coefficient of variation of interspike intervals greater than 0.5. how and when?

7
Abstract. Using Stein’s model with and without reversal potentials, we investigated the mechanism of production of spike trains with a CV (ISI) (standard deviation/mean interspike interval) greater than 0.5, as observed in the visual cortex. When the attractor of the deterministic part of the dynamics is below the firing threshold, spike generation results primarily from random fluctuations. Using computer simulation for a range of membrane decay times and with other model parameters set to values appropriate for the visual cortex, we demonstrate that CV (ISI) is then usually greater than 0.5; if the attractor is above the threshold, spike generation is mainly due to deterministic forces, and CV (ISI) is then usually lower than 0.5. The critical value of the inhibitory postsynaptic potential (IPSP) rate at which CV (ISI) becomes greater than 0.5 is determined, resulting in specifications of how neurones might adjust their synaptic inputs to elicit irregular spike trains. 1 Introduction In vivo cell recordings show that most neurones fire irregularly. For example, the coecient of variation (CV), of interspike intervals (ISIs), of neurones in the visual cortex of the monkey is greater than 0.5 [28]. A comparison between in vitro and in vivo experiments strongly supports the assertion that the irregularity results from input from other neurones, both inhibitory and excitatory [14]. How does the variability of the elicited spike train relate to the characteristics of the synaptic input? This is a fundamental question and one of the central themes in computational neuroscience [17, 30]. A better understanding of the origins of irregularity in spike trains will help us to elaborate general principles of neuronal circuitry [20, 21, 25–27] and to test whether rate or timing coding is the fundamental mode of neural communication [12, 15, 27]. In this paper, using the leaky integrate-and-fire model (Stein’s model), we explore how values of CV (ISI) (abbreviated to CV in the remainder of the paper) greater than 0.5 occur for physiologically plausible pa- rameter regions. The firing mechanisms in these model neurones can operate in two ways. The first occurs when the passage of membrane potential to the firing thresh- old is essentially driven by deterministic forces (as de- fined below in Sect. 2.4), possibly supplemented or modified by random fluctuations; that is to say, without these fluctuations the neurone would fire regularly. In this case CV is usually less than 0.5. The other case occurs when the threshold is crossed primarily as a result of random fluctuations; in this case, without the fluctu- ations firing would not occur at all, and CV is normally greater than 0.5. This provides an answer to the ques- tion: how does CV greater than 0.5 occur? Using this theory, we also estimate how large an inhibitory post- synaptic potential (IPSP) rate (or alternatively what number of active inhibitory synapses) is needed for CV to reach 0.5. This answers the question: when does a CV greater than 0.5 occur? We provide evidence for these relationships for a range of parameter values using computer simulations of Stein’s model. To confirm our results using a slightly more realistic model, we consider a model with reversal potentials. Numerical simulations provide the same results as for the leaky integrator without reversal potentials. The number of inhibitory input synapses or the IPSP rate at which CV becomes greater than 0.5 closely matches the number at which random fluctuations are essential for threshold crossings to occur. In recent years there has been much research on the output variability of single neurone models, see for ex- ample [8] and references therein. In particular (see also [23]), Troyer and Miller [32] have pointed out that if postspike voltage is reset to a value higher than the resting potential, the integrate-and-fire model is capable of producing eerent spike trains with a CV greater than 0.5. In the present paper and [8] where the behaviour of the perfect integrate-and-fire model was analysed, we show that there is a wide range of parameters within Biol. Cybern. 80, 291–297 (1999) Coecient of variation of interspike intervals greater than 0.5. How and when? Jianfeng Feng, David Brown Computational Neuroscience Laboratory, The Babraham Institute, Cambridge CB2 4AT, UK Received: 25 June 1998 = Accepted in revised form: 16 December 1998 Correspondence to: J. Feng (e-mail: [email protected])

Upload: jianfeng-feng

Post on 15-Jul-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Coefficient of variation of interspike intervals greater than 0.5. How and when?

Abstract. Using Stein's model with and without reversalpotentials, we investigated the mechanism of productionof spike trains with a CV (ISI) (standard deviation/meaninterspike interval) greater than 0.5, as observed in thevisual cortex. When the attractor of the deterministicpart of the dynamics is below the ®ring threshold, spikegeneration results primarily from random ¯uctuations.Using computer simulation for a range of membranedecay times and with other model parameters set tovalues appropriate for the visual cortex, we demonstratethat CV (ISI) is then usually greater than 0.5; if theattractor is above the threshold, spike generation ismainly due to deterministic forces, and CV (ISI) is thenusually lower than 0.5. The critical value of theinhibitory postsynaptic potential (IPSP) rate at whichCV (ISI) becomes greater than 0.5 is determined,resulting in speci®cations of how neurones might adjusttheir synaptic inputs to elicit irregular spike trains.

