spatially structured oscillations in a two-dimensional

17
J Comput Neurosci (2010) 28:193–209 DOI 10.1007/s10827-009-0199-6 Spatially structured oscillations in a two-dimensional excitatory neuronal network with synaptic depression Zachary P. Kilpatrick · Paul C. Bressloff Received: 10 April 2009 / Revised: 14 October 2009 / Accepted: 16 October 2009 / Published online: 29 October 2009 © Springer Science + Business Media, LLC 2009 Abstract We study the spatiotemporal dynamics of a two-dimensional excitatory neuronal network with synaptic depression. Coupling between populations of neurons is taken to be nonlocal, while depression is taken to be local and presynaptic. We show that the network supports a wide range of spatially structured oscillations, which are suggestive of phenomena seen in cortical slice experiments and in vivo. The particular form of the oscillations depends on initial conditions and the level of background noise. Given an initial, spa- tially localized stimulus, activity evolves to a spatially localized oscillating core that periodically emits target waves. Low levels of noise can spontaneously generate several pockets of oscillatory activity that interact via their target patterns. Periodic activity in space can also organize into spiral waves, provided that there is some source of rotational symmetry breaking due to external stimuli or noise. In the high gain limit, no oscillatory behavior exists, but a transient stimulus can lead to a single, outward propagating target wave. Keywords Excitatory neuronal network · Synaptic depression · Oscillations · Spiral waves · Stationary bumps Action Editor: Bard Ermentrout Z. P. Kilpatrick · P. C. Bressloff Department of Mathematics, University of Utah, Salt Lake City, UT 84112-0090, USA P. C. Bressloff (B ) Mathematical Institute, University of Oxford, 24-29 St. Giles’, Oxford OX1 3LB, UK e-mail: [email protected] 1 Introduction Spatially localized oscillations arise both in vivo and in vitro and may be observed experimentally using multi-electrode arrays or voltage-sensitive dye imaging (Wu 2008). Such organizing activity in the brain has been purported to play a role in sensory perception (Ermentrout and Kleinfeld 2001; Lakatos et al. 2007), memory—both working and long term (Klimesch 1999), and pathological events like epilepsy (Milton and Jung 2003). Whether or not large-scale brain oscilla- tions are epiphenomena or have functional significance remains an open question in many cases. However, both experiment and modeling continue to devote ef- forts to understanding the mechanisms that generate and sustain oscillations (Buszaki and Draguhn 2004). When neocortical or hippocampal in vitro slices are treated with an inhibitory neurotransmitter antagonist such as bicuculline, effectively eliminating inhibition, a localized current stimulus evokes population activity. Such activity may take the form of a spatially localized group of neurons whose population activity oscillates around 1–10 Hz (Wu et al. 1999; Milton and Jung 2003; Shusterman and Troy 2008); during each oscillation cycle the population may emit elevated activity that propagates as a traveling pulse (Wu et al. 1999; Chervin et al. 1988; Wu 2008) or a spiral wave (Huang et al. 2004; Schiff et al. 2007). Spiral waves provide a mecha- nism for spatially organizing extensive episodes of pe- riodic activity, effectively reducing the dimensionality of the dynamics (Schiff et al. 2007). Since inhibitory connectivity is pharmacologically blocked, any negative feedback in the network is likely to arise at the single cell level due to mechanisms such as synaptic depres- sion (Abbott et al. 1997; Bart et al. 2005; Matveev and

Upload: others

Post on 30-Nov-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209DOI 10.1007/s10827-009-0199-6

Spatially structured oscillations in a two-dimensionalexcitatory neuronal network with synaptic depression

Zachary P. Kilpatrick · Paul C. Bressloff

Received: 10 April 2009 / Revised: 14 October 2009 / Accepted: 16 October 2009 / Published online: 29 October 2009© Springer Science + Business Media, LLC 2009

Abstract We study the spatiotemporal dynamics ofa two-dimensional excitatory neuronal network withsynaptic depression. Coupling between populations ofneurons is taken to be nonlocal, while depression istaken to be local and presynaptic. We show that thenetwork supports a wide range of spatially structuredoscillations, which are suggestive of phenomena seenin cortical slice experiments and in vivo. The particularform of the oscillations depends on initial conditionsand the level of background noise. Given an initial, spa-tially localized stimulus, activity evolves to a spatiallylocalized oscillating core that periodically emits targetwaves. Low levels of noise can spontaneously generateseveral pockets of oscillatory activity that interact viatheir target patterns. Periodic activity in space can alsoorganize into spiral waves, provided that there is somesource of rotational symmetry breaking due to externalstimuli or noise. In the high gain limit, no oscillatorybehavior exists, but a transient stimulus can lead to asingle, outward propagating target wave.

Keywords Excitatory neuronal network ·Synaptic depression · Oscillations · Spiral waves ·Stationary bumps

Action Editor: Bard Ermentrout

Z. P. Kilpatrick · P. C. BressloffDepartment of Mathematics, University of Utah,Salt Lake City, UT 84112-0090, USA

P. C. Bressloff (B)Mathematical Institute, University of Oxford,24-29 St. Giles’, Oxford OX1 3LB, UKe-mail: [email protected]

1 Introduction

Spatially localized oscillations arise both in vivo andin vitro and may be observed experimentally usingmulti-electrode arrays or voltage-sensitive dye imaging(Wu 2008). Such organizing activity in the brain hasbeen purported to play a role in sensory perception(Ermentrout and Kleinfeld 2001; Lakatos et al. 2007),memory—both working and long term (Klimesch1999), and pathological events like epilepsy (Milton andJung 2003). Whether or not large-scale brain oscilla-tions are epiphenomena or have functional significanceremains an open question in many cases. However,both experiment and modeling continue to devote ef-forts to understanding the mechanisms that generateand sustain oscillations (Buszaki and Draguhn 2004).

When neocortical or hippocampal in vitro slices aretreated with an inhibitory neurotransmitter antagonistsuch as bicuculline, effectively eliminating inhibition, alocalized current stimulus evokes population activity.Such activity may take the form of a spatially localizedgroup of neurons whose population activity oscillatesaround 1–10 Hz (Wu et al. 1999; Milton and Jung 2003;Shusterman and Troy 2008); during each oscillationcycle the population may emit elevated activity thatpropagates as a traveling pulse (Wu et al. 1999; Chervinet al. 1988; Wu 2008) or a spiral wave (Huang et al.2004; Schiff et al. 2007). Spiral waves provide a mecha-nism for spatially organizing extensive episodes of pe-riodic activity, effectively reducing the dimensionalityof the dynamics (Schiff et al. 2007). Since inhibitoryconnectivity is pharmacologically blocked, any negativefeedback in the network is likely to arise at the singlecell level due to mechanisms such as synaptic depres-sion (Abbott et al. 1997; Bart et al. 2005; Matveev and

Page 2: Spatially structured oscillations in a two-dimensional

194 J Comput Neurosci (2010) 28:193–209

Wang 2000; Tabak et al. 2000; Tsodyks et al. 1998) andspike frequency adaptation (Benda and Herz 2003).

A variety of sensory stimuli have been linked tooscillations in vivo. For example, a number of studiesof vertebrate and invertebrate olfactory bulbs havefound that odor stimuli can elicit oscillations (Lamet al. 2000; Delaney et al. 1994). Stimuli can alsoevoke oscillations and waves in visual cortex (Singerand Gray 1995; Roelfsema et al. 1997; Xu et al. 2007;Benucci et al. 2007; Han et al. 2008), rat barrel cortex(Petersen et al. 2003), and auditory cortex (Lakatoset al. 2007). Spatiotemporal activity is not only a neuralcorrelate of sensory stimuli, but is also associated withvarious forms of memory. For example, the encodingof new information as well as the retrieval of long-term memory is reflected by the period of oscillations(Klimesch 1999), and the recall of a previous memory isoften accompanied by an increase in oscillatory power(Sederberg et al. 2003). On the other hand, station-ary bumps of persistent spatial activity that neitherpropagate nor oscillate have been seen during workingmemory tasks (Wang 1999).

