231
UNIVERSIDAD AUTONOMA DE MADRID Facultad de Ciencias Departamento de F´ ısica Te´ orica Computational consequences of Short Term Synaptic Depression Jaime de la Rocha V´ azquez Memoria de Tesis Presentada en la Facultad de Ciencias de la Universidad Aut´ onoma de Madrid para optar al grado de Doctor en Ciencias F´ ısicas Trabajo dirigido por estor Parga Carballeda Madrid, 26 de noviembre de 2002

 · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

  • Upload
    others

  • View
    15

  • Download
    0

Embed Size (px)

Citation preview

Page 1:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

UNIVERSIDAD AUTONOMA DE MADRIDFacultad de Ciencias

Departamento de Fısica Teorica

Computational consequences of Short TermSynaptic Depression

Jaime de la Rocha Vazquez

Memoria de Tesis

Presentada en la Facultad de Ciencias

de la Universidad Autonoma de Madrid

para optar al grado de Doctor en Ciencias Fısicas

Trabajo dirigido porNestor Parga Carballeda

Madrid, 26 de noviembre de 2002

Page 2:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 3:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Dedico esta tesis a mis padres, Paloma y Manolo,

por todo el apoyo y el carino recibido, y por todo

lo que me han esenado.

Tambien a Monica, por compartir su vida conmigo.

Page 4:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 5:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Contents

Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

Acknowledgments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

Preface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

1 Introduction: the synaptic function 1

1.1 Anatomical description of a synapse. . . . . . . . . . . . . . . . . . . . . 1

1.2 Synaptic transmission and synaptic dynamics. . . . . . . . . . . . . . . . 3

1.2.1 Unreliability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2.2 Short term depression. . . . . . . . . . . . . . . . . . . . . . . . 5

1.2.3 Facilitation and other mechanisms. . . . . . . . . . . . . . . . . . 7

1.2.4 Univesicular release hypothesis. . . . . . . . . . . . . . . . . . . 8

1.2.5 Synaptic diversity. . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.3 Objectives and overview of this work. . . . . . . . . . . . . . . . . . . . . 10

2 A model of synaptic depression and unreliability 13

2.1 Experimental motivation of the model. . . . . . . . . . . . . . . . . . . . 13

2.2 Methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.2.1 Model of one synaptic contact with many docking sites. . . . . . . 15

2.2.1.1 Vesicle Release. . . . . . . . . . . . . . . . . . . . . . 15

2.2.1.2 Vesicles Recovery. . . . . . . . . . . . . . . . . . . . . 16

2.2.2 Input spike-train statistics. . . . . . . . . . . . . . . . . . . . . . 18

2.2.3 Statistics of the synaptic response. . . . . . . . . . . . . . . . . . 22

2.2.4 Population of synapses. . . . . . . . . . . . . . . . . . . . . . . . 28

2.3 Results: synaptic response statistics. . . . . . . . . . . . . . . . . . . . . 30

2.3.1 Single docking site:N0 = 1 . . . . . . . . . . . . . . . . . . . . . 31

2.3.1.1 Poisson input . . . . . . . . . . . . . . . . . . . . . . . 32

2.3.1.2 Correlated input. . . . . . . . . . . . . . . . . . . . . . 34

2.3.2 Multiple docking sites:N0 > 1 . . . . . . . . . . . . . . . . . . . 38

2.4 Tables of symbols. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

i

Page 6:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

ii Contents

3 Information transmission through synapses with STD 47

3.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

3.2 Methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.2.1 Fisher Information. . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.2.2 Mutual Information. . . . . . . . . . . . . . . . . . . . . . . . . . 52

3.2.3 Mutual Information in a population of synapses. . . . . . . . . . . 53

3.2.4 Optimization with Metabolic considerations in the recovery rate. . 55

3.2.5 Numerical methods. . . . . . . . . . . . . . . . . . . . . . . . . . 57

3.3 Results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

3.3.1 Rate dependence of information measures. . . . . . . . . . . . . . 58

3.3.1.1 Dependence of the Fisher information onν . . . . . . . . 58

3.3.1.2 Dependence of the Mutual Information onf(ν) . . . . . 62

3.3.2 Optimization of the recovery time constantτv . . . . . . . . . . . . 65

3.3.2.1 Optimizingτv with the Fisher Information. . . . . . . . 65

3.3.2.2 Optimizingτv with the Mutual Information. . . . . . . . 69

3.3.2.3 Several vesicles:N0 ≥ 1 . . . . . . . . . . . . . . . . . 71

3.3.2.4 Dependence ofτopt on the other parameters. . . . . . . . 78

3.3.2.5 Metabolic considerations. . . . . . . . . . . . . . . . . 82

3.3.3 Optimization of the release probabilityU . . . . . . . . . . . . . . 85

3.3.3.1 OptimizingU with the Fisher Information . . . . . . . . 85

3.3.3.2 OptimizingU with the Mutual Information. . . . . . . . 90

3.3.4 Optimization of the distribution of synaptic parameters. . . . . . . 93

3.4 Conclusions and Discussion. . . . . . . . . . . . . . . . . . . . . . . . . 96

4 Synaptic current produced by synchronized neurons 103

4.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .103

4.2 Parameterization of the afferent current. . . . . . . . . . . . . . . . . . . 105

4.3 Afferent spike trains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

4.4 Statistics of the synaptic releases. . . . . . . . . . . . . . . . . . . . . . . 109

4.4.1 Dynamics of one synaptic contact between two neurons. . . . . . 109

4.4.2 Several synaptic contacts between two neurons. . . . . . . . . . . 113

4.4.3 Release correlations among two synapses from different neurons. . 117

4.5 Statistics of the total afferent current. . . . . . . . . . . . . . . . . . . . . 118

4.5.1 The mean of the afferent current. . . . . . . . . . . . . . . . . . . 119

4.5.2 Correlations of the current. . . . . . . . . . . . . . . . . . . . . . 120

4.5.2.1 Auto-correlation in single contacts. . . . . . . . . . . . 121

Page 7:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Contents iii

4.5.2.2 Cross-correlation between pairs of contacts with the same

input train . . . . . . . . . . . . . . . . . . . . . . . . . 121

4.5.2.3 Cross-correlation between pairs of contacts with corre-

lated input trains. . . . . . . . . . . . . . . . . . . . . . 123

4.5.2.4 Total current correlation. . . . . . . . . . . . . . . . . . 124

5 The response of a LIF neuron with depressing synapses 129

5.1 The Leaky Integrate and fire (LIF) neuron. . . . . . . . . . . . . . . . . . 129

5.2 The analytical calculation of the output rate of a LIF neuron. . . . . . . . 130

5.2.1 The diffusion approximation. . . . . . . . . . . . . . . . . . . . . 131

5.2.2 The solution ofνout for awhite noiseinput . . . . . . . . . . . . . 133

5.2.3 Perturbative solution ofνout for a correlated input. . . . . . . . . . 134

5.2.4 Several input populations. . . . . . . . . . . . . . . . . . . . . . . 135

5.3 Results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .136

5.3.1 The saturation ofµ and the subthreshold regime. . . . . . . . . . . 138

5.3.2 The modulation of the variance. . . . . . . . . . . . . . . . . . . 141

5.3.2.1 VaryingM , with CM constant . . . . . . . . . . . . . . 141

5.3.2.2 VaryingM , with MJ constant . . . . . . . . . . . . . . 147

5.3.3 The output rate of a LIF neuron. . . . . . . . . . . . . . . . . . . 152

5.3.3.1 Output rate atCM fixed . . . . . . . . . . . . . . . . . . 152

5.3.3.2 Saturations eliminates synchrony. . . . . . . . . . . . . 154

5.3.3.3 Output rate atMJ fixed . . . . . . . . . . . . . . . . . . 156

5.3.4 Information beyond saturation of the mean current. . . . . . . . . 161

5.4 An interesting experiment. . . . . . . . . . . . . . . . . . . . . . . . . . . 163

5.5 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .164

A The exponentially correlated input 167

B Computing ρiri(t|ν) for any renewal input 173

C The calculation of the population distribution D(U,N0) 179

D Computation of the output whenN0 = 1 183

D.1 Computation of the p.d.f. of the IRIsρiri(t) . . . . . . . . . . . . . . . . . 183