1 Introduction

In vivo cell recordings show that most neurones ®reirregularly. For example, the coe�cient of variation(CV), of interspike intervals (ISIs), of neurones in thevisual cortex of the monkey is greater than 0.5 [28].A comparison between in vitro and in vivo experimentsstrongly supports the assertion that the irregularityresults from input from other neurones, both inhibitoryand excitatory [14]. How does the variability of theelicited spike train relate to the characteristics of thesynaptic input? This is a fundamental question and oneof the central themes in computational neuroscience [17,30]. A better understanding of the origins of irregularityin spike trains will help us to elaborate general principlesof neuronal circuitry [20, 21, 25±27] and to test whetherrate or timing coding is the fundamental mode of neuralcommunication [12, 15, 27].

In this paper, using the leaky integrate-and-®re model(Stein's model), we explore how values of CV (ISI)(abbreviated to CV in the remainder of the paper)greater than 0.5 occur for physiologically plausible pa-rameter regions. The ®ring mechanisms in these modelneurones can operate in two ways. The ®rst occurs whenthe passage of membrane potential to the ®ring thresh-old is essentially driven by deterministic forces (as de-®ned below in Sect. 2.4), possibly supplemented ormodi®ed by random ¯uctuations; that is to say, withoutthese ¯uctuations the neurone would ®re regularly. Inthis case CV is usually less than 0.5. The other caseoccurs when the threshold is crossed primarily as a resultof random ¯uctuations; in this case, without the ¯uctu-ations ®ring would not occur at all, and CV is normallygreater than 0.5. This provides an answer to the ques-tion: how does CV greater than 0.5 occur? Using thistheory, we also estimate how large an inhibitory post-synaptic potential (IPSP) rate (or alternatively whatnumber of active inhibitory synapses) is needed for CVto reach 0.5. This answers the question: when does a CVgreater than 0.5 occur? We provide evidence for theserelationships for a range of parameter values usingcomputer simulations of Stein's model.

To con®rm our results using a slightly more realisticmodel, we consider a model with reversal potentials.Numerical simulations provide the same results as forthe leaky integrator without reversal potentials. Thenumber of inhibitory input synapses or the IPSP rate atwhich CV becomes greater than 0.5 closely matches thenumber at which random ¯uctuations are essential forthreshold crossings to occur.

In recent years there has been much research on theoutput variability of single neurone models, see for ex-ample [8] and references therein. In particular (see also[23]), Troyer and Miller [32] have pointed out that ifpostspike voltage is reset to a value higher than theresting potential, the integrate-and-®re model is capableof producing e�erent spike trains with a CV greater than0.5. In the present paper and [8] where the behaviour ofthe perfect integrate-and-®re model was analysed, weshow that there is a wide range of parameters within

Biol. Cybern. 80, 291±297 (1999)

Coe�cient of variation of interspike intervals greater than 0.5.How and when?

Jianfeng Feng, David Brown

Computational Neuroscience Laboratory, The Babraham Institute, Cambridge CB2 4AT, UK

Received: 25 June 1998 =Accepted in revised form: 16 December 1998

Correspondence to: J. Feng (e-mail: [email protected])

Page 2: Coefficient of variation of interspike intervals greater than 0.5. How and when?

which the integrate-and-®re model ®res irregularly, witha CV greater than 0.5. We focus on Stein's model with orwithout reversal potentials, but without the device usedby Troyer and Miller. A major purpose of the presentpaper is to clarify the underlying mechanisms of gener-ation of irregular spike trains for the integrate-and-®remodel. Our work thus di�ers from some earlier studies[32] both in the model used and the purpose of the in-vestigation. Furthermore, to the best of our knowledge,no one has systematically carried out an analysis (con-®rmed by numerical simulations) of CV by linking it tothe position of the equilibrium of the deterministic partof the single neurone dynamics, as we do here.

