform into centers surrounded by a foliation of periodic librating orbits (Q\ellroman_ℓ Q\ellroman_ℓ, Fig. 4 LABEL:sym-kappa-grow-III). Vice versa, the stable and unstable nodes have become saddles with heteroclinic connections (green orbits in Fig. 4 LABEL:sym-kappa-grow-III) separating the librations from a foliation of rotations (S S). Consequently, this region is bi-stable between co-existence of Q\ellroman_ℓ Q\ellroman_ℓ and S S states.

Finally, traversing saddle-node bifurcation SN1 located at η=0𝜂0\eta=0italic_η = 0 in the regime of positive coupling, κ>0𝜅0\kappa>0italic_κ > 0, while increasing η𝜂\etaitalic_η, the fixed points coalesce at the origin in a saddle-node bifurcation, leaving behind a spiking state (S S) only. This scenario is similar to the ones observed for 12<κ012𝜅0\frac{1}{2}<\kappa\leq 0divide start_ARG 1 end_ARG start_ARG 2 end_ARG < italic_κ ≤ 0.

III.3 Dynamics for non-identical neurons (η1η2subscript𝜂1subscript𝜂2\eta_{1}\neq\eta_{2}italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT)

III.3.1 Bifurcations of fixed points and limit cycles

We now move on to the case of non-identical excitabilities (η1η2subscript𝜂1subscript𝜂2\eta_{1}\neq\eta_{2}italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT). In what follows, we will use the parametrization ησ=(η1+η2)/2subscript𝜂𝜎subscript𝜂1subscript𝜂22\eta_{\sigma}=(\eta_{1}+\eta_{2})/2italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = ( italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) / 2, ηδ=(η2η1)/2subscript𝜂𝛿subscript𝜂2subscript𝜂12\eta_{\delta}=(\eta_{2}-\eta_{1})/2italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = ( italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) / 2. With this parametrization, the identical case can be seen as the limiting case with ηδ=0subscript𝜂𝛿0\eta_{\delta}=0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0, η=ησ𝜂subscript𝜂𝜎\eta=\eta_{\sigma}italic_η = italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT; similarly, we will here fix an ηδ0subscript𝜂𝛿0\eta_{\delta}\neq 0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT ≠ 0, and look at bifurcations in (ησ(\eta_{\sigma}( italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT,κ)\kappa)italic_κ ) space. Without loss of generality, we will assume that ηδ>0subscript𝜂𝛿0\eta_{\delta}>0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT > 0 (η2>η1subscript𝜂2subscript𝜂1\eta_{2}>\eta_{1}italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT). Further, we will use ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1 throughout much of our analysis. The stability diagram and phase portraits in Figs. 5 to 10 summarize the possible states and bifurcations.

As for the identical case, we consider first the situation of uncoupled neurons (κ=0𝜅0\kappa=0italic_κ = 0). Since ι1=η1η2=ι2subscript𝜄1subscript𝜂1subscript𝜂2subscript𝜄2\iota_{1}=\eta_{1}\neq\eta_{2}=\iota_{2}italic_ι start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_ι start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, the two neurons will undergo SNIC bifurcations at different thresholds: ησ+ηδ=η2=ι2=0subscript𝜂𝜎subscript𝜂𝛿subscript𝜂2subscript𝜄20\eta_{\sigma}+\eta_{\delta}=\eta_{2}=\iota_{2}=0italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT + italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_ι start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 and ησηδ=η1=ι1=0subscript𝜂𝜎subscript𝜂𝛿subscript𝜂1subscript𝜄10\eta_{\sigma}-\eta_{\delta}=\eta_{1}=\iota_{1}=0italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT - italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_ι start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 for neurons 2 and 1, respectively. The non-identical uncoupled system can thus have 3 states: When ησ<ηδsubscript𝜂𝜎subscript𝜂𝛿\eta_{\sigma}<-\eta_{\delta}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT < - italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT both neurons are quiescent (Q Q), when ηδ<ησ<ηδsubscript𝜂𝛿subscript𝜂𝜎subscript𝜂𝛿-\eta_{\delta}<\eta_{\sigma}<\eta_{\delta}- italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT < italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT < italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT, neuron 1 is quiescent while neuron two is spiking (Q S) and when ηδ<ησsubscript𝜂𝛿subscript𝜂𝜎\eta_{\delta}<\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT < italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT both neurons are spiking (S S). These transitions are shown in Fig. 5.

\begin{overpic}[width=433.62pt]{figures/nonadaptive-asym/stability-diagram} \put(96.0,4.0){\Large$\eta_{\sigma}$} \put(4.0,96.0){\Large$\kappa$} \put(36.0,71.0){SN\textsubscript{1b} } \put(25.0,56.5){SN\textsubscript{1a} } % \put(70.0,13.0){SN\textsubscript{2} } \put(80.0,37.0){LPC} \put(51.5,38.0){Cusp} \put(43.0,48.0){k} \put(47.5,48.0){l} \put(51.5,48.0){m} \put(47.5,24.5){n} \put(53.0,24.5){o} \put(57.0,24.5){p} \put(30.0,69.0){q} \put(30.0,76.0){r} \put(22.0,35.0){\Large Q Q } \put(70.0,65.0){\Large S S } \put(22.0,83.0){\large Q Q } \put(70.0,25.0){\large Q S } \put(56.0,15.0){\large 2Q Q } \end{overpic}
Figure 5: Stability diagram for non-identical neurons in the non-adaptive system Eq. 6 with a=0𝑎0a=0italic_a = 0 and ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1. Saddle-node bifurcations are shown as solid lines, and limit points of (LPC) cycles as a dashed line. Regions are colored based on their attractors: Q Q is blue, Q S is pink and S S is orange. Phase portraits for the points LABEL:asym-close0-I-LABEL:asym-pos-II are shown in Figs. 6, 7 and 8

For two coupled neurons (κ0𝜅0\kappa\neq 0italic_κ ≠ 0), the input to neuron 1 ι1subscript𝜄1\iota_{1}italic_ι start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT will vary with time , and so the Q S state does — strictly speaking — not exist. Instead, we can observe the state Q\ellroman_ℓ S, where neuron 1 is “chasing” a moving stable phase. However, looking topologically at the phase space of the system, the Q\ellroman_ℓ S state is not distinct from the Q S state, and we will therefore treat them as the same type of state.

Next, we focus on coupling κ𝜅\kappaitalic_κ close to 0, specifically less than the minimum value of the upper component of SN1b (κ<|ηδ|+2|ηδ|+12𝜅subscript𝜂𝛿2subscript𝜂𝛿12\kappa<\lvert\eta_{\delta}\rvert+\sqrt{2\lvert\eta_{\delta}\rvert}+\frac{1}{2}italic_κ < | italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT | + square-root start_ARG 2 | italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT | end_ARG + divide start_ARG 1 end_ARG start_ARG 2 end_ARG) and greater than the value at the cusp point where SN1a and SN2 meet (κ>(1+|ηδ|)/2𝜅1subscript𝜂𝛿2\kappa>-(1+\lvert\eta_{\delta}\rvert)/2italic_κ > - ( 1 + | italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT | ) / 2). In this region, the same bifurcations occur when varying ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT as in the uncoupled case. However, since the neurons now are coupled, we must consider the bifurcations of individual neurons as bifurcations of the entire system.

\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/kappa-% close0/phase-plot-eta1-m0p25-eta2-m0p05-kappa-0p1} \put(5.0,86.0){\ref{asym-close0-I})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/kappa-% close0/phase-plot-eta1-m0p15-eta2-0p05-kappa-0p1} \put(5.0,86.0){\ref{asym-close0-II})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/kappa-% close0/phase-plot-eta1-m0p05-eta2-0p15-kappa-0p1} \put(5.0,86.0){\ref{asym-close0-III})} \end{overpic}
Figure 6: Phase portraits for the non-adaptive system with a=0𝑎0a=0italic_a = 0, κ=0.1𝜅0.1\kappa=0.1italic_κ = 0.1, ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1, and ησ=0.15,0.05,0.05subscript𝜂𝜎0.150.050.05\eta_{\sigma}=-0.15,-0.05,0.05italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = - 0.15 , - 0.05 , 0.05, corresponding to the parameter points LABEL:asym-close0-I-LABEL:asym-close0-III in Fig. 5, respectively. (Nullclines are shown as gray dashed lines, stable/unstable/saddle fixed points are shown as filled/empty/half-filled circles.)

For small ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, we have two saddles and a stable and an unstable node (Q Q, Fig. 6 LABEL:asym-close0-I). As ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT is increased and passes the lower component of SN1, the two pairs of left and right fixed points respectively coalesce in a saddle-node bifurcation on θ2=0subscript𝜃20\theta_{2}=0italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0. What is left is a stable and an unstable limit cycle that crosses θ2=πsubscript𝜃2𝜋\theta_{2}=\piitalic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_π (corresponding to a spike), but where only θ1subscript𝜃1\theta_{1}italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT librates (Q S, Fig. 6 LABEL:asym-close0-II). Further increasing ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, the two limit cycles coalesce in a limit point of cycles (LPC), and for larger ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT there are only periodic orbits where both neurons spike (S S, Fig. 6 LABEL:asym-close0-III).

Moving our attention to the region below the cusp point where κ<(1+|ηδ|)/2𝜅1subscript𝜂𝛿2\kappa<-(1+\lvert\eta_{\delta}\rvert)/2italic_κ < - ( 1 + | italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT | ) / 2, starting at small ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT there are 4 fixed points (stable/unstable, 2 saddles) (Q Q, Fig. 7 LABEL:asym-neg-I).

\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/kappa-% neg/phase-plot-eta1-m0p1-eta2-0p1-kappa-m1p} \put(5.0,86.0){\ref{asym-neg-I})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/kappa-% neg/phase-plot-eta1-0p-eta2-0p2-kappa-m1p} \put(5.0,86.0){\ref{asym-neg-II})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/kappa-% neg/phase-plot-eta1-0p1-eta2-0p3-kappa-m1p} \put(5.0,86.0){\ref{asym-neg-III})} \end{overpic}
Figure 7: Phase portraits for the non-adaptive system with a=0𝑎0a=0italic_a = 0, κ=1𝜅1\kappa=-1italic_κ = - 1 and ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1, and ησ=0,0.1,0.2subscript𝜂𝜎00.10.2\eta_{\sigma}=0,0.1,0.2italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = 0 , 0.1 , 0.2, corresponding to the parameter points LABEL:asym-neg-I-LABEL:asym-neg-III in Fig. 5, respectively. (Nullclines are shown as gray dashed lines, stable/unstable/saddle fixed points are shown as filled/empty/half-filled circles.)