D.2 Computation of the correlation function of the IRIs. . . . . . . . . . . . . 185

D.3 Release firing rateνr and coefficient of variationCViri . . . . . . . . . . . 186

E Computation of the conditioned probability 〈pv(β|α)〉 189

Page 8:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

iv Contents

F Computation of the conditioned probability 〈pv(j|i)〉 193

Bibliography 195

List of Figures 213

List of Tables 217

Page 9:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Agradecimientos

Quiero dar las gracias, en primer lugar, a Nestor, por haber estado siempre ahı y haberme

ensenado tanto. En segundo lugar, a mis companeros del grupo: a Jose por su ayuda en los

ultimos momentos; a Ruben por todo lo aprendido juntos; a Gonzalo por tan provechosos

comentarios sobre la tesis ; a Angel por su sabia lentitud y su talante tan generoso; a Alfonso

por ensenarme que las cuentas pueden ser bonitas, y a apasionarme por este trabajo. A mis

contemporaneos Enrique, Alex, Ernesto, David T. y el bueno de Yago con el que tantos anos

he compartido despacho. Tengo especial gratitud para Natxo por toda su ayuda y por su

carino, y para Stephane por todo lo compartido.

Gracias tambien a mis padres por su apoyo incondicional y a mis hermanos Manolo,

Marta y Miguel. A los amigos de siempre, Isra, Marcos, Cucho, Manus, Juan, Vıctor e Ivan.

Tambien a Alberto, Yago, Dani y Pardo. A mis companeros de carrera, Ivan, Estela, Raul,

Elena, etc.

Gracias en especial a Monica, por susplannings, sus esfuerzos, sus cuidados y sus mi-

mos.

v

Page 10:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 11:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Preface

The nervous systems of an organism collects information about its environment, processes

it, and computes a behavioral response. Its survival in its enviromment relies on its capa-

bility to perform this task in an optimal manner. An interesting idea is that the evolution,

viewed as the natural selection of the organisms which best adapt to the environment, can be

interpreted as an optimization of the strategies the animal uses to overcome the difficulties

that its survival prompts.

A key optimization of the system would be to obtain an efficientinternal representa-

tion of the sensory world. This information is represented in the brain by the activity of the

neurons, that is, the emission of action potentials or spikes. In particular, the activity rate

of specific populations of neurons is believed to convey, at least, part of this information

[Adrian, 1926, Werner and Mountcastle, 1965, Tolhurst et al., 1983, Tolhurst, 1989, Britten

et al., 1992, Tovee et al., 1993]. The way this information is processed and the computa-

tions which are performed are still a puzzle for neuroscientists. Some operations seem to be

performed by cortical microcircuits, whereas the capability of individual neurons to accom-

plish computations is still an open question [Koch and Segev, 2000]. Whatever the case, if

information is encoded in the firing frequency, neurons receiving those spikes downstream

need to have an efficient and robust internal representation of those pre-synaptic rates. The

striking feature about this description is that the pre-synaptic spikes reach the neuron through

a noisy dynamical channel: the chemical synapse.

The motivation of my work is to study whether or not the dynamical properties of the

synapses improve the internal representation of the pre-synaptic information. I will focus on

synaptic unreliabilityandshort-term depressionand I will address the question of whether

their existence can be postulated in terms of an optimization criterion. In other words, I will

optimize the transmission of information through a model synapse by tuning its parameters,

and check if the resulting model resembles the biology described in the experiments. An

additional working hypothesis is that this could occur if the information is encoded in the

spike trains in a redundant fashion so that synapses may optimize the internal representation

by filtering out that redundancy [Barlow, 1961, Atick, 1992, Nadal and Parga, 1994, Dan

vii

Page 12:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

viii Contents

et al., 1996, Nadal et al., 1998, Goldman et al., 2002].

In order to perform the optimization of the synaptic channel, I will quantify the trans-

fer of information across the synapse by using Information Theory. Information Theory

[Shannon, 1948] is a mathematical framework which has been widely used in Computational

Neuroscience to quantify how much information is encoded in neuronal activity [Borst and

Theunissen, 1999]. In this thesis it will be used to quantify the goodness of the representa-

tion provided by the synaptic responses (or transmitter releases) about the pre-synaptic firing

rate.

Neurons, on the other hand, must not only obtain an efficient representation of the in-

coming information, but process it and eventually carry out simple computations. In the last

years the notion of what neurons can do has evolved enormously. The idea that neurons are

just simple current integrators has been overcome by showing that neurons are provided with

the necessary biophysical machinery (such as dynamic synapses, active conductances, etc)

to process the pre-synaptic activity in a rather sophisticated manner. The role of synaptic dy-

namics in increasing the computational capabilities of neurons has just started to be explored

[Abbott et al., 1997, Tsodyks and Markram, 1997, Lisman, 1997, Senn et al., 1998, Chance

et al., 1998a, Maass and Zador, 1999, Matveev and Wang, 2000a, Natschlager and Maass,

2001, Maass and Markram, 2002, Fuhrmann et al., 2002, de la Rocha et al., 2002]. For exam-

ple, an important consequence of short-term depression is that it constraints the range of rates

that can be transmitted to the post-synaptic neuron: since the recovery of synaptic resources

takes about a few hundreds of millisecondsindependentlyof the pre-synaptic frequency, it

sets an upper bound to the synaptic activity rate [Abbott et al., 1997, Tsodyks and Markram,

1997]. Beyond this limit, the synapse saturates and the mean post-synaptic current becomes

independent of presynaptic rate. However, neurons are driven not only by mean synaptic ac-

tivity but also by the fluctuations of this activity around the mean [Amit and Tsodyks, 1991,

Shadlen and Newsome, 1994, 1998]. Thus, the stochasticity of synaptic transmission and the

correlations of the input spike trains can play a crucial role in driving the neuronal response.

I will explore the implications of short-term depression and the stochasticity of transmission

when the presynaptic population shows correlated activity. I will show how both aspects may

combine, yielding to a non-monotonic transfer function, which enables the neuron to extract

information about the input rate, even beyond the saturation of the mean synaptic current.

Page 13:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Contents ix

Part of the work presented in this thesis has been published or presented in the following

articles and conferences:

• Jaime de la Rocha, Angel Nevado and Nestor Parga.

Information transmission by stochastic synapses with short term depression: neural

coding and optimization.

Neucomputing, 44:85-90 (2002).

• Jaime de la Rocha, Angel Nevado and Nestor Parga

Information transmission by stochastic synapses with short-term depression: neural

coding and optimization.

Oral communication in the Tenth Annual Computational Neuroscience Meeting.

(CNS2001. Monterey - Pacific Grove, USA. June 30 - July 5, 2001)

• Nestor Parga, Jaime de la Rocha and Angel Nevado.

Information transmition of correlated spike trains through stochastic synapses with

short-term depression.

Poster communication in the 31th Annual Meeting of the Society for Neuroscience

(San Diego. December, 2001).

Society for Neuroscience Abstracts 27.

• Nestor Parga, Jaime de la Rocha and Angel Nevado.

Information processing by depressing synapses.

Oral communication in the Program on Dynamics of Neural Networks: From bio-

physics to behavior. (Jul-Dec 2001).

Institute for Theoretical Physics, University of Santa Barbara, USA.

• Nestor Parga, Jaime de la Rocha and Angel Nevado

Optimizing information processing by depressing synapses

Oral communication in the Trimester on Neuroscience and Computation.

Centre Emile Borel - Institut Henri Poincare, Paris. February 2002

• Ruben Moreno, Jaime de la Rocha and Nestor Parga.

Response of a leaky integrate and fire neuron when stimulated with synapses showing

short-term depression.

Poster communication in the 32th Annual Meeting of the Society for Neuroscience

(Orlando. November, 2002).

Society for Neuroscience Abstracts28: 752.6

Page 14:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

x Contents

• Ruben Moreno, Jaime de la Rocha, Alfonso Renart and Nestor Parga.