2 Models

2.1 Stein's model

The basic idea of Stein's model [7, 13, 34] is thatneurones are integrate-and-®re devices charged withincoming excitatory postsynaptic potentials (EPSPs) andIPSPs. The inter-arrival times of single EPSPs and IPSPsare exponentially distributed with rate NEkE and NIkI,respectively, where NE�NI� is the number of a�erent,excitatory (inhibitory) synapses and kE�kI� is the rate ofEPSPs (IPSPs) propagating along each synapse. Themembrane potential Vt at time t is governed by

dVt � ÿ 1

c�Vt ÿ Vrest�dt � a � dNE

t ÿ b � dN It �1�

with V0 � Vrest, where a > 0 and b > 0 are the magnitudeof single EPSPs and IPSPs, Vrest is the resting potential,1=c the decay rate and NE

t ;NIt are Poisson processes with

rate NEkE and NIkI. Once the membrane potentialcrosses the threshold potential, Vth, a spike is elicited,and Vt is reset to Vrest. We take Vrest � ÿ50 mV, Vth is20 mV above the resting potential and a � b � 0:5 mV[30] for Stein's model.

2.2 Lower bounds on membrane potential

There is, however, a substantial problem for the modelde®ned by (1). We know that the voltage of singleneurones is bounded from below, about 10 mV belowthe resting potential, but Vt visits any negative value witha positive probability. There are several ways to preventthis from happening. One is simply to suppose themembrane potential is V �t � max�Vt; Vrest�. The advan-tage of this modi®cation is that all analytical results inthe literature on the ®rst exit time of Vt are valid for V �t[13, 34]. Another way is to impose a lower boundarycondition for Vt and thus obtain a new process vt [30]which is de®ned in the following way.

dv�n�1�t � ÿ 1c �v�n�1�t ÿ Vrest�dt � a � dNE

t ÿ b � dN It

v�n�1�sn�1 � Vrest

(�2�

Let s1 � 0, n � 0, and de®ne a sequence of stoppingtimes by

sn�1 � infft : v�n�t < Vrest; t � sngvt is then de®ned as

Pn v�n�t Ifsn<t<sn�1g, where I is the

indicator function. According to the de®nition above, itis easily seen that vt � Vrest. Obviously, vt and Vt or V �tare di�erent. Suppose that at time t0; Vt0 � Vrest; andVt < Vrest for t0 < t < t1. Vt evolves according to (1) andtherefore below the resting potential; for t0 < t < t1; V �tremains at its resting potential, and any incoming EPSPshave no e�ect. Furthermore, the ®rst passage times of Vtand V �t are identical. vt, however, responds to anyincoming EPSPs, and therefore vt is di�erent from Vt,and their ®rst passage times might be di�erent. V �t mightspend a very long time at the resting potential and notrespond to any incoming signals at all. The de®nition ofvt overcomes this drawback. Interestingly from numer-ical results (see Fig. 1), we conclude that if we concen-trate on CV, which is the purpose of this paper, thebehaviour of these two models is very close, and henceall theoretical results about CV of Vt are good approx-imations to those of vt. In the remainder of this paper,we therefore con®ne ourselves to Vt.

The other, more biologically realistic modi®cation isto include reversal potentials in Stein's model [23, 34]

dZt � ÿ Zt ÿ Vrest

cdt � �a�VE ÿ Zt�dNE

t � �b�VI ÿ Zt�dN It

�3�Zt is now a birth-and-death process with boundaries VE

and VI where VI < Vrest < Vth < VE. Once Zt is below Vrest,the decay term Zt ÿ Vrest will increase membrane poten-tial Zt; whereas when Zt is above Vrest, the decay term willreduce it.

2.3 Deterministic and stochastic components

Usually a discrete process like Stein's model (a birth-and-death process [6]) is hard to deal with theoretically,and so various approximations have been sought. Thedi�usion process [9] serves such a purpose [24, 34]:

dwt � ÿ 1

c�wt ÿ Vrest�dt � l dt � r dBt �4�

with w0 � Vrest, where Bt is standard Brownian motionand

l � aNEkE ÿ bNIkI r2 � a2NEkE � b2NIkI �5�More generally, a stochastic dynamical system can beexpressed as

dXt � b�Xt�dt � r�Xt�dMt �6�where b; r 2 C1�IR� and Mt is an L2-martingale. Wede®ne the following dynamics

dEXt � Eb�Xt�dt �7�as the deterministic part of Xt where E is expectation [3].Note that Vt;wt and Zt are all of form (6). For example,

292

Page 3: Coefficient of variation of interspike intervals greater than 0.5. How and when?

