15
Macroscopic Description for Networks of Spiking Neurons Ernest Montbrió, 1 Diego Pazó, 2 and Alex Roxin 3 1 Center for Brain and Cognition, Department of Information and Communication Technologies, Universitat Pompeu Fabra, 08018 Barcelona, Spain 2 Instituto de Física de Cantabria (IFCA), CSIC-Universidad de Cantabria, 39005 Santander, Spain 3 Centre de Recerca Matemàtica, Campus de Bellaterra, Edifici C, 08193 Bellaterra, Spain (Received 30 December 2014; published 19 June 2015) A major goal of neuroscience, statistical physics, and nonlinear dynamics is to understand how brain function arises from the collective dynamics of networks of spiking neurons. This challenge has been chiefly addressed through large-scale numerical simulations. Alternatively, researchers have formulated mean-field theories to gain insight into macroscopic states of large neuronal networks in terms of the collective firing activity of the neurons, or the firing rate. However, these theories have not succeeded in establishing an exact correspondence between the firing rate of the network and the underlying microscopic state of the spiking neurons. This has largely constrained the range of applicability of such macroscopic descriptions, particularly when trying to describe neuronal synchronization. Here, we provide the derivation of a set of exact macroscopic equations for a network of spiking neurons. Our results reveal that the spike generation mechanism of individual neurons introduces an effective coupling between two biophysically relevant macroscopic quantities, the firing rate and the mean membrane potential, which together govern the evolution of the neuronal network. The resulting equations exactly describe all possible macroscopic dynamical states of the network, including states of synchronous spiking activity. Finally, we show that the firing-rate description is related, via a conformal map, to a low-dimensional description in terms of the Kuramoto order parameter, called Ott-Antonsen theory. We anticipate that our results will be an important tool in investigating how large networks of spiking neurons self-organize in time to process and encode information in the brain. DOI: 10.1103/PhysRevX.5.021028 Subject Areas: Biological Physics, Interdisciplinary Physics, Nonlinear Dynamics I. INTRODUCTION The processing and coding of information in the brain necessarily imply the coordinated activity of large ensem- bles of neurons. Within sensory regions of the cortex, many cells show similar responses to a given stimulus, indicating a high degree of neuronal redundancy at the local level. This suggests that information is encoded in the population response and, hence, can be captured via macroscopic measures of the network activity [1]. Moreover, the collective behavior of large neuronal networks is particu- larly relevant given that current brain measurement techniques, such as electroencephalography or functional magnetic resonance imaging, provide data that are neces- sarily averaged over the activity of a large number of neurons. The macroscopic dynamics of neuronal ensembles has been extensively studied through computational models of large networks of recurrently coupled spiking neurons, including Hodgkin-Huxley-type conductance-based neu- rons [2], as well as simplified neuron models; see, e.g., Refs. [35]. In parallel, researchers have sought to develop statistical descriptions of neuronal networks, mainly in terms of a macroscopic observable that measures the mean rate at which neurons emit spikes, the firing rate [620]. These descriptions, called firing-rate equations (FREs), have been proven to be extremely useful in understanding general computational principles underlying functions such as memory [21,22], visual processing [2325], motor control [26], or decision making [27]. Despite these efforts, to date there is no exact theory linking the dynamics of a large network of spiking neurons with that of the firing rate. Specifically, current macro- scopic descriptions do not offer a precise correspondence between the microscopic dynamics of the individual neurons, e.g., individual membrane potentials, and the firing-rate dynamics of the neuronal network. Indeed, FREs are generally derived through heuristic arguments that rely on the underlying spiking activity of the neurons being asynchronous and hence uncorrelated. As such, firing-rate descriptions are not sufficient to describe Published by the American Physical Society under the terms of the Creative Commons Attribution 3.0 License. Further distri- bution of this work must maintain attribution to the author(s) and the published articles title, journal citation, and DOI. PHYSICAL REVIEW X 5, 021028 (2015) 2160-3308=15=5(2)=021028(15) 021028-1 Published by the American Physical Society

Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

Macroscopic Description for Networks of Spiking Neurons

Ernest Montbrió,1 Diego Pazó,2 and Alex Roxin31Center for Brain and Cognition, Department of Information and Communication Technologies,

Universitat Pompeu Fabra, 08018 Barcelona, Spain2Instituto de Física de Cantabria (IFCA), CSIC-Universidad de Cantabria, 39005 Santander, Spain

3Centre de Recerca Matemàtica, Campus de Bellaterra, Edifici C, 08193 Bellaterra, Spain(Received 30 December 2014; published 19 June 2015)

A major goal of neuroscience, statistical physics, and nonlinear dynamics is to understand how brainfunction arises from the collective dynamics of networks of spiking neurons. This challenge has beenchiefly addressed through large-scale numerical simulations. Alternatively, researchers have formulatedmean-field theories to gain insight into macroscopic states of large neuronal networks in terms of thecollective firing activity of the neurons, or the firing rate. However, these theories have not succeeded inestablishing an exact correspondence between the firing rate of the network and the underlying microscopicstate of the spiking neurons. This has largely constrained the range of applicability of such macroscopicdescriptions, particularly when trying to describe neuronal synchronization. Here, we provide thederivation of a set of exact macroscopic equations for a network of spiking neurons. Our results revealthat the spike generation mechanism of individual neurons introduces an effective coupling between twobiophysically relevant macroscopic quantities, the firing rate and the mean membrane potential, whichtogether govern the evolution of the neuronal network. The resulting equations exactly describe all possiblemacroscopic dynamical states of the network, including states of synchronous spiking activity. Finally, weshow that the firing-rate description is related, via a conformal map, to a low-dimensional description interms of the Kuramoto order parameter, called Ott-Antonsen theory. We anticipate that our results will bean important tool in investigating how large networks of spiking neurons self-organize in time to processand encode information in the brain.

DOI: 10.1103/PhysRevX.5.021028 Subject Areas: Biological Physics,Interdisciplinary Physics,Nonlinear Dynamics

I. INTRODUCTION

The processing and coding of information in the brainnecessarily imply the coordinated activity of large ensem-bles of neurons. Within sensory regions of the cortex, manycells show similar responses to a given stimulus, indicatinga high degree of neuronal redundancy at the local level.This suggests that information is encoded in the populationresponse and, hence, can be captured via macroscopicmeasures of the network activity [1]. Moreover, thecollective behavior of large neuronal networks is particu-larly relevant given that current brain measurementtechniques, such as electroencephalography or functionalmagnetic resonance imaging, provide data that are neces-sarily averaged over the activity of a large number ofneurons.The macroscopic dynamics of neuronal ensembles has

been extensively studied through computational models of

large networks of recurrently coupled spiking neurons,including Hodgkin-Huxley-type conductance-based neu-rons [2], as well as simplified neuron models; see, e.g.,Refs. [3–5]. In parallel, researchers have sought to developstatistical descriptions of neuronal networks, mainly interms of a macroscopic observable that measures the meanrate at which neurons emit spikes, the firing rate [6–20].These descriptions, called firing-rate equations (FREs),have been proven to be extremely useful in understandinggeneral computational principles underlying functions suchas memory [21,22], visual processing [23–25], motorcontrol [26], or decision making [27].Despite these efforts, to date there is no exact theory

linking the dynamics of a large network of spiking neuronswith that of the firing rate. Specifically, current macro-scopic descriptions do not offer a precise correspondencebetween the microscopic dynamics of the individualneurons, e.g., individual membrane potentials, and thefiring-rate dynamics of the neuronal network.Indeed, FREs are generally derived through heuristic

arguments that rely on the underlying spiking activity of theneurons being asynchronous and hence uncorrelated. Assuch, firing-rate descriptions are not sufficient to describe

Published by the American Physical Society under the terms ofthe Creative Commons Attribution 3.0 License. Further distri-bution of this work must maintain attribution to the author(s) andthe published article’s title, journal citation, and DOI.

PHYSICAL REVIEW X 5, 021028 (2015)

2160-3308=15=5(2)=021028(15) 021028-1 Published by the American Physical Society

Page 2: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

network states involving some degree of spike synchroni-zation. Synchronization is, however, a ubiquitous phe-nomenon in the brain, and its potential role in neuronalcomputation is the subject of intense research [14,28–34].Hence, the lack of firing-rate descriptions for synchronousstates limits the range of applicability of mean-fieldtheories to investigate neuronal dynamics.Here, we propose a method to derive the FREs for

networks of heterogeneous, all-to-all coupled quadraticintegrate-and-fire (QIF) neurons, which is exact in thethermodynamic limit, i.e., for large numbers of neurons.We consider an ansatz for the distribution of the neurons’membrane potentials that we denominate the Lorentzianansatz (LA). The LA solves the corresponding continuityequation exactly, making the system amenable to theoreti-cal analysis. Specifically, for particular distributions of theheterogeneity, the LA yields a system of two ordinarydifferential equations for the firing rate and mean mem-brane potential of the neuronal population. These equationsfully describe the macroscopic states of the network—including synchronized states, and represent the firstexample of an exact firing-rate description for a networkof recurrently connected spiking neurons. Finally, we showhow the LA transforms, via a conformal mapping, intothe so-called Ott-Antonsen ansatz (OA) that is usedextensively to investigate the low-dimensional dynamicsof large populations of phase oscillators in terms of theKuramoto order parameter [35].