Response of spiking neurons to correlated inputs.

Physical Review Letters, 89 (28), 288101 (2002)

Page 15:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Chapter 1

Introduction: the synaptic function

1.1 Anatomical description of a synapse

Neurons in the nervous system (NS) communicate with each other throughsynapses.

The synapse is the connection where two or more neurons establish a functional contact,

that is, a “touching” point where a communication channel is built. There are two basic

types of synapses:electricalandchemical. In the first type, also calledgap junctions, neu-

rons exchange ions that cross the plasma membrane of both cells through a special kind of

channels called connexons. The second, and more common type of synapses in mammals

are the chemical synapses in which neurons “talk” to each other by means of chemical sub-

stances, theneurotransmitters. In the most common case, an axonal fiber forms a cavity

called synaptic bouton, or pre-synaptic terminal, which is located very close to a dendritic

spine1, to the dendritic shaft or to the soma (cell body) of the post-synaptic cell. The pre- and

post-synaptic membranes are separated by the synaptic cleft which is approximately20− 50

nm wide. The pre-synaptic bouton is filled with dozens of little membrane-enclosed sacks

(50 nm of diameter), eventually filled with transmitter, which are called synaptic vesicles.

At a certain number of specific locations of the synapse, both pre- and post-synaptic neurons

develop the necessary machinery which mediates release and reception of transmitter. These

particular areas within the synapse, which look very dense under the electronic microscope,

are called synapticspecializations(see micrograph in figure1.1 and picture in figure1.2).

The number of specializations in a synaptic bouton, depends on the synapse type and on its

size. Glutamatergic synapses onto CA1 pyramidal neurons in the hippocampus often display

one or two [Sorra and Harris, 1993, Schikorski and Stevens, 1997]. Other type of larger

boutons may have many separate specializations (from15 − 20 in the case of the contacts

1A small bag of membrane that protrudes from the dendrites of some cells and receives synaptic input [Bear

et al., 1996].

1

Page 16:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2 Chapter 1: Introduction: the synaptic function

Figure 1.1: Electron micrograph of a synapse from the stratum radiatum in CA1 in the

hippocampus of an adult mouse. Three active zones are delimited by arrows at both pre-

and post-synaptic side. Head arrows indicate the presence of a docked vesicle. Scale bar:

0.5 µm. (Taken from [Schikorski and Stevens, 1997])

between spindle afferents and spinocerebellar tract cells [Walmsley, 1991], to hundreds in

the neuromuscular junctions [del Castillo and Katz, 1954a, Katz and Miledi, 1968] or the

calyx of Held synapse [Held, 1893, Ryugo et al., 1996] in the auditory pathway).

At the pre-synaptic side of the specialization, also defined asactive zone, in the plasma

membrane we can find: i) voltage-gated calcium channels; ii) a series of proteins, with a

filament shape, which are responsible for keeping some vesicles attached ordockedat the

membrane (see head arrows in the micrograph in fig.1.1). Upon the influx of calcium (see

yellow trace in fig.1.2), these proteins can make one of thedockedvesicles fuse its membrane

with the cell membrane. In this fusion operation, calledexocytosis, the transmitter content

of the vesicle is released into the synaptic cleft (see red trace in fig.1.2). The signal that

triggers the release, i.e. the elevation of the concentration of intracellular calcium[Ca2+]i, is

elicited by the arrival of an action potential (AP) to the pre-synaptic terminal. The released

neurotransmitter diffuses throughout the synaptic cleft eventually reaching the post-synaptic

membrane, where it binds to specific receptors. This causes the influx of a post-synaptic

current (PSC) into the membrane of the post-synaptic cell, via directly coupled channels or

activation of second messenger pathways. Depending on the type of synapse, the transmitter

released may cause a depolarization of the post-synaptic cell by means of an excitatory post-

synaptic current or EPSC (e.g. if the transmitter is glutamate), or an hyper-polarization by

means of an inhibitory post-synaptic current (IPSC), e.g. if the transmitter is GABA.

Page 17:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

1.2. Synaptic transmission and synaptic dynamics 3

Figure 1.2: Description of the morphology of a synapse and the mechanism of vesicular

release. Top: A synaptic bouton with three synaptic specializations (or active zones) is

shown. Vesicles are painted in red.Bottom: Detailed description of an active zone. Many

vesicles are separated from the membrane (reserve pool) while the one on the right (touching

the plasma membrane) represents a docked vesicle. On the center an exocytosis event is

taking place: influx of calcium (shown in yellow) has caused a nearby docked vesicle to fuse

with the membrane and release its transmitter content (shown in red) to the synaptic cleft.

(Taken from [Walmsley et al., 1998])

1.2 Synaptic transmission and synaptic dynamics

Although the basic picture of how synaptic transmission occurs is partly described by

the previous explanation, and it is believed to be similar for most synapses in both inver-

tebrates and vertebrates, many important mechanisms involved in synaptic transmission are

not completely understood. Here, we will discuss briefly the main properties of synaptic

transmission which are relevant for this work.

Page 18:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4 Chapter 1: Introduction: the synaptic function

1.2.1 Unreliability

Central synapses are usually very unreliable [Hessler et al., 1993, Rosenmund et al.,

1993, Allen and Stevens, 1994, Stevens and Wang, 1994]. Only a fraction of the incoming

AP’s elicit measurable responses at the post-synaptic terminal. Thus, transmission has been

modeled as an stochastic process with a certain probability of release,pr [del Castillo and

Katz, 1954a, Boyd and Martin, 1956].

Katz and colleagues [del Castillo and Katz, 1954a, Katz, 1969] discovered the quantal

nature of transmission at the neuromuscular junction of the frog. They first described the

synaptic transmission as a probabilistic process and initiated what has become a widely used

method in the study of synapses,quantal analysis[del Castillo and Katz, 1954a, Boyd and

Martin, 1956]. This method assumes that the pre-synaptic terminal hasn release sites, where

individual vesicles may independently fuse to the membrane with a uniform probabilityp.

The morphological correlate of a release site (which is an heuristic definition coming from

quantal analysis) can be an active zone if, at each of them, only a single vesicle release

can occur upon the arrival of an AP (see section1.2.4below about the uni-vesicular release

hypothesis). If on the contrary, one assumes that multiple vesicles can be released at a single

active zone, then the number of release sites within an active zone would be bigger than one.

The method of quantal analysis is completely specified by giving a third parameterQ, which

represents the mean amplitude of an EPSC elicited by the release of a single vesicle. Thus,

the number of releases upon arrival of a spike follows a binomial distribution of parameters

n andp. Via the analysis of the histogram of the synaptic response amplitude (the size of the

PSC), one can estimate the three parametersn, p andQ [Tuckwell, 1988].

In synapses like the neuromuscular junction, where the number of release sites (or the

maximum number of vesicles which simultaneously undergo exocytosis) is of the order of

several hundreds, unreliability produces fluctuations from stimulus to stimulus of the re-

sponse amplitude. On the other hand, the effect of unreliability in synapses with only one or

two release sites is much more noticeable. A probability of transmission of e.g.0.1 implies

that, on average, nine out of ten spikes would fail to transmit any information. Whether this

represents a limitation of the brain wet-ware or if, on the contrary, it represents a functional

advantage for the NS is still and open and interesting issue [Smetters and Zador, 1996].

This last possibility seems more plausible in the light of observations showing failure-free

synapses outside [Paulsen and Heggelund, 1994, 1996, Bellingham et al., 1998] and inside

[Stratford et al., 1996] the brain. If the NS is able to produce reliable contacts but most of

them are unreliable, the latter are likely to be advantageous in some sense. Recently [Levy

and Baxter, 2002] have argued that unreliability may be an optimal way of transmission

under energy constraints. In contrast, other studies which explored the impact of synaptic

Page 19:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

1.2. Synaptic transmission and synaptic dynamics 5

failures in information transmission without metabolic constraints, have shown that unrelia-

bility severely reduces the information content of the output [Zador, 1998, Fuhrmann et al.,

2002]. In addition it has been shown that the release probabilitypr is subject to plasticity

[Markram and Tsodyks, 1996a] so that it could be involved in long-term synaptic changes

related with experience and learning.