Stein's model with Poisson process inputs can be writtenin this formulation, with deterministic componentyt � EVt given by

dyt � ÿ 1

c�yt ÿ Vrest�dt � l dt �8�

and stochastic component �r � 1�d �Mt � a�dNE

t ÿ NEkE dt� ÿ b�dN It ÿ NIkI dt� �9�

When we say a state y is an attractor of the deterministicpart of Vt we mean that y is an attractor of the dynamics(8).

3 Results: Stein's model

We ®rst consider Stein's model and then move (in thenext section) to a model with reversal potentials. Forconvenience of discussion, we have ®xed a few param-eters: NE � 100 � NI (see [30] for a discussion of thischoice), kE � 100Hz [1], and c lies within the range20:2� 14:6ms [22]. Hence the rate of total excitatoryinputs is 10; 000 Hz, which is approximately equivalentto 300� 33 Hz due to the properties of the Poissonprocess, i.e. N 0E � 300 and k0E � 33. In [30] the authorsemploy N 0E � 300 in their simulations. Furthermore, inour simulations, c takes the values 5.6, 10.1, 20.2,34.8 ms. As pointed out in the previous section, it is

hard to ®nd an informative analytical formula for the®rst exit time of Vt from �ÿ1; Vth�.

3.1 Predicting CV

A direct check on the dynamics de®ned by (1) shows thatthere are two essentially di�erent cases:

1. Vth > lc� Vrest, i.e. the threshold is above the positionof the attractor y of the deterministic part given by

ÿ�y ÿ Vrest�=c� l � 0 �10�The process starting from Vrest will initially be drivenby the deterministic force ÿ�yt ÿ Vrest�=c dt � l dt toapproach lc� Vrest, while also ¯uctuating with arandom motion d �Mt about this trajectory. Once, themembrane potential reaches lc� Vrest, the determin-istic force ± tending either to depolarize or hyperpo-larize the membrane potential ± becomes zero. Therandom ¯uctuations cause the membrane potential to¯uctuate around lc� Vrest and threshold crossingsoccur purely as a result of these random ¯uctuations.We then expect the e�erent spike train to be irregular.

2. Vth < lc� Vrest. In this case, the membrane potentialwould cross the threshold as a result of the deter-ministic force ÿ�yt ÿ Vrest�=c dt � l dt alone, but itstrajectory is modi®ed by random ¯uctuations. Weexpect then that the e�erent spike trains would bequite regular.

Fig. 1. Mean ®ring time and coe�cient of variation (CV) for the two models Vt (lines) and vt (points), kE � 100. We numerically simulate CV and®ring time of vt and Vt at kI � 10; 20; . . . ; 100 for NE � NI � 100. When the ratio r � kI=kE is small both ®ring time and CV of Vt and vt areclose. kd

I is de®ned as the point at which CV equals 0.5 and indicated by arrows, from left to right, for comparison with Fig. 2

293

Page 4: Coefficient of variation of interspike intervals greater than 0.5. How and when?

Note that this analysis does not involve the well-knowntheory of random perturbations (see [3] for a completedescription, [29, 35, 37]) where the random forcebecomes small, and the trajectories of the randomprocesses concentrate around the deterministic trajecto-ries. Then, only with a small probability does therandom process deviate from the deterministic process.As a consequence its CV approaches zero. In the modelwe consider here, the variation of the random term isalways greater than a

�����������NEkEp

. This large random forcewill always introduce appreciable and signi®cant `noise'into the model.

How does the nature of the synaptic inputs relate towhether �Vth ÿ Vrest� < lc or > lc? According to thede®nition of l, we see that the threshold condition�Vth ÿ Vrest� � lc is equivalent to kE ÿ kI � �Vth ÿ Vrest�=�acNE�. Let us denote ka

I � kE ÿ �Vth ÿ Vrest�=�acNE� asthe critical point, the critical level of inhibitory input, atwhich the dynamical attractor equals the threshold.Starting from kI � 0 (see Fig. 2), i.e. input with onlyexcitatory synapses, we see that the position of attractorw is above the threshold. For c � 5:60 ms, ka

I � 29 is thecritical point(indicated by arrow in Fig. 2): for kI < 29the attractor is situated above the threshold; and viceversa for kI > 29. ka

I � 61; 80; 89 are the critical pointsfor c � 10:1 ms, c � 20:2 ms and c � 34:8 ms, respec-tively (indicated by arrows in Fig. 2). In particular, whenan exact balance between excitatory and inhibitory in-

puts is attained, the attractor y is situated at the restingpotential, i.e. there is no deterministic input from syn-aptic input at all.