Oscillations can also be the signature of certain brainpathologies such as epilepsy (Milton and Jung 2003).Electrophysiology has been used to study epilepsy inhumans as well as animal models, and seizures areusually accompanied by measurable structured popu-lation activity. Trauma or developmental malfunctioncan lead to reduced regions of inhibition, axonal sprout-ing, or synaptic reorganization of excitatory circuitry(Dudek and Spitz 1997). Such regions are prime can-didates for epileptic seizure foci. Any incurring excita-tory input may be sufficient to create high frequencyoscillations in the population activity of these patches ofcortex (McNamara 1994). The nature of such structuredpopulation activity as recorded by electroencephala-gram can indicate the nature of the seizure mechanism(Lee et al. 2006). As in cortical slice studies, someseizures have hallmark electrical activity traces consist-ing of focused localized synchronous oscillations thatemit traveling pulses (Schiff et al. 2005).

In light of the above examples, it is important tounderstand the mechanisms behind spatially structuredoscillations in large scale neuronal networks due totheir functional and pathological implications. A num-ber of organizing mechanisms for such spatiotempo-ral activity have been suggested, including a singlepacemaker oscillator exciting successive neighbors inan excitable network, or coupled oscillators propagat-ing gradual phase delays in space (Ermentrout andKleinfeld 2001; Wu 2008). Therefore, activity that prop-agates away from a focused region of high frequencyoscillations may either travel faster than the character-

istic time-scale set by the oscillating region, accordingto dynamics of an excitable medium, or at a speedset by the period of the oscillating core if the restof the medium is oscillatory as well. Conceivably, thismay establish a dynamical systems explanation for thewide range in speed at which seizures spread acrossthe cortex, which can be anywhere from 0.05 mm/s to10 cm/s (Milton and Jung 2003).

Recently Troy and Shusterman (2007) have shownhow a pulse emitting oscillating core can occur inan excitatory neuronal network based on a neuralfield model with linear negative feedback (Pinto andErmentrout 2001a). They first considered a two-dimensional network operating in an excitable regimewith a homogeneous low activity (Down) state thatcorresponded to a stable fixed point of the space-clamped dynamics. Linearization about the fixed pointyielded a Jacobian with complex eigenvalues so thatthe fixed point was a stable focus. In this regime,Troy and Shusterman showed that after breaking therotational invariance of an initially generated targetwave, reentry of the free ends of the resulting rotorinto the center of the domain generated subsequentwaves. In the same system, spiral waves were generatedby periodically breaking rotational invariance with aninhibitory stimulus. In a subsequent study they showedthat the same system, with large enough negative feed-back, supports spatially localized oscillations that pe-riodically emit traveling pulses (Shusterman and Troy2008). In this regime the space-clamped model exhibitsbistability, in which a stable Down state coexists witha stable limit cycle. However, since the linear form ofadaptation used in these studies is not directly relatedto physiological models of adaptation, it is difficult toascertain whether or not the levels of feedback requiredare reasonable.

In this paper, we show how various forms of spatiallystructured oscillations, including spiral waves and pulseemitters, can occur in a two dimensional excitatoryneuronal network with a physiologically based form ofnonlinear negative feedback, namely, synaptic depres-sion. Previously, we considered the combined effectsof synaptic depression and spike frequency adaptationon the spatiotemporal dynamics of a one-dimensionalexcitatory network (Kilpatrick and Bressloff 2009). Weshowed that synaptic depression tends to dominate thedynamics, providing a mechanism for generating spa-tially localized oscillations. Here we extend our analysisto two dimensions and to the case of noisy networks.For simplicity, we ignore the effects of spike frequencyadaptation since they tend to be relatively weak. Thebasic two-dimensional model is introduced in Section 2.We then analyze the space-clamped version of the

Page 3: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209 195

model, and show that it supports a stable limit cyclein the absence of noise (Section 3). However, in thepresence of additive white noise, the parameter regimeover which oscillations can occur can be significantlywidened. In Section 4 we present a number of numer-ical simulations illustrating various two-dimensionalspatiotemporal activity patterns supported by the fullnetwork model. Depending on the initial conditions,we show that in the absence of noise, network activitycan evolve as a pulse-emitting oscillating core or as aspiral wave. Furthermore, addition of a small amountof spatially uncorrelated noise to a quiescent networkcan drive pockets of the system superthreshold andlead to discrete locations of pulse-emitting cores. Onthe other hand, large amounts of noise lead to bulkoscillations which can disrupt any spatially structuredactivity. We also show that when a radially symmetricstimulus is applied to the network in the presence ofnoise, spiral waves can be generated due to symmetrybreaking, similar to the organized activity found inmammalian cortical slice (Huang et al. 2004; Schiff et al.2007). Finally, we study the system in the high-gainlimit (Section 5). In this case, oscillations do not exist inthe deterministic system, but depression is a sufficientmechanism for generating outward propagating targetwaves following a brief stimulus.

2 Neural network model with synaptic depression

We consider a neuronal network model which includessynaptic depression (Abbott et al. 1997; Bart et al.2005; Matveev and Wang 2000; Tabak et al. 2000;Tsodyks et al. 1998; Zucker and Regehr 2002). Asopposed to the usual Pinto-Ermentrout formulation ofnegative feedback (Pinto and Ermentrout 2001a; Foliasand Bressloff 2004; Troy and Shusterman 2007; Troy2008) in spatially extended neural fields, here we takenegative feedback to depend on output firing rate

∂u(r, t)∂t

= −u(r, t) + w ∗ (qf (u)) (r, t) (2.1a)

∂q(r, t)∂t

= 1 − q(r, t)α

− βq(r, t) f (u(r, t)), (2.1b)

where

w ∗ (qf (u)) (r, t) =∫

R2w

(∣∣r − r′∣∣) q(r′, t

)f(u

(r′, t

))dr′

where r and r′ are spatial positions in the two-dimensional plane R

2. Equation (2.1a) describes theevolution of the synaptic current or drive u(r, t) in the

presence of synaptic depression, which takes the formof a synaptic scaling factor q(r, t) evolving accordingto Eq. (2.1b). The factor q(r, t) can be interpreted asa measure of available presynaptic resources, whichare depleted at a rate β f (Tsodyks et al. 1998; Bartet al. 2005; Tabak et al. 2000), and are recovered ona timescale specified by the constant α (experimentallyshown to be 200–1500 ms (Abbott et al. 1997; Tsodyksand Markram 1997; Tsodyks et al. 1998; Stevens andWesseling 1998)). In a previous study, we considereda one-dimensional model with synaptic depression andadaptation and showed that adaptation has a relativelysmall effect on the dynamics (Kilpatrick and Bressloff2009). Therefore, we focus on synaptic depression here.It will be convenient in the following to fix parametersso that f is interpreted as the fraction of the maximumfiring rate, that is 0 ≤ f ≤ 1.

Most of our analysis and numerics are carried outfor a general piecewise linear firing rate function thatattains saturation as pictured in Fig. 1:

f (u) =⎧⎨⎩

0, u ∈ (−∞, θ),

σ (u − θ), u ∈ (θ, θ + σ−1),

1, u ∈ [θ + σ−1, ∞).

(2.2)

Here σ specifies the slope of the linear regime. (Onecould also consider a firing rate function with a stepfollowed by linear increase (Guo and Chow 2005a, b)).Note that taking the firing rate to be a linear function

-0.1 0 0.1 0.2 0.3 0.4 0.5

0

0.2

0.4

0.6

0.8

1

u

f(u) step

piecewise linear

sigmoid

θ s = 0.175

θ = 0.05

Fig. 1 Comparison of the step, piecewise linear, and sigmoidfiring rate functions. Parameter values are θ = 0.05 and σ = 4.The sigmoid function has the same slope and value as the piece-wise linear function at their mean values. When compared tothe sigmoid function, it is apparent that the piecewise linearfunction’s true threshold is more accurately given by θs = 0.175,rather than θ , the point at which nonzero firing occurs

Page 4: Spatially structured oscillations in a two-dimensional

196 J Comput Neurosci (2010) 28:193–209

close to threshold is consistent with the observationthat spike frequency adaptation tends to linearize thefiring frequency-input current curve (Ermentrout 1998;Wang 1998). In the limit that σ → ∞, we recover theHeaviside step function used in Amari’s original workon scalar networks (Amari 1977) and most analyt-ical studies of the Pinto-Ermentrout model (Pinto andErmentrout 2001a, b; Shusterman and Troy 2008; Foliasand Bressloff 2004, 2005a, b; Troy and Shusterman2007; Troy 2008):

f (u) = H(u − θ) ={

0, u ∈ (−∞, θ),

1, u ∈ [θ, ∞).(2.3)