The causes of unreliability in the transmission have been intensively investigated inin

vitro experiments [Rosenmund et al., 1993, Hessler et al., 1993, Allen and Stevens, 1994].

Although the mechanisms are not entirely clear, several experiments have shed light over

several points: first, synaptic unreliability is not due to a failure of nerve impulses to arrive

to the synapse [Allen and Stevens, 1994]. Second, release probability correlates with the

number of morphologically identified vesicle docking sites [Schikorski and Stevens, 1997],

which are special positions in the active zone where vesicles dock before releasing the trans-

mitter. The number of docking sites, in turn is positively correlated with the size of the

active zone [Schikorski and Stevens, 1997], which implies that release probability is larger

at larger synaptic specializations. In addition, several features of the calcium channels af-

fect the probability of release: the number and type of voltage-gated calcium channels at the

pres-ynaptic terminal; the spacing between docking sites and calcium channels together with

the extracellular calcium concentration (see [Atwood and Karunanithi, 2002] for a review).

On the other hand, variations of temperature affect very little the release probability [Allen

and Stevens, 1994]. Despite many experimental observations of release failure in culture

neurons or in slice preparations, the probabilistic nature of transmission on central synapses

still needs to be shownin vivo.

1.2.2 Short term depression

At many synapses, pre-synaptic activity dynamically affects synaptic strength (see e.g.

[Magelby, 1987, Fisher et al., 1997, Zucker, 1989, Zador and Dobrunz, 1997] or [von Gers-

dorff and Borst, 2002, Zucker and Regehr, 2002] for recent reviews). Thus, the amplitude of

the post-synaptic response is not a static quantity but, besides fluctuations due to noise, it is

an activity dependent magnitude: recent activity can either decrease or enhance the efficacy

of the synapse. All these changes, which have a time scale ranging from milliseconds to at

most a few minutes, are referred to asshort-term plasticity(STP).

Short-term depression is one of the most common expressions of these changes. At

synapses with this type of modulation, pre-synaptic activity produces a decrease in synaptic

strength [Magelby, 1987]. The most common explanation for this decrease in efficacy is

the reduction of transmitter release due to depletion of synaptic vesicles (see [Liley and

North, 1953, Hubbard, 1963, Stevens and Wang, 1995] and [Zucker, 1996, Neher, 1998,

Page 20:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

6 Chapter 1: Introduction: the synaptic function

Schneggenburger et al., 2002] for reviews). There are different models of vesicle depletion

with different degrees of physiological detail [Dobrunz and Stevens, 1997, Weis et al., 1999,

Matveev and Wang, 2000b, Trommershauser, 2000]. A visual common general assumption

is that, at a single active zone, only a small pool of the pre-synaptic vesicles, thereadily

releasable pool(RRP) [Rosenmund and Stevens, 1996], are prepared to be immediately

released upon a sudden increase of[Ca2+]i produced by the arrival of an AP. The size of

this functional pool (i.e. the RRP) correlates with the morphological definition of the pool

of docked vesicles [Schikorski and Stevens, 2001], i.e. the set of vesicles which, when seen

through the electronic microscope, appear tethered to the cell membrane2. In the micrograph

shown in fig. 1.1 the head arrows indicate what is believed to be a docked vesicle at a

hippocampal synapse. A key finding, as described by Dobrunz and Stevens, is that the

release probabilitypr covaries with the number of vesicles in the RRP across a population

of synapses [Dobrunz and Stevens, 1997], and at single synapses [Dobrunz, 2002]. In other

words, the more vesicles waiting ready to be released, the more likely it is that the exocytosis

of one of them takes place.

Once a pre-synaptic pulse triggers the exocytosis of one of theseready-for-releasevesi-

cles, it takes some time until it undergoesendocytosis(the mechanism by which the vesi-

cle becomes separated from the plasma membrane ) and it is refilled with transmitter again

[Murthy and Stevens, 1998, Sun et al., 2002]. In the meanwhile, after the release, the number

of vesicles at the RRP is transientlyreduced, resulting in a transitory decrease of the trans-

mission probability. This means that the synapse is momentarily depressed. If one leaves

the contact rest for a period of time of the order of a second, that vesicle or new ones which

were already filled of transmitter but not ready for release (commonly described as belong-

ing to the reserve pool) may dock making the transmission probability return to its rest value

[Dobrunz and Stevens, 1997, Matveev and Wang, 2000b].

Little is known about how the RRP recovers. In most cases it has been observed that

the transmission probability recovers following a single exponential in both, hippocam-

2More specifically, the RRP is believed to be a subset of the docked pool, i.e. to be docked is a necessary

but not sufficient condition to beready to go[Atwood and Karunanithi, 2002]. More sophisticated versions

of vesicle turnover models [Matveev and Wang, 2000b, Trommershauser, 2000] define a third pool of vesicles

(besides the RRP and the reserve pool), usually called theprimedpool or pre-primed pool. Vesicles in the

reserve pool first approach the membrane and “dock”. In a subsequent step, they areprimed(the configuration

of the proteins which hold the vesicle attached to the membrane changes) so that the vesicle is then ready to be

fused when[Ca2+]i increases. While the RRP has been estimated to be formed by about a dozen vesicles, the

primed pool is supposed to be composed of only one or two vesicles (see [Zucker, 1996, Zenisek et al., 2000]

for biophysical details and [Weis et al., 1999, Matveev and Wang, 2000b] for modeling). However, due to

analytical calculation constraints, we will not make that distinction in this thesis. Therefore, the termsdocked

andprimedwill denote the vesicles in the RRP.

Page 21:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

1.2. Synaptic transmission and synaptic dynamics 7

pal [Stevens and Tsujimoto, 1994, Dobrunz and Stevens, 1997], and neocortical synapses

[Markram et al., 1998a, Wang and Kaczmarek, 1998, Finnerty et al., 1999, Varela et al.,

1999, Petersen, 2002], with a time constant which ranges between half a second and several

seconds. Sometimes two exponentials have been used to fit a fast recovery (∼ 0.5s) along

with a slower recovery (∼ 5 − 10 s) [Varela et al., 1997]. Although in the first genera-

tion of models recovery was described as an activity independent process, several findings

have shown that the[Ca2+]i affects the rate at which vesicles become primed, i.e. the RRP

[Dittman and Regehr, 1998]. Since[Ca2+]i is basically modulated by pre-synaptic activity,

some works have experimentally shown that the stimulus input rate affects the recovery rate

[Wang and Kaczmarek, 1998, Stevens and Wesseling, 1998, Neher and Sakaba, 2001]. For

that reason, some models of vesicle depletion have included this calcium dependency of the

recovery rate in the vesicle dynamics [Weis et al., 1999, Dittman et al., 2000].

There are, however, other underlying mechanisms apart from vesicle depletion, which

could also be involved in short-term depression [Dobrunz et al., 1997, Hanse and Gustafs-

son, 2002, Zucker and Regehr, 2002]. In the pre-synaptic side the release of transmitter into

the synaptic cleft can lead to activation of auto-receptors which are usually inhibitory. This

would result in a negative feedback and a reduction of future transmitter release [Zucker and

Regehr, 2002]. Other mechanisms have been proposed such as the so-calledinactivationof

exocytosis machinery [Hsu et al., 1996, Dobrunz et al., 1997], which differs from the previ-

ous ones in that now the probability of vesicle fusion diminishes after the production of an

exocytosis event even when[Ca2+]i is maintained at a high concentration [Hsu et al., 1996].

On the post-synaptic side of the synapse, the desensitization of post-synaptic receptors can

also lead to a use-dependent decrease of the response (see [Jones and Westbrook, 1996] for

a review).