II. MODEL DESCRIPTION

Hodgkin-Huxley-type neuronal models can be broadlyclassified into two classes, according to the nature of theirtransition to spiking in response to an injected current[36,37]. Neuronal models with so-called class I excitabilitygenerate action potentials with arbitrarily low frequency,depending on the strength of the applied current. Thisoccurs when a resting state disappears through a saddle-node bifurcation. In contrast, in neurons with class IIexcitability, the action potentials are generated with a finitefrequency. This occurs when the resting state loses stabilityvia a Hopf bifurcation. The QIF neuron is the canonicalmodel for class I neurons, and, thus, generically describestheir dynamics near the spiking threshold [5,38,39]. Ouraim here is to derive the FREs corresponding to aheterogeneous all-to-all coupled population of N QIFneurons. The correspondence is exact in the thermody-namic limit, i.e., when N → ∞ (this convergence wasstudied recently in Ref. [40]).The microscopic state of the population of QIF neurons

is given by the membrane potentials fVjgj¼1;…;N , whichobey the following ordinary differential equations [5]:

_Vj ¼ V2j þ Ij; if Vj ≥ Vp; then Vj ← Vr: ð1Þ

Here, the overdot denotes the time derivative and Ijrepresents an input current. Each time a neuron’s mem-brane potential Vj reaches the peak value Vp, the neuronemits a spike and its voltage is reset to the value Vr. In ouranalysis, we consider the limit Vp ¼ −Vr → ∞. Thisresetting rule captures the spike reset as well as therefractoriness of the neuron. Without loss of generality,we rescale the time and the voltage in Eq. (1) to absorb anycoefficients that would have appeared in the first two terms.The form for the input currents is

Ij ¼ ηj þ JsðtÞ þ IðtÞ; ð2Þ

where the external input has a heterogeneous, quenchedcomponent ηj as well as a common time-varying compo-nent IðtÞ, and the recurrent input is the synaptic weight Jtimes the mean synaptic activation sðtÞ, which is written as

sðtÞ ¼ 1

N

XNj¼1

Xkntkj<t

Zt

−∞dt0aτðt − t0Þδðt0 − tkjÞ: ð3Þ

Here, tkj is the time of the kth spike of the jth neuron, δðtÞ isthe Dirac delta function, and aτðtÞ is the normalizedsynaptic activation caused by a single presynaptic spikewith time scale τ, e.g., aτðtÞ ¼ e−t=τ=τ.

A. Continuous formulation

In the thermodynamic limit N → ∞, we drop the indicesin Eqs. (1) and (2) and denote ρðVjη; tÞdV as the fraction ofneurons with membrane potentials between V and V þ dVand parameter η, at time t. Accordingly, parameter ηnow becomes a continuous random variable distributedaccording to a probability distribution function gðηÞ.The total voltage density at time t is then givenby

R∞−∞ ρðVjη; tÞgðηÞdη.

The conservation of the number of neurons leads to thefollowing continuity equation:

∂tρþ ∂V ½ðV2 þ ηþ Jsþ IÞρ� ¼ 0; ð4Þ

where we explicitly include the velocity given byEqs. (1) and (2).

III. RESULTS

The continuity equation (4) without temporal forcingIðtÞ ¼ 0 has a trivial stationary solution. For each value ofη, this solution has the form of a Lorentzian function:ρ0ðVjηÞ ∝ ðV2 þ ηþ JsÞ−1. Physically, the Lorentziandensity means that firing neurons with the same η valuewill be scattered with a density inversely proportional totheir speed [Eq. (1)]; i.e., they will accumulate at slowplaces and thin out at fast places on the V axis. In addition,for those η values corresponding to quiescent neurons, the

ERNEST MONTBRIÓ, DIEGO PAZÓ, AND ALEX ROXIN PHYS. REV. X 5, 021028 (2015)

021028-2

Page 3: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

density ρ0 collapses at the rest state in the form of a Diracdelta function.Next, we assume that, independently of the initial

condition, solutions of Eq. (4) generically converge to aLorentzian-shaped function, so that all relevant dynamicsoccur inside that lower-dimensional space. This fact ismathematically expressed by the following LA for theconditional density functions:

ρðVjη; tÞ ¼ 1

π

xðη; tÞ½V − yðη; tÞ�2 þ xðη; tÞ2 ; ð5Þ

which is a Lorentzian function with time-dependent half-width xðη; tÞ and center at yðη; tÞ. In the following, weassume that the LA Eq. (5) completely describes themacroscopic dynamics of the network of QIF neuronsand postpone the mathematical justification of its validity toSec. III E.

A. Macroscopic observables: Firing rateand mean membrane potential

The half-width xðη; tÞ of the LA has a particularly simplerelation with the firing rate of the neuronal population (i.e.,the number of spikes per unit time). Indeed, the firing ratefor each η value at time t, rðη; tÞ, can be computed bynoting that neurons fire at a rate given by the probabilityflux at infinity: rðη; tÞ ¼ ρðV → ∞jη; tÞ _VðV → ∞jη; tÞ.The limit V → ∞ on the right-hand side of this equationcan be evaluated within the LA, and gives the simpleidentity

xðη; tÞ ¼ πrðη; tÞ: ð6ÞThe (total) firing rate rðtÞ is then

rðtÞ ¼ 1

π

Z∞

−∞xðη; tÞgðηÞdη: ð7Þ

Additionally, the quantity yðη; tÞ is, for each value of η, themean of the membrane potential:

yðη; tÞ ¼ p:v:Z

−∞ρðVjη; tÞVdV: ð8Þ

Here, we take the Cauchy principal value, defined asp:v:

R∞−∞ fðxÞdx ¼ limR→∞RR−R fðxÞdx, to avoid the other-

wise ill-defined integral. The mean membrane potential isthen

vðtÞ ¼Z

−∞yðη; tÞgðηÞdη: ð9Þ

B. Firing-rate equations

Substituting the LA Eq. (5) into the continuityequation (4), we find that, for each value of η, variables

x and y must obey two coupled equations, which can bewritten in complex form as

∂twðη; tÞ ¼ i½ηþ JsðtÞ − wðη; tÞ2 þ IðtÞ�; ð10Þ

where wðη; tÞ≡ xðη; tÞ þ iyðη; tÞ. Closing this equationrequires expressing the mean synaptic activation sðtÞ asa function of wðη; tÞ. The simplest choice is to take the limitof infinitely fast synapses, τ → 0 in Eq. (3), so that we getan equality with the firing rate sðtÞ ¼ rðtÞ. This allows forthe system of QIF neurons [Eqs. (1)–(3)] to be exactlydescribed by Eqs. (10) and (7), an infinite set of integro-differential equations if gðηÞ is a continuous distribution.Equation (10) is useful for general distributions gðηÞ (see

Appendix B), but a particularly sharp reduction in dimen-sionality is achieved if η is distributed according to aLorentzian distribution of half-width Δ centered at η̄:

gðηÞ ¼ 1

π

Δðη − η̄Þ2 þ Δ2

: ð11Þ

Note that this distribution accounts for the quenchedvariability in the external inputs. The fact that it isLorentzian is unrelated to the LA for the density ofmembrane potentials. Assuming Eq. (11), the integrals inEqs. (7) and (9) can be evaluated closing the integralcontour in the complex η plane and using the residuetheorem [41]. Notably, the firing rate and the meanmembrane potential only depend on the value of w atthe pole of gðηÞ in the lower half η plane:

πrðtÞ þ ivðtÞ ¼ wðη̄ − iΔ; tÞ:

As a result, we only need to evaluate Eq. (10) atη ¼ η̄ − iΔ, and thereby obtain a system of FREs com-posed of two ordinary differential equations:

_r ¼ Δ=π þ 2rv; ð12aÞ

_v ¼ v2 þ η̄þ Jrþ IðtÞ − π2r2: ð12bÞ

This nonlinear system describes the macroscopic dynamicsof the population of QIF neurons in terms of the populationfiring rate r and mean membrane potential v.It is enlightening to compare the mean-field model

Eq. (12) with the equations of the spiking neurons. Notethat Eq. (12b) resembles Eqs. (1) and (2) for the individualQIF neuron, but without spike resetting. Indeed, themacroscopic firing-rate variable r enters as a negativefeedback term in Eq. (12b) and impedes the explosivegrowth of the mean membrane potential v.This negative feedback, combined with the coupling

term on the right-hand side of Eq. (12a), describes theeffective interaction between the firing rate and meanmembrane potential at the network level. Therefore, the

MACROSCOPIC DESCRIPTION FOR NETWORKS OF … PHYS. REV. X 5, 021028 (2015)

021028-3

Page 4: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

FREs (12) describe the effect of the single-cell spikegeneration and reset mechanism at the network level.In the following, we examine the dynamics described by

Eq. (12) and show that they fully reproduce the macro-scopic dynamics of the network of QIF neurons, evenduring episodes of strong spike synchrony.