3.2 Testing the predictions

Corresponding to the cases discriminated above, wecarried out numerical simulations of Stein's model todetermine whether the predictions about CV are correct.We ®nd that the threshold (ka

I ) at which CV is greaterthan 0.5 almost coincides with the threshold at which Vth

becomes less than Vrest � lc as shown in numericalresults (Fig. 1b).We summarise the numerical results inTable 1. More speci®cally, we have the following result:

kdI ÿ 10 < ka

I < kdI � 10 �11�

In conclusion, we see that CV is greater than 0.5 mainlydue to random forces, the increase in which is caused bythe increase in inhibitory synaptic inputs. When inhib-itory inputs and excitatory inputs are poorly balanced,output CV is less than 0.5, whereas a good balanceensures a CV greater than 0.5. CV is thus a goodindicator of the underlying dynamics: CV > 0:5 impliesthat the attractor is almost always below the threshold;CV < 0:5 that the attractor is almost always above thethreshold. Finally, it is impossible to assert ka

I � kdI since

when the attractor is close to the threshold, prediction ofCV becomes uncertain because the `noise' term d �Mt isappreciable.

4 The model with reversal potentials

In this section we assume that VI � ÿ60 mV andVE � 50 mV, values which match experimental dataand, are used in the literature [23]. Again as in theliterature [23], we impose a local balance condition onthe magnitude of each excitating PSP (EPSP) and IPSPby �a�VE ÿ Vrest� � �b�Vrest ÿ VI� � 1 mV, i.e. starting fromthe resting potential, a single EPSP or a single IPSP willdepolarize or hyperpolarize the membrane potential1 mV above or below the resting potential. We alsoconsider the case in which these expressions equal0.5 mV.

4.1 Predicting CV in the presence of reversal potentials

The deterministic part of Zt de®ned by (3) can be writtenas

Fig. 2. Position of attractor y, a linear function of NI, de®ned by (10)vs inhibitory postsynaptic potential (IPSP) rate, for NI � 100 andtotal (EPSP) rate=10,000 Hz. When there are only excitatory inputs,y is above the threshold (20 mV above resting potential which equalsÿ30 mV) no matter what c is, for su�ciently strong input intensity; abetter balance between excitatory and inhibitory inputs pulls it downto the resting potential ÿ30 mV. The values of ka

I at which theattractor is equal to the threshold are indicated by arrows for di�erentc � 5:6; 10:1; 20:2 and 34:8 ms (from left to right), by comparisonwith Fig. 1b

Table 1. kaI and �kd

I ÿ 10; kdI � 10� for c � 5:60; 10:1; 20:2; 34:8 ms

in Stein's model

c (ms) kaI �kd

I ÿ 10; kdI � 10�

5.60 29 [24, 44]10.1 61 [50, 70]20.2 80 [64, 84]34.8 89 [70, 90]

294

Page 5: Coefficient of variation of interspike intervals greater than 0.5. How and when?

dEZt � ÿEZt ÿ Vrest

cdt

� ��a�VE ÿ EZt�NEkE � �b�VI ÿ EZt�NIkI�dt �12�and the equilibrium state is given by

z � Vrest=c� �aNEkEVE � �bNIkIVI

1=c� �aNEkE � �bNIkI�13�

As before we de®ne kaI as the value satisfying

Vth � Vrest=c� �aNEkEVE � �bkaINIVI

1=c� �aNEkE � �bkaINI

We have a similar situation (Fig. 3) as describedabove for Stein's model: a strong imbalance betweenexcitatory and inhibitory inputs, i.e. an equivalent largedeterministic input ensures the attractor z is above thethreshold Vth; a better balance between excitatory andinhibitory inputs moves the attractor towards the restingpotential, here ÿ50 mV.

4.2 Testing the predictions

Numerical results are summarised in Table 2 for Vth �ÿ25 mV (25 mV above the resting potential),Vth � ÿ30 mV (20 mV above the resting potential)and Vth � ÿ35 mV (15 mV above the resting potential).We see that a much better prediction than that of Stein's

model (without reversal potentials) is obtained. Forexample, when Vth � ÿ30 mV we have kd

I � kaI � kd

I � 2.In other words in Stein's model with reversal potentials,kdI is constrained to be closer to ka

I than in the samemodel without reversal potentials.