1.2.3 Facilitation and other mechanisms

Short-term facilitation is one of the mechanisms by which pre-synaptic activity produces

an enhancement of the synaptic response [Fisher et al., 1997, Zucker and Regehr, 2002]. It

has a time scale of hundreds of milliseconds. One common way of assessing facilitation is

paired-pulses facilitationwhich occurs when the synapse is stimulated with a pair of con-

secutive pulses, and the response amplitude to the second is larger than the response to the

first. If one separates the two pulses by a large enough time window (several hundreds of

milliseconds) the effect disappears, and the response to both pulses is equal on average. An

AP may produce facilitation at some synapses, regardless of whether vesicle release occurs

or not [del Castillo and Katz, 1954b]. The effect is believed to happen at some point af-

ter the AP arrival and before transmitter release. The biophysical substrate of facilitation

Page 22:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

8 Chapter 1: Introduction: the synaptic function

is thought to be the existence of residual calcium [Katz and Miledi, 1968]: after an action

potential invades the terminal, a calcium influx elevates the concentration[Ca2+]i which in

turnmayproduce vesicle exocytosis. If a second AP reaches the terminal before[Ca2+]i has

returned to normal levels, the new incoming calcium will add to the residual one, resulting

in an enhancement of the probability of release [Zucker, 1996].

Facilitation and depression effects interact: if facilitation produces an increase of release

probability, but the number of releasable vesicles is small, depletion of the vesicle pool is

faster. Thus the train of responses elicited by repetitive stimulation of a facilitating synapse

may show, first, an enhancement of the amplitude of the first EPSCs due to facilitation,

and later, a decrease due to short-term depression, perhaps produced by vesicle depletion

[Dittman et al., 2000, Gupta et al., 2000].

Other forms of synaptic enhancement with larger time scales are the so-calledaugmen-

tation andpost-tetanic potentiation[Zucker and Regehr, 2002]. In these processes each AP

may produce an increase of synaptic strength by only1− 15%. However, since they last for

5− 10 seconds in the case of augmentation, and from30 secs. to several minutes in the case

of post-tetanic augmentation, in episodes of prolonged periods of high pre-synaptic activity,

their effects can be important, and even relevant inworking memorytasks [Hempel et al.,

2000].

1.2.4 Univesicular release hypothesis

So far, we have not discussed the exact relation between the number of vesicles in the

RRP and the probability of release. In particular, assuming that at a given active zone sev-

eral vesicles are ready for release, how many of them can undergo exocytosis upon arrival

of an AP? Given the observation that at individual active zones in synapses onto spinal mo-

toneurons the response amplitude has little variability, Redman and colleagues proposed that

synaptic transmission occurs in an all-or-none manner [Edwards et al., 1976]. This little

variability can be due either to the saturation of the post-synaptic receptors (which would

be already produced by the content of a single vesicle) or to the occurrence ofat mostone

release per active zone (see e.g. [Triller and Korn, 1982] or [Redman, 1990, Walmsley et al.,

1998] for review). This last possibility gained increasing support by further experiments

[Triller and Korn, 1982, Stevens and Wang, 1995] and has become a commonly used hy-

pothesis. A plausible explanation is the following: after one vesicle undergoes exocytosis,

the energy barrier of fusion for the rest of the primed vesicles increases, resulting in an effec-

tive lateral inhibition that prevents them to be released. After a short period of time (∼ 5 ms.)

the fusion energy returns to its resting value. However the calcium concentration[Ca2+]i has

already decreased to its stationary value and no more vesicles undergo exocytosis until a new

Page 23:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

1.2. Synaptic transmission and synaptic dynamics 9

AP arrives [Stevens and Wang, 1995, Dobrunz et al., 1997].

The univesicular release hypothesis does not imply, however, that multi-modal response

histograms cannot occur. This kind of histogram would appear if the synaptic connection

consists of several synapses, or if at each synapse there are several specializations (since the

hypothesis only states that at each active zone at most one release may occur, existing com-

plete independence between active zones). Thus, this hypothesis leads to an identification of

the morphological concept of active zone (or synaptic specialization) and to the functional

definition of release site.

On the other hand there is evidence that multi-vesicular release may occur at some

synapses (usually under high release probability conditions), so that the homogeneity in the

response amplitude should be attributed to receptor saturation [Tong and Jahr, 1994, Auger

et al., 1998, Oertner et al., 2002].

In this thesis, except when explicitely specified, the uni-vesicular release hypothesis will

be assumed. Hence, we will use indistinctly the terms synaptic specialization, active zone

and release site. We will reserve the term synaptic connection to refer to the whole set of

specializations established by two neurons.

1.2.5 Synaptic diversity

A third important aspect is the huge diversity that exists among synapses in the nervous

system (see e.g. [Atwood and Karunanithi, 2002] for a recent review). The first notorious

difference between synaptic connections in the brain is their size and the number of con-

tacts they have [Walmsley et al., 1998]: while some connections in the nervous system have

hundreds of contacts (usually in the peripheral nervous system like in the neuromuscular

junctions, or in the brain-stem like the giant “end-bulbs of Held”, etc.), in the cortex neurons

are commonly connected by a few small boutons (5−10 [Markram et al., 1997a]), sometimes

only one [Sorra and Harris, 1993], containing usually a single specialization. Furthermore,

specializations vary greatly in shape and size. Since the larger the active zone the more dock-

ing sites it has, larger active zones exhibit a larger probability of release, while small ones

are more unreliable [Walmsley, 1991, Ryugo et al., 1996, Schikorski and Stevens, 1997].

This variability in size may partially explain the heterogeneity in release probability among

release sites from the same pre-synaptic fiber [Rosenmund et al., 1993, Hessler et al., 1993,

Murthy et al., 1997]. Nevertheless this cannot be the only factor, since, as it has been shown

in neocortical slices, this probability is subject to pairing plasticity [Markram and Tsodyks,

1996a].

Other forms of synaptic differentiation arise from the diversity in the expression of ac-

tivity dependent mechanisms like short-term depression or facilitation (see above1.2.2and

Page 24:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

10 Chapter 1: Introduction: the synaptic function

1.2.3). The first interesting observation is that different cells innervating a given neuron

evoke different post-synaptic responses. For instance, climbing fibers innervating a given

cerebellar Purkinje cell elicit a depressing response, while parallel fibers innervating the

same cell elicit a facilitating response [Dittman et al., 2000]. This type of diversity is even

more prominent in interneuron synapses [Gupta et al., 2000]. On the other hand, different

branches of the same axon may contact different post-synaptic cells showing a variety of

distinct characteristics at each bouton [Markram et al., 1998a].

In the first part of this thesis we will be concerned with the heterogeneity of the release

probability of a population of synapses innervating a given neuron and the functional impli-

cations of this variability. The functional relevance of the number of contacts constituting a

connection will also be assessed in the second half of this work.

1.3 Objectives and overview of this work

During the last years there has been a strong interest in the functional relevance of the

different synaptic properties, such as unreliability [Zador, 1998, Levy and Baxter, 2002,

Senn et al., 2002a], short-term depression [Tsodyks and Markram, 1997, Abbott et al., 1997,

O’Donovan and Rinzel, 1997, Chance et al., 1998b, Senn et al., 1998, Adorjan and Ober-

mayer, 1999, Matveev and Wang, 2000b, Natschlager et al., 2001, Senn et al., 2002b, Gold-

man et al., 2002], facilitation [Lisman, 1997, Matveev and Wang, 2000a, Fuhrmann et al.,

2002] and heterogeneity [Markram and Tsodyks, 1996b, Markram et al., 1998b]. However,

there are still many open questions regarding how these synaptic features influence the com-

putational capacity of neurons. The objective of this thesis is to investigate what is the effect

of short-term depression and unreliability in, firstly, the transmission of information from the

pre- to the post-synaptic neuron, and secondly the computational capabilities of neurons.

In particular we want to study how this kind of synapses transform auto-correlations

in the pre-synaptic spike trains and whether this transformation can lead an efficient rep-

resentation of the input rate on the synaptic responses. We would like to test whether the

optimal values of the synaptic parameters, such as the release probability or the recovery

time constant, which maximize the information transfer, coincide with those observed in the

experiments. This optimization can be achieved in a number of different ways depending on

the choice of the information measure (e.g. information transmission, parameter estimation,

information per unit time, per response, taking into account metabolic constraints). Thus,

we will check which are the optimal values of these parameters under different optimization

hypothesis and we will check whether the experimental data validates one over the others.