C. Analysis of the firing-rate equations

To begin with the analysis of Eq. (12), we first note thatin the absence of forcing, IðtÞ ¼ 0, the only attractors ofEq. (12) are fixed points. Figure 1(a) shows a phasediagram of the system as a function of the mean externaldrive η̄ and synaptic weight J, both normalized by the widthof the input distribution [42]. There are three qualitativelydistinct regions of the phase diagram: (1) a single stablenode corresponding to a low-activity state (white), (2) asingle stable focus (spiral) generally corresponding to ahigh-activity state (gray), and (3) a region of bistabilitybetween low and high firing rate [cyan; see a phase portraitof this region in Fig. 1(d)]. Comparison of a samplebifurcation diagram of the fixed points from numericalsimulation of networks of QIF neurons with that obtainedfrom the FREs (12) shows an excellent correspondence; seeFigs. 1(b) and 1(c).A similar phase diagram can be readily reproduced by

traditional heuristic firing-rate models, with one significantqualitative difference: the presence of a stable focus—andhence, damped oscillations. Specifically, in the gray regionof the phase diagram in Fig. 1(a), the system undergoesoscillatory decay to the stable fixed point. This oscillatorydecay occurs as well for the high-activity state over a largeextent of the region of bistability (cyan); see, e.g., Fig. 1(d).The presence of damped oscillations at the macroscopic

level reflects the transitory synchronous firing of a fraction

of the neurons in the ensemble. This behavior is common inspiking neuron models with weak noise, and is not capturedby traditional firing-rate models (see, e.g., Ref. [20]).

D. Analysis of the firing-rate equations:Nonstationary inputs

To show that the FREs (12) fully describe the macro-scopic response of the population of QIF neurons to time-varying stimuli (up to finite-size effects), we consider twotypes of stimulus IðtÞ: (1) a step function and (2) asinusoidal forcing. In both cases, we simulate the fullsystem of QIF neurons and the FREs (12).Figure 2 shows the system’s response to the two different

inputs. In both cases, the system is initially (t < 0) in abistable regime and set in the low-activity state, withparameters corresponding to those of Fig. 1(d). The left-hand panels of Fig. 2 show the response of the system to astep current applied at t ¼ 0. The applied current issuch that the system abandons the bistable region—seeFig. 1(a)—and approaches the high-activity state, which isa stable focus. This is clearly reflected in the time seriesrðtÞ; vðtÞ, where the rate equations (12) exactly predict thedamped oscillations exhibited by the mean field of the QIFneurons. The raster plot in Fig. 2(e) shows the presence ofthe oscillations, which is due to the transitory synchronousfiring of a large fraction of neurons in the population.Finally, at t ¼ 30, the current is removed and the systemconverges—again, showing damped oscillations—to thenew location of the (focus) fixed point, which clearlycoexists with the stable node where it was originallyplaced (t < 0).The right-hand panels of Fig. 2 show the response of the

model to a periodic current, which drives the system fromone side of the bistable region to the other. As a result, we

-10 -5 0

η/Δ

-3

0

v/Δ1/

2

-10 -5 0

η/Δ0

10

20

J/Δ1/

2 -10 -5 00

1

r/Δ1/

2

0 0.5 1 1.5 2

r/Δ1/2

-2

-1

0

1

v/Δ1/

2

Bistability

Stable node

Stablefocus

(a) (b)

(c)

(d)

FIG. 1. Analysis of the steady states of FREs (12). (a) Phase diagram: In the wedge-shaped cyan-shaded region there is bistabilitybetween a high- and a low-activity state. The boundary of the bistability region is the locus of a saddle-node bifurcation which is exactlyobtained in parametric form: ðη̄; JÞSN ¼ ½−π2r2 − 3Δ2=ð2πrÞ2; 2π2rþ Δ2=ð2π2r3Þ�. To the right of the dashed line, defined byη̄f ¼ −½J=ð2πÞ�2 − ðπΔ=JÞ2, there is a stable focus (shaded regions). (b) r − η̄ and (c) v − η̄ bifurcation diagrams for J=Δ1=2 ¼ 15.Square symbols: Results obtained from numerical simulations of QIF neurons (see Appendix A for details). (d) Phase portrait of thesystem in the bistable region [η̄=Δ ¼ −5; J=Δ1=2 ¼ 15, triangle in (a)] with three fixed points: a stable focus (with its basin of attractionshaded), a stable node, and a saddle point.

ERNEST MONTBRIÓ, DIEGO PAZÓ, AND ALEX ROXIN PHYS. REV. X 5, 021028 (2015)

021028-4

Page 5: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

observe periodic bursting behavior when the system visitsthe stable focus region of the phase diagram, Fig. 1(a).To further illustrate the potential of the FREs (12) to

predict and investigate complex dynamics in ensembles ofspiking neurons, we present here a simple situation wherethe system of QIF neurons exhibits macroscopic chaos.This is observed by increasing the frequency ω of thesinusoidal driving, so that the system cannot trivially followthe stable fixed point at each cycle of the applied current.Figure 3(a) shows a phase diagram obtained using

Eq. (12). The shaded regions indicate parameter valueswhere the rate model has either a single chaotic attractor (inblack) or a chaotic attractor coexisting with a periodic orbit(in cyan). A trajectory on the chaotic attractor is depicted inFig. 3(b), and the clearly irregular time series of the firingrate is shown in orange in Fig. 3(d). Using the sameparameters, we perform numerical simulations of the QIFneurons [Eqs. (1) and (2)], finding a similar attractor andirregular dynamics as in the rate model; see Figs. 3(c)and 3(d). To obtain the time series shown in Fig. 3(d), werun the QIF neurons numerically and, after a long transient,at time t ¼ 0, the rate model Eq. (12) is initiated with thevalues of r and v obtained from the population of QIFneurons. The time series of the two systems, which areinitially close, rapidly diverge, reflecting the chaotic natureof the system.Finally, to illustrate this chaotic state at the microscopic

level, Fig. 3(e) shows the raster plot for 300 randomlychosen neurons, corresponding to the time series inFig. 3(d). The irregular firing of neurons in Fig. 3(e) hassome underlying structure that may be visualized orderingthe same set of neurons according to their intrinsic currentsas ηk < ηkþ1, with k ∈ ½1; 300�; see Fig. 3(f). Clearly, themaxima of the firing rate coincide with the synchronousfirings of clusters of neurons with similar η values. The size

of these clusters is highly irregular in time, in concomitancewith the chaotic behavior.

E. Validity of the Lorentzian ansatz

Thus far, we have shown that the LA Eq. (5) solves thecontinuity equation (4) and confirmed that these solutionsagree with the numerical simulations of the original systemof QIF neurons [Eqs. (1) and (2)]. Here, we further clarifywhy the LA holds for ensembles of QIF neurons.Transforming the voltage of the QIF neuron into a phase

via Vj ¼ tanðθj=2Þ, the system [Eqs. (1) and (2)] becomesan ensemble of “theta neurons” [38]:

_θj ¼ð1− cosθjÞþð1þ cosθjÞ½ηjþJsðtÞþ IðtÞ�: ð13Þ

In the new phase variable θ ∈ ½0; 2πÞ, the LA Eq. (5)becomes

~ρðθjη; tÞ ¼ 1

2πRe

�1þ αðη; tÞeiθ1 − αðη; tÞeiθ

�; ð14Þ

where the function αðη; tÞ is related to wðη; tÞ as

αðη; tÞ ¼ 1 − wðη; tÞ1þ wðη; tÞ : ð15Þ

Equations (5) and (14) are two representations of the so-called Poisson kernel on the half-plane and on the unit disk,respectively. These representations are related via Eq. (15),which establishes a conformal mapping from the half-planeReðwÞ ≥ 0 onto the unit disk jαj ≤ 1. In the next section,we show that variables r and v can be related, via the sameconformal map, to the Kuramoto order parameter, which isa macroscopic measure of phase coherence [43,44].

-2

0

2

v

0

3

-10 0 10 20 30 40 50 60 70 80time

-303

-10 0 10 20 30 40time

0

3

I(t)

-2

0

2

0

3

r0

300

Neu

ron

inde

x

030

0

(a) (b)

(d)

(f)

(h)

(c)

(e)

(g)

FIG. 2. The transient dynamics of an ensemble of QIF model neurons [Eqs. (1) and (2)] are exactly described by the FREs (12).(a),(b) Instantaneous firing rate and (c),(d) mean membrane potential of the QIF neurons and the FREs are depicted in black and orange,respectively. (e),(f) Raster plots of 300 randomly selected QIF neurons of the N ¼ 104 in the ensemble. (a),(c),(e),(g) At time t ¼ 0,a current I ¼ 3 is applied to all neurons, and set to zero again at t ¼ 30; stimulus IðtÞ shown in (g). (b),(d),(f),(h) At time t ¼ 0,a sinusoidal current is applied to all neurons IðtÞ ¼ I0 sinðωtÞ, with I0 ¼ 3, ω ¼ π=20; stimulus IðtÞ shown in (h). Parameters:J ¼ 15, η̄ ¼ −5;Δ ¼ 1.

MACROSCOPIC DESCRIPTION FOR NETWORKS OF … PHYS. REV. X 5, 021028 (2015)

021028-5

Page 6: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

The key observation supporting the applicability of theLA is the fact that Eq. (14) turns out to be the ansatzdiscovered in 2008 by Ott and Antonsen [35]. According tothe OA theory, in the thermodynamic limit, the dynamics ofthe class of systems

∂tθðη; tÞ ¼ Ωðη; tÞ þ Im½Hðη; tÞe−iθ� ð16Þ

generally converges to the OA manifold [Eq. (14)]. In ourcase, for Eq. (13), we have Ωðη; tÞ ¼ 1þ ηþ Jsþ I andHðη; tÞ ¼ ið−1þ ηþ Jsþ IÞ. Thus far, the convergenceof Eq. (16) to the OA manifold has been only proven forHðη; tÞ ¼ HðtÞ. This includes the well-known Kuramoto

and Winfree models [45,46], but not Eq. (13). However,there are theoretical arguments [47] that strongly suggestthat the OA manifold is also attracting for η-dependent H,as numerically confirmed by a number of recent papersusing theta neurons [48–50] and other phase-oscillatormodels [51–55].