We will use such a function in order to study targetwaves and stationary bumps (see Section 5). One im-portant point, as illustrated in Fig. 1, is that the thresh-old θ in Eq. (2.2) is not necessarily analogous to thethreshold θ in Eq. (2.3). A region above threshold inthe piecewise linear case might still be considered “off”if it only sustains a low level of activity. This is furtherreinforced by comparison with a sigmoid function inwhich threshold is taken to be the point at which thefiring rate is half its maximum value:

f (u) = 1/(1 + exp(−σ(u − θs))). (2.4)

The homogeneous weight distribution w(|r − r′|) de-fines the strength of the synaptic connections betweenneurons at r and r′, depending only on the distance be-tween two cells. Typical excitatory weight functions aremonotone decreasing functions such as the exponentialand the Gaussian. We will take w to be given by adifference of modified Bessel functions of the secondkind: (Folias and Bressloff 2004; Laing 2005; Owenet al. 2007)

w(r) = 2w0

3πd(K0(r/d) − K0(2r/d)) , (2.5)

where w0 determines the strength of the synaptic con-nections. The factor 2/3π ensures that Eq. (2.5) is avery good fit to the exponential weight function

w(r) = w0

2πde−r/d

The expansion in terms of Bessel functions is particu-larly convenient because it allows us to transform thesystem (2.1) into a fourth order PDE, which is compu-tationally less expensive to simulate (Laing and Troy2003; Laing 2005; Troy and Shusterman 2007; Owenet al. 2007) (see Section 4). Finally, we fix the temporaland spatial scales of the network by setting τ = 1 andd = 1. The membrane time constant is typically around10 ms, whereas the range of synaptic connections withincortex is of the order 1 mm. We also fix synaptic

strength by setting w0 = 1. The effects of varying w0 arebriefly discussed at the end of Section 3.

3 Oscillations in the space-clamped system

Previous modeling studies of space-clamped neuronalnetworks with synaptic depression showed the exis-tence of oscillations in the case of excitatory/inhibitorynetworks (Tsodyks et al. 1998) or for a purely excitatorynetwork with noise (Bart et al. 2005). Tabak et al.showed that an excitatory network with depressioncould support regular oscillations and bursting, using analternative form for the neural field equation as well asdifferent gains and thresholds for each variable’s activa-tion function (Tabak et al. 2000). In our study, we findthat saturation of the activation function is sufficient tostabilize limit cycles using the same activation functionfor both the activity and depression variables.

3.1 Phase plane for piecewise linear firing rate

As a means of determining the oscillatory behavior ofthe system, we examine the equilibria of the space-clamped system (Bart et al. 2005; Tabak et al. 2000;Tsodyks et al. 1998)

u(t) = −u(t) + q(t) f (u(t)),

αq(t) = 1 − q(t) − αβq(t) f (u(t)), (3.1)

where f is the piecewise linear activation function (2.2)shown in Fig. 1. We carry out the stability analysis ofphase space using the piecewise linear function becauseexplicit analytical expressions can be derived for thefixed points. However, these results extend to the casewhere there is a smooth transition from the linear to thesaturated portion of the firing rate.

To calculate equilibria of (3.1), we consider the pos-sible solutions on the three domains of the piecewisefunction f (u). We find that there is a low activity orDown state on the lower domain (u < θ) for θ > 0 suchthat (u, q) = (0, 1). The stability of this Down state isdetermined by the eigenvalues of the Jacobian

J (0, 1) =(−1 0

0 −1/α

)(3.2)

and is therefore stable for all realistic parameters. Astable Down state exists in the network for any f with ahard threshold, that is f (u) = 0 for u < θ . Without thiscondition, it is possible that the Down state may desta-bilize or vanish due to a nonzero firing rate existing forzero synaptic drive.

Page 5: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209 197

We find additional equilibria by solving (3.1) on themiddle and upper domains of f . On the middle domain(θ ≤ u ≤ θ + σ−1), where f (u) = σ(u − θ), we have

u = σ(u − θ)q, (3.3)

q = 1/(1 + σαβ(u − θ)), (3.4)

θ ≤ u ≤ θ + σ−1, (3.5)

which has solutions

u = σ + σαβθ − 1 ± √D

2σαβ(3.6)

q = 2

1 + σ − σαβθ ± √D

(3.7)

D = (σ + σαβθ − 1)2 − 4σ 2αβθ

provided D ≥ 0 and condition (3.5) is satisfied. Stabilityis determined by the eigenvalues of the Jacobian

J (u, q) =(−1 + σq σ(u − θ)

−βσq −(1/α + βσ(u − θ))

). (3.8)

We find that for a wide range of parameters, the mid-dle domain contains two equilibria, one of which is asaddle and the other is a stable or unstable focus. Thelatter corresponds to a high activity or Up state. Forsufficiently fast depression, destabilization of the Upstate can lead to the formation of a stable limit cyclevia a subcritical Hopf bifurcation as pictured in Fig. 2.In parameter regimes where the focus equilibrium

does not exist, the Up state occurs on the upper domain(u > θ + σ−1), where f (u) = 1, and is given by

u = 1/(1 + αβ), (3.9)

q = 1/(1 + αβ), (3.10)

Its stability is determined by the eigenvalues of theJacobian

J (u, q, a) =(−1 1

0 −(1/α + β)

), (3.11)

which guarantees that such an Up state is always stable.In Fig. 3 we show a simulation of the space-clamped

network for a choice of parameters that supports alimit cycle. The parameter value for synaptic depressiontime constant α is taken to be within the physiologicalrange 200–1500 ms (Abbott et al. 1997; Tsodyks et al.1998). Notice that both variables oscillate at a period ofroughly 40 time units or 400 ms, which correlates wellwith the scale of epileptiform events (Buszaki 2006;McNamara 1994; Milton and Jung 2003; Shustermanand Troy 2008). This also implies that the timescaleof oscillations is roughly set by the time constant ofsynaptic depression. Notice that as opposed to selfsustained oscillations in the Pinto-Ermentrout model(Shusterman and Troy 2008), the equilibrium focusin our model is associated with the Up rather thanthe Down or rest state. As stated, these results easilyextend to the case where f is a smooth, saturatingfunction above the threshold value u = θ . In particular,since limit cycles are structurally stable solutions incontinuous systems, oscillations persist when f (u) ismodified by smoothing out the corner at u = θ + 1/σ .However, without a hard threshold at u = θ , we havenot witnessed the same types of dynamics as presented

0.02 0.03 0.04 0.05 0.06 0.07 0.08

0

0.1

0.2

0.3

0.4

u

stable focus

unstable focus

saddle

stable node

limit cycle

β

(a) 0.12

0.1

0.08

0.06

0.04

0.02

20 30 40 50 60 70 80 90 100α

β

(b)

Fig. 2 Piecewise linear firing rate function. (a) Bifurcation dia-gram showing fixed points u of the system (3.1) as a functionof β for α = 80. (b) Stability diagram for the space-clampedsystem (3.1) showing regions in parameter space where the Up

state is a stable focus (black), an unstable focus surrounded by astable limit cycle (grey), or an unstable focus without a limit cycle(white). Other parameters are θ = 0.01, and σ = 4

Page 6: Spatially structured oscillations in a two-dimensional

198 J Comput Neurosci (2010) 28:193–209

0 20 40 60 80 100 120 140 160 1800

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

t

u

q

(a)

u

q

limit cycle

q nullcline

u nullcline

(b)

0 0.05 0.1 0.15 0.2 0.25 0.30.1

0.15

0.2

0.25

0.3

0.35

0.4

Fig. 3 (a) Numerical simulation of (3.1) using the parametersθ = 0.01, σ = 4, α = 80, and β = 0.05 given the initial condition(u, q) = (1, 1). The synaptic input u and fraction of availableresources q are plotted as a function of time t. Oscillations

lock to a period roughly determined by the time constant α.(b) Corresponding phase-plane plot of q versus u (dashed line)showing that the system supports a stable limit cycle

here for the piecewise linear f . If oscillations exist, astable Down state does not, which we show in an analy-sis of the system with a sigmoidal firing rate function(2.4).