We also want to assess the functional relevance of the synaptic diversity by computing the

Page 25:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

1.3. Objectives and overview of this work 11

information transmitted through an heterogeneous population of synapses, and comparing

its performance with different distributions.

Furthermore, we will compute the response firing rate of aleaky integrate-and-fire neuron

(LIF) when synapses are modeled as depressing stochastic channels with an arbitrary number

of contacts. We will analytically compute the transfer function when the pre-synaptic neu-

rons fire synchronously, and make numerical simulations of a LIF spike generator to validate

the analytical predictions. We will study how the synchronous firing of a pre-synaptic popu-

lation of neurons results in a non-monotonic behavior of the fluctuations of the input current,

which eventually decrease as synaptic saturation takes place. This provides a mechanism to

obtain a non-monotonic transfer function if the target cell works in a regime driven by the

current fluctuations. It will be shown that, in this regime, the response of the cell conveys

information about the input rate beyond the occurrence of saturation of the mean current.

The quantification of the information transmission in a model which considers these

synaptic features will help to elucidate whether or not their existence improves or worsens the

transmission of messages between cells. For example it may explain why synapses are not

failure-free channels. Moreover, a theoretical approach to this issue provides a framework to

investigate what sort of computations neurons can, or cannot, perform when a certain realistic

biological property is considered. For instance, the fact that the release of neurotransmitter

often occurs in an asynchronous manner with respect to the arrival of the AP [Hagler and

Goda, 2001], may constitute a strong limitation to the hypothesis that the precise time of

each AP conveys information. In addition, the asynchronous release would impose severe

limitations to a coordinated code in which, for instance, the synchronous arriving of AP

constitutes a coding strategy.

The thesis is organized as follows: In chapter2 a model of synaptic transmission is pre-

sented. It consists in a stochastic model of vesicle turnover, with which we can analytically

compute the statistical properties of the responses. The output statistics are calculated for

input spike trains with auto-correlations. The model considers only a pre-synaptic terminal

with a single specialization which has several docking sites. Vesicle dynamics are stochastic

and release is unreliable. At the end of the chapter, a model of the heterogeneity found in

hippocampal synapses of CA3-CA1 connections is proposed.

In chapter3, using the response statistics of a single mono-synaptic connection calculated

in chapter2, we compute the information they convey about the input firing rate for a general

family of correlated inputs. In particular, we are interested in finding optimal values for the

synaptic parameters which maximize the information transmission and estimation of input

parameters. At the level of a population of inputs, an optimization is performed over the

distribution of values of the synaptic parameters.

Page 26:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

12 Chapter 1: Introduction: the synaptic function

In chapter4 we compute the synaptic current produced by an ensemble of pre-synaptic

cells onto a target neuron when the connections between them are formed by and array of

synaptic contacts (or specializations) individually modeled as in chapter2. Besides, cross-

correlations among pre-synaptic units are included in a simplead hocmanner.

In chapter5 we compute the output of a leaky integrate-and-fire neuron (both numerically

and analytically) when the total current obtained in the previous chapter isinjectedinto it.

We also study the effect of the spatial correlations in the output introduced either by multi-

contact connections or by synchronous firing.

The conclusions of the work are presented at the end of chapters3 and5, where a brief

discussion is also laid out.

Page 27:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Chapter 2

A model of synaptic depression and

unreliability

2.1 Experimental motivation of the model

As previously introduced in chapter1, synaptic transmission is a complex process which

occurs in an unreliable way (see section1.2.1and references [Hessler et al., 1993, Rosen-

mund et al., 1993, Allen and Stevens, 1994, Stevens and Wang, 1994]). It also involves rapid

(from ∼ 100 ms. to several seconds) use-dependent mechanisms which are generally re-

ferred to asshort-term plasticity[von Gersdorff and Borst, 2002, Zucker and Regehr, 2002].

In this thesis we will focus on short-term synaptic depression [Magelby, 1987] and on the

stochastic nature of synaptic transmission. We propose a model of vesicle depletion, where

release occurs in an unreliable way, mainly based on experiments performed on pyramidal

cell synapses, from mouse hippocampal slices (or culture hippocampal cells) by Stevens, Do-

brunz and colleagues [Stevens and Wang, 1994, Murthy et al., 1997, Dobrunz and Stevens,

1997, Dobrunz et al., 1997, Dobrunz, 2002] and by Hanse & Gustaffson [Hanse and Gustafs-

son, 2001a,b]. These works have studied the connections between CA3 and CA1, which are

believed to be mono-synaptic, with a single active zone in most cases (around70 − 90%)

[Sorra and Harris, 1993, Schikorski and Stevens, 1997]. This allows the experimentalist to

measure the properties of an individual active zone, something usually difficult in neocor-

tical synapses, where the number of synapses between two cells is usually larger than one

[Markram et al., 1997a, Gupta et al., 2000]. Using the method ofminimal stimulation1 they

1The method of minimal stimulation [Raastad et al., 1992, Allen and Stevens, 1994, Dobrunz and Stevens,

1997] consists of applying a stimulus in a given pathway and record intracellularly from a neuron to which

those fibers project. Because the stimulation is extracellular it is not possible to tell how many axons are being

excited. However if the stimulus is gradually decreased until a final “step” is reached before the response in

13

Page 28:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

14 Chapter 2: A model of synaptic depression and unreliability

extracellularly activate a putative synapse and record the synaptic currents produced. Stim-

ulating with a single pulse they can detect, with great accuracy, whether a vesicle release

occurs and, repeating the stimulation over many trials , compute the probability of release.

In addition, they can deplete the readily releasable pool (RRP) using a high frequency train,

while counting the releases which take place until complete depletion is achieved. This

number gives the size of the RRP. The first important conclusion of their first experiments

was that release probabilitypr covaries with the number of docked vesicles [Schikorski and

Stevens, 1997] across a population of synapses. In other words,Schikorski and Stevens

[1997] found that those synapses which optically showed more vesicles adjacent to the cell

membrane, had a higher release probabilitypr. The mean number ofvisualizeddocked vesi-

cles was∼ 10. Dobrunz and Stevens[1997], using physiological methods, found that the

release probability correlated with the number of vesiclesN in the RRP across a population

of synapses. They proposed a functional relation for the release probability of an individual

synapse,pr(N), which approximately fitted the data obtained from a population of synapses

[Dobrunz and Stevens, 1997, Murthy et al., 2001]. A possible explanation of the results is

the following: if all readily releasable vesicles have the same and independent probability of

undergoing exocytosis,pv, but only one of them can release its content (see section1.2.4),

the release probability will be one minus the probability that all of them fail to fuse:

pr(N) = 1− (1− pv)N (2.1.1)

This expression is non-linear in N and tends to one as the number of releasable vesicles is in-

creased. RecentlyDobrunz[2002] confirmed that this relation also holds for a single synapse

(and not only across a population). She also showed thatpv should be the average of the in-

dividual fusion probabilities, in the likely case that they are heterogeneous. In addition, in

facilitating synapses, residual calcium may increasepv, and an activity independent descrip-

tion of the functionpr(N) would no longer be valid. Likewise, if the distribution of fusion

probabilitiespv was bimodal, (like in the calyx synapses [Wu and Borst, 1999, Neher and

Sakaba, 2001, Schneggenburger et al., 2002] where a subset of the RRP has higher fusion

probability than the rest of the pool), a description of the release probabilities with a single

function pr(N) would not apply. This partition of the RRP into two pools also seems in

order [Matveev and Wang, 2000b] if one wants to obtain the largepaired pulseddepression

(defined as the ratio between the magnitude of the averaged response of the second and the

the voltage of the target cell disappears, it is likely that at that level of stimulation only asinglefiber is being

excited, in such a way that the depolarization of the whole patched cell is due to a single synapse. This method

presents several limitations over the dual whole cell recordings in which one can be sure that a single connection

(perhaps with several synapses) is mediating the excitation of the target cell [Thomson and Deuchars, 1997,

Markram et al., 1997a, Gupta et al., 2000].