From numerical results (see Figs. 1b and 4), we seethat one of the di�erences between Stein's model withand without reversal potentials lies in the former beingmuch less sensitive to the decay rate 1=c [19]. This couldbe understood from (4) and (2). If we rewrite (4) in theform

Fig. 3. Position of attractor z de®ned by (13) vs IPSP rate kI, kE � 100with �a�VE ÿ Vrest� � �b�Vrest ÿ VI� � 1 mV, and NE � 100 � NI. Whenthere are only excitatory inputs, z is above the threshold (ÿ30 mV); abetter balance between excitatory and inhibitory inputs moves ittowards the resting potential ÿ50 mV. ka

I ± at which the position ofattractor is equal to the threshold ± is indicated by arrows forc � 5:6; 10:1; 20:2 and 34:8 ms (from left to right)

Table 2. kaI and �kd

I ÿ 10; kdI � 10� for c � 5:6; 10:1; 20:2 and

34.8 ms, Vth � ÿ25, )30 and )35 mV. Results for the case whenVth � ÿ30 mV are plotted in detail in Figs. 3 and 4

c (ms) kaI �kd

I ÿ 10; kdI � 10�

5.60 9 [0, 20]10.1 15 [5, 25]20.2 18 [8, 28]34.8 19 [9, 29]

Vth � ÿ25mV5.60 15 [5, 25]10.1 20 [11, 31]20.2 23 [12, 32]34.8 25 [13, 33]

Vth � ÿ30mV5.60 23 [12, 32]10.1 28 [14, 34]20.2 31 [14, 34]34.8 32 [16, 36]

Vth � ÿ35mV

Fig. 4. CV in simulations of the reversal potentials model forthe case �a�VE ÿ Vrest� � �b�Vrest ÿ VI� � 1 mV, NE � 100 � NI,Vth � ÿ30 mV, calculated at kI � 10; 20; � � � ; 100 for c � 5:60;10:1; 20:2 and 34:8 ms, and kE � 100. kd

I is de®ned as the pointat which CV equals 0.5, indicated by arrows, for comparison withFig. 3

295

Page 6: Coefficient of variation of interspike intervals greater than 0.5. How and when?

dZt �ÿ Zt ÿ Vrest

cdt ÿ �a�Zt ÿ Vrest�dNE

t ÿ �b�Zt ÿ Vrest�dN It

� �a�VE ÿ Vrest�dNEt � �b�VI ÿ Vrest�dN I

t

�ÿ �Zt ÿ Vrest� 1c dt � �a dNE

t � �b dN It

� �� �a�VE ÿ Vrest�dNE

t � �b�VI ÿ Vrest�dN It �14�

we see that in the model with reversal potentials, thereare three `leakage' terms: the ®rst is �Zt ÿ Vrest�=c dt, thesecond and the third are �a�Zt ÿ Vrest�dNE

t and�b�Zt ÿ Vrest�dN I

t , respectively. The latter two terms arestronger than the ®rst term (see the range of c at the endof Sect. 2), i.e. the model with reversal potentials tendsto forget much faster than Stein's model without reversalpotentials and, at the same time, to ensure its activity isless sensitive to the decay rate. Another advantage of(14) is that we can apply the comparison theorem instochastic di�erential equations. We expect that appli-cation of this theorem would provide a theoreticalanswer to a current debate about whether the tail of theoutput interspike interval distribution is long or short[37, 34].

The EPSP and IPSP sizes used in the simulationsabove for the model with reversal potentials are 1 mVwhen Zt � Vrest. However, we use 0.5 mV for the modelwithout reversal potentials. Of course, they are notcompletely comparable since for the model with reversalpotentials, the term �a�VE ÿ Vt� (�b�Vt ÿ VI�) changewhen Vt moves away from Vrest: In particular, whenVt � VI, then the term �b�Vt ÿ VI� vanishes. However, wealso consider the case when �a�VE ÿ Vrest� � a �

�b�Vrest ÿ VI� � b � 0:5 mV. In Figs. 5 and 6 we shownumerical simulations for this case. It is readily seen thatkI at which the position of attractor is the same as thethreshold coincides well with the value at which CVequals 0.5.