F. Firing rate and Kuramoto order parameter

There exists a mapping between the macroscopic var-iables r and v and the so-called Kuramoto order parameterZ. Equation (15) relates, for each value of η, the firing rateand the mean membrane potential, both contained in w,with the uniformity of the phase density, measured byα—note that α ¼ 0 in Eq. (14) yields a perfectly uniformdensity. The Kuramoto order parameter is obtained byintegrating α over the whole population as

ZðtÞ ¼Z

−∞gðηÞ

Z2π

0

~ρðθjη; tÞeiθdθdη

¼Z

−∞gðηÞα�ðη; tÞdη; ð17Þ

where we assume that the density ~ρ is the OA ansatzEq. (14). The quantity Z can be seen as the center of massof a population of phases distributed across the unitcircle eiθ.For a Lorentzian distribution of currents, gðηÞ, we may

derive the exact formula (see Appendix B)

Z ¼ 1 −W�

1þW� ; ð18Þ

relating the Kuramoto order parameter Z with the firing-rate quantity W ≡ πrþ iv. Figure 4 illustrates this con-formal mapping of the right half-plane (r ≥ 0) onto the unitdisk (jZj ≤ 1).

0

1

2

r

0 10 20 30 40 50 60time

030

0O

rder

ed n

euro

n in

dex

2-4-

η

8

12

16

20

J0

300

Neu

ron

inde

x

r-2

0

2

v

0 1 2r

-2

0

2

v

(a) (b)

(c)

(d)

(e)

(f)

FIG. 3. Firing-rate model predicts the existence of macroscopicchaos in ensemble QIF neurons [Eq. (1)] with an injected periodiccurrent IðtÞ ¼ I0 sinðωtÞ. (a) Phase diagram with the regionswhere chaos is found, obtained using the rate model Eq. (12). Inthe black-shaded region, there is only one chaotic attractor,whereas in the cyan region the chaotic attractor coexists with aperiodic attractor. Dotted lines correspond to the bistabilityboundary of Fig. 1(a), depicted to facilitate comparison. (b) Cha-otic trajectory obtained simulating the FREs (12); the Lyapunovexponent is λ ¼ 0.183…. (c) Chaotic trajectory obtained from theQIF neurons—parameters corresponding to the (yellow) trianglesymbol in (a). (d) Time series for the rate model (orange) and QIFmodel (black). (e) Raster plot corresponding to 300 randomlyselected neurons. (f) Same raster plot as in (e), ordering neuronsaccording to their intrinsic current ηk. Parameters: I0 ¼ 3, ω ¼ πand η̄ ¼ −2.5, J ¼ 10.5 [triangle symbol in (a)].

FIG. 4. The conformal map Eq. (18) transforms the right half-plane onto the unit disk. This transformation and its inversedefine a one-to-one mapping between the Kuramoto orderparameter Z and the macroscopic quantities of the populationof QIF neurons: r, firing rate; v, mean membrane potential. Notethat in the limit Z ¼ eiΨ [full coherence, ~ρðθÞ ¼ δðθ −ΨÞ], werecover v ¼ ið1 − eiΨÞ=ð1þ eiΨÞ ¼ tanðΨ=2Þ, as the originalmapping V ¼ tanðθ=2Þ between QIF and theta neurons dictates.

ERNEST MONTBRIÓ, DIEGO PAZÓ, AND ALEX ROXIN PHYS. REV. X 5, 021028 (2015)

021028-6

Page 7: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

IV. CONCLUSIONS

We present a method for deriving firing-rate equationsfor a network of heterogeneous QIF neurons, which is exactin the thermodynamic limit. To our knowledge, the result-ing system of ordinary differential equations (12) repre-sents the first exact FREs of a network of spiking neurons.We emphasize that the derivation of the FREs does not

rely on assumptions of weak coupling, separation of timescales, averaging, or any other approximation. Rather, theassumptions underlying the validity of Eq. (12) are that(i) the QIF neurons are all-to-all connected, (ii) hetero-geneity is quenched, and (iii) inputs are distributed accord-ing to a Lorentzian distribution.The last two assumptions may seem particularly restric-

tive. Concerning assumption (iii), it must be stressed thatfor arbitrary distributions of quenched heterogeneity, theLA Eq. (5) remains generally valid, and, therefore, Eq. (10)can be used to discern the stability of the network states(see Appendix C). Furthermore, in Appendix C, we alsoshow numerical simulations using uniform- and Gaussian-distributed inputs that reveal very similar macroscopicdynamics in response to time-varying inputs. Even relaxingassumption (ii), numerical simulations with identical QIFneurons driven by external independent Gaussian noisesources show qualitative agreement with the FREs (12) (seeAppendix D). In summary, the choice of a quenchedLorentzian distribution gðηÞ is, thus, a mere mathematicalconvenience, whereas the insights gained from the resultingfiring-rate description are valid more generally.Our derivation represents a sharp departure from and a

major advance compared to previous studies in severalregards. First, in the past it has been possible to calculateonly the approximate firing rate of networks of spikingneurons for stationary states, or for weak deviations of suchstates [9–12,15,19,20]. The equations that result from thesederivations are difficult to solve, and typically requirespecial numerical methods. In contrast, FREs (12) canbe easily analyzed and simulated, and exactly reproduce thebehavior of the spiking network far from any fixed pointand for arbitrary external currents. Second, firing-ratedescriptions traditionally assume that the activity of apopulation of neurons is equivalent to a set of uncorrelatedstochastic processes with a given rate; see, e.g.,Refs. [13,14]. However, in simulations of spiking networks,it is well known that the response of the network tononstationary inputs generally involves some degree ofspike synchrony. Recent theoretical work has sought toimprove on classical rate models, which act as a low-passfilter of inputs, by fitting them with linear filters extractedfrom the corresponding Fokker-Planck equation for thenetwork [19,20]. These filters tend to generate dampedoscillations when the external noise level is not too high,reflecting the presence of spike synchrony [15,20]. TheFREs (12) also capture this phenomenon and reveal that theunderlying mechanism is an interplay between firing rate

and subthreshold voltage. Furthermore, because theseequations are exact, we can explore the full nonlinearresponse of the network, such as the generation of chaoticstates of synchronous bursting, as shown in Fig. 3.Recently, it has been shown that not only Kuramoto-like

models, but a much wider class of phase models, evolve inthe OA manifold defined by Eq. (14). Specifically, net-works of pulse-coupled oscillators [46] and theta neurons[48–50] allow for an exact, low-dimensional description interms of the Kuramoto order parameter [Eq. (17)].Although these later works use a finite-width phase-coupling function that differs from the synaptic coupling[Eq. (3)], the obtained low-dimensional description interms of the Kuramoto order parameter is analogous toours, but in a different space [56]. Indeed, we show that theOA ansatz Eq. (14) is related, via the nonlinear trans-formation of variables Eq. (15), to the LA ansatz Eq. (5).Remarkably, this transformation establishes an exact cor-respondence between the Kuramoto order parameter and anovel, biophysically meaningful macroscopic observablethat describes the firing rate and mean membrane potentialof the neuronal network. Interestingly, the low-dimensionaldescription in terms of firing rates seems to be a morenatural description for networks of spiking neurons, com-pared to that in terms of the Kuramoto order parameter. Thefiring-rate equations (12) take a surprisingly simple form inthe LA manifold, which makes them a valuable tool toexplore and understand the mechanisms governing themacroscopic dynamics of neuronal networks.Finally, since the OA ansatz is the asymptotic solution

for systems of the form in Eq. (16), applying the change ofvariables V ¼ tanðθ=2Þ automatically implies that the LAshould hold for populations governed by

_Vj ¼ Aðηj; tÞ þ Bðηj; tÞVj þ Cðηj; tÞV2j ; ð19Þ

where A, B, and C are related to Ω and H asA ¼ ½Ωþ ImðHÞ�=2, B ¼ −ReðHÞ, C ¼ ½Ω − ImðHÞ�=2.Notably, Eq. (19) defines a wide family of ensembles ofQIF neurons.Therefore, the LA is actually valid for more general

networks beyond the one investigated here—ηj in Eq. (19)may also be a vector containing several forms of disorder.As particularly relevant cases, in Appendix E, we providethe FREs governing the dynamics of an excitatory networkwith both heterogeneous inputs η and synaptic weights J, aswell as a pair of interacting excitatory and inhibitorypopulations of QIF neurons. In addition, according toEq. (19), the LA is also valid if synapses are modeledas conductances, in which case the reversal potentials maybe distributed as well. Moreover, the role of gap junctionsor synaptic kinetics may be considered within the sameframework.

MACROSCOPIC DESCRIPTION FOR NETWORKS OF … PHYS. REV. X 5, 021028 (2015)

021028-7

Page 8: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

ACKNOWLEDGMENTS

D. P. and A. R. acknowledge support by MINECO(Spain) under the Ramón y Cajal program. E. M. and A.R. acknowledge support from a grants from the SpanishMinistry of Economics and Competitiveness PSI2013-42091 and BFU2012-33413.