Page 29:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.2. Methods 15

first pulses) observed in neocortical slices [Markram and Tsodyks, 1996a, Varela et al., 1997,

Thomson, 1997] and in hippocampal slices [Hanse and Gustafsson, 2001a] (see [Trommer-

shauser, 2000, Matveev and Wang, 2000b] for examples of this specialized pool model and

[Schneggenburger et al., 2002] for a recent review on calyx synapses).

In our work we will not consider these extensions of the single RRP model [Dobrunz

and Stevens, 1997]. Thus, our model will be an adaptation of the one proposed by Dobrunz

and Stevens with a single RRP, which we will also call primed pool or docked pool. We will

not consider facilitation or any other activity-dependent changes, because our purpose is to

analyze the effects that depression alone has on the transmission of information.

2.2 Methods

2.2.1 Model of one synaptic contact with many docking sites

The model of a single active zone, or synaptic contact, is completely described by: i) the

release probability functionpr(N), which gives the probability that one vesicle releases its

content, when there areN of them in the RRP. ii) The functionPrec(n, t|N), which defines

the probability that anyn out ofN empty docking sites are refilled by new vesicles during a

time window of lengtht.

2.2.1.1 Vesicle Release

The experimental data about the probability of releasepr for different synapses was ap-

proximately fitted with the functionpr(N) defined by equation2.1.1(see fig. 5 in [Dobrunz

and Stevens, 1997]. This function ofN shows saturation and a non-linear behavior. In

order to make further calculations analytically tractable, we have adapted this exponential

dependence, capturing both aspects, non-linearity and saturation, into a simpler new func-

tion which reads,

pr(N) = UΘ(N −Nth) N = 0 . . . N0 (2.2.1)

whereΘ(N) is the Heaviside function. Thus, if the RRP sizeN is smaller than a certain

threshold valueNth, the probability of release is zero, and when the number of vesicles in

the pool exceedsNth, the probability becomesU . The last parameterN0 sets the maximum

number of vesicles that the RRP can hold, which in our description represents the number of

morphologically separate docking sites that together constitute the synaptic contact2. Thus,

2The question of whether the number of morphological vesicle docking sites constitutes the maximum size

of the RRP or on the contrary the latter pool is a functional subset of the first (see [Schikorski and Stevens,

Page 30:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

16 Chapter 2: A model of synaptic depression and unreliability

p_r

N

Figure 2.1: Release probability as a function of the number of vesicles in the RRP: points

are experimental data of the release probability to the first pulse of a burst, from different

CA-CA1 hippocampal synapses (taken from [Dobrunz and Stevens, 1997]). Dashed line is

the fit with the exponential model by Dobrunz & Stevens (1997) (see eq.2.1.1). Solid black

line is an extended version of that model (see [Dobrunz and Stevens, 1997]). Solid red line

represents the step-like model used in this thesis, with the parametersNth = 5, N0 = 13 and

U = 0.7.

depletion of vesicles below the threshold levelNth will produce depression in the synapse,

because there will be a time interval, before more vesicles join the RRP, during which the

release probability is zero. Figure2.1 shows the experimental data taken from fig. 5 of

[Dobrunz and Stevens, 1997] together with our step-like release probability function. Al-

though this function is an oversimplified model, the figure illustrates that it accounts for the

qualitative behavior of the data.

2.2.1.2 Vesicles Recovery

Besides the dependence of the release probability on the number of ready-releasable

vesicles, one must define the stochastic dynamics of the vesicles recovery. Much less is

known about the recovery of vesicles than about their release. The only experimental data

2001]) is beyond the aim of this model. However, even though in some synapses docked vesicles must undergo

a priming step before they become ready for fusion [Sudholf, 2000], in hippocampal synapses it has been

shown that the number of visually docked vesicles is roughly the same as the number of vesicles in the RRP

[Schikorski and Stevens, 2001, Murthy et al., 2001].

Page 31:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.2. Methods 17

r p (N) Release:

N 0N − N

Unavaliable Pool

Avaliable Pool

Recovery: 1/τ

Figure 2.2: Schematic picture of the synaptic dynamics seen as a system composed of two

pools of vesicles: the avaliable pool represents the (RRP). The unavailable pool represents

the set of empty docking sites. An action potential evokes the release of a vesicle with

probabilitypr(N), shown in the picture as a transition of a vesicle from the avaliable to the

unavailable pool. The recovery of the vesicles takes place with a rateN0−Nτv

, i.e. each empty

docking site recovers with a constant rate1τv

.

is that the release probability recovers exponentially [Dobrunz and Stevens, 1997, Markram

and Tsodyks, 1996a]. Given this constraint, we model the recovery from the point of view

of the structural docking sites instead of from the vesicles perspective. Hence, we assume

that each of theN0 − N empty docking sites may be filled at any time with a constant

and homogeneous probability1/τv, which means that the amount of time that any docking

site stays empty is an independent random variable, exponentially distributed, with mean

τv. This type of dynamics implies two things: i) since thesingle siterecovery rate is not

affected by the number of empty sites, the number of vesicles in the recovery pool (vesicles

filled with transmitter but not docked) should be much larger3 thanN0 ; ii) the recovery rate

1/τv is independent of pre-synaptic activity ; iii) docking sites are equally distributed across

the synaptic bouton, i.e. there is no preference to refill certain empty sites over the others.

Because we have assumed that the refilling of each empty site is an independent process, we

can now easily write the probability thatn, out of theN0 −N empty docking sites, recover

within a time window∆ as:

Prec(n, ∆|N0 −N) =

N0 −N

n

(1− e−∆/τv)n (e−∆/τv)(N0−N−n) (2.2.2)

3This is a good approximation because the number of docking sites is around5 while the total number of

vesicles in the synaptic bouton is about200 in hippocampal synapses [Schikorski and Stevens, 1997]

Page 32:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

18 Chapter 2: A model of synaptic depression and unreliability

which is just a binomial distribution of parametersN0 − N and(1 − e−∆/τv). This second

parameter is the probability that a vesicle fills an empty site in the time interval∆.

Here it must be noticed that, when in this model we setN0 = 1, that is, the maximum

number of primed vesicles is one, we recover the model introduced bySenn et al.[2001] and

later used in several papers [Goldman et al., 2002, Fuhrmann et al., 2002].

2.2.2 Input spike-train statistics

Our first assumption will be that the generation of action potentials (AP) by cortical

neurons can be modeled as random events, that is, a spike train can be treated as a stochastic

point process. This hypothesis is based on the experimental observation that cortical neurons

in vivo fire irregularly [Softky and Koch, 1993], seemingly randomly. The consideration of

the AP as a point event is a good approximation if one takes into account the stereotyped

shape they present and their very short duration (∼ 1 ms) in comparison with the rest of the

temporal scales we will consider in our model.

The statistics of these spike trains in the brain are often well approximated by renewal

processes [van Vreeswijk, 2000], which means that consecutive inter-spike-intervals (ISI’s)

can be considered as samples of a random variable, which areindependentlydrawn from a

given probability distribution function (p.d.f.)ρisi(t) [Cox, 1962]. A Poisson process is the

most typical example of renewal process, whose ISI’s follow and exponential distribution.

Nevertheless, the fact that the correlation between consecutive ISI’s is zero, does not imply

that the correlation between spike times are zero. The occurrence of a spike may increase

(or decrease) the probability of observing another one immediately afterwards, with respect

to the case in which nothing is known about the previous activity. The Poisson process has

no memory in this sense, that is, previous spikes do not convey any information about what

is coming next.

Positive autocorrelations among spike times have often been observed inin vivo ex-

periments [Mandl, 1993, Zohary et al., 1994, Bair et al., 1994, Lisman, 1997, Fenton and

Muller, 1998, Matveev and Wang, 2000a, Goldman et al., 2002]. A common case of positive

auto-correlations arebursts, which are groups of spikes which occur in close succession. In

hippocampus, for instance, pyramidal cells fire what are calledcomplexspikes [Kandel and

Spencer, 1991, Ranck, 1973], that are brief bursts of two to nine spikes with ISI’s ranging

from 2 to 10 ms. Positive autocorrelations can express aredundantcoding of sensory (or

other kind) of information [Dan et al., 1996, Baddeley et al., 1997]. Since the information

coming from the sensory world is very redundant in time [Dan et al., 1996], neuronal activity

in sensory areas may show autocorrelations arising from the encoding of such redundancy.