3.2 Phase plane for sigmoidal firing rate

We use a numerical root finding algorithm to identifythe equilibria of the system (3.1) in the case where fis the sigmoidal function (2.4). It is possible to findphysiologically reasonable parameter regimes wherelimit cycles exist, but they do not appear to coincidewith a stable Down state, as in the piecewise linear fcase. We show an example of one such limit cycle inFig. 4, where transitions between low and high activitystates occur at the knees of the u-nullcline. There isno such mechanism in the limit cycle present in thepiecewise linear system. This distinction suggests thatthe loss of a hard threshold may change the overalltopology of dynamics within the network. Rather thanfinding an excitable regime with limit cycles about the

Up state, we find either a purely oscillatory regime, oran excitable regime with no limit cycles.

3.3 Space-clamped system with noise

A previous study of the space-clamped system (3.1)with f given by (2.2) considered parameter regimesin which the subcritical Hopf bifurcation of the Upstate only produced an unstable limit cycle (Bart et al.2005). In this case the authors showed that oscillationscould be generated in the presence of additive noise,which switched the system between the Up and Downstates (see also Holcman and Tsodyks 2006). It followsthat noise enlarges the parameter regime over whichself-sustained oscillations can occur. We illustrate theeffects of additive noise by simulating the system

u(t) = −u(t) + q(t) f (u(t)) + γ ν(t),

αq(t) = 1 − q(t) − αβq(t) f (u(t)), (3.12)

Fig. 4 Sigmoidal firing ratefunction. (a) Bifurcationdiagram showing fixed pointsu of the system (3.1) as afunction of β for α = 50. (b)Phase-plane plot of q versus u(dashed line) showing thesystem supports a limit cycle.Other parameters areθs = 0.15, and σ = 20

(a)

u

0.1 0.20

0.1

0.2

0.3

0.4

0.5

stable focus

unstable focusstablefocus

limit cycle

0.150.5 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.40

0.1

0.2

0.3

0.4

0.5

0.6

u

q

limit cycle

q nullclineu nullcline

(b)

β

Page 7: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209 199

where f is the piecewise linear function (2.2), ν(t) isa Gaussian white noise process such that 〈ν(t)〉 = 0,and 〈ν(s)ν(t)〉 = 2δ(t − s); γ is the noise strength. Wesimulated the system (3.12) using an Euler-Maruyamascheme for stochastic differentiation with a timestep�t = 10−6. The nature of the noise-induced oscilla-tions depends upon whether the Up state is a stableor unstable focus. In the case of a stable focus, eventhough oscillations are damped out eventually in thedeterministic system, noise is sufficient to repeatedlydrive the system between the Up and Down states,along analogous lines to Bart et al. (2005). However,the oscillations tend to be rather irregular as illustratedin Fig. 5.

More regular noise-induced oscillations occur in thecase of an unstable focus. Equation (3.12) now repre-sent an excitable system with only a stable Down state,in which noise periodically drives the system abovethreshold, leading to an elevated firing rate that thenrelaxes back down as synaptic depression is activated.An example simulation is shown in Fig. 6, which neatlyillustrates the regularity of the noise-induced oscilla-tions. This is an example of a well known stochas-tic property of excitable systems, namely, coherenceresonance (Linder et al. 2004). That is, there existsan optimal level of noise with respect to the degreeof regularity of the induced oscillations; if the levelof noise is too high then this completely washes outany oscillations. We conclude that noise extends the

0 50 100 150 200 250 300t

u

q

0

0.2

0.4

0.6

0.8

1

Fig. 5 Numerical simulation of the space-clamped system (3.12)in which background noise drives the system between Up andDown states. The horizontal dashed line denotes the input cur-rent value at which activity is half of its maximal value (θs =0.135). Firing rate is taken to be piecewise linear function (2.2).Parameters are α = 60, β = 0.06, γ = 0.02, θ = 0.01, and σ = 4.In the absence of noise, the Up state is a stable focus

0 50 100 150 200 250 300t

0

0.2

0.4

0.6

0.8

1

Fig. 6 Same as Fig. 5 except α = 80 so that now the Up state isan unstable focus in the absence of noise

parameter range over which the space-clamped sys-tem supports oscillations to include regions where theunderlying deterministic system supports a stable orunstable focus without a stable limit cycle. This alsohas important implications for the effects of noise in thespatially extended system (see Section 4).

Bart et al. (2005) showed that changing the synapticstrength w0 can also alter the stability of the Up stateof the system (3.1), whilst keeping all other parametersfixed. Stronger synapses (higher w0) stabilize the Upstate, while weaker synapses (lower w0) destabilizeit. Consistent with these observations, we found thatchanging w0 alters the parameter ranges of α and β

over which a stable limit cycle exists. That is, increasingw0 shifts the region in which limit cycles exist to highervalues of α. On the other hand, decreasing w0 allows forlimit cycles to exist for lower values of α, but the rangeof β values over which they exist is much narrower.Thus, superthreshold activity in a network with weaksynapses is much more easily overridden by synapticdepression. In our simulations we take w0 = 1.

4 Oscillations in the spatially extended model

We now consider the implications of the existenceof deterministic and noise-induced oscillations in thespace-clamped model for spatially structured oscilla-tions in the full model (2.1). Using numerical sim-ulations, we demonstrate that the two dimensionalnetwork supports a spatially localized oscillating corethat emits target waves each cycle, as well as spiralwaves. The results presented here can also be generated

Page 8: Spatially structured oscillations in a two-dimensional

200 J Comput Neurosci (2010) 28:193–209

for a system with smoother forms of f . However, as thesimulations are qualitatively similar, we only show re-sults for the piecewise linear case. As in previous stud-ies of two-dimensional neural field models, we carryout a transformation of our system for more efficientcomputation (Laing and Troy 2003; Laing 2005; Troyand Shusterman 2007; Owen et al. 2007). That is, weconvert the integro-differential equation system (2.1)to a fourth order PDE using two-dimensional Fouriertransforms. This is possible due to the fact that theFourier transform of the weight distribution w(r) givenby Eq. (2.5) is a rational function. Discretizing theresulting differential operators leads to sparse matrices,as opposed to full matrices arising from an integraloperator.

Numerical simulations are thus performed on thefollowing system, which is equivalent to Eqs. (2.1) and(2.5):[∇4 − A∇2 + B

](ut + u) = Mqf (u),

qt = 1 − qα

− βqf (u) (4.1)

over the domain ⊂ R2. Here, the fourth order oper-

ator, L = ∇4 − A∇2 + B, arises as the denominator ofthe two-dimensional Fourier transform of our modifiedBessel weight function (2.5), which is given by

w(ρ) = 2

(1

ρ2 + 1− 1

ρ2 + 22

)

= 2/π

ρ4 + 5ρ2 + 4, (4.2)

where · denotes the two-dimensional Fourier trans-form. In this case, A = 5, B = 4, and M = 2/π , but wemay adjust these parameters by considering a rescal-ing of w. We solve the system (4.1) numerically on aCartesian grid of 1000 × 1000, applying homogeneousDirichlet and Neumann boundary conditions. For thefourth order operator, we employed a second orderfinite difference method to construct a matrix version ofL. The time derivative was approximated using forwardEuler with a timestep of �t = 0.01, which was smallenough so that shrinking it further did not changeresults.

4.1 Pulse emitter

Similar to our previous study of a one-dimensionalnetwork (Kilpatrick and Bressloff 2009), we find thatin parameter regimes where a stable limit cycle existsin the space-clamped system, the corresponding two-dimensional network supports a spatially localized os-cillating core that periodically emits traveling pulses.

All that is necessary to induce such behavior is an initialcondition of the form

(u(r, 0), q(r, 0)) = (χe−(x2+y2)/ζ 2, 1), (4.3)

where χ and ζ parameterize the amplitude and spatialconstant of the initial state. We seek to characterize theevolving activity in the limit cycle regime, especially theperiod of oscillation and the speed of emitted pulses. InFig. 7, we show an example of a pulse-emitting core,which oscillates at a frequency of roughly 3Hz. Pulsesare emitted each cycle, and travel at a speed of roughly30 cm/s, which is determined by the period of the oscil-lations; the latter is set by the time constant of synapticdepression. The initial emission of spreading activityappears as a traveling front which propagates from theregion activated by the input current into the surround-ing region of zero activity; it travels at the same speedas the subsequent target waves. The front convertseach region of the network into an oscillatory statethat is phase-shifted relative to the core, resulting inthe appearance of a radially symmetric target pattern.Since our network has solely excitatory connections, wecan consider it to be akin to disinhibited neocortical orhippocampal slices (Chervin et al. 1988; Wu et al. 1999;Wu 2008) or regions of cortex or hippocampus whereexcitatory circuitry dominates due to some pathology(Buszaki 2006; Dudek and Spitz 1997). Interestingly,the speed of the simulated waves matches the time-scale of fast seizure spread in cortex (Milton and Jung2003).