APPENDIX A: NUMERICAL SIMULATIONS

To numerically simulate the population of QIF neurons[Eq. (1)], we use the Euler method, with time stepdt ¼ 10−4. The population has N ¼ 104 neurons, andthe Lorentzian distribution Eq. (11) is deterministicallygenerated using ηj¼ η̄þΔarctan½π=2ð2j−N−1Þ=ðNþ1Þ�,where j ¼ 1;…; N and Δ ¼ 1.The time it takes for the membrane potential of a QIF

neuron (with Ij > 0) to reach infinity from a given positivevalue of the membrane potential is arctanð ffiffiffiffi

Ijp

=VpÞ=ffiffiffiffiIj

p.

ForffiffiffiffiIj

p≪ Vp, this expression can be approximated as

arctan ð ffiffiffiffiIj

p=VpÞffiffiffiffi

Ijp ≈

1

Vp:

In simulations, we consider Vp ¼ −Vr ¼ 100, and usethe previous approximation. Thus, the time for the neuronsto reach infinity from Vp is 1=Vp ≈ 10−2. Additionally, thetime taken from minus infinity to Vr is 1=Vr ≈ 10−2.Numerically, once a neuron’s membrane potential Vj

satisfies Vj ≥ Vp the neuron is reset to −Vj and held therefor a refractory time 2=Vj. The neurons produce a spikewhen Vj reaches infinity, i.e., a time interval 1=Vj aftercrossing Vp. The exact time of the spike cannot beevaluated exactly in numerical simulations, since dt isfinite. However, simulations agree very well with thetheory, provided that 1=Vp ≫ dt.To evaluate the mean membrane potential v, the pop-

ulation average is computed discarding those neurons in therefractory period. The choice Vp ¼ −Vr is not needed, butin this way the observed mean membrane potential v agreeswith the theory, consistent with the definition in Eq. (8).The mean synaptic activation Eq. (3) is evaluated usingthe Heaviside step function aτðtÞ ¼ Θðτ − tÞ=τ, withτ ¼ 10−3. The instantaneous firing rates are obtainedbinning time and counting spikes within a slidingtime window of size δt ¼ 2 × 10−2 (Figs. 2,6–8) andδt ¼ 4 × 10−2 (Fig. 3).

APPENDIX B: PROOF OF EQ. (18)

The inverse of Eq. (18) reads

W ¼ 1 − Z�

1þ Z� : ðB1Þ

Recalling Eqs. (7) and (9), we write the macroscopicfield W as

WðtÞ ¼Z

−∞wðη; tÞgðηÞdη;

where w≡ xþ iy. Inserting the conformal mappingw ¼ ð1 − αÞ=ð1þ αÞ—the inverse of Eq. (15)—we get

WðtÞ ¼Z

−∞

�1 − αðη; tÞ1þ αðη; tÞ

�gðηÞdη:

Using the geometric series formula, and grouping thepowers of α, we may evaluate the integrals, obtaining

WðtÞ ¼ 1 − 2Z�ðtÞ þ 2Z�2ðtÞ − 2Z�

3ðtÞ þ � � � ; ðB2Þ

where the Zm are the generalized order parameters [57]:

ZmðtÞ ¼Z

−∞gðηÞ

Z2π

0

~ρðθjη; tÞeimθdθdη

¼Z

−∞gðηÞ½α�ðη; tÞ�mdη:

Since, for Lorentzian gðηÞ, ZmðtÞ¼½α�ðη¼ η̄−iΔ;tÞ�m ¼ZðtÞm, we can revert the power series in Eq. (B2) and obtainthe result in Eq. (B1).

APPENDIX C: RESULTS FOR ARBITRARYDISTRIBUTIONS OF CURRENTS

In this Appendix, we compare the results in the maintext, obtained using a Lorentzian distribution of currentsgðηÞ, with other distributions. The results are qualitativelysimilar, as evidenced by Figs. 5–7 (cf. Figs. 1–3).Note that, if a particular distribution gðηÞ has 2n poles

(all of them off the real axis). one can readily obtain theFREs consisting of n complex-valued ordinary differentialequations. Even if gðηÞ does not fulfill this condition,as in the case of Gaussian or uniform distributions, theLorentzian ansatz still holds (see below).Moreover, it is possible to efficiently simulate the

dynamics of a population by integrating Eq. (10) for asample of η values that approximate a particular gðηÞ [58].However, if gðηÞ is a discrete distribution (or has a discretepart), as in the case of the Dirac delta functiongðηÞ ¼ δðη − η̄Þ, the so-called Watanabe-Strogatz theoryapplies [47,60,61], and the “Lorentzian manifold” is nolonger attracting since the dynamics is extremely degen-erate, akin to a Hamiltonian system.

1. Steady states

Considering IðtÞ ¼ 0 in Eqs. (1) and (2), it is possible tofind the solutions with steady firing rate rðtÞ ¼ r0 forarbitrary distributions of currents.

ERNEST MONTBRIÓ, DIEGO PAZÓ, AND ALEX ROXIN PHYS. REV. X 5, 021028 (2015)

021028-8

Page 9: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

As we point out in the main text, the stationary solutionof the continuity equation (4) for those neurons that areintrinsically active (η > −Jr0) must be inversely propor-tional to their speed; that is,

ρ0ðVjηÞ ¼cðηÞ

V2 þ ηþ Jr0;

where the normalizing constant is cðηÞ ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiηþ Jr0

p=π. The

firing rate for each value of η is then

r0ðηÞ ¼ ρ0ðV → ∞jηÞ _VðV → ∞jη; tÞ ¼ cðηÞ;

and integration over all the firing neurons gives theself-consistent condition for r0:

r0 ¼Z

−Jr0cðηÞgðηÞdη; ðC1Þ

which is valid for any distribution gðηÞ [an analogousexpression was obtained in Eq. (45) of Ref. [40]]. ForLorentzian gðηÞ, the integral in Eq. (C1) can be evaluated,and solving the resulting equation for r0, one obtains thesteady states in agreement with the result of setting_r ¼ _v ¼ 0 in the FREs (12).

2. Linear stability analysis of steady states

Assuming that the LA captures the actual dynamics ofthe population of QIF neurons, one can also investigate thestability of steady states for arbitrary distributions gðηÞ.Indeed, the evolution of infinitesimal perturbations from asteady state,

w0ðηÞ≡ x0ðηÞ þ iy0ðηÞ ¼� ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

ηþ Jr0p

if η > −Jr0−i ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi−η − Jr0

pif η ≤ −Jr0;

is determined by the linearization of Eq. (10):

∂tδwðη; tÞ ¼ iðJδr − 2w0δwÞ; ðC2Þ

where δrðtÞ ¼ ð2πÞ−1 R∞−∞½δwðη; tÞ þ δw�ðη; tÞ�gðηÞdη.To find the discrete spectrum of eigenvalues [62], we use

the ansatz δwðη; tÞ ¼ bðηÞeλt in Eq. (C2) and obtain

½λþ 2w0ðηÞi�bðηÞeλt ¼iJ2π

Z∞

−∞½bðηÞeλt þ b�ðηÞeλ�t�gðηÞdη:

Solving the equation for beλt, we find

bðηÞeλt ¼ 1

2w0ðηÞ − iλJ2π

Z∞

−∞½bðηÞeλt þ b�ðηÞeλ�t�gðηÞdη:

After summing the complex conjugate at each side of theequation, we multiply by gðηÞ and integrate over η. This

-2 -1 0

η/γ

0

10

20J/

γ1/2

-3 -2 -1 0

η/σ

0

10

20

J/σ1/

2

(a) (b)

FIG. 5. Phase diagrams for (a) uniform and (b) Gaussiandistributions of currents gðηÞ. The cyan-shaded region representsthe bistable region, with the solid lines corresponding to saddle-node bifurcations analytically obtained from Eqs. (C5) and (C6).Square symbols are estimates of the bifurcations’ loci obtained bydirect numerical simulations of N ¼ 104 QIF neurons. Trianglesymbols indicate parameter values used in numerical simulationsof Figs. 6 and 7.

-6-4-202

v

0

3

-10 0 10 20 30 40 50 60 70 80time

-1.50

1.5

-2

0

2

0

3

r0

300

Neu

ron

inde

x

030

0

-10 0 10 20 30 40 50 60 70 80time

-101

I(t)

(a) (e)

(f)

(g)

(h)

(b)

(c)

(d)

FIG. 6. Numerical simulations of the excitatory network of QIF neurons, Eqs. (1) and (2). with (a)–(d) uniform and (e)–(h) Gaussiandistributions of currents. As in Fig. 2, at time t ¼ 0, an external sinusoidal current IðtÞ ¼ I0 sinðωtÞ—shown in (d) and (h)—is applied toall neurons. (a),(e) Time series of the firing rate and (b),(f) the mean membrane potential of all cells. In (c) and (g), we depict the rasterplots of 300 randomly chosen neurons. Parameters correspond to the black triangle symbols in Fig. 5: (a)–(d) η̄ ¼ −1, J ¼ 8, γ ¼ 1,I0 ¼ 1; (e)–(h) η̄ ¼ −2, J ¼ 10, σ ¼ 1, I0 ¼ 1.5.

MACROSCOPIC DESCRIPTION FOR NETWORKS OF … PHYS. REV. X 5, 021028 (2015)

021028-9

Page 10: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

allows one to cancel out the integralsR ðbeλt þ b�eλ�tÞgdη.