Increasing ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT across SN1, four more fixed points (stable/unstable spirals, 2 saddles) which we denote as “inner fixed points” are created in a saddle-node bifurcation on θ2=0subscript𝜃20\theta_{2}=0italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 (Fig. 7 LABEL:asym-neg-II). Thus, these region exhibits bi-stability with two fixed points corresponding to quiescence, and so we denote the region with (2Q Q). Further increasing ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT across SN2, the inner and outer fixed points coalesce in each quadrant in a saddle-node bifurcation, and leaves a stable and an unstable limit cycle the only spike in neuron 2 (Q S, Fig. 7 LABEL:asym-neg-III). Finally, as in the case for κ𝜅\kappaitalic_κ closer to 0, for large ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT the two limit cycles coalesce and disappear in an LPC; thus, only states where both neurons spike (S S) are left.

Considering larger κ>0𝜅0\kappa>0italic_κ > 0, for ησ<|ηδ|subscript𝜂𝜎subscript𝜂𝛿\eta_{\sigma}<\lvert\eta_{\delta}\rvertitalic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT < | italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT |, there is a further saddle-node bifurcation SN1b bordering the S S-region.

\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/kappa-% pos/phase-plot-eta1-m0p6-eta2-m0p4-kappa-1p} \put(85.0,86.0){\ref{asym-pos-I})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/kappa-% pos/phase-plot-eta1-m0p6-eta2-m0p4-kappa-1p1} \put(30.0,86.0){\ref{asym-pos-II})} \end{overpic}
Figure 8: Phase portraits for the non-adaptive system with a=0𝑎0a=0italic_a = 0, ησ=0.5subscript𝜂𝜎0.5\eta_{\sigma}=-0.5italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = - 0.5 and ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1 with κ=1,1.1𝜅11.1\kappa=1,1.1italic_κ = 1 , 1.1, corresponding to the parameter points LABEL:asym-pos-I and LABEL:asym-pos-II in Fig. 5, respectively. (Nullclines are shown as gray dashed lines, stable/unstable/saddle fixed points are shown as filled/empty/half-filled circles.)

Crossing SN1b from inside the S S-region by increasing κ𝜅\kappaitalic_κ, 4 fixed points (stable/unstable node, 2 saddles) appear in a saddle-node bifurcation on θ1=0subscript𝜃10\theta_{1}=0italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0, see Fig. 8 panels LABEL:asym-pos-I) and LABEL:asym-pos-II).

III.3.2 Mode-locking

The S S region (orange shading in Fig. 5) features mode-locking regions, where the spiking activity of the two neurons locks into ratios of fixed frequency. These mode-locking regions are defined by a constant average ratio, R𝑅Ritalic_R, of spikes n1(t)subscript𝑛1𝑡n_{1}(t)italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) and n2(t)subscript𝑛2𝑡n_{2}(t)italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) occurring in neurons 1 and 2 over a time period t𝑡titalic_t. Formally, we express this ratio as

R=limtn1(t)n2(t)=limt0tδ(θ1(s)π)\dls0tδ(θ2(s)π)\dls,𝑅subscript𝑡subscript𝑛1𝑡subscript𝑛2𝑡subscript𝑡superscriptsubscript0𝑡𝛿subscript𝜃1𝑠𝜋\dl𝑠superscriptsubscript0𝑡𝛿subscript𝜃2𝑠𝜋\dl𝑠R=\lim_{t\to\infty}\frac{n_{1}(t)}{n_{2}(t)}=\lim_{t\to\infty}\frac{\int_{0}^{% t}\delta(\theta_{1}(s)-\pi)\dl s}{\int_{0}^{t}\delta(\theta_{2}(s)-\pi)\dl s},italic_R = roman_lim start_POSTSUBSCRIPT italic_t → ∞ end_POSTSUBSCRIPT divide start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) end_ARG = roman_lim start_POSTSUBSCRIPT italic_t → ∞ end_POSTSUBSCRIPT divide start_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_δ ( italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_s ) - italic_π ) italic_s end_ARG start_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_δ ( italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_s ) - italic_π ) italic_s end_ARG , (16)

where δ𝛿\deltaitalic_δ is the Dirac distribution on the circle 𝕋𝕋\mathbb{T}blackboard_T. This ratio R𝑅Ritalic_R equals the inverse fraction of the periods Tksubscript𝑇𝑘T_{k}italic_T start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT of the two nodes. Thus, orbits can only be periodic (and possibly be locked with the corresponding frequency mode) whenever R𝑅Ritalic_R is a rational number, R=p:q:𝑅𝑝𝑞R=p:qitalic_R = italic_p : italic_q with p,q𝑝𝑞p,q\in\mathbb{N}italic_p , italic_q ∈ blackboard_N.

\begin{overpic}[width=433.62pt]{figures/nonadaptive-asym/stability-diagram-% ratios} \put(96.0,4.0){\Large$\eta_{\sigma}$} \put(4.0,96.0){\Large$\kappa$} \put(28.0,58.0){s} \put(30.0,63.0){{\color[rgb]{1,1,1}\definecolor[named]{pgfstrokecolor}{rgb}{% 1,1,1}\pgfsys@color@gray@stroke{1}\pgfsys@color@gray@fill{1}t}} \put(30.0,66.0){{\color[rgb]{1,1,1}\definecolor[named]{pgfstrokecolor}{rgb}{% 1,1,1}\pgfsys@color@gray@stroke{1}\pgfsys@color@gray@fill{1}u}} \put(22.0,35.0){\Large Q Q } \put(70.0,65.0){\Large S S } \put(22.0,83.0){\large Q Q } \put(70.0,25.0){\large Q S } \put(56.0,15.0){\large 2Q Q } \end{overpic}
Figure 9: Stability diagram for the non-adaptive system Eq. 6 with a=0𝑎0a=0italic_a = 0, ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1. The S S region exhibits mode-locking for the spiking frequency. We only highlighted a selection of the frequency ratios, R=n/(n+1),nformulae-sequence𝑅𝑛𝑛1𝑛R=n/(n+1),\,n\in\mathbb{N}italic_R = italic_n / ( italic_n + 1 ) , italic_n ∈ blackboard_N. For a more complete view on the underlying structure, see the staircase shown in Fig. 11. Furthermore the ratio R𝑅Ritalic_R was estimated numerically and cannot resolve every mode-locking ratio. Phase portraits for the points LABEL:asym-locking-I-LABEL:asym-locking-III are shown in Fig. 10.

The mode-locking regions form Arnol’d tongues that emanate from the limit of zero coupling with κ=0𝜅0\kappa=0italic_κ = 0. The period of uncoupled neurons k=1,2𝑘12k=1,2italic_k = 1 , 2 with κ=0𝜅0\kappa=0italic_κ = 0, Tk=π/ηksubscript𝑇𝑘𝜋subscript𝜂𝑘T_{k}=\pi/\sqrt{\eta_{k}}italic_T start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_π / square-root start_ARG italic_η start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG , is easily obtained analytically [40], and we obtain

R=T2T1=η1η2=ησηδησ+ηδ.𝑅subscript𝑇2subscript𝑇1subscript𝜂1subscript𝜂2subscript𝜂𝜎subscript𝜂𝛿subscript𝜂𝜎subscript𝜂𝛿R=\frac{T_{2}}{T_{1}}=\frac{\sqrt{\eta_{1}}}{\sqrt{\eta_{2}}}=\frac{\sqrt{\eta% _{\sigma}-\eta_{\delta}}}{\sqrt{\eta_{\sigma}+\eta_{\delta}}}.italic_R = divide start_ARG italic_T start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG = divide start_ARG square-root start_ARG italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG end_ARG start_ARG square-root start_ARG italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG end_ARG = divide start_ARG square-root start_ARG italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT - italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT end_ARG end_ARG start_ARG square-root start_ARG italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT + italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT end_ARG end_ARG . (17)

Thus, Arnol’d tongues have their origin at κ=0𝜅0\kappa=0italic_κ = 0 with excitability values η1subscript𝜂1\eta_{1}italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and η2subscript𝜂2\eta_{2}italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, consistent with the condition R𝑅R\in\mathbb{Q}italic_R ∈ blackboard_Q according to (17).

However, to classify the neuron’s spiking activity by their spiking ratio R𝑅Ritalic_R for non-zero coupling (κ0𝜅0\kappa\neq 0italic_κ ≠ 0), we need to approximate the spiking ratios by numerically integrating trajectories (limit cycles) and measuring the ratio n1(T):n2(T):subscript𝑛1𝑇subscript𝑛2𝑇n_{1}(T):n_{2}(T)italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_T ) : italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_T ) for a large time T𝑇Titalic_T. To detect the mode-locking regions shown in Fig. 9, we estimated spiking ratios that lie within a tolerance of 0.003. A variety of ratios occur, but we specifically chose to only highlight ratios of the form n:(n+1):𝑛𝑛1n:(n+1)italic_n : ( italic_n + 1 ) where n𝑛n\in\mathbb{N}italic_n ∈ blackboard_N, since those ratios correspond to the widest regions and, thus, better reveal the parameter dependence in the stability diagram.

For negative and smaller positive values of ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, the n:(n+1):𝑛𝑛1n:(n+1)italic_n : ( italic_n + 1 ) regions cover almost all of the parameter-space, while they almost abruptly become much thinner around a certain small positive value of ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT. In Fig. 10, we show the phase portraits for the modes 00, 1:2:121:21 : 2 and 2:3:232:32 : 3.