4.2 Spiral waves

Several experimental and modeling studies of two-dimensional cortex reveal the existence of spiral waves(Huang et al. 2004; Laing 2005; Milton and Jung 2003;Schiff et al. 2005, 2007). Such self-sustained activitycan often be classified by a constant angular velocity(Huang et al. 2004; Schiff et al. 2005). When identi-fied using voltage sensitive dye, one finds such activ-ity patterns have true phase singularities about whichthe spiral organizes. One may think of such spatiallystructured activity as a network property manifestingthe recovery period necessary for groups of neurons.Therefore, sections of cortex about the phase singu-larity alternate between Down and Up states, givingample time for sections to recover during the Downstate.

Spiral waves have been generated in previous studiesof neural field models with linear adaptation, in whichthe neuronal network acts like an excitable medium(Laing 2005; Troy and Shusterman 2007). The oscilla-tions necessary for the generation of spiral waves arise

Page 9: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209 201

t = 60

t = 150

-200 -100 0 100 200x

y

200

100

0

-100

-200

t = 0 t = 30

y

200

100

0

-100

-200

t = 75 t = 90

-200 -100 0 100 200x

-200 -100 0 100 200x

Fig. 7 Snapshots of the solution u(x, y, t) to the fourth orderPDE (4.1), following a stimulus specified by Eq. (4.3) at t = 0,where χ = 1 and ζ = 25. Initially, an activated state spreadsradially outward, across the entire medium as a traveling front.Then, the localized oscillating core of activity emits a target wave

with each oscillation cycle. Eventually, these target waves fill thedomain. Each target wave can be considered as a phase shiftin space of the oscillation throughout the medium; they travelwith the same speed as the initial front. Parameters are α = 80,β = 0.05, σ = 4

from the Down state of the network being a stablefocus. Laing used the rotational symmetry of the spiralwaves to generate equations for the activity profile andangular velocity of a spiral on a disc domain (Laing2005). Troy and Shusterman generated spiral wavesby continually breaking the symmetry of target wavesin the network (Troy and Shusterman 2007). In ourmodel, we find that spiral wave patterns can be inducedby breaking the rotational symmetry of pulse emittersolutions. More specifically, we chose an initial condi-tion where the target pattern produced by the emitterhas the top and bottom halves of its domain phaseshifted. The network then evolves into two counter-rotating spirals on the left and right halves of thedomain as shown in Fig. 8. Closer inspection of oneof these spirals reveals that it has a fixed center aboutwhich activity rotates indefinitely as shown in Fig. 9.

Huang and others showed that spiral waves gener-ated in cortical slices are a way for oscillating activityto organize spatially in a smooth and isotropic medium

(Huang et al. 2004). They found the waves persisted forup to 30 cycles and rotated at a rate of roughly 10 cyclesper second. Also, the phase singularity at the center ofa spiral wave experiences a reduction in oscillation am-plitude due to the mixing of all phases in a small region.Certainly, the spiral waves we have found in our systempersist for a long time, but it seems that the rotationrate is slightly slower at roughly 2 Hz. Of course this isdue in part to the time constant of synaptic depression.As we have shown in our previous work, includingspike frequency adaptation can increase the frequencyof oscillations (Kilpatrick and Bressloff 2009).

4.3 Noise-induced oscillations

As in the space-clamped system, it is interesting toconsider the effects of noise on the two-dimensionalspatially extended network. In a recent study of therole of additive Gaussian noise on Turing instabili-ties in neural field equations, Hutt et al. found that

Page 10: Spatially structured oscillations in a two-dimensional

202 J Comput Neurosci (2010) 28:193–209

Fig. 8 Snapshots of a solutionu(x, y, t) to the fourth orderPDE (4.1) revealing thecounter-rotation of two spiralwaves on either side of thedomain. These weregenerated with an initialcondition where the targetpattern of Fig. 7 had the topand bottom halves of thedomain phase shifted.Parameters are α = 80,β = 0.05, σ = 4, and θ = 0.01

y

200

100

0

-100

-200

t = 0 t = 21

-200 -100 0 100 200x

-200 -100 0 100 200x

0.3

0.2

0.1

0

noise delays the onset of pattern formation (Hutt et al.2008). Also, Laing et al. have shown that in a neuralfield model with linear adaptation, moving bumps areslowed by the introduction of an additive noise term(Laing et al. 2007). Here we show that in addition to

modulating spatiotemporal activity patterns that existin the deterministic system, noise also gives rise to newdynamics.

Following a previous study of neural field modelswith additive noise (Laing et al. 2007), we introduce

y

60

30

0

-30

-60

t = 0 t = 9 t = 18

t = 27 t = 36 t = 45

-120 -90 -60 -30 0x

y

60

30

0

-30

-60

-120 -90 -60 -30 0x

-120 -90 -60 -30 0x

Fig. 9 A zoomed in version of the rotating left spiral wavepictured in Fig. 8. The period of the spiral wave oscillation isroughly the same as the period of the oscillation in the space-clamped system. All patches of neurons are oscillating at the

same frequency, but phase-shifted as coordinates are rotatedabout the central phase singularity. Parameters are α = 80, β =0.05, σ = 4, and θ = 0.01

Page 11: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209 203

a Gaussian white noise term to each equation of adiscretized version of the fourth order PDE (4.1):

Lh

(uk+1

ij − ukij

�t+ uij + ημij(t)

)= Mqij f (uij),

qk+1ij − qk

ij

�t= 1 − qij

α− βqij f (uij), (4.4)

where i = 1, ..., Nx and j = 1, ..., Ny, Lh is the finite dif-ference version of the linear operator given in Eq. (4.1),uij and qij are discrete values of u and q at (x, y) =(xi, y j), each μij evolves independently as 〈μij(t)〉 = 0and 〈μij(t)μij(s)〉 = δ(t − s), and η is the variance of ourwhite noise term.

In the case of low levels of spatially incoherentGaussian noise, we find that small pockets of the net-work spontaneously form spatially localized oscillatorswhich then interact with one another via the targetwaves that propagate from their cores. We picture this

in Fig. 10 for η = 0.005. Therefore, as in the space -clamped case, noise provides a mechanism for gener-ating oscillations in a situation where the deterministicsystem would remain quiescent. If the noise level isincreased then it tends to disrupt these oscillating cores,which provides a symmetry breaking mechanism forthe generation of spiral waves as illustrated in Fig. 11.Following induction of a spatially localized oscillationusing a Gaussian stimulus of the form (4.3), we findthat the oscillating core begins to be broken up bythe noise such that the two halves of the core oscillateantisynchronously. A semi-ring wave then propagatesfrom the bottom to the top of the domain (first threesnapshots in Fig. 11), and breaks up into two spiralwaves as it reaches the boundary of the core (fourthsnapshot). Background oscillations absorb the two spi-ral waves and the ring-wave is reinitiated (final twosnapshots). At even higher levels of noise any spatiallystructured activity in the network is disrupted and theentire network exhibits bulk oscillations. Indeed, it haspreviously been shown that there can be an optimal

y

200

100

0

-100

-200

t = 30 t = 90 t = 120

y

200

100

0

-100

-200

t = 150 t = 180 t = 210

-200 -100 0 100 200x

-200 -100 0 100 200x

-200 -100 0 100 200x

Fig. 10 Snapshots of a solution u(x, y, t) to the noisy system(4.4) in the absence of stimulation. Background noise initiatesspatially localized oscillating cores at discrete sites in the medium.

Target waves emitted by the various oscillating regions collide,disrupting the spatial structure of the oscillations. Parameters areα = 80, β = 0.05, η = 0.005, σ = 4, and θ = 0.01

Page 12: Spatially structured oscillations in a two-dimensional

204 J Comput Neurosci (2010) 28:193–209

y

80

40

0

-40

-80

t = 558 t = 564 t = 570

y

80

40

0

-40

-80

t = 576 t = 582 t = 588

-80 -40 0 40 80 -80 -40 0 40 80 -80 -40 0 40 80

Fig. 11 Snapshots of a solution u(x, y, t) to the noisy system (4.4),following a stimulus specified by Eq. (4.3) at t = 90, where χ = 1and ζ = 60. As activity develops, the background noise starts tobreak the symmetry of an oscillating central core. Rather thana contiguous oscillation, two halves of the core oscillate anti-synchronously. A semi-ring wave propagates from the bottom

to the top of the medium, and breaks into two spiral waves asit collides with the boundary of the core. This exemplifies noiseinduced spiral waves, which are sustained for relatively long timeintervals. Parameters are α = 80, β = 0.05, η = 0.01, σ = 4, andθ = 0.01

level of noise for the generation of spiral waves in ex-citable media (Jung and Mayer-Kress 1995). Note thatan alternative mechanism to noise for generating spiralwaves is to introduce random network inhomogeneities(quenched disorder), as shown in the case of a two-dimensional integrate-and-fire network (Milton et al.1993).