5 Discussion

Inputs in the present paper are exclusively Poissonprocess [8], which indicates that all conclusions aboveare true if we ®x input rates kE � 100 Hz, kI � 100 Hzand NE � 100 and replace kI by NI as a parameter. Inrecent years there has been much research devoted toanswering the following question (see [8] and referencestherein): whether there is a region of NI, in agreementwith anatomical data, in which the model generates CVgreater than 0.5. We note here that in the model withreversal potentials, CV is greater than 0.5 for mostvalues of NI, and similar phenomena have been observedin the literature [18, 23, 33]. For example, whenc � 20:2 ms and Vth � ÿ30 mV, as soon asNI > Nd

I � 25, i.e. NI=NE > 0:25, CV is greater than0.5. In our analysis we have considered the case kI � kE.A more e�cient IPSP input will reduce Nd

I further,remembering that the typical composition of corticaltissue is about NI=NE � 1=6 � 0:17 [1]. Therefore, interms of simple statistical quantities like CV, there is nocontradiction between the output of theoretical modelsand experimental data [28]. Furthermore, to produce aCV greater than 0.5, it is not necessary to go to theextreme case: an exact balance between excitatory andinhibitory inputs as discussed in [30, 36] is not needed.

Fig 5. Position of attractor z de®ned by (13) vs IPSP rate kI,NE � 100 � NI, kE � 100 Hz and �a�VE ÿ Vrest� � a � 0:5mV ��b�Vrest ÿ VI� � b. kI at which the position of attractor is the same asthe threshold is indicated by arrows for c � 5:6; 10:1; 20:2 and34:8 ms (from left to right)

Fig. 6. CV vs IPSP rate kI, kE � 100 Hz, and �a�VE ÿ Vrest� � a �0:5mV � �b�Vrest ÿ VI� � b. kI at which CV equals 0.5 is indicated byarrows for di�erent c � 5:6; 10:1; 20:2 and 34:8 ms (from left to right),for comparison with Fig. 5

296

Page 7: Coefficient of variation of interspike intervals greater than 0.5. How and when?

Many analytical, numerical and simulation studieshave attempted to predict how CV of e�erent spiketrains greater than 0.5 might arise for the integrate-and-®re model [11, 18, 19, 28, 30, 32, 33, 35]. Some of thesestudies are particularly concerned with the accuracy andutility of di�usion approximations (see Sect. 2.3). None,so far as we are aware, relate CV to the position of theattractor of the deterministic part of the dynamics as wehave done here. In this paper based upon the simplestmodel of a single neurone (Stein's model) and theequivalent model with reversal potentials, we ®nd thatwhen the position of the attractor of the deterministicpart is below the threshold, CV is greater than 0.5; andvice versa when the attractor is above the threshold. Inother words, in the former case, neuronal ®ring is thendue to purely stochastic forces. There has been muchresearch concerned with the role of noise in neurody-namics [3±5, 10, 16]. Here for the ®rst time, by analysingthe mechanisms of neuronal activity in a very simplecanonical model, we clarify another role played by sto-chastic forces. A separate question of course is whyneurones employ stochastic forces rather than the morereliable deterministic forces to cross the threshold andhence to process information. The question may only beclearly answered when we more fully understand thebrain.

Acknowledgements. We are grateful to Martin Baxter and ananonymous referee for their valuable comments. The work re-ported in this paper was ®nancially supported by the BBSRC andthe Royal Society.

References

1. Abbott LF, Varela JA, Sen K, Nelson SB (1997) Synaptic de-pression and cortical gain control. Science 275:220±223

2. Abeles M (1990) Corticonics. Cambridge University Press,Cambridge, UK

3. Albeverio S, Feng J, Qian M (1995) Role of noises in neuralnetworks. Phys Rev E 52:6593±6606

4. Arieli A, Sterkin A, Grinvald A, Aertsen AD (1996) Dynamicsof ongoing activity: explanation of the large variability inevoked cortical responses Science 273:1868±1871

5. Bezrukov SM, Vodyanoy I (1997) Stochastic resonance in non-dynamical systems without response thresholds. Nature385:319±321

6. Feng J (1996) The hydrodynamic limit for the reaction di�usionequation ± an approach in terms of the GPV method. J TheorProbab 9:285±299

7. Feng J (1997) Behaviours of spike output jitter in the integrate-and-®re model. Phys Rev Lett 79:4505±4508

8. Feng J, Brown D (1998) Impact of temporal variation and thebalance between excitation and inhibition on the output of theperfect integrate-and-®re model. Biol Cybern 78:369±376