\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/% locking/phase-plot-eta1-m0p6-eta2-m0p4-kappa-0p55} \put(5.0,86.0){\ref{asym-locking-I})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/% locking/phase-plot-eta1-m0p6-eta2-m0p4-kappa-0p65} \put(5.0,86.0){\ref{asym-locking-II})} \end{overpic}
\begin{overpic}[width=130.08731pt]{figures/nonadaptive-asym/phase-plots/% locking/phase-plot-eta1-m0p6-eta2-m0p4-kappa-0p75} \put(5.0,86.0){\ref{asym-locking-III})} \end{overpic}
Figure 10: Phase portraits for the non-adaptive system with a=0𝑎0a=0italic_a = 0, ησ=0.5subscript𝜂𝜎0.5\eta_{\sigma}=-0.5italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = - 0.5 and ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1 with κ=0.55,0.65,0.75𝜅0.550.650.75\kappa=0.55,0.65,0.75italic_κ = 0.55 , 0.65 , 0.75, corresponding to the parameter points LABEL:asym-locking-I-LABEL:asym-locking-III in Fig. 9, respectively. (Nullclines are shown as gray dashed lines, stable/unstable/saddle fixed points are shown as filled/empty/half-filled circles.)

Note that the mode with 0:1=0:0100:1=00 : 1 = 0 corresponds to the Q S state we already discussed further above.

\begin{overpic}[width=433.62pt]{figures/nonadaptive-asym/bifurcation-diagram-% proportion-eta0-composite} \put(5.0,63.0){\Large$R$} \put(94.0,4.0){\Large$\kappa$} \put(-1.0,59.0){$1:1$} \put(-1.0,46.5){$3:4$} \put(-1.0,41.7){$2:3$} \put(-1.0,33.0){$1:2$} \put(4.0,6.4){$0$} \put(6.5,0.0){$0$} \put(48.0,0.0){$0.5$} \put(92.0,0.0){$1$} \end{overpic}
Figure 11: Numerical estimation of frequency ratios of spikes R𝑅Ritalic_R for varying κ𝜅\kappaitalic_κ with ησ=0subscript𝜂𝜎0\eta_{\sigma}=0italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = 0. Plateaus of frequency-locking corresponding to the colored regions in Fig. 9 can be seen, and form a devil’s staircase.

Measuring the spike ratio with ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT fixed while increasing the coupling strength κ𝜅\kappaitalic_κ, Fig. 11 reveals a Devil’s staircase formed by the mode-locking regions. The n:(n+1):𝑛𝑛1n:(n+1)italic_n : ( italic_n + 1 ) modes are visible as steps in the staircase, but we also observe many other locked modes which appear smaller in size.

III.3.3 Comparison of dynamics occurring for identical versus non-identical neurons

While the non-adaptive system features three control parameters (ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, ηδsubscript𝜂𝛿\eta_{\delta}italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT and κ𝜅\kappaitalic_κ), the special case of identical excitabilities is represented by the parameter plane with ηδ=0subscript𝜂𝛿0\eta_{\delta}=0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0. We have seen that a variety of dynamic structures observed for that plane, including librations around center points and bi-stability, disappear in the non-identical case. These structures are not structurally robust towards perturbations such as breaking the system symmetry of ηδ=0subscript𝜂𝛿0\eta_{\delta}=0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0, and the only structures that are left are fixed points, limit cycles and saddle-node bifurcations. Notably, mode-locking regions are entirely absent for identical neurons. Although, in some sense one might be tempted to say that the spiking orbits in the S S region are locked; but this interpretation is problematic since the center orbits are not attractive and thus in reality provide no locking mechanism.

IV Analysis for adaptive coupling (a>0𝑎0a>0italic_a > 0)

We now consider the dynamics for adaptive coupling with a>0𝑎0a>0italic_a > 0. For simplicity we set the baseline coupling b=0𝑏0b=0italic_b = 0 and the time scale of adaptation to ε=0.01𝜀0.01\varepsilon=0.01italic_ε = 0.01. Since the dynamics for identical excitabilities in the non-adaptive case turned out to be structurally non-robust (Section III.3.3), we focus on non-identical excitabilities with ηδ=(η2η1)/2=0.1subscript𝜂𝛿subscript𝜂2subscript𝜂120.1\eta_{\delta}=(\eta_{2}-\eta_{1})/2=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = ( italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) / 2 = 0.1. The remaining control parameters are then the average excitability ησ=(η1+η2)/2subscript𝜂𝜎subscript𝜂1subscript𝜂22\eta_{\sigma}=(\eta_{1}+\eta_{2})/2italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = ( italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) / 2 and the adaptivity parameter a𝑎aitalic_a.

IV.1 Regimes of quiescent neurons (Q Q) and of one spiking neuron (Q S)

We first focus on the two states of two quiescent neurons (Q Q) and of one single spiking neuron (Q S). The corresponding stability regions are shown in Fig. 12 with distinct gray shadings.

\begin{overpic}[width=433.62pt]{figures/coev-asym/stability-diagram} \put(96.0,5.0){\Large$\eta_{\sigma}$} \put(4.0,63.0){\Large$a$} \put(50.0,7.0){BT} \put(21.0,30.0){LPC\textsubscript{1} } \put(59.0,30.0){LPC\textsubscript{2} } \put(18.0,17.0){\large Q Q } \put(49.0,17.0){\large Q S } \put(83.5,56.5){SN} \put(83.5,50.25){Hopf} \put(83.5,44.0){LPC} \end{overpic}
Figure 12: Stability diagram for the model with adaptation. Parameters are ε=0.01𝜀0.01\varepsilon=0.01italic_ε = 0.01, b=0𝑏0b=0italic_b = 0, and ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1.

The non-adaptive case with a=0𝑎0a=0italic_a = 0 corresponds in the asymptotic time limit to the uncoupled system (κ=0𝜅0\kappa=0italic_κ = 0) studied in Section III.3; namely, the system undergoes the transition called SN1 from Q Q to Q S in a SNIC bifurcation at ησ=|ηδ|=0.1subscript𝜂𝜎subscript𝜂𝛿0.1\eta_{\sigma}=-\lvert\eta_{\delta}\rvert=-0.1italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = - | italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT | = - 0.1. However, note that two changes occur at the SNIC bifurcation: firstly, increasing ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, two fixed points coalesce and disappear at the SNIC; secondly, while the associated fixed points previously lived on an invariant and stable curve in phase space, this curve transforms into a limit cycle.

For non-zero adaptivity a0𝑎0a\neq 0italic_a ≠ 0, the SNIC bifurcation disappears and instead a region of bi-stability between Q Q and Q S states is formed, as shown in the bifurcation diagram in Fig. 13 for a small adaptivity value a=0.01𝑎0.01a=0.01italic_a = 0.01.

\begin{overpic}[width=433.62pt]{figures/coev-asym/bifurcation-diagram-a-0p01-% quasicont} \put(96.0,6.0){\Large$\eta_{\sigma}$} \put(3.0,63.0){\Large$\kappa$} \put(84.5,36.8){$\max_{t}(\kappa)$} \put(84.5,29.2){$\min_{t}(\kappa)$} \put(10.0,51.0){\large Q Q } \put(54.0,34.0){\large Q S } \end{overpic}
Figure 13: Bifurcation diagram displaying minima and maxima detected in time, i.e., mint(κ)subscript𝑡𝜅\min_{t}(\kappa)roman_min start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_κ ) and maxt(κ)subscript𝑡𝜅\max_{t}(\kappa)roman_max start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_κ ) for a=0.01𝑎0.01a=0.01italic_a = 0.01 (ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1, ε=0.01𝜀0.01\varepsilon=0.01italic_ε = 0.01 and b=0𝑏0b=0italic_b = 0) when performing quasi-continuation from both the left and right. The maxima are red lines and the minima blue lines.

The two topological changes in phase space now occur in different bifurcations; the fixed points coalesce in a saddle-node bifurcation (SN, blue line in Figs. 12 and 14), while the limit cycle is born from a limit point of cycles (LPC1, purple line in Figs. 12 and 14).

\begin{overpic}[width=433.62pt]{figures/coev-asym/stability-diagram-small-a} \put(96.0,4.0){\Large$\eta_{\sigma}$} \put(5.0,63.0){\Large$a$} \put(37.0,47.5){BT} \put(34.0,24.0){BT} \put(55.0,15.0){\large Q Q } \put(75.0,25.0){\large Q S } \put(81.0,54.5){SN} \put(81.0,48.25){Hopf} \put(81.0,42.0){LPC} \end{overpic}
Figure 14: Stability diagram of attractors of the co-evolutionary system for small a𝑎aitalic_a, ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1.

For a0.0195𝑎0.0195a\approx 0.0195italic_a ≈ 0.0195, a Bogdanov–Takens bifurcation BT (see Fig. 14) occurs on the saddle-node curve SN. Above the BT point, the quiescent (Q Q) state loses stability in a subcritical Hopf bifurcation H (green line in Figs. 12 and 14); the resulting unstable Q Q state is annihilated in SN. Above the BT, the bi-stable region between Q Q and Q S is therefore bounded by the LPC1 and H curves. Thus, for a𝑎aitalic_a-values above BT, the saddle-node bifurcation SN no longer influences the stability regions, as it merely annihilates a fixed rendered unstable by the Hopf bifurcation (The associated saddle-node curves undergo two cusp bifurcations which we for simplicity did not highlight.). Note that — even for larger a𝑎aitalic_a — the bi-stable region is very thin and is therefore shown enlarged in the inset in Fig. 14.

Lastly, we observe that the single neuron spiking state Q S is annihilated for large ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT in another limit point of cycles (LPC2, Fig. 5). Note that this boundary only denotes where the Q S ends; the S S states are sometimes bi-stable with the Q S states. We discuss the nature and parameter range for S S states in the following section.

IV.2 Regime of spiking neurons (S S)

We now turn our attention to the regime in which two neurons are spiking (S S). Since the S S states display mode-locking (Section III.3), the interesting question arises of how these modes are affected by adaptive dynamics when a>0𝑎0a>0italic_a > 0. As for the non-adaptive case, we are concerned with the spiking ratio R𝑅Ritalic_R of S S states, in particular periodic oscillations with R𝑅R\in\mathbb{Q}italic_R ∈ blackboard_Q, for which mode-locking is present.