5 High-gain limit

In order to analyze the existence and stability of spa-tially structured solutions in neuronal networks, thehigh-gain limit of a sigmoid-like firing rate function isoften considered, whereby f reduces to the Heavisidefunction (2.3) with a discontinuity at the threshold u =θ (Amari 1977; Pinto and Ermentrout 2001a). Althoughan excitatory neuronal network with synaptic depres-sion as given by Eq. (2.1) no longer supports oscilla-tory solutions in the high-gain limit, stationary pulses(bumps) and target wave solutions can be found. How-

ever, the presence of a Heaviside function in the dy-namics of the synaptic depression variable means thatthe resulting dynamical system is piecewise smooth,which considerably complicates the analysis of the sta-bility of waves and bumps. In this section, we carryout a phase-plane analysis of the space-clamped system,show that stable target wave solutions exist, and deriveequations for the existence of stationary bumps. Theissue of stability will be addressed elsewhere (see alsoKilpatrick and Bressloff 2009).

5.1 Phase-plane analysis

The space-clamped system with a Heaviside firing ratefunction takes the form

u(t) = −u(t) + q(t)H(u(t) − θ),

αq(t) = 1 − q(t) − αβq(t)H(u(t) − θ). (5.1)

In order to calculate equilibria of (5.1), we considerthe possible solutions on the two domains of the step

Page 13: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209 205

function H(u − θ). We find that there is always a lowactivity or Down state on the lower domain (u < θ)for θ > 0 such that (u, q) = (0, 1). The stability of thisDown state is determined by the eigenvalues of theJacobian

J (0, 1) =(−1 0

0 −1/α

)(5.2)

and is therefore stable for all realistic parameters. Asstated in our analysis of the system with the piecewiselinear firing rate function, this stable Down state indeedshould exist for all systems possessing an f with a hardthreshold.

In the upper domain (u > θ), an equilibrium is givenby the system

0 = −u + q, (5.3)

0 = (1 − q)/α − βq, (5.4)

implying a fixed point (u,q) = (1/(1 + αβ), 1/(1 + αβ))will exist, provided θ < 1/(1 + αβ). Its stability is deter-mined by the eigenvalues of the Jacobian

J (u, q, a) =(−1 1

0 −(1/α + β)

), (5.5)

which guarantees that such an Up state is always stable.Therefore, as stated, we have a bistable system as longas θ < 1/(1 + αβ), as pictured in Fig. 12. Additive noisecould then be a mechanism for switching the systembetween its Up and Down states. However, if θ >

1/(1 + αβ), only the Down state exists, which physicallymeans that in this case synaptic depression curtailsrecurrent excitation to the point that no sustained ac-tivity is possible. In the special case θ = 1/(1 + αβ), an

0 0.2 0.3 0.4 0.5 0.60

0.2

0.4

0.6

0.8

1

u

q

0.1

q nullcline

q nullcline

u nullclineu nullcline

Fig. 12 Phase plane plot of the space-clamped system (5.1) finthe case θ < 1/(1 + αβ) for which there exist two stable fixedpoints. Parameters are α = 50, β = 0.05, and θ = 0.1

equilibrium exists at u = q = θ , provided that we takeH(0) = 1. However, the piecewise smooth nature of thedynamics needs to be taken into account in order todetermine the stability of the fixed point. That is, thefixed point is stable with respect to perturbations δu > 0but unstable with respect to perturbations δu < 0. Thusstability cannot be established simply by linearizingabout the fixed point. While this special case is non-generic in the space-clamped system, it foreshadows po-tential problems in the study of the stability of spatiallystructured solutions of the full system (2.1). This is dueto the fact that one has to consider perturbations atthreshold crossing points x where u(x, t) = θ . We willdiscuss this issue further, following a study of spatiallystructured solutions.

5.2 Target waves

As shown by numerical simulations in the case of apiecewise linear firing rate, spatially structured oscil-lations can generate expanding target waves via prop-agating phase shifts in an oscillatory medium. Here,we show that in the high-gain limit, target waves arisein the context of an excitable medium. We studiedthe existence of traveling pulses in a one-dimensionalexcitatory network with synaptic depression in a pre-vious study (Kilpatrick and Bressloff 2009). Travelingpulses of this type represent a homoclinic orbit in theprojected space of the traveling wave coordinate, ratherthan phase shifts of an existing limit cycle, as in anoscillatory medium. Implications of these two differentdynamical systems scenarios may be useful in determin-ing the mechanism that generates traveling waves inexperiment. For example, in studies of disinhibited cor-tical slice, localized stimuli may lead to either travelingplane waves, which are transient, or spiral waves, whichare persistent (Huang et al. 2004).

We find that by simulating the two-dimensional spa-tially extended system with the Heaviside step firingrate function, a localized stimulus can lead to outwardlypropagating target waves. In Fig. 13, we show snap-shots of such a simulation where the initial condition istaken to be a Gaussian stimulus of the synaptic drive,as specified by Eq. (4.3). In the case of a piecewiselinear firing rate function, radially symmetric stimulilead to an oscillating core that recurrently generatedtarget waves. Here, a single radially symmetric targetwave is generated, after which, the system returns toa quiescent state. The structure of the solution reflectsthe underlying symmetries of the weight function (2.5)and the stimulus. Numerical simulations suggest thatthese target waves are stable.

Page 14: Spatially structured oscillations in a two-dimensional

206 J Comput Neurosci (2010) 28:193–209

y

30

20

0

-10

-20

10

y

30

20

0

-10

-20

10

10-30 -20 0 20 30xx

-1010-30 -20 0 20 30-10 10-30 -20 0 20 30x

-10

Fig. 13 Snapshots of a solution u(x, y, t) to the fourth orderPDE (4.1) showing a single outward propagating target wave,following a stimulus specified by Eq. (4.3) at t = 0, where χ = 1

and ζ = 25. The firing rate function is taken to be a Heavisidefunction, f (u) = H(u − θ). Parameters are α = 50, β = 0.4, andθ = 0.1

5.3 Standing bumps

Finally, we extend our previous analysis of bumps ina one-dimensional network (Kilpatrick and Bressloff2009) in order to derive conditions for the existenceof standing bumps in the two-dimensional network(2.1) with a Heaviside firing rate function. We canassume radially symmetric bumps since the correspond-ing weight distribution (2.5) is itself radially symmetric.Consider a radially symmetric stationary bump solutionof Eq. (2.1) such that u(r, t) = U(r), q(r, t) = Q(r) andU(r) crosses the threshold θ at the unique radius r = a.The solution is taken to be superthreshold within thedomain r < a and subthreshold otherwise, and a is iden-tified as the radial width of the bump. Given reasonableboundary conditions, we thus have

U(r) ≷ θ, for r ≶ a, (5.6)

{U(r), Q(r))} → {0, 1} , as r → ∞. (5.7)

Substituting such a solution into Eq. (2.1) gives

U(r) =∫U

Q(r′)w(|r − r′|)dr′, (5.8)

Q(r) = (1 + αβ�(a − r))−1, (5.9)

where U = {r = (r, θ) : r ≤ a} is the domain on whichthe bump is superthreshold and

�(r) ={

1, r ≥ 0,

0, r < 0.(5.10)

If we express (5.9) as

Q(r) =⎧⎨⎩

1

1 + αβ, r ≤ a,

1, r > a,

(5.11)

then we can substitute (5.11) back into (5.8) to yield

(1 + αβ)U(r) = �(a, r), (5.12)

where

�(a, r) =∫ 2π

0

∫ a

0w

(∣∣r − r′∣∣) r′dr′dθ ′. (5.13)

Page 15: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209 207

We can calculate the double integral in (5.13) usingFourier transform and Bessel function identities, as ina previous study (Folias and Bressloff 2004). Thus, wefind that

�(a, r) = 2πa∫ ∞

0w(ρ)J0(rρ)J1(aρ)dρ, (5.14)

where w(ρ) is the two-dimensional Fourier transformof w, and Jν(z) is a Bessel function of the first kind.