Then, imposing self-consistency, we find

1 ¼ J2π

Z∞

−∞

�1

2w0ðηÞ − iλþ c:c:

�gðηÞdη: ðC3Þ

Local bifurcations are associated with marginal stability ofthe fixed points: ReðλÞ → 0. In correspondence with theresults obtained for a Lorentzian distribution in the maintext, we expect bifurcations of saddle-node type associatedwith λ ¼ 0. Therefore, from Eq. (C3) we find that the lociof the saddle-node bifurcations are located at

JSN ¼ 2πR∞−∞

x0ðηÞjw0ðηÞj2 gðηÞdη

: ðC4Þ

Since x0 ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiηþ Jr0

pfor η > −Jr0, and zero otherwise, we

can restrict the integration over η to the range ð−Jr0;∞Þ.Now, multiplying Eq. (C4) by Eq. (C1), we get, afterdefining ξ ¼ JSNr0, two equations that permit us to find thelocus of the saddle-node bifurcations systematically forarbitrary distributions of currents:

JSN ¼ 2πR∞0 η−1=2gðη − ξÞdη ; ðC5Þ

with ξ obtained solving

ξ ¼ 2R∞0 η1=2gðη − ξÞdηR∞

0 η−1=2gðη − ξÞdη : ðC6Þ

3. Bistability region for uniformand Gaussian distributions

Equations (C5) and (C6) are particularly easy to solve foruniform distributions of currents:

gðηÞ ¼� 1

2γ for jη − η̄j < γ

0 otherwise:

After defining the rescaled parameters ~η ¼ η̄=γ and~J ¼ J=

ffiffiffiγ

p, we find two branches of saddle-node bifurca-

tions emanating from the cusp point at ~η ¼ −1=3:

~Jð1ÞSN ¼ 2πffiffiffiffiffiffiffiffiffiffiffiffiffi3~ηþ 3

p ;

~Jð2ÞSN ¼ 2πffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi~ηþ 1þ 2

ffiffiffiffiffiffiffiffiffiffiffiffi13þ ~η2

qr−

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi~η − 1þ 2

ffiffiffiffiffiffiffiffiffiffiffiffi13þ ~η2

qr :

These functions are plotted in Fig. 5(b). Numerical sim-ulations using the network of QIF neurons confirm thecorrectness of these boundaries (see the square symbolsin Fig. 5).For Gaussian distributions,

gðηÞ ¼ 1ffiffiffiffiffiffi2π

pσe−ðη−η̄Þ2=ð2σ2Þ;

the solutions of Eqs. (C5) and (C6) can be numericallyfound. The results are shown in Fig. 5(b). Again, there is aperfect agreement between Eqs. (C5) and (C6)—derivedassuming the Lorentzian Ansatz—and the numerical esti-mations obtained simulating the network of QIF neurons.

4. Excitatory networks with external periodic currents

Firing-rate equations (12) predict the existence of astable focus in the shaded region of Fig. 1. Trajectoriesattracted to this fixed point display oscillations in the firingrate and mean membrane potential due to the transientsynchronous firing of the QIF neurons.When an external periodic current is injected to all

neurons in the network, the spiral dynamics around thefixed point is responsible for the bursting behaviorobserved in Fig. 2, as well as for the emergence of themacroscopic chaos shown in Fig. 3. Here, we investigatewhether similar phenomena occur when an external

0

1r

0 20 40 60time

030

0O

rder

ed n

euro

n in

dex

030

0N

euro

n in

dex

(a)

(b)

(c)

FIG. 7. Chaotic state in a network ofN ¼ 104 QIF neurons withGaussian-distributed currents. Network parameters correspond tothe yellow triangle in the phase diagram of Fig. 5(b). As in Fig. 3,neurons receive a common periodic current IðtÞ ¼ I0 sinðωtÞ offrequency ω ¼ π. Parameters: η̄ ¼ −1, J ¼ 8, σ ¼ 1, I0 ¼ 1.5.

ERNEST MONTBRIÓ, DIEGO PAZÓ, AND ALEX ROXIN PHYS. REV. X 5, 021028 (2015)

021028-10

Page 11: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

periodic current of the same frequency is injected in anexcitatory network with either uniform or Gaussian-distributed currents.We introduce a sinusoidal forcing IðtÞ ¼ I0 sinðωtÞ

observing a behavior qualitatively identical to the onereported in the main text with Lorentzian gðηÞ. Under alow-frequency forcing, the system displays periodic burst-ing, see Fig. 6, provided the parameters are set inside thebistable region (see the black triangles in Fig. 5).Furthermore, note that the range of firing rates andmean membrane potentials in Fig. 6 are similar to thoseof Fig. 2—for Lorentzian distributions of currents.As for the simulations shown in Fig. 3, we next increase

the frequency of the injected current up to ω ¼ π toinvestigate the emergence of macroscopic chaos in anetwork with Gaussian-distributed currents. Figure. 7shows the emergence of an apparently chaotic state. Theobserved dynamics is similar to that of Fig. 3, whichsuggests it is chaotic. This type of chaos persists in thethermodynamic limit N → ∞. On top of this, highlydimensional but weakly chaotic dynamics is probably alsopresent due to “residual” finite-size effects.

APPENDIX D: IDENTICAL QIF NEURONSDRIVEN BY INDEPENDENT NOISE TERMS

Now, we compare the results obtained above and in themain text for quenched heterogeneity gðηÞ with the resultsfor identical neurons gðηÞ ¼ δðη − η̄Þ driven by noise.Specifically, the input currents are now taken as

Ij ¼ η̄þ JrðtÞ þ ξjðtÞ; ðD1Þ

where ξj are independent white noise terms with expectedvalues hξjðtÞi ¼ 0, and hξjðtÞξkðt0Þi ¼ 2Dδjkδðt − t0Þ.In the thermodynamic limit, the density ρðV; tÞ obeys the

Fokker-Planck equation:

∂tρþ ∂V ½ðV2 þ η̄þ JrÞρ� ¼ D∂2Vρ: ðD2Þ

The Lorentzian ansatz is not a solution of this equation, butwe demonstrate here that the phenomena observed withquenched heterogeneity also arise with independent noisesources. This qualitative similarity at the macroscopic levelbetween quenched Lorentzian heterogeneity and Gaussiannoise has been noted in previous work [63].A subtle point in Eq. (D2) is that its nondimensionaliza-

tion entails a different (compared to Δ) rescaling with D:~V¼V=D1=3, ~η ¼ η=D2=3, ~J ¼ J=D1=3, ~t ¼ tD1=3 (implying~r ¼ r=D1=3).Numerical simulations reveal the existence of a region of

bistability, see Fig. 8(a), analogous to the one observed forquenched heterogeneity, cf. Figs. 1(a) and 5. Obviously,true bistability holds only in the thermodynamic limit,while what we observe are rather exceedingly long resi-dence times close to each fixed point (see Fig. 8 caption fordetails).Additionally, in order to investigate and compare the

dynamical behavior of the identical noise-driven neuronswith that of the FREs (12), an external current of intensity

-2

0

2

v

-10 0 10 20 30 40time

0

3

I(t)

0

3

r0

300

Neu

ron

inde

x

-10 -5 0

η/D2/3

0

5

10

15

20

J/D

1/3

(a)

(c)

(d)

(e)

(b)

FIG. 8. Numerically obtained phase diagram (a) and time series (b)–(d) for a network of identical QIF neurons driven by independentwhite noise terms. The square symbols in (a) represent the boundaries that enclose the region of bistability between a high- and a low-activity state. The boundaries have been obtained using a network of 104 neurons, and by numerical continuation of a low-activitysolution (filled squares) and a high-activity solution (empty squares). Specifically, using a noise intensityD ¼ 1 and a particular value ofJ, the system is initialized either at η̄ ¼ −5 or at η̄ ¼ 0 and, after t ¼ 10 time units, parameter η̄ is decreased or increased an amount0.025 or 0.1, respectively. This continuation is made until the relative change of two successive values of the averaged firing rate (timeaveraged over the last time unit for each parameter value) is larger than 50%. Other parameters for the numerical simulations are thesame as in all other figures, and are described in Appendix A. The triangle symbol indicates the parameter value,ðη̄=D2=3; J=D1=3Þ ¼ ð−5; 15Þ, corresponding to the numerical simulations shown in black in the right-hand panels—for the simulations.we use D ¼ 1. To facilitate the comparison with the FREs (12), the orange curves show the time series of the FREs, using the sameparameters ðη̄; JÞ ¼ ð−5; 15Þ, but with Δ ¼ 1.

MACROSCOPIC DESCRIPTION FOR NETWORKS OF … PHYS. REV. X 5, 021028 (2015)

021028-11

Page 12: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

I0 ¼ 3 is injected to all neurons at time t ¼ 0, like in Fig. 2.In Figs. 8(b) and 8(c), the time series of the firing rate andthe mean membrane potential clearly display dampedoscillations after the injection of the current, confirmingthe existence of a stable focus, exactly as observed in theFREs (12). The existence of a stable focus reflects thepresence of transient spike synchrony in the network, asseen by the raster plot in Fig. 8(d). Remarkably, the rasterplot and the time series closely resemble those of Fig. 2.The resemblance is not just qualitative, but rather there is anear quantitative fit between the network of QIF neuronsdriven by Gaussian noise and the FREs (12), which arederived assuming quenched Lorentzian heterogeneity(orange curves).

APPENDIX E: MODEL GENERALIZATIONS

To illustrate the potential of the LA for investigatingmore sophisticated networks, here we provide the low-dimensional FREs corresponding to an excitatory networkof QIF neurons with independently distributed currents andsynaptic weights and to two interacting populations ofexcitatory and inhibitory QIF neurons with distributedcurrents.