As we vary the adaptivity with (a>0𝑎0a>0italic_a > 0), the Arnol’d tongues associated to each mode-locking regions become apparent, see Fig. 15(a).

\begin{overpic}[width=433.62pt]{figures/coev-asym/stability-diagram-ratio-many% } \put(-1.0,60.0){a)} \put(96.0,5.0){\Large$\eta_{\sigma}$} \put(4.0,63.0){\Large$a$} \put(15.0,21.4){$n$:$3n$} \put(15.0,15.2){$n$:$2n$} \put(15.0,9.0){$2n$:$3n$} \put(9.0,50.0){LPC\textsubscript{1} } \put(42.0,58.0){LPC\textsubscript{2} } \put(30.0,10.0){\large Q Q } \put(54.0,10.0){\large Q S } \end{overpic}
\begin{overpic}[width=433.62pt]{figures/coev-asym/stability-diagram-ratio-1t2} \put(-1.0,60.0){b)} \put(5.0,42.8){ \leavevmode\hbox to222.33pt{\vbox to0.4pt{\pgfpicture% \makeatletter\hbox{\hskip 0.2pt\lower-0.2pt\hbox to0.0pt{\pgfsys@beginscope% \pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}% {0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to% 0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {{}{}}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setdash{3.0pt,3.0pt}{0.0pt}% \pgfsys@invoke{ }\color[rgb]{.5,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}% {.5,.5,.5}\pgfsys@color@gray@stroke{.5}\pgfsys@invoke{ }% \pgfsys@color@gray@fill{.5}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{% rgb}{.5,.5,.5}{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@lineto{221.9315pt}{0.0pt}% \pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}% \pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}% \lxSVG@closescope\endpgfpicture}}} \put(5.0,111.7){ \leavevmode\hbox to222.33pt{\vbox to0.4pt{\pgfpicture% \makeatletter\hbox{\hskip 0.2pt\lower-0.2pt\hbox to0.0pt{\pgfsys@beginscope% \pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}% {0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to% 0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{{}}{} {{}{}}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setdash{3.0pt,3.0pt}{0.0pt}% \pgfsys@invoke{ }\color[rgb]{.5,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}% {.5,.5,.5}\pgfsys@color@gray@stroke{.5}\pgfsys@invoke{ }% \pgfsys@color@gray@fill{.5}\pgfsys@invoke{ }\definecolor[named]{pgffillcolor}{% rgb}{.5,.5,.5}{}\pgfsys@moveto{0.0pt}{0.0pt}\pgfsys@lineto{221.9315pt}{0.0pt}% \pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}% \pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}% \lxSVG@closescope\endpgfpicture}}} \put(96.0,5.0){\Large$\eta_{\sigma}$} \put(4.0,63.0){\Large$a$} \put(16.0,21.4){LPC} \put(16.0,15.2){PD} \put(16.0,9.0){$n$:$2n$} \put(9.0,50.0){LPC\textsubscript{1} } \put(42.0,58.0){LPC\textsubscript{2} } \put(30.0,10.0){\large Q Q } \put(54.0,10.0){\large Q S } \end{overpic}
Figure 15: Stability diagrams for the co-evolutionary adaptive model with ε=0.01𝜀0.01\varepsilon=0.01italic_ε = 0.01, b=0𝑏0b=0italic_b = 0, and ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1. Different aspects of interest are highlighted in panels (a) and (b). (a) S S states are limit cycles which mode-lock with various ratios R𝑅Ritalic_R. Regions corresponding to ratios 1:3:131:31 : 3, 1:2:121:21 : 2 and 2:3:232:32 : 3 are highlighted. Selected bifurcations curves shown in Fig. 12 are displayed for guidance (black lines). (b) Bifurcations for the mode-locking region with R=n:2n=1:2:𝑅𝑛2𝑛1:2R=n:2n=1:2italic_R = italic_n : 2 italic_n = 1 : 2 (highlighted in yellow). Bifurcations related to limit cycles with this ratio R𝑅Ritalic_R are shown as colored curves: limit point of cycles (LPC) bounding the n:2n:𝑛2𝑛n:2nitalic_n : 2 italic_n limit cycles are shown in purple; associated period doubling (PD) bifurcations are shown in light blue. The dashed gray lines indicate the parameter sweep for the diagrams in Fig. 16.

Interestingly, we observe that the tongues are widening as the adaptivity is increased—this indicates a relationship between the adaptivity a𝑎aitalic_a, and the coupling strength, κ𝜅\kappaitalic_κ, which typically is associated with the widening off mode-locking regions and their associated Arnol’d tongues. Furthermore, multi-stability is possible between the different S S mode-locking regimes, as well as between the Q S state and certain S S modes. The data in Fig. 15(a) was generated by estimating R𝑅Ritalic_R numerically for initial conditions on a grid with even spacing (10, 10, and 7 grid points along the coordinates θ1subscript𝜃1\theta_{1}italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, θ2subscript𝜃2\theta_{2}italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, κ𝜅\kappaitalic_κ, respectively), and therefore the regions identified are not exhaustive.

To further understand the Arnol’d tongues, we focus on the mode with R=1:2:𝑅12R=1:2italic_R = 1 : 2. Numerical continuation with MatCont reveals that the tongue is bounded on both sides by LPCs (purple lines in Fig. 15(b)). For larger adaptivity a𝑎aitalic_a, numerical continuation also shows that the S S limit cycle may undergo a series of period doubling bifurcations (light blue lines in Fig. 15(b)). Shown are the first three bifurcations of a series of period doubling bifurcations, forming a cascade that ultimately leads to chaos as decrease ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, as seen in Fig. 16(b). Throughout this cascade and far into the chaotic region, the ratio R𝑅Ritalic_R remains constant (see Fig. 16(a); however, at some critical value of ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT (indicated by an arrow), there appears to be a crisis of (chaotic) attractors, which possibly even are associated to other resonant modes. As a consequence, the fixed spiking ratio R𝑅Ritalic_R is compromised and sharp mode-locking regions appear only in small windows of intermittency. However, inspection Fig. 16(b) suggests that chaotic regions between intermittencies feature a blurred mode-locking ratios, i.e., there are no sharp mode-locking steps due to chaos. As we decrease ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT further, the chaotic motion subsides together with the spiking modes, i.e., the spiking ratio drops to R=0𝑅0R=0italic_R = 0. We leave a more detailed analysis for the chaotic region and its associated bifurcations for a future study.

\begin{overpic}[width=433.62pt]{figures/coev-asym/quasicontinuation-right2left% } \put(-1.0,94.0){a)} \put(-1.0,45.0){b)} \put(60.0,94.0){$\downarrow$} \put(60.0,34.0){$\downarrow$} \put(6.0,48.0){\large$\kappa$} \put(95.0,5.0){\Large$\eta_{\sigma}$} \put(6.0,95.0){\large$R$} \put(95.0,52.0){\Large$\eta_{\sigma}$} \put(83.5,41.0){$\max_{t}(\kappa)$} \put(83.5,36.7){$\min_{t}(\kappa)$} \end{overpic}
Figure 16: Spiking ratios R𝑅Ritalic_R (a) and the bifurcation diagram (b) for varying ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT with adaptivity a=1𝑎1a=1italic_a = 1 (gray dashed line in Fig. 15) and fixed parameters ηδ=0.1subscript𝜂𝛿0.1\eta_{\delta}=0.1italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0.1, ε=0.01𝜀0.01\varepsilon=0.01italic_ε = 0.01 and b=0𝑏0b=0italic_b = 0. The bifurcation diagram in panel (b) reports local minima and maxima in time, mint(κ)subscript𝑡𝜅\min_{t}(\kappa)roman_min start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_κ ) and maxt(κ)subscript𝑡𝜅\max_{t}(\kappa)roman_max start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_κ ). Both diagrams are produced via quasi-continuation from right to left.

V Discussion

We have investigated the dynamics of a pair of Theta neurons with non-autaptic coupling, i.e., there is only coupling between the two neurons without self-interaction. The first part of the study concerned the case of stationary coupling strengths, and the second part includes an ad-hoc adaptation rule for the coupling strength [34, 36], which effectively leads to the emergence of co-evolutionary network dynamics [29, 30].

The non-adaptive case, a=0𝑎0a=0italic_a = 0, leads already to a surprisingly wide range of dynamical phenomena, when comparing (6) to the Kuramoto model with N=2𝑁2N=2italic_N = 2 oscillators [36]. The principal reason for this is that the Kuramoto model inherently exhibits a phase shift invariance; thus, a two oscillator model can be reduced to a one dimensional system

ϕ˙˙italic-ϕ\displaystyle\dot{\phi}over˙ start_ARG italic_ϕ end_ARG =ω+Ksinϕ,absent𝜔𝐾italic-ϕ\displaystyle=\omega+K\sin{\phi},= italic_ω + italic_K roman_sin italic_ϕ , (18)

describing the evolution of the phase difference ϕ(t)italic-ϕ𝑡\phi(t)italic_ϕ ( italic_t ) and frequency difference ω𝜔\omegaitalic_ω of the two oscillator nodes, respectively. Thus, the system exhibits only two simple behaviors, drift (K<|ω|𝐾𝜔K<|\omega|italic_K < | italic_ω |) and mode-locking (K>|ω|𝐾𝜔K>|\omega|italic_K > | italic_ω |), where only a 1:1 frequency ratio between oscillators is possible. Due to the excitable nature of the Theta neuron, they do not possess a phase shift symmetry; consequently, the system cannot be reduced and is genuinely two dimensional.

For the case of identical neurons, ηδ=0subscript𝜂𝛿0\eta_{\delta}=0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0, attractive states of quiescence (Q Q), librations (Q\ellroman_ℓ Q\ellroman_ℓ) and spiking in both neurons (S S) are observed. These states transition through various bifurcation scenarios of the saddle-node type and display regions of bi-stability. The identical nature of the neurons implies symmetric dynamics, i.e., the two neurons display identical behavior. It is useful to see the parameter plane defined by ηδ=0subscript𝜂𝛿0\eta_{\delta}=0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT = 0 as the limiting collection of degenerate bifurcations for the more general case of non-identical neurons. Indeed, librations (around centers) and the spiking rotations are foliated in phase space; however, these dynamics are not structurally robust.