To illustrate the parameter dependence of stationarybumps, we consider the concrete example of a weightfunction w given by the difference of modified Besselfunctions (2.5), which has the Fourier transform (4.2).The integral (5.14) can then be evaluated explicitly bysubstituting (4.2) into (5.14), setting r = a, and using theidentity

a∫ ∞

0

1

ρ2 + s2J0(aρ)J1(aρ)dρ = a

sI1(sa)K0(sa),

where Iν is the modified Bessel function of the firstkind. Thus, the condition for existence of a stationarybump of radius a is given by

(1 + αβ)θ = �(a), (5.15)

with

�(a) ≡ �(a, a)

= 4

3

(aI1(a)K0(a) − a

2I1(2a)K0(2a)

). (5.16)

Relations between bump radius a and depressionstrength β are shown in Fig. 14. Numerical simulationssuggest all such bumps are unstable, so that some formof lateral inhibition is required in order to stabilizethe bumps. Alternatively bumps could be stabilized byglobal divisive inhibition (Wu et al. 2008; Fung et al.2008)

While bump existence calculations are straightfor-ward in the case of a Heaviside firing rate function,bump stability calculations are not, due to the piece-wise smooth nature of the depression dynamics. Fol-lowing previous studies of bump stability (Pinto andErmentrout 2001b; Folias and Bressloff 2004; Owenet al. 2007), one could formally linearize the neuralfield equations. However, as we have recently shownin the case of one-dimensional bumps (Kilpatrick andBressloff 2009), considerable care has to be takenin evaluating terms arising from perturbations of thebump boundary. It turns out one needs to keep track ofthe sign of such perturbations, analogous to what hap-pens when θ = 1/(1 + αβ) in the space-clamped system.The details of this analysis will be presented elsewhere.

0 0.02 0.04 0.06 0.08 0.1 0.120

1

2

3

4

5

6

7

8

β

rad

ius

a

κ = 0.05κ = 0.1

κ = 0.2

Fig. 14 Bump radius a as a function of depression strength β fordifferent values of threshold θ , while α = 80

6 Discussion

In this paper, we analyzed the spatiotemporal dynamicsof a two-dimensional excitatory neuronal network withsynaptic depression. We showed that there is an exten-sive parameter range over which spatially structuredoscillations are supported. With the inclusion of noisein the model this range is widened even further. Wefound that application of a localized current input as aninitial condition to the network leads to a localized re-gion of synchronous activity repeatedly emitting targetwaves. This type of activity has been linked to epilepticseizures (Buszaki 2006; Milton and Jung 2003), memory(Klimesch 1999), and sensory input (Lakatos et al. 2007;Lam et al. 2000; Roelfsema et al. 1997; Singer and Gray1995). Additionally, breaking the symmetry of targetwave emitting solutions either using external stimu-lation or noise generated spiral waves. Disinhibitedmammalian cortical slices also support spiral waves,and it has been postulated that such activity allowsperiodic activity to be organized in spatially extendedpopulations of neurons (Huang et al. 2004; Schiff et al.2007). Finally, we showed that in the high-gain limit,the network acts like an excitable rather than oscilla-tory neural medium in which solitary target waves canpropagate but stationary bumps are unstable.

Although the effects of short-term depression withand without noise have been studied in a variety ofcontexts (Abbott et al. 1997; Tsodyks et al. 1998;Matveev and Wang 2000; Tabak et al. 2000; Bart et al.2005; Holcman and Tsodyks 2006), its role in spatiallystructured networks has not previously been consid-ered. Just as synaptic depression provides a negative

Page 16: Spatially structured oscillations in a two-dimensional

208 J Comput Neurosci (2010) 28:193–209

feedback mechanism for generating moderately fastwave (1–10 Hz) oscillations, it is possible that synapticfacilitation of excitatory to inhibitory synapses playsa role in producing slow wave oscillations (0.1–1 Hz),as recently suggested by a modeling study of a space-clamped network (Melamed et al. 2008). We hope topursue this in a future study.

Acknowledgements This publication was based on work sup-ported in part by the National Science Foundation (DMS-0813677) and by Award No KUK-C1-013-4 made by KingAbdullah University of Science and Technology (KAUST). PCBwas also partially supported by the Royal Society-Wolfson Foun-dation. We would like to thank Carlo Laing for helpful con-versations regarding numerical simulations. We also thank BardErmentrout for highlighting issues regarding bump stability cal-culations in the high-gain limit.

References

Abbott, L. F., Varela, J. A., Sen, K., & Nelson, S. B. (1997).Synaptic depression and cortical gain control. Science, 275,220–224.

Amari, S. (1977). Dynamics of pattern formation in lateral-inhibition type neural fields. Biological Cybernetics, 27,77–87.

Bao, S., Chang, E. F., Davis, J. D., Gobeske, K. T., & Merzenich,M. M. (2003). Progressive degradation and subsequent re-finement of acoustic representations in the adult auditorycortex. Journal of Neuroscience, 23, 10765–10775.

Bart, E., Bao, S., & Holcman, D. (2005). Modeling the sponta-neous activity of the auditory cortex. Journal of Computa-tional Neuroscience, 19, 357–378.

Benda, J., & Herz, A. V. M. (2003). A universal model for spike-frequency adaptation. Neural Computation, 15, 2523–2564.

Benucci, A., Frazor, R. A., & Carandini, M. (2007). Standingwaves and traveling waves distinguish two circuits in visualcortex. Neuron, 55, 103–117.

Buszaki, G. (2006). Rhythms of the brain. Oxford: Oxford Uni-versity Press.

Buszaki, G., & Draguhn, A. (2004). Neuronal oscillation in corti-cal networks. Science, 304, 1926–1929.

Chervin, R. D., Pierce, P. A., & Connors, B. W. (1988). Peri-odicity and directionality in the propagation of epileptiformdischarges across neocortex. Journal of Neurophysiology, 60,1695–1713.

Delaney, K., Gelperin, A., Fee, M., Flores, J., Gervais, R., &Tank, D. (1994). Waves and stimulus-modulated dynamics inan oscillating olfactory network. Proceedings of the NationalAcademy of Sciences of the United States of America, 91,669–73.

Dudek, F. E., & Spitz, M. (1997). Hypothetical mechanismsfor the cellular and neurophysiological basis of secondaryepileptogenesis: Proposed role for synaptic reorganization.Journal of Clinical Neurophysiology, 14, 90–101.

Ermentrout, G. B. (1998). Linearization of f-I curves by adapta-tion. Neural Computation, 10, 1721–1729.

Ermentrout, G. B., & Kleinfeld, D. (2001). Traveling electricalwaves in cortex: Insights from phase dynamics and specula-tion on a computational role. Neuron, 29, 33–44.

Folias, S. E., & Bressloff, P. C. (2004). Breathing pulses in an exci-tatory neural network. SIAM Journal on Applied DynamicalSystems, 3, 378–407.

Folias, S. E., & Bressloff, P. C. (2005a). Breathers in two-dimensional neural media. Physical Review Letters, 95,208107.

Folias, S. E., & Bressloff, P. C. (2005b). Stimulus-locked travelingwaves and breathers in an excitatory neural network. SIAMjournal on Applied Mathematics, 65, 2067–2092.

Fung, C. C. A., Wong, K. Y. M., & Wu, S. (2008). Dynam-ics of neural networks with continuous attractors. EPL, 84,18002.

Guo, Y., & Chow, C. C. (2005a). Existence and stability of stand-ing pulses in neural networks: I. Existence. SIAM Journal onApplied Dynamical Systems, 4, 217–248.

Guo, Y., & Chow, C. C. (2005b). Existence and stability of stand-ing pulses in neural networks: II. Stability. SIAM Journal onApplied Dynamical Systems, 4, 249–281.

Han, F., Caporale, N., & Dan, Y. (2008). Reverberation of recentvisual experience in spontaneous cortical waves. Neuron, 60,321–327.

Hansel, D., & Sompolinsky, H. (2001). Methods in neuronal mod-eling. In Modeling feature selectivity in local cortical circuits(2nd Ed., ch. 13, pp. 499–568). Cambridge: MIT.