1. Excitatory population with heterogeneouscurrents and synaptic weights

As a first example, let us assume that both the currents ηand the synaptic weights J are distributed—this type ofheterogeneity was also considered in Ref. [40]. The inputcurrents then read [66]

Ij ¼ ηj þ JjrðtÞ þ IðtÞ:

For simplicity, we additionally assume that η and J aredistributed independently, with a joint distributionpðη; JÞ ¼ gðηÞhðJÞ. In the simplest situation ofLorentzian gðηÞ and hðJÞ,

gðηÞ ¼ Δ=πðη − η̄Þ2 þ Δ2

; hðJÞ ¼ Γ=πðJ − J̄Þ2 þ Γ2

;

the problem is extremely simplified using the Lorentzianansatz, which now trivially reads

ρðVjη; J; tÞ ¼ 1

π

xðη; J; tÞ½V − yðη; J; tÞ�2 þ xðη; J; tÞ2 :

The firing rate and mean membrane potential are deter-mined only by the value of w≡ xþ iy at the poles in thelower half-planes:

rðtÞ þ ivðtÞ ¼ZZ

wðη; J; tÞgðηÞhðJÞdηdJ

¼ wðη̄ − iΔ; J̄ − iΓ; tÞ:

Finally, evaluating Eq. (10) at the poles (η ¼ η̄ − iΔ,J ¼ J̄ − iΓ), we get the exact FREs:

_r ¼ Δ=π þ Γr=π þ 2rv; ðE1aÞ

_v ¼ v2 þ η̄þ J̄rþ IðtÞ − π2r2: ðE1bÞ

These equations simply contain an extra term þΓr=πcompared to Eq. (12). Figure 9 shows the bistabilityboundaries obtained from Eq. (E1) for different ratios ofΓ and Δ, and IðtÞ ¼ 0. Note that the region of bistabilityshifts to lower values of η̄=Δ and to higher values of J̄=

ffiffiffiffiΔ

pas the level of heterogeneity in the synaptic coupling Γ isincreased.

2. Firing-rate equations for a pair ofexcitatory-inhibitory populations

The microscopic state of each population of QIFneurons is characterized by the membrane potentials

fVðe;iÞj gj¼1;…;N , which obey the following ordinary differ-

ential equations:

_Vðe;iÞj ¼ ðVðe;iÞ

j Þ2þ Iðe;iÞj ; if Vðe;iÞj ≥ Vp; then Vðe;iÞ

j ← Vr:

Here, Vðe;iÞj represents the membrane potential of neuron j

in either the excitatory ðeÞ or the inhibitory ðiÞ population.The external currents for the excitatory and inhibitorypopulations are, respectively,

IðeÞj ¼ ηðeÞj þ JeesðeÞðtÞ − JiesðiÞðtÞ þ IðeÞðtÞ;IðiÞj ¼ ηðiÞj þ JeisðeÞðtÞ − JiisðiÞðtÞ þ IðiÞðtÞ;

where the synaptic weights are Jee; Jii; Jie; Jei. Finally, themean synaptic activation for each population is

05-01-η/Δ

0

5

10

15

20

J/ Δ

1/2 Γ/ Δ1/2= 0

Γ/ Δ1/2= 1

Γ/Δ1/2= 5

FIG. 9. Phase diagram for an excitatory population withheterogeneous currents and synaptic weights obtained usingthe FREs (E1).

ERNEST MONTBRIÓ, DIEGO PAZÓ, AND ALEX ROXIN PHYS. REV. X 5, 021028 (2015)

021028-12

Page 13: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

sðe;iÞðtÞ ¼ 1

N

XNj¼1

Xknðtkj Þðe;iÞ<t

Zt

−∞dt0aτðt − t0Þδðt0 − ðtkjÞðe;iÞÞ:

Here, ðtkjÞðe;iÞ is the time of the kth spike of the jth neuron ineither the excitatory ðeÞ or the inhibitory ðiÞ population.Additionally, δðtÞ is the Dirac delta function, and aτðtÞ isthe normalized synaptic activation caused by a single pre-synaptic spike with time scale τ, e.g., aτðtÞ ¼ e−t=τ=τ.It is straightforward to apply the LA and the method

described in the main text to obtain the firing-rate equationscorresponding to the two-population model. Consideringthe limit of infinitely fast synapses, τ → 0, we getsðe;iÞðtÞ ¼ rðe;iÞðtÞ. Finally, assuming the Lorentzian dis-tributions of currents for both populations,

ge;iðηÞ ¼1

π

Δe;i

ðη − η̄e;iÞ2 þ Δ2e;i;

we obtain the firing-rate equations:

_rðeÞ ¼ Δe=π þ 2rðeÞvðeÞ;

_vðeÞ ¼ ðvðeÞÞ2 þ η̄e þ JeerðeÞ − JierðiÞ þ IðeÞðtÞ − ðπrðeÞÞ2;_rðiÞ ¼ Δi=π þ 2rðiÞvðiÞ;

_vðiÞ ¼ ðvðiÞÞ2 þ η̄i þ JeirðeÞ − JiirðiÞ þ IðiÞðtÞ − ðπrðiÞÞ2:

[1] B. B. Averbeck, P. E. Latham, and A. Pouget, Neuralcorrelations, population coding and computation, Nat.Rev. Neurosci. 7, 358 (2006).

[2] A. L. Hodgkin and A. F. Huxley, A quantitative descriptionof membrane current and its application to conduction andexcitation in nerve, J. Physiol. 117, 500 (1952).

[3] R. Brette, M. Rudolph, T. Carnevale, M. Hines, D. Beeman,J. M. Bower, M. Diesmann, A. Morrison, P. H. Goodman,F. C. Harris, Jr. et al., Simulation of networks of spikingneurons: A review of tools and strategies, J. Comput.Neurosci. 23, 349 (2007).

[4] E. M. Izhikevich and G. M. Edelman, Large-scale model ofmammalian thalamocortical systems, Proc. Natl. Acad. Sci.U.S.A. 105, 3593 (2008).

[5] E. M. Izhikevich, Dynamical systems in neuroscience(MIT Press, Cambridge, MA, 2007).

[6] H. R. Wilson and J. D. Cowan, Excitatory and inhibitoryinteractions in localized populations of model neurons,Biophys. J. 12, 1 (1972).

[7] S.-i. Amari, A method of statistical neurodynamics,Kybernetik 14, 201 (1974).

[8] W. J. Freeman, Mass action in the nervous system(Academic Press, New York, 1975).

[9] W. Gerstner, Time structure of the activity in neural networkmodels, Phys. Rev. E 51, 738 (1995).

[10] S. Fusi and M. Mattia, Collective behavior of networks withlinear (VLSI) integrate-and-fire neurons, Neural Comput.11, 633 (1998).

[11] D. J. Amit and N. Brunel, Model of global spontaneousactivity and local structured activity during delay periods inthe cerebral cortex, Cereb. Cortex 7, 237 (1997).

[12] N. Brunel, Dynamics of sparsely connected networks ofexcitatory and inhibitory spiking neurons, J. Comput.Neurosci. 8, 183 (2000).

[13] P. Dayan and L. F. Abbott, Theoretical neuroscience (MITPress, Cambridge, MA, 2001).

[14] W. Gerstner and W.M. Kistler, Spiking neuron models:Single neurons, populations, plasticity (CambridgeUniversity Press, Cambridge, England, 2002).

[15] O. Shriki, D. Hansel, and H. Sompolinsky, Rate models forconductance-based cortical neuronal networks, NeuralComput. 15, 1809 (2003).

[16] S. Coombes, Waves, bumps, and patterns in neural fieldtheories, Biol. Cybern. 93, 91 (2005).

[17] G. Deco, V. K. Jirsa, P. A. Robinson, M. Breakspear, andK. Friston, The dynamic brain: From spiking neurons toneural masses and cortical fields, PLoS Comput. Biol. 4,e1000092 (2008).

[18] G. B. Ermentrout and D. H. Terman, Mathematical founda-tions of neuroscience (Springer, New York, 2010), Vol. 64.

[19] S. Ostojic and N. Brunel, From spiking neuron models tolinear-nonlinear models, PLoS Comput. Biol. 7, e1001056(2011).

[20] E. S. Schaffer, S. Ostojic, and L. F. Abbott, A complex-valued firing-rate model that approximates the dynamicsof spiking networks, PLoS Comput. Biol. 9, e1003301(2013).

[21] J. J. Hopfield, Neurons with graded response have collec-tive computational properties like those of two-stateneurons, Proc. Natl. Acad. Sci. U.S.A. 81, 3088 (1984).

[22] G. Mongillo, O. Barak, and M. Tsodyks, Synaptic theory ofworking memory, Science 319, 1543 (2008).

[23] R. Ben-Yishai, R. Lev Bar-Or, and H. Sompolinsky, Theoryof orientation tuning in visual cortex, Proc. Natl. Acad. Sci.U.S.A. 92, 3844 (1995).

[24] D. Hansel and H. Sompolinsky, in Methods in neuronalmodelling: From ions to networks, edited by C. Koch andI. Segev (MIT Press, Cambridge, MA, 1998), pp. 499–567.

[25] R. Moreno-Bote, J. Rinzel, and N. Rubin, Noise-inducedalternations in an attractor network model of perceptualbistability, J. Neurophysiol. 98, 1125 (2007).