For non-identical neurons, ηδ0subscript𝜂𝛿0\eta_{\delta}\neq 0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT ≠ 0, centers transform into spirals and the librations are thus not any longer present; furthermore, the broken symmetry implies that neurons adhere to distinct dynamics — e.g., we observe states where only the one neuron with larger excitability) is spiking (Q S). In contrast to the identical case, bi-stable configurations are entirely absent. In the regime where both neurons are excited and spiking (S S), the neuron with higher excitability appears to always spike at a higher rate. While the Kuramoto model (18) only allows for a single locking mode, non-identical excitabilities break the symmetry of the system resulting in a fan of locking regions with varying spiking ratios R𝑅R\in\mathbb{Q}italic_R ∈ blackboard_Q. As characteristic time scales of the individual phases are varied, either via κ𝜅\kappaitalic_κ or ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, non-identical excitabilities allow spiking neurons to resonate in modes corresponding to various frequency ratios. The resulting Arnol’d tongues form a devil’s stair case, and, as we continuously increase the coupling strength, κ𝜅\kappaitalic_κ, the associated mode locking regions decrease their width, while the frequency ratio asymptotically converges towards R1𝑅1R\to 1italic_R → 1 — i.e., a locking state, which most closely resembles the frequency locked state of the Kuramoto model.

Several new dynamical phenomena/features appear as we allow the coupling to adapt with a>0𝑎0a>0italic_a > 0. The system is now three dimensional, thus inadvertently implying the possibility for richer dynamics. First, regions of bi-stability are observed between quiescent (Q Q) and single-neuron spiking (Q S). Second, the mode-locking regions gain in width as the adaptivity parameter a𝑎aitalic_a is increased. For sufficiently large adaptivity, Arnol’d tongues begin to overlap, even multiple of them, so that topologically several modes of limit-cycles can exist at the same time — in other words, the steps in the stair case begin to overlap. Thus, a very high multiplicity of mode-locking configurations is offered by the system to reside in. As a𝑎aitalic_a is further increased, oscillations in each locking region (with fixed R𝑅Ritalic_R) undergo a period doubling cascade to chaos with quasi-periodic orbits. In some parameter regions, chaotic attractors even appear to collide in more complicated bifurcation scenarios. As a result, the stair case appears to get ’blurred’ by chaotic dynamics (see Fig. 16).

It is interesting to compare the non-adaptive and the adaptive dynamics as observed in different oscillator models. As is the case in the present study, adaptive Kuramoto oscillators with N=2𝑁2N=2italic_N = 2 display a widening of the locking region with increasing adaptivity a𝑎aitalic_a, together with bi-stability of drift and locked oscillator dynamics. While the Kuramoto model only offers a single control parameter with the frequency difference ω𝜔\omegaitalic_ω, in the Theta neuron model we may chose from two parameters, the average and difference of excitabilities, ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT and ηδsubscript𝜂𝛿\eta_{\delta}italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT, respectively. To investigate the (possible) widening of locking regions, we chose ησsubscript𝜂𝜎\eta_{\sigma}italic_η start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT as the control parameter, since fixing the value ηδ0subscript𝜂𝛿0\eta_{\delta}\neq 0italic_η start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT ≠ 0 already guarantees the heterogeneity necessary to observe locking. In simple terms, the nature of adaptive coupling, leading to co-evolutionary network dynamics, leaves more ’space’ for mode-locking: the coupling may adapt in a way that accommodates better for heterogeneity, i.e., non-identical frequencies in the Kuramoto model and excitabilities in the Theta neuron model, respectively. Consequently, the coupling adapts while maintaining stability of the locked mode and thus ’stretches’ the width of the locking region. This effect is seen in both models. Finally, it is interesting to note — in a certain sense — a structural similarity between the adaptation rule in Eq. (4) and the phase dynamics of the Kuramoto model in Eq. (18): in either of the two models, locked states imply a fixed (but adaptive) relationship between phase difference and coupling strength.

An open challenge presents itself with the question as of what dynamics result from the adaptive Theta neuron model (6) for large N𝑁Nitalic_N. A recent study performed a careful bifurcation analysis for the adaptive Kuramoto model [37] under the two assumptions that ϵitalic-ϵ\epsilonitalic_ϵ be small and that N𝑁N\to\inftyitalic_N → ∞. Using an (approximative) dimensional reduction of the system by averaging the rows of the associated coupling matrix, the subsequent analysis was based on studying a self-consistency equation. A similar approach for the system considered here seems momentarily unreachable, since the formulation of a self-consistency equation for Theta-neurons first would need to be shown to be feasible. However, the approximate dimensional reduction via row-averaging the coupling matrix is certainly feasible and may yield interesting insights into the dynamics. Generally speaking, a formidable challenge remains to develop other means for the dimensional reduction of adaptive oscillator networks and co-evolutionary network dynamics.

A difficulty in studying systems with N>2𝑁2N>2italic_N > 2 oscillators is to classify the mode-locking states in a meaningful way — indeed, large oscillator systems are known to exhibit a multitude of frequency clusters [34]. Following the approach of this study, i.e., using the adaptivity parameter a𝑎aitalic_a to perturb a way from non-adaptive case, it makes sense to consider methods from the analysis of stationary networks with non-uniform coupling corresponding to the non-adaptive limit where a=0𝑎0a=0italic_a = 0. Indeed, the adaptive nature of the coupling may lead to simple stationary structures corresponding to submatrices (blocks) of uniform coupling within the coupling matrix, similar to models of coupled oscillators arranged in M𝑀Mitalic_M populations [27]. Indeed, such an approach was recently used [49] to study the effect of time-dependent or adaptive coupling in models with subpopulations, though with (non-adaptive) fixed size. Other authors have used methods from representation theory to determine frequency clusters in oscillator systems with stationary coupling [50]. It will be interesting to see if such methods may be adopted to adaptive network models.

Finally, to make further progress in the mathematical theory of co-evolutionary network dynamics there is a certain need to clarify the relationship between the ad-hoc adaptation rule used in this study (which can be regarded as a truncated Fourier series) and more realistic plasticity models used in computational neuroscience such as spike-time-dependent plasticity [51, 38]. While the Theta neuron model lends itself for a simplistic formulation of the adaptation rule on the phase domain, while standard plasticity models for the equivalent quadratic integrate and fire model are formulated on the time domain, i.e., spikes are integrated on the time rather than the phase domain. While some work has been formulated to relate the two dynamic representations [52], a rigorous mathematical treatise appears to be missing. Alternatively, it could also be worthwhile to compare the dynamics resulting from the model studied here with the adaptive corollary of a QIF model in the large N𝑁Nitalic_N limit. Finally, the adaptation rule in (4) formulates a symmetric response in terms of the phase difference; synaptic regulation more naturally adheres to an asymmetric response, respecting the different roles of pre- and post-synaptic dynamics. A future study could address the effect of such a model.

Acknowledgements.
We wish to acknowledge R. Cestnik and C. Bick for helpful comments and discussions. We gratefully acknowledge financial support from the Royal Swedish Physiographic Society of Lund.

Appendix A Solutions of a quadratic in an interval

Here we present a derivation of the number of (fixed point) solutions to Eq. 13. We begin with a general lemma on solutions of a quadratic in an interval:

Lemma A.1.

The polynomial

p(x)=ax2+bx+1𝑝𝑥𝑎superscript𝑥2𝑏𝑥1p(x)=ax^{2}+bx+1italic_p ( italic_x ) = italic_a italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_b italic_x + 1 (19)

has the following number and locations of solutions in the interval x[0,1]𝑥01x\in[0,1]italic_x ∈ [ 0 , 1 ]:

Criteria Solutions
a+b+1<0𝑎𝑏10a+b+1<0italic_a + italic_b + 1 < 0 1 solution x(0,1)𝑥01x\in(0,1)italic_x ∈ ( 0 , 1 )
a+b+1=0𝑎𝑏10a+b+1=0italic_a + italic_b + 1 = 0, a1𝑎1a\leq 1italic_a ≤ 1 1 solution x=1𝑥1x=1italic_x = 1
a>1𝑎1a>1italic_a > 1 2 sols. x1(0,1)subscript𝑥101x_{1}\in(0,1)italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∈ ( 0 , 1 ), x2=1subscript𝑥21x_{2}=1italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1
a+b+1>0𝑎𝑏10a+b+1>0italic_a + italic_b + 1 > 0, b2𝑏2b\geq-2italic_b ≥ - 2 No solution
b<2𝑏2b<-2italic_b < - 2, 4a>b24𝑎superscript𝑏24a>b^{2}4 italic_a > italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT No solution
4a=b24𝑎superscript𝑏24a=b^{2}4 italic_a = italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 1 solution x(0,1)𝑥01x\in(0,1)italic_x ∈ ( 0 , 1 )
4a<b24𝑎superscript𝑏24a<b^{2}4 italic_a < italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 2 solutions x1,x2(0,1)subscript𝑥1subscript𝑥201x_{1},x_{2}\in(0,1)italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ∈ ( 0 , 1 )

Note that all a,b𝑎𝑏a,b\in\mathbb{R}italic_a , italic_b ∈ blackboard_R are covered in the table above.

Proof.

The proof goes through a set of assumptions on a𝑎aitalic_a and b𝑏bitalic_b that cover all a,b𝑎𝑏a,b\in\mathbb{R}italic_a , italic_b ∈ blackboard_R. Where the assumptions does not map one-to-one with the criteria in the table, an explanation is given.

Assume that p(1)=a+b+1<0𝑝1𝑎𝑏10p(1)=a+b+1<0italic_p ( 1 ) = italic_a + italic_b + 1 < 0.

Since p(0)=1>0𝑝010p(0)=1>0italic_p ( 0 ) = 1 > 0, by the intermediate value theorem, there must be at least one solution in (0,1)01(0,1)( 0 , 1 ). Because p𝑝pitalic_p is quadratic, there can not be two roots in the interval and while the function has different signs on the boundary. x=1𝑥1x=1italic_x = 1 is not a root. Thus there is only one root x(0,1)𝑥01x\in(0,1)italic_x ∈ ( 0 , 1 ).

Assume that p(1)=a+b+1=0𝑝1𝑎𝑏10p(1)=a+b+1=0italic_p ( 1 ) = italic_a + italic_b + 1 = 0 and a0𝑎0a\leq 0italic_a ≤ 0.

x=1𝑥1x=1italic_x = 1 is a root. Also,

p(x)>0,x(0,1).formulae-sequence𝑝𝑥0for-all𝑥01p(x)>0,\,\forall x\in(0,1).italic_p ( italic_x ) > 0 , ∀ italic_x ∈ ( 0 , 1 ) . (20)

x=1𝑥1x=1italic_x = 1 is thus the only root.