Holcman, D., & Tsodyks, M. (2006). The emergence of Up andDown states in cortical networks. PLoS Computational Bi-ology, 2, 174–181.

Huang, X., Troy, W. C., Yang, Q., Ma, H., Laing, C. R., Schiff,S. J., et al. (2004). Spiral waves in disinhibited mammalianneocortex. Journal of Neuroscience, 24, 9897–9902.

Hutt, A., Longtin, A., & Schimansky-Geier, L. (2008). Additivenoise-induces Turing transitions in spatial systems with ap-plication to neural fields and the Swift-Hohenberg equation.Physica D, 237, 755–773.

Jung, P., & Mayer-Kress, G. (1995). Spatiotemporal stochasticresonance in excitable media. Physical Review Letters, 74,2130–2133.

Kilpatrick, Z. P., & Bressloff, P. C. (2009). Effects of adap-tation and synaptic depression on spatiotemporal dy-namics of an excitatory neuronal network. Physica D.doi:10.1016/j.physd.2009.06.003.

Klimesch, W. (1999). EEG alpha and theta oscillations reflectcognitive and memory performance: A review and analysis.Brain Research Reviews, 29, 169–195.

Laing, C. R. (2005). Spiral waves in nonlocal equations. SIAMJournal on Applied Dynamical Systems, 4, 588–606.

Laing, C. R., Frewen, T. A., & Kevrekidis, I. G. (2007). Coarse-grained dynamics of an activity bump in a neural field model.Nonlinearity, 20, 2127–2146.

Laing, C. R., & Troy, W. C. (2003). PDE methods for nonlocalmodels. SIAM Journal on Applied Dynamical Systems, 2,487–516.

Lakatos, P., Chen, C. M., O’Connell, M. N., Mills, A., &Schroeder, C. E. (2007). Neuronal oscillations and multi-sensory interaction in primary auditory cortex. Neuron, 53,279–292.

Lam, Y. W., Cohen, L. B., Wachowiak, M., & Zochowski, M. R.(2000). Odors elicit three different oscillations in the turtleolfactory bulb. Journal of Neuroscience, 20, 749–62.

Lee, U., Kim, S., & Jung, K. Y. (2006). Classification of epilepsytypes through global network analysis of scalp electroen-cephalograms. Physical Review E, 73, 041920.

Linder, B., Garcia-Ojalvo, J., Neiman, A., & Schimansky-Geier,L. (2004). Effects of noise in excitable systems. Physics Re-ports, 393, 321–424.

Page 17: Spatially structured oscillations in a two-dimensional

J Comput Neurosci (2010) 28:193–209 209

Matveev, V., & Wang, X. J. (2000). Implications of all-or-nonesynaptic transmission and short-term depression beyondvesicle depletion: A computational study. Journal of Neuro-science, 20, 1575–1588.

McNamara, J. O. (1994). Cellular and molecular basis of epilepsy.Journal of Neuroscience, 14, 3412–3425.

Melamed, O., Barak, O., Silberberg, G., Markram, H., &Tsodyks, M. (2008). Slow oscillations in neural networkswith facilitating synapses, Journal of Computational Neuro-science, 25, 308–316.

Milton, J., & Jung, P. (2003). Epilepsy as a dynamic disease.Berlin: Springer.

Milton, J. G., Chu, P. H., & Cowan, J. D. (1993). Spiral wavesin integrate-and-fire neural etworks. In S. J. Hanson, J. D.Cowan, & C. L. Giles (Eds.), Advances in neural informationn processing systems (pp. 1001–1007). San Mateo: MorganKaufmann.

Owen, M. R., Laing, C. R., & Coombes, S. (2007). Bumps andrings in a two-dimensional neural field: splitting and rota-tional instabilities. New Journal of Physics, 9, 378.

Petersen, C. C. H., Grinvald, A., & Sakmann, B. (2003). Spa-tiotemporal dynamics of sensory responses in layer 2/3of rat barrel cortex measured in vivo by voltage-sensitivedye imaging combined with whole-cell voltage recordingsand neuron reconstructions. Journal of Neuroscience, 23(4),1298–1309.

Pinto, D. J., & Ermentrout, G. B. (2001a). Spatially structured ac-tivity in synaptically coupled neuronal networks: I. Travelingfronts and pulses. SIAM journal on Applied Mathematics, 62,206–225.

Pinto, D. J., & Ermentrout, G. B. (2001b). Spatially structured ac-tivity in synaptically coupled neuronal networks: II. Lateralinhibition and standing pulses. SIAM Journal on AppliedMathematics, 62, 226–243.

Prechtl, J. C., Cohen, L. B., Pesaran, B., Mitra, P. P., & Kleinfeld,D. (1997). Visual stimuli induce waves of electrical activityin turtle cortex. Proceedings of the National Academy ofSciences of the United States of America, 94, 7621–7626.

Roelfsema, P. R., Engel, A. K., Konig, P., & Singer, W. (1997).Visuomotor integration is associated with zero time-lag syn-chronization among cortical areas. Nature, 385, 1157–1161.

Rubin, J., & Bose, A. (2004). Localized activity patterns in exci-tatory neuronal networks. Network, 15, 133–158.

Schiff, S. J., Huang, X., & Wu, J. Y. (2007). Dynamical evolu-tion of spatiotemporal patterns in mammalian middle cortex.Physical Review Letters, 98, 178102.

Schiff, S. J., Sauer, T., Kumar, R., & Weinstein, S. L. (2005). Neu-ronal spatiotemporal pattern discrimination: The dynamicalevolution of seizures. Neuroimage, 28, 1043–1055.

Sederberg, P. B., Kahana, M. J., Howard, M. W., Donner, E. J., &Madsen, J. R. (2003). Theta and gamma oscillations during

encoding predict subsequent recall. Journal of Neuroscience,23, 10809–10814.

Shusterman, V., & Troy, W. C. (2008). From baseline toepileptiform activity: A path to synchronized rhythmic-ity in large-scale neural networks. Physical Review E, 77,061911.

Singer, W., & Gray, C. M. (1995). Visual feature integrationand the temporal correlation hypothesis. Annual Review ofNeuroscience, 18, 555–586.

Stevens, C., & Wesseling, J. (1998). Activity-dependent modula-tion of the rate at which synaptic vesicles become availableto undergo exocytosis. Neuron, 21, 415–424.

Tabak, J., Senn, W., O’Donovan, M. J., & Rinzel, J. (2000). Mod-eling of spontaneous activity in developing spinal cord us-ing activity-dependent depression in an excitatory network.Journal of Neuroscience, 20, 3041–3056.

Troy, W. C. (2008). Traveling waves and synchrony in an ex-citable large-scale neuronal network with asymmetric con-nections. SIAM Journal on Applied Dynamical Systems, 7,1247–1282.

Troy, W. C., & Shusterman, V. (2007). Patterns and featuresof families of traveling waves in large-scale neuronal net-works. SIAM Journal on Applied Dynamical Systems, 6,263–292.

Tsodyks, M. S., & Markram, H. (1997). The neural code betweenneocortical pyramidal neurons depends on neurotransmitterrelease probability. Proceedings of the National Academy ofSciences of the United States of America, 94, 719–723.

Tsodyks, M. S., Pawelzik, K., & Markram, H. (1998). Neuralnetworks with dynamic synapses. Neural Computation, 10,821–835.

Wang, X. J. (1998). Calcium coding and adaptive temporal com-putation in cortical pyramidal neurons. Journal of Neuro-physiology, 79, 1549–1566.

Wang, X. J. (1999). Synaptic basis of cortical persistent activity:the importance of NMDA receptors to working memory.Journal of Neuroscience, 19, 9587–9603.

Wu, J. Y. (2008). Propagating waves of activity in the neocortex:What they are, what they do. Neuroscientist, 14, 487–502.

Wu, J. Y., Guan, L., & Tsau, Y. (1999). Propagating activa-tion during oscillations and evoked responses in neocorticalslices. Journal of Neuroscience, 19, 5005–5015.

Wu, S., Hamaguchi, K., & Amari, S.-I. (2008). Dynamics andcomputation of continuous attractors. Neural Computation,20, 994–1025.

Xu, W., Huang, X., Takagaki, K., & Wu, J. Y. (2007). Compres-sion and reflection of visually evoked cortical waves. Neuron,55, 119–129.

Zucker, R. S., & Regehr, W. G. (2002). Short-term synaptic plas-ticity. Annual Review of Physiology, 64, 355–405.