[26] K. Zhang, Representation of spatial orientation by theintrinsic dynamics of the head-direction cell ensemble:A theory, J. Neurosci. 16, 2112 (1996).

[27] K.-F. Wong and X.-J. Wang, A recurrent networkmechanism of time integration in perceptual decisions,J. Neurosci. 26, 1314 (2006).

[28] M. Abeles, Corticonics: Neural Circuits of the CerebralCortex (Cambridge University Press, Cambridge, England,1991).

[29] F. Varela, J.-P. Lachaux, E. Rodriguez, and J. Martinerie,The brainweb: Phase synchronization and large-scaleintegration, Nat. Rev. Neurosci. 2, 229 (2001).

[30] A. K. Engel and W. Singer, Temporal binding and theneural correlates of sensory awareness, Trends Cogn. Sci.5, 16 (2001).

MACROSCOPIC DESCRIPTION FOR NETWORKS OF … PHYS. REV. X 5, 021028 (2015)

021028-13

Page 14: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

[31] L. M. Ward, Synchronous neural oscillations and cognitiveorocesses, Trends Cogn. Sci. 7, 553 (2003).

[32] P. Fries, A mechanism for cognitive dynamics: Neuronalcommunication through neuronal coherence, Trends Cogn.Sci. 9, 474 (2005).

[33] G. Buzsaki, Rhythms of the Brain (Oxford University Press,New York, 2006).

[34] J. Fell and N. Axmacher, The role of phase synchronizationin memory processes, Nat. Rev. Neurosci. 12, 105 (2011).

[35] E. Ott and T. M. Antonsen, Low dimensional behavior oflarge systems of globally coupled oscillators, Chaos 18,037113 (2008).

[36] A. L. Hodgkin, The local electric changes associated withrepetitive action in a non-medullated axon, J. Physiol. 107,165 (1948).

[37] J. Rinzel and B. Ermentrout, in Methods in NeuronalModelling: From Ions to Networks, edited by C. Kochand I. Segev (MIT Press, Cambridge, MA, 1998),pp. 251–291.

[38] B. Ermentrout and N. Kopell, Parabolic bursting in anexcitable system coupled with a slow oscillation, SIAM J.Appl. Math. 46, 233 (1986).

[39] P. E. Latham, B. J. Richmond, P. G. Nelson, and S.Nirenberg, Intrinsic dynamics in neuronal networks. I.Theory, J. Neurophysiol. 83, 808 (2000).

[40] M. A. Buice and C. C. Chow, Dynamic finite size effects inspiking neural networks, PLoS Comput. Biol. 9, e1002872(2013).

[41] We make an analytic continuation of wðη; tÞ from real ηinto complex-valued η ¼ ηr þ iηi. This is possible intothe lower half-plane ηi < 0, since this guarantees that thehalf-width xðη; tÞ remains positive zero: ∂txðη; tÞ¼−ηi > 0at x ¼ 0. Therefore, we close the integrals in Eqs. (7)and (9) with an arc jηjeiϑ, with jηj → ∞ and ϑ ∈ ð−π; 0Þ.This contour encloses one pole of the Lorentzian distribu-tion [Eq. (11)], which, written in partial fractions, reads asfollows: gðηÞ ¼ ð2πiÞ−1½ðη − η̄ − iΔÞ−1 − ðη − η̄þ iΔÞ−1�.

[42] The number of effective parameters can be reduced bynondimensionalizing the system as ~η ¼ η̄=Δ, ~J ¼ J=

ffiffiffiffiΔ

p,

ð~r; ~vÞ ¼ ðr; vÞ= ffiffiffiffiΔ

p, ~t ¼ ffiffiffiffi

Δp

t, and ~Ið~tÞ ¼ Iðt= ffiffiffiffiΔ

p Þ=Δ.[43] Y. Kuramoto, Chemical Oscillations, Waves, and

Turbulence (Springer-Verlag, Berlin, 1984).[44] A. S. Pikovsky, M. G. Rosenblum, and J. Kurths,

Synchronization, A Universal Concept in NonlinearSciences (Cambridge University Press, Cambridge,England, 2001).

[45] E. Ott and T. M. Antonsen, Long time evolution of phaseoscillator systems, Chaos 19, 023117 (2009); E. Ott, B. R.Hunt, and T. M. Antonsen, Comment on “long timeevolution of phase oscillators systems”, Chaos 21,025112 (2011).

[46] D. Pazó and E. Montbrió, Low-Dimensional Dynamics ofPopulations of Pulse-Coupled Oscillators, Phys. Rev. X 4,011009 (2014).

[47] A. Pikovsky and M. Rosenblum, Dynamics of hetero-geneous oscillator ensembles in terms of collective varia-bles, Physica (Amsterdam) 240D, 872 (2011).

[48] T. B. Luke, E. Barreto, and P. So, Complete classification ofthe macroscopic behavior of a heterogeneous network oftheta neurons, Neural Comput. 25, 3207 (2013).

[49] P. So, T. B. Luke, and E. Barreto, Networks of theta neuronswith time-varying excitability: Macroscopic chaos, multi-stability, and final-state uncertainty, Physica (Amsterdam)267D, 16 (2014).

[50] C. R. Laing, Derivation of a neural field model from anetwork of theta neurons, Phys. Rev. E 90, 010901(2014).

[51] E. Montbrió and D. Pazó, Shear Diversity PreventsCollective Synchronization, Phys. Rev. Lett. 106, 254101(2011).

[52] D. Pazó and E. Montbrió, The Kuramoto modelwith distributed shear, Europhys. Lett. 95, 60007(2011).

[53] E. Montbrió and D. Pazó, Collective synchronization inthe presence of reactive coupling and shear diversity,Phys. Rev. E 84, 046206 (2011).

[54] D. Iatsenko, S. Petkoski, P. V. E. McClintock, and A.Stefanovska, Stationary and Traveling Wave States of theKuramoto Model with an Arbitrary Distribution ofFrequencies and Coupling Strengths, Phys. Rev. Lett.110, 064101 (2013).

[55] D. Iatsenko, P. V. E. McClintock, and A. Stefanovska,Oscillator glass in the generalized Kuramoto model:Synchronous disorder and two-step relaxation, Nat.Commun. 5, 4188 (2014).

[56] However, note that the phase diagrams in Luke et al. [48] donot fully agree with ours and show additional instabilities.These bifurcations are associated with the finite width of thepulses, as it occurs in the Winfree model [46].

[57] H. Daido, Onset of cooperative entrainment in limit-cycleoscillators with uniform all-to-all interactions: Bifurcationof the order function, Physica (Amsterdam) 91D, 24(1996).

[58] Equation (10) then becomes a reversible system [59] ofordinary differential equations with the invariance undertime reversal and w → w�, e.g., Eq. (12) for Δ ¼ 0 isreversible, with invariance under t → −t, v → −v. It isadvisable to approximate Eq. (10) by a set of ordinarydifferential equations (one for each η value) with additionalsmall terms breaking time reversibility. This is achieved byincluding a small term in each ordinary differential equation,like Δ in Eq. (12a), so that gðηÞ is actually approximated bya sum of very sharp Lorentzian functions.

[59] J. A. G. Roberts and G. R.W. Quispel, Chaos andtime-reversal symmetry. Order and chaos in reversibledynamical systems, Phys. Rep. 216, 63 (1992).

[60] S. Watanabe and S. H. Strogatz, Constant of motion forsuperconducting Josephson arrays, Physica (Amsterdam)74D, 197 (1994).

[61] A. Pikovsky and M. Rosenblum, Partially integrabledynamics of hierarchical populations of coupled oscillators,Phys. Rev. Lett. 101, 264103 (2008).

[62] A subtle question is that, as it occurs in the Kuramotomodel, there exists a continuous spectrum of eigenvalues atλ ¼ −2w0ðηÞi and the complex conjugate. Indeed, for ηvalues such that neurons are actively firing [y0ðηÞ ¼ 0], theeigenvalues lie exactly on the imaginary axis. Apparently,this seems to forbid the exponential stability obtained fromEq. (12) for the Lorentzian distribution. The same math-ematical conundrum was encountered and solved in the

ERNEST MONTBRIÓ, DIEGO PAZÓ, AND ALEX ROXIN PHYS. REV. X 5, 021028 (2015)

021028-14

Page 15: Macroscopic Description for Networks of Spiking …materias.df.uba.ar/dnla2019c1/files/2019/03/PhysRevX.5...Kuramoto order parameter [35]. II. MODEL DESCRIPTION Hodgkin-Huxley-type

Kuramoto model [63–65], which suggests that similarmechanisms are operating here.

[63] S. H. Strogatz and R. E. Mirollo, Stability of incoherence ina population of coupled oscillators, J. Stat. Phys. 63, 613(1991).

[64] S. H. Strogatz, R. E. Mirollo, and P. C. Matthews,Coupled Nonlinear Oscillators below the Synchronization

Threshold: Relaxation by Generalized Landau Damping,Phys. Rev. Lett. 68, 2730 (1992).

[65] H. Chiba and I. Nishikawa, Center manifold reduction forlarge populations of globally coupled phase oscillators,Chaos 21, 043103 (2011).

[66] Note that considering stronger forms of heterogeneity forthe synaptic weights may generally not lead to Eq. (19).

MACROSCOPIC DESCRIPTION FOR NETWORKS OF … PHYS. REV. X 5, 021028 (2015)

021028-15