Assume that p(1)=a+b+1=0𝑝1𝑎𝑏10p(1)=a+b+1=0italic_p ( 1 ) = italic_a + italic_b + 1 = 0 and 0<a10𝑎10<a\leq 10 < italic_a ≤ 1.

x=1𝑥1x=1italic_x = 1 is a root and p𝑝pitalic_p has a minimum in

xmin=b2a=1+a2a1,subscript𝑥min𝑏2𝑎1𝑎2𝑎1x_{\text{min}}=-\frac{b}{2a}=\frac{1+a}{2a}\geq 1,italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = - divide start_ARG italic_b end_ARG start_ARG 2 italic_a end_ARG = divide start_ARG 1 + italic_a end_ARG start_ARG 2 italic_a end_ARG ≥ 1 , (21)

so

p(x)>0,x(0,1).formulae-sequence𝑝𝑥0for-all𝑥01p(x)>0,\,\forall x\in(0,1).italic_p ( italic_x ) > 0 , ∀ italic_x ∈ ( 0 , 1 ) . (22)

Thus x=1𝑥1x=1italic_x = 1 is the only root.

Assume that p(1)=a+b+1=0𝑝1𝑎𝑏10p(1)=a+b+1=0italic_p ( 1 ) = italic_a + italic_b + 1 = 0 and 1<a1𝑎1<a1 < italic_a.

x2=1subscript𝑥21x_{2}=1italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 is a root and p𝑝pitalic_p has a minimum in

xmin=b2a=1+a2a.subscript𝑥min𝑏2𝑎1𝑎2𝑎x_{\text{min}}=-\frac{b}{2a}=\frac{1+a}{2a}.italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = - divide start_ARG italic_b end_ARG start_ARG 2 italic_a end_ARG = divide start_ARG 1 + italic_a end_ARG start_ARG 2 italic_a end_ARG . (23)

Since

0<xmin<10subscript𝑥min10<x_{\text{min}}<10 < italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT < 1 (24)

and

p(xmin)=(a1)24a<0,𝑝subscript𝑥minsuperscript𝑎124𝑎0p(x_{\text{min}})=-\frac{(a-1)^{2}}{4a}<0,italic_p ( italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT ) = - divide start_ARG ( italic_a - 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_a end_ARG < 0 , (25)

by the intermediate value theorem there is another root x1(0,xmin)(0,1)subscript𝑥10subscript𝑥min01x_{1}\in(0,x_{\text{min}})\subset(0,1)italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∈ ( 0 , italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT ) ⊂ ( 0 , 1 ).

Assume that p(1)=a+b+1>0𝑝1𝑎𝑏10p(1)=a+b+1>0italic_p ( 1 ) = italic_a + italic_b + 1 > 0 and a0𝑎0a\leq 0italic_a ≤ 0.

Then one root x1<0subscript𝑥10x_{1}<0italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 0 and the other root x2>1subscript𝑥21x_{2}>1italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 1 (if a=0𝑎0a=0italic_a = 0, only one of these roots exist). Thus there are no roots in [0,1]01[0,1][ 0 , 1 ]. Note that the assumption lies in only one region in the table, since b>a11>2𝑏𝑎112b>-a-1\geq-1>-2italic_b > - italic_a - 1 ≥ - 1 > - 2.

Assume that a+b+1>0𝑎𝑏10a+b+1>0italic_a + italic_b + 1 > 0, a>0𝑎0a>0italic_a > 0 and b0𝑏0b\geq 0italic_b ≥ 0.

The minimum

xmin=b2a0,subscript𝑥min𝑏2𝑎0x_{\text{min}}=-\frac{b}{2a}\leq 0,italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = - divide start_ARG italic_b end_ARG start_ARG 2 italic_a end_ARG ≤ 0 , (26)

and thus

p(x)>0,x[0,1].formulae-sequence𝑝𝑥0for-all𝑥01p(x)>0,\,\forall x\in[0,1].italic_p ( italic_x ) > 0 , ∀ italic_x ∈ [ 0 , 1 ] . (27)
Assume that p(1)=a+b+1>0𝑝1𝑎𝑏10p(1)=a+b+1>0italic_p ( 1 ) = italic_a + italic_b + 1 > 0, a>0𝑎0a>0italic_a > 0 and 2a<b<02𝑎𝑏0-2a<b<0- 2 italic_a < italic_b < 0.

The minimum

xmin=b2a(0,1).subscript𝑥min𝑏2𝑎01x_{\text{min}}=-\frac{b}{2a}\in(0,1).italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = - divide start_ARG italic_b end_ARG start_ARG 2 italic_a end_ARG ∈ ( 0 , 1 ) . (28)

Thus there will be no, one or two solutions in (0,1)01(0,1)( 0 , 1 ) when

p(xmin)𝑝subscript𝑥min\displaystyle p(x_{\text{min}})italic_p ( italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT ) =1b24a>04a>b2iffabsent1superscript𝑏24𝑎04𝑎superscript𝑏2\displaystyle=1-\frac{b^{2}}{4a}>0\iff 4a>b^{2}= 1 - divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_a end_ARG > 0 ⇔ 4 italic_a > italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (29)
p(xmin)𝑝subscript𝑥min\displaystyle p(x_{\text{min}})italic_p ( italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT ) =1b24a=04a=b2iffabsent1superscript𝑏24𝑎04𝑎superscript𝑏2\displaystyle=1-\frac{b^{2}}{4a}=0\iff 4a=b^{2}= 1 - divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_a end_ARG = 0 ⇔ 4 italic_a = italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (30)
p(xmin)𝑝subscript𝑥min\displaystyle p(x_{\text{min}})italic_p ( italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT ) =1b24a<04a<b2iffabsent1superscript𝑏24𝑎04𝑎superscript𝑏2\displaystyle=1-\frac{b^{2}}{4a}<0\iff 4a<b^{2}= 1 - divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_a end_ARG < 0 ⇔ 4 italic_a < italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (31)

respectively.

Assume that p(1)=a+b+1>0𝑝1𝑎𝑏10p(1)=a+b+1>0italic_p ( 1 ) = italic_a + italic_b + 1 > 0, a>0𝑎0a>0italic_a > 0 and b2a𝑏2𝑎b\leq-2aitalic_b ≤ - 2 italic_a.

The minimum

xmin=b2a1,subscript𝑥min𝑏2𝑎1x_{\text{min}}=-\frac{b}{2a}\geq 1,italic_x start_POSTSUBSCRIPT min end_POSTSUBSCRIPT = - divide start_ARG italic_b end_ARG start_ARG 2 italic_a end_ARG ≥ 1 , (32)

and thus

p(x)>0,x[0,1].formulae-sequence𝑝𝑥0for-all𝑥01p(x)>0,\,\forall x\in[0,1].italic_p ( italic_x ) > 0 , ∀ italic_x ∈ [ 0 , 1 ] . (33)

Note that 2b>2a2b22𝑏2𝑎2𝑏22b>-2a-2\geq b-22 italic_b > - 2 italic_a - 2 ≥ italic_b - 2 or equivalently b2𝑏2b\geq-2italic_b ≥ - 2. ∎

The coefficients in Eq. 13 are

a𝑎\displaystyle aitalic_a =2κν1ν2absent2𝜅subscript𝜈1subscript𝜈2\displaystyle=-2\kappa\frac{\nu_{1}}{\nu_{2}}= - 2 italic_κ divide start_ARG italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG (34a)
b𝑏\displaystyle bitalic_b =ν1absentsubscript𝜈1\displaystyle=\nu_{1}= italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (34b)

when |ν1||ν2|subscript𝜈1subscript𝜈2\lvert\nu_{1}\rvert\leq\lvert\nu_{2}\rvert| italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | ≤ | italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | and

a𝑎\displaystyle aitalic_a =2κν2ν1absent2𝜅subscript𝜈2subscript𝜈1\displaystyle=-2\kappa\frac{\nu_{2}}{\nu_{1}}= - 2 italic_κ divide start_ARG italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG (35a)
b𝑏\displaystyle bitalic_b =ν2absentsubscript𝜈2\displaystyle=\nu_{2}= italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (35b)

when |ν2||ν1|subscript𝜈2subscript𝜈1\lvert\nu_{2}\rvert\leq\lvert\nu_{1}\rvert| italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | ≤ | italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT |. We apply the lemma separately for the identical and non-identical case. In the case of identical neurons, we also have to consider the case ν1=ν2=η+2κ1=0subscript𝜈1subscript𝜈2𝜂2𝜅10\nu_{1}=\nu_{2}=\eta+2\kappa-1=0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_η + 2 italic_κ - 1 = 0 in order to cover the entire parameter space. The number of fixed points in different regions can be seen in Tables 1 and 2.