9. Feng J, Lei G, Qian M (1992) Second-order algorithms forSDE. J Comput Math 10:376±387

10. Ferster D (1996) Is neural noise just a nuisance? Science273:1812

11. Hanson FB, Tuckwell HC (1983) Di�usion approximation forneuronal activity including synaptic reversal potentials. J TheorNeurobiol 2:127±153

12. Hertz J (1999) Reading the information in the outcome ofneural computation. In Building blocks for intelligent systems(in press)

13. Holden AV (1976) Models of the stochastic activity of neu-rones. Springer, Berlin, Heidelberg, New York

14. Holt GR, Softky WR, Koch C, Douglas RJ (1996) Comparisonof discharge variability in vitro and in Vivo in cat visual cortexneurons. J Neurophysiol 75:1806±1814

15. Hop®eld JJ, Herz AVM (1995) Rapid local synchronization ofaction potentials: towards computation with coupled integrate-and-®re networks. Proc Natl Acad Sci USA 92:6655±6662

16. Jung P, Wiesenfeld K (1997) Too quite to hear a whisper Na-ture 385:291

17. Koch C (1997) Computation and the single neurone Nature385:207±210

18. La nsky P, Musila M (1991) Variable initial depolarization inStein's neuronal model with synaptic reversal potentials. BiolCybern 64:285±291

19. La nsky P, Sacerdote L, Tomassetti F (1995) On the comparisonof Feller and Ornstein-Uhlenbeck models for neural activity.Biol Cybern 73:457±465

20. Mainen ZF, Sejnowski TJ (1995) Reliability of spike timing inneocortical neurones. Science 268:1503±1506

21. Mainen ZF, Sejnowski TJ (1996) In¯uence of dendritic struc-ture on ®ring pattern in model neocortical neurones. Nature382:363±366

22. McCormick DA, Connors BW, Lighthall JW, Prince DA (1985)Cooperative electrophysiology of pyramidal and sparsely spinystellate neurons of the neocortex. J Neurophysiol 54:782±805

23. Musila M, La nsky P (1994) On the interspike intervals calcu-lated from di�usion approximations for Stein's neuronal modelwith reversal potentials. J Theor Biol 171:225±232

24. Ricciardi LM, Sato S (1990) Di�usion process and ®rst-pas-sage-times problems. Lectures in applied mathematics and in-formatics. Manchester University Press, Manchester

25. Rieke F, Warland D, Ruyter van Steveninck R, de Bialek W(1997) Spikes ± exploring the neural code. MIT Press, Cam-bridge, Mass

26. Ruyter van Steveninck RR de, Lewen GD, Strong SP, KoberleR, Bialek W (1997) Reproducibility and variability in neuralspike trains. Science 275:1805±1808

27. Sejnowski TJ (1995) Time for a new neural code? Nature323:21±22

28. Softky W, Koch C (1993) The highly irregular ®ring of cortical-cells is inconsistent with temporal integration of randomEPSPs. J Neurosci 13:334±350

29. Smith CE (1992) A heuristic approach to stochastic models ofsingle neurons. In: McKenna T, David J, Zorrnetzer SF (eds)Single neuron computation. Academic Press, San Diego, 561±588

30. Shadlen MN, Newsome WT (1994) Noise, neural codes andcortical organization Curr Opin Neurobiol 4:569±579

31. Thomson AM (1997) Activity-dependent properties of synaptictransmission at two classes of connections made by rat neo-cortical pyramidal. J Physiol 502:131±147

32. Troyer TW, and Miller KD (1997) Physiological gain leads tohigh ISI variability in a simple model of a cortical regularspiking cell. Neural Comput 9:733±745

33. Tuckwell HC (1979) Synaptic transmission in a model for sto-chastic neural activity. J Theor Biol 77:65±81

34. Tuckwell HC (1988) Stochastic processes in the neurosciences.Society for Industrial and Applied Mathematics, Philadelphia

35. Tuckwell HC, Richter W (1978) Neuronal interspike time dis-tributions and the estimation of neurophysiological and ne-uroanatomical parameters. J Theor Biol 71:167±183

36. Vreeswijk C van, Sompolinsky H (1996) Chaos in neuronalnetworks with balanced excitatory and inhibitory activity.Science 274:1724±1726

37. Wan FYM, Tuckwell HC (1982) Neuronal ®ring and inputsvariability. J Theor neurobiol 1:197±218

38. Wilbur WJ, Rinzel J (1983) A theoretical basis of large coe�-cient of variation and bimodality in neuronal interspike intervaldistribution. J Theor Biol 105:345±368

297