Criteria Num. fixed points
η<0𝜂0\eta<0italic_η < 0, η+2κ10𝜂2𝜅10\eta+2\kappa-1\neq 0italic_η + 2 italic_κ - 1 ≠ 0 4 fixed points
η+2κ1=0𝜂2𝜅10\eta+2\kappa-1=0italic_η + 2 italic_κ - 1 = 0 Continuum where 2κxy=12𝜅𝑥𝑦12\kappa xy=12 italic_κ italic_x italic_y = 1
η=0𝜂0\eta=0italic_η = 0, κ<12𝜅12\kappa<-\frac{1}{2}italic_κ < - divide start_ARG 1 end_ARG start_ARG 2 end_ARG 5 fixed points
κ12𝜅12\kappa\geq-\frac{1}{2}italic_κ ≥ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG 1 fixed point
η>0𝜂0\eta>0italic_η > 0, κ<(1+η)22𝜅superscript1𝜂22\kappa<-\frac{\left(1+\sqrt{\eta}\right)^{2}}{2}italic_κ < - divide start_ARG ( 1 + square-root start_ARG italic_η end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG 8 fixed points
κ=(1+η)22𝜅superscript1𝜂22\kappa=-\frac{\left(1+\sqrt{\eta}\right)^{2}}{2}italic_κ = - divide start_ARG ( 1 + square-root start_ARG italic_η end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG 4 fixed points
κ>(1+η)22𝜅superscript1𝜂22\kappa>-\frac{\left(1+\sqrt{\eta}\right)^{2}}{2}italic_κ > - divide start_ARG ( 1 + square-root start_ARG italic_η end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG No fixed points
Table 1: Fixed points for identical excitabilities, η=η1=η2𝜂subscript𝜂1subscript𝜂2\eta=\eta_{1}=\eta_{2}italic_η = italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.
Criteria Num. fixed points
ν1>0subscript𝜈10\nu_{1}>0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > 0, ν1ν22κν1+ν2<0subscript𝜈1subscript𝜈22𝜅subscript𝜈1subscript𝜈20\nu_{1}\nu_{2}-2\kappa\nu_{1}+\nu_{2}<0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 2 italic_κ italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < 0 4 fixed points
ν1ν22κν1+ν2=0subscript𝜈1subscript𝜈22𝜅subscript𝜈1subscript𝜈20\nu_{1}\nu_{2}-2\kappa\nu_{1}+\nu_{2}=0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 2 italic_κ italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 2 fixed points
ν2<0subscript𝜈20\nu_{2}<0italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < 0, ν1ν22κν2+ν1>0subscript𝜈1subscript𝜈22𝜅subscript𝜈2subscript𝜈10\nu_{1}\nu_{2}-2\kappa\nu_{2}+\nu_{1}>0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 2 italic_κ italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > 0 4 fixed points
ν1ν22κν2+ν1=0subscript𝜈1subscript𝜈22𝜅subscript𝜈2subscript𝜈10\nu_{1}\nu_{2}-2\kappa\nu_{2}+\nu_{1}=0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 2 italic_κ italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0, ν1+2κν20subscript𝜈12𝜅subscript𝜈20\nu_{1}+2\kappa\nu_{2}\leq 0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 italic_κ italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ 0 2 fixed points
ν1+2κν2>0subscript𝜈12𝜅subscript𝜈20\nu_{1}+2\kappa\nu_{2}>0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 italic_κ italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0 6 fixed points
ν2<2subscript𝜈22\nu_{2}<-2italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < - 2, ν1ν22κν2+ν1<0subscript𝜈1subscript𝜈22𝜅subscript𝜈2subscript𝜈10\nu_{1}\nu_{2}-2\kappa\nu_{2}+\nu_{1}<0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 2 italic_κ italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 0, ν1ν2+8κ<0subscript𝜈1subscript𝜈28𝜅0\nu_{1}\nu_{2}+8\kappa<0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 8 italic_κ < 0 8 fixed points
ν1ν2+8κ=0subscript𝜈1subscript𝜈28𝜅0\nu_{1}\nu_{2}+8\kappa=0italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 8 italic_κ = 0 4 fixed points
otherwise No fixed points
Table 2: Fixed points for non-identical excitabilities, η1η2subscript𝜂1subscript𝜂2\eta_{1}\neq\eta_{2}italic_η start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ italic_η start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.

Appendix B Details of numerics

Below follows some notes on the numerics used in the paper.

B.1 Numerical scheme for finding attractors

For a parameter value, initial conditions are sampled throughout the phase space. For each such initial condition, the model is integrated using an explicit 5/4 Runge–Kutta method for some time Ttransientsubscript𝑇transientT_{\text{transient}}italic_T start_POSTSUBSCRIPT transient end_POSTSUBSCRIPT where no states except the final one is stored. Then, the integration is restarted with this state as the initial value and integrated using an explicit 5/4 Runge–Kutta method for some time Tstablesubscript𝑇stableT_{\text{stable}}italic_T start_POSTSUBSCRIPT stable end_POSTSUBSCRIPT, and the state in every time step is recorded. Based on the assumption that all transient behavior occurred during the first Ttransientsubscript𝑇transientT_{\text{transient}}italic_T start_POSTSUBSCRIPT transient end_POSTSUBSCRIPT part of the integration, the initial condition is categorized as being in the basin of attraction for a type of attractor based on the following criteria:

  1. 1.

    If the state has changed more than 2π2𝜋2\pi2 italic_π in any θksubscript𝜃𝑘\theta_{k}italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, the orbit is labeled a rotation in those θksubscript𝜃𝑘\theta_{k}italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT.

  2. 2.

    If the state has changed less than some tolerance δf.p.subscript𝛿f.p.\delta_{\text{f.p.}}italic_δ start_POSTSUBSCRIPT f.p. end_POSTSUBSCRIPT in the 2-norm, the orbit is labeled a fixed point.

  3. 3.

    Otherwise, the orbit is labeled a libration.

The parameter value is then marked as having at least one attractor of that given type. After the process has been repeated for all initial conditions, the found types of attractors for the parameter value can then be recorded.

The method can miss types of attractors if the sampling is not dense enough; at least one sample per basin of attraction is required to get correct results. The method can also falsely identify some types of orbits if the time Ttransientsubscript𝑇transientT_{\text{transient}}italic_T start_POSTSUBSCRIPT transient end_POSTSUBSCRIPT is shorter than the duration of transient behavior. False identification can also happen if the time Tstablesubscript𝑇stableT_{\text{stable}}italic_T start_POSTSUBSCRIPT stable end_POSTSUBSCRIPT is shorter than the period of a rotation.

Also note that this scheme ignores how many distinct attractors of a given type exist, as well as differentiating between periodic and quasi-periodic rotations.

B.2 Modification of co-evolutionary system for MatCont

Bifurcation curves of the co-evolutionary system were investigated using the numerical continuation software MatCont [53]. MatCont requires that the dynamical system has a real vector space msuperscript𝑚\mathbb{R}^{m}blackboard_R start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT as the phase space. The embedding θk(cosθk,sinθk)=(xk,yk)maps-tosubscript𝜃𝑘subscript𝜃𝑘subscript𝜃𝑘subscript𝑥𝑘subscript𝑦𝑘\theta_{k}\mapsto\left(\cos\theta_{k},\sin\theta_{k}\right)=\left(x_{k},y_{k}\right)italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ↦ ( roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , roman_sin italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = ( italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) of the periodic phase variables is used to achieve this. The vector field is extended in the radial direction by

rk˙=1rk˙subscript𝑟𝑘1subscript𝑟𝑘\dot{r_{k}}=1-r_{k}over˙ start_ARG italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG = 1 - italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT (36)

where xk=rkcosθksubscript𝑥𝑘subscript𝑟𝑘subscript𝜃𝑘x_{k}=r_{k}\cos\theta_{k}italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and yk=rksinθksubscript𝑦𝑘subscript𝑟𝑘subscript𝜃𝑘y_{k}=r_{k}\sin\theta_{k}italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT such that the unit circle is attractive. Thus the node dynamics used in MatCont are

xk˙=\diff(rkcosθk)t=rk˙cosθkrksin(θk)θk˙=(1rk)cosθkykθk˙˙subscript𝑥𝑘\diffsubscript𝑟𝑘subscript𝜃𝑘𝑡˙subscript𝑟𝑘subscript𝜃𝑘subscript𝑟𝑘subscript𝜃𝑘˙subscript𝜃𝑘1subscript𝑟𝑘subscript𝜃𝑘subscript𝑦𝑘˙subscript𝜃𝑘\displaystyle\begin{split}\dot{x_{k}}&=\diff*{\left(r_{k}\cos\theta_{k}\right)% }{t}\\ &=\dot{r_{k}}\cos\theta_{k}-r_{k}\sin(\theta_{k})\dot{\theta_{k}}\\ &=(1-r_{k})\cos\theta_{k}-y_{k}\dot{\theta_{k}}\end{split}start_ROW start_CELL over˙ start_ARG italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG end_CELL start_CELL = ∗ ( italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) italic_t end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = over˙ start_ARG italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_sin ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) over˙ start_ARG italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = ( 1 - italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT - italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over˙ start_ARG italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG end_CELL end_ROW (37a)
yk˙=\diff(rksinθk)t=rk˙sinθk+rkcos(θk)θk˙=(1rk)sinθk+xkθk˙˙subscript𝑦𝑘\diffsubscript𝑟𝑘subscript𝜃𝑘𝑡˙subscript𝑟𝑘subscript𝜃𝑘subscript𝑟𝑘subscript𝜃𝑘˙subscript𝜃𝑘1subscript𝑟𝑘subscript𝜃𝑘subscript𝑥𝑘˙subscript𝜃𝑘\displaystyle\begin{split}\dot{y_{k}}&=\diff*{\left(r_{k}\sin\theta_{k}\right)% }{t}\\ &=\dot{r_{k}}\sin\theta_{k}+r_{k}\cos(\theta_{k})\dot{\theta_{k}}\\ &=(1-r_{k})\sin\theta_{k}+x_{k}\dot{\theta_{k}}\end{split}start_ROW start_CELL over˙ start_ARG italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG end_CELL start_CELL = ∗ ( italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) italic_t end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = over˙ start_ARG italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG roman_sin italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_cos ( italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) over˙ start_ARG italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = ( 1 - italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) roman_sin italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over˙ start_ARG italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG end_CELL end_ROW (37b)

where

rk=xk2+yk2subscript𝑟𝑘superscriptsubscript𝑥𝑘2superscriptsubscript𝑦𝑘2r_{k}=\sqrt{x_{k}^{2}+y_{k}^{2}}italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = square-root start_ARG italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (38)

and

cosθk=xkrk.subscript𝜃𝑘subscript𝑥𝑘subscript𝑟𝑘\cos\theta_{k}=\frac{x_{k}}{r_{k}}.roman_cos italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = divide start_ARG italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG . (39)

References

References

  • Strogatz [2001] S. H. Strogatz, “Exploring complex networks,” nature 410, 268–276 (2001).
  • Barabási and Oltvai [2004] A.-L. Barabási and Z. N. Oltvai, “Network biology: understanding the cell’s functional organization.” Nature Reviews. Genetics 5, 101–113 (2004).
  • Blinder et al. [2013] P. Blinder, P. S. Tsai, J. P. Kaufhold, P. M. Knutsen, H. Suhl, and D. Kleinfeld, “The cortical angiome: an interconnected vascular network with noncolumnar patterns of blood flow,” Nature neuroscience 16, 889–897 (2013).
  • Kirst et al. [2020] C. Kirst, S. Skriabine, A. Vieites-Prado, T. Topilko, P. Bertin, G. Gerschenfeld, F. Verny, P. Topilko, N. Michalski, M. Tessier-Lavigne, et al.“Mapping the fine-scale organization and plasticity of the brain vasculature,” Cell 180, 780–795 (2020).
  • Roth-Nebelsick et al. [2001] A. Roth-Nebelsick, D. Uhl, V. Mosbrugger, and H. Kerp, “Evolution and function of leaf venation architecture: a review,” Annals of Botany 87, 553–566 (2001).
  • Ronellenfitsch et al. [2015] H. Ronellenfitsch, J. Lasser, D. C. Daly, and E. Katifori, “Topological phenotypes constitute a new dimension in the phenotypic space of leaf venation networks,” PLoS computational biology 11, e1004680 (2015).
  • Sporns [2013] O. Sporns, “Structure and function of complex brain networks,” Dialogues in clinical neuroscience 15, 247–262 (2013).
  • Lynn and Bassett [2019] C. W. Lynn and D. S. Bassett, “The physics of brain network structure, function and control,” Nature Reviews Physics 1, 318–332 (2019).
  • Pagani and Aiello [2014] G. A. Pagani and M. Aiello, “Power grid complex network evolutions for the smart grid,” Physica A: Statistical Mechanics and its Applications 396, 248–266 (2014).
  • Rohden et al. [2012] M. Rohden, A. Sorge, M. Timme, and D. Witthaut, “Self-organized synchronization in decentralized power grids,” Physical review letters 109, 064101 (2012).
  • Tu [2000] Y. Tu, “How robust is the internet?” Nature 406, 353–354 (2000).
  • Kaluza et al. [2010] P. Kaluza, A. Kölzsch, M. T. Gastner, and B. Blasius, “The complex network of global cargo ship movements.” Journ. Roy. Society Interface 7, 1093–103 (2010).
  • Xie and Levinson [2007] F. Xie and D. Levinson, “Measuring the structure of road networks,” Geographical analysis 39, 336–356 (2007).
  • Strogatz [2004] S. Strogatz, “Sync: The emerging science of spontaneous order,”   (2004).
  • Rosenblum and Kurths [2003] M. Rosenblum and J. Kurths, Synchronization: a universal concept in nonlinear science (Cambridge University Press, 2003).
  • Kempter, Gerstner, and van Hemmen [1999] R. Kempter, W. Gerstner, and L. J. van Hemmen, “Hebbian learning and spiking neurons,” Physical Review E 59, 4498–4514 (1999).
  • Jacobsen, Mulvany, and Holstein-Rathlou [2008] J. C. B. Jacobsen, M. J. Mulvany, and N.-H. Holstein-Rathlou, “A mechanism for arteriolar remodeling based on maintenance of smooth muscle cell activation.” American journal of physiology. Regulatory, integrative and comparative physiology 294, R1379–R1389 (2008).
  • Martens and Klemm [2017] E. A. Martens and K. Klemm, “Transitions from trees to cycles in adaptive flow networks,” Frontiers in Physics 5, 1–10 (2017)1711.00401 .
  • Mestre et al. [2020] H. Mestre, T. Du, A. M. Sweeney, G. Liu, A. J. Samson, W. Peng, K. N. Mortensen, F. F. Stæger, P. A. Bork, L. Bashford, et al.“Cerebrospinal fluid influx drives acute ischemic tissue swelling,” Science 367, eaax7171 (2020).
  • Ginelli and Chaté [2010] F. Ginelli and H. Chaté, “Relevance of metric-free interactions in flocking phenomena,” Physical review letters 105, 168103 (2010).
  • Couzin [2009] I. D. Couzin, “Collective cognition in animal groups,” Trends in cognitive sciences 13, 36–43 (2009).
  • Rosenblum and Pikovsky [2003] M. Rosenblum and A. Pikovsky, “Synchronization: from pendulum clocks to chaotic lasers and chemical oscillators,” Contemporary Physics 44, 401–416 (2003).
  • Luke, Barreto, and So [2013] T. B. Luke, E. Barreto, and P. So, “Complete classification of the macroscopic behavior of a heterogeneous network of theta neurons,” Neural computation 25, 3207–3234 (2013).
  • Montbrió, Pazó, and Roxin [2015] E. Montbrió, D. Pazó, and A. Roxin, “Macroscopic description for networks of spiking neurons,” Physical Review X 5, 021028 (2015).
  • Strogatz [2000] S. H. Strogatz, “From kuramoto to crawford: exploring the onset of synchronization in populations of coupled oscillators,” Physica D: Nonlinear Phenomena 143, 1–20 (2000).
  • Acebrón et al. [2005] J. A. Acebrón, L. L. Bonilla, C. J. Pérez Vicente, F. Ritort, and R. Spigler, “The kuramoto model: A simple paradigm for synchronization phenomena,” Reviews of modern physics 77, 137–185 (2005).
  • Bick et al. [2020] 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,” The Journal of Mathematical Neuroscience 10, 9 (2020).
  • Cestnik and Martens [2024] R. Cestnik and E. A. Martens, “Integrability of a globally coupled complex riccati array: Quadratic integrate-and-fire neurons, phase oscillators, and all in between,” Physical Review Letters 132, 057201 (2024).
  • Gross and Blasius [2008] T. Gross and B. Blasius, “Adaptive coevolutionary networks: a review,” Journal of the Royal Society Interface 5, 259–271 (2008).
  • Berner et al. [2023] R. Berner, T. Gross, C. Kuehn, J. Kurths, and S. Yanchuk, “Adaptive dynamical networks,” Physics Reports 1031, 1–59 (2023).
  • Kuehn et al. [2015] C. Kuehn et al.Multiple time scale dynamics, Vol. 191 (Springer, 2015).
  • Kuramoto and Kuramoto [1984] Y. Kuramoto and Y. Kuramoto, Chemical turbulence (Springer, 1984).
  • Seliger, Young, and Tsimring [2002] P. Seliger, S. C. Young, and L. S. Tsimring, “Plasticity and learning in a network of coupled phase oscillators,” Physical Review E 65, 041906 (2002).
  • Berner et al. [2019] R. Berner, J. Fialkowski, D. Kasatkin, V. Nekorkin, S. Yanchuk, and E. Schöll, “Hierarchical frequency clusters in adaptive networks of phase oscillators,” Chaos 29, 103134 (2019)arXiv:1904.06927 .
  • Berner et al. [2021] R. Berner, S. Vock, E. Schöll, and S. Yanchuk, “Desynchronization transitions in adaptive networks,” Physical Review Letters 126, 028301 (2021).
  • Jüttner and Martens [2023] B. Jüttner and E. A. Martens, “Complex dynamics in adaptive phase oscillator networks,”  053106 (2023), 10.1063/5.01331902209.10514 .
  • [37] R. Cestnik and E. A. Martens, “Adaptive dynamics of the Kuramoto model in the continuum limit,”  , 103134arXiv:1904.06927 .
  • [38] M. Yamakou, J. Zhu, and E. A. Martens, “Inverse stochastic resonance in adaptive small-world neural networks,”  , 103134arXiv:1904.06927 .
  • Ermentrout and Kopell [1986] G. B. Ermentrout and N. Kopell, “Parabolic bursting in an excitable system coupled with a slow oscillation,” SIAM journal on applied mathematics 46, 233–253 (1986).
  • Gutkin [2022] B. Gutkin, “Theta neuron model,” in Encyclopedia of computational neuroscience (Springer, 2022) pp. 3412–3419.
  • Ermentrout [2008] B. Ermentrout, “Ermentrout-kopell canonical model,” Scholarpedia 3, 1398 (2008).
  • Pietras and Daffertshofer [2019] B. Pietras and A. Daffertshofer, “Network dynamics of coupled oscillators and phase reduction techniques,” Physics Reports 819, 1–105 (2019).
  • Coombes and Byrne [2018] S. Coombes and Á. Byrne, “Next generation neural mass models,” in Nonlinear dynamics in computational neuroscience (Springer, 2018) pp. 1–16.
  • Pietras, Pikovsky et al. [2023] B. Pietras, A. Pikovsky, et al.“Exact finite-dimensional description for networks of globally coupled spiking neurons,” Physical Review E 107, 024315 (2023).
  • Schmidt et al. [2018] H. Schmidt, D. Avitabile, E. Montbrió, and A. Roxin, “Network mechanisms underlying the role of oscillations in cognitive tasks,” PLoS computational biology 14, e1006430 (2018).
  • Ariaratnam and Strogatz [2001] J. T. Ariaratnam and S. H. Strogatz, “Phase diagram for the winfree model of coupled nonlinear oscillators,” Physical Review Letters 86, 4278 (2001).
  • Note [1] This boundary is numerically determined by considering orbits starting near the outer saddle and determining whether the resulting orbit remains oscillatory (libration) or converges to the stable node.
  • Note [2] The location of this bifurcation was determined numerically by sampling many initial conditions and determining whether trajectories only converge to a quiescent fixed point (stable node Q Q), or whether some trajectories correspond to rotations (S S) in phase space.
  • Duchet, Bick, and Byrne [2023] B. Duchet, C. Bick, and Á. Byrne, “Mean-field approximations with adaptive coupling for networks with spike-timing-dependent plasticity,”  35, 1481–1528 (2023).
  • Sorrentino et al. [2016] F. Sorrentino, L. M. Pecora, A. M. Hagerstrom, T. E. Murphy, and R. Roy, “Complete characterization of the stability of cluster synchronization in complex dynamical networks,” Science advances 2, e1501737 (2016).
  • Gerstner and Kistler [2002] W. Gerstner and W. M. Kistler, “Mathematical formulations of hebbian learning,” Biol. Cybern. 87 (2002), 10.1007/S00422-002-0353-Y.
  • Lücken et al. [2016] L. Lücken, O. V. Popovych, P. A. Tass, and S. Yanchuk, “Noise-enhanced coupling between two oscillators with long-term plasticity,” Physical Review E 93, 032210 (2016).
  • Dhooge et al. [2008] A. Dhooge, W. Govaerts, Y. A. Kuznetsov, H. G. E. Meijer, and B. Sautois, “New features of the software matcont for bifurcation analysis of dynamical systems,” Mathematical and Computer Modelling of Dynamical Systems 14, 147–175 (2008).
  • Klemm and Martens [2023] K. Klemm and E. A. Martens, “Bifurcations in adaptive vascular networks: Toward model calibration,” Chaos: An Interdisciplinary Journal of Nonlinear Science 33 (2023).