Besides, sensory acquisition mechanisms such as saccadic eye movements, whisking, etc.

Page 33:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.2. Methods 19

may also produce autocorrelations [Goldman et al., 2002]. Thus, positive autocorrelations

have been observed in the visual areas of awake monkeys such as V1 [Matveev and Wang,

2000a, Goldman et al., 2002] and MT [Bair et al., 1994]. Autocorrelations have also been

measured in other higher processing areas such as prefrontal cortex [Matveev and Wang,

2000a].

We will consider input trains governed by stationary renewal processes, that is, theρisi(t)

will not change in time, with non-zero autocorrelation between the spike times. For the sake

of simplicity, we assume that this auto-correlation is exponential, which will turn out to have

an accessible analytical treatment. As it is shown in AppendixA, any renewal process with

exponential correlations has an ISI distribution which can be written,

ρisi(t) = (1− ε)β1e−β1t + εβ2e

−β2t , t > 0 (2.2.3)

This function is normalized by construction ifβ1, β2 > 0. In order to be a p.d.f. it must also

be non-negative, condition which imposes two more constraints in the values that the term

[β1, β2, ε] can take (see AppendixA ). One can also parameterize this input in terms of three

physical quantities: the spike rateν, defined as the inverse of the mean ISI; the coefficient

of variation of the ISI’sCVisi, defined as the ratio of the standard deviation to the mean

of the ISI; and the time scale of the correlationsτc, which is the characteristic time of the

exponential decay. The dependence of[ν, CVisi, τc] on [β1, β2, ε] follows a complex mapping

which is notone-to-onedue to the symmetryβ1 ↔ β2, β2 ↔ β1, ε ↔ 1 − ε (see Appendix

A).

Let us now define the two-point correlation functionC(t, t′) defined as the probability

density of finding one event (e.g. a spike or a release) at timet and another one at time

t′. For a stationary renewal process this function only depends on the time difference, i.e.

C(t, t′) = C(t′−t). It can be computed, for a renewal process, using the following expression:

C(t− t′) = C(t) = (2.2.4)

= r(δ(t) + ρ(t) +

∫ t

0dxρ(x)ρ(t− x) +

∫ t

0dx∫ t

0dyρ(x)ρ(y)ρ(t− x− y) + . . .

)Here,ρ(t) denotes the p.d.f. of either the ISI’s (ρisi) or the time interval between synaptic

responses/releases, which we will refer to as inter-response-interval or IRI, (ρiri). For this

expression to be well defined, we need to defineρ(t) for t < 0 as zero. The magnituder

is the constant event rate (ν for spikes andνr for releases). Each of the terms of the sum

represents the probability density of finding two events, at time zero and at timet, when

there has been zero, one, two,... event-intervals between them, respectively. We now connect

this correlation and normalize it byr and, assumingt > 0, we discard the delta function.

Page 34:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

20 Chapter 2: A model of synaptic depression and unreliability

The result reads

Cc(t) =C(t)− r2

r(2.2.5)

This new functionCc(t) is often called the connectedconditionalrate: added to the rate

r, it gives the frequency of events at timet given there was one at time zero. However,

by abuse of terminology, we shall also refer toCc(t) as the correlation function, because it

represents how much more probable it is to find an event at timet, given that there was one

at time zero, in comparison to the uncorrelated case.

We now come back to the input model introduced above (eq.2.2.3) and compute, by

means of the Laplace transform, the correlationCc(t) (AppendixA). If we write it as a

function of the physical parameters[ν, CVisi, τc] it takes the simple form

Cc(t) =CV 2

isi − 1

2τc

e−t/τc (2.2.6)

It is now straightforward to describe this correlation in terms of the physical parameters: in

the first place, the correlation does not depend on the spike rateν (because it is connected).

The value ofCVisi determines the total area under the correlation function: ifCVisi > 1

the function is positive for allt, and negative forCVisi < 1. At CVisi = 1 the correlation

becomes zero for allt and one recovers exactly a Poisson process4. Finally, τc determines

the temporal extent of the correlation. We stress again the fact that in the three-dimensional

[ν, CVisi, τc] space, the admissible region is not the whole subspace determined byν > 0,

CVisi > 0, τc > 0. If CVisi ≥ 1 (positive correlation) then any value ofν > 0 andτc > 0 is

allowed. However, in the interval0 < CVisi < 1 the conditionτc >(1−CV 2

isi)

νmust hold to

ensure thatρisi is a p.d.f. (see AppendixA).

As previously mentioned, we are mostly interested in positive input auto-correlations,

that isCVisi > 1. In this regime0 < ε < 1 and it is possible to interpret the input process

ρisi(t) as a combination of two Poisson processes with ratesβ1 andβ2. At each ISI realization

one of these two underlying Poisson processes is chosen with probabilities(1 − ε) and ε

respectively. Fig.2.3 shows some input examples for different values of[ν, CVisi, τc]. It

can be observed that, for small values ofτc (∼ 10ms) and large values ofCVisi (∼ 2), the

train shows a bursty structure. In fact, a third description of the input as aburst generator

is possible5: if β2 ∼ 100 − 500 Hz andβ1 β2, then following the description of the two

Poisson processes, it can be shown that the mean number of spikes within a burst,Nb reads:

Nb =ε

(1− ε)2+ 1 ' CV 4

isi − 1

4+ 1 (2.2.7)

4At this limit, one of the exponentials of the ISI distributionρisi(t) vanishes becauseε becomes either one

or zero, resulting in a Poisson process with rateβ2 or β1, respectively.5Although a renewal process may seem incompatible with a process of bursts, if the notion of burst is

generalized by assuming a large variability in the number of AP’s within it, then the description is conceivable.

Page 35:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.2. Methods 21

0 25 50 75 100(ms)

0

0.05

0.1

0.15

ρ isi(t

| ν)

−100 −50 0 50 100 (ms)

0

0.02

0.04

0.06

0.08

0.1

0.12

Cc (t

)

||| | | | |||| || | ||| |||| |||||| | | || |

|| | | | | | || || ||| ||| | | | | ||

| ||| ||||| |||| || | ||0 250 500 750 1000

(ms)

CV= 1

CV= 1.3

CV= 1.8

Figure 2.3: Examples of input spike trains with exponential autocorrelations.Top left: Dis-

tribution of ISI’s given the rateν, ρisi(t|ν). Top right: Connected conditional rate (also

called correlation function, given by eq.2.2.6). Bottom plots Examples of spike train re-

alizations: vertical bars represent the position of the action potentials on a temporal axis.

Three values ofCVisi have been chosen for all plots:CVisi = 1 (green lines),CVisi = 1.3

(red lines) andCVisi = 1.8 (blue lines). It can be observed that the number of spikes within

a burst increases asCVisi increases (andν andτc are kept constant). In all cases:ν = 10 Hz,

τc = 10 ms.

Page 36:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

22 Chapter 2: A model of synaptic depression and unreliability

and the mean duration of the bursts,Tb, is just the mean interval between spikes of a burst,1β2

, times the mean number of intervals within a burstNb − 1:

Tb =ε

β2(1− ε)2(2.2.8)

The third parameter is the mean interval between bursts (which can have onlyoneAP) which

is simply 1β1

. In this way, the burst generatorρisi(t) would be expressed as a function of the

term [β1, Tb, Nb].

Our objective in the following sections will be to determine the information about the

input rateν conveyed by the synaptic responses. Thus, the firing rateν is the coding variable

of the input signal, so that it takes values from an ensemble with a certain p.d.f. that will

be denoted asf(ν). To be rigorous we must therefore change the notation ofρisi(t) we just

introduced. Hence for a fixed value ofν we will write ρisi(t|ν) while if want to refer to

the distribution of the ISI considering the whole input ensemble we will denote it asρisi(t)

which is defined as:

ρisi(t) ≡∫

f(ν) ρisi(t|ν)dν (2.2.9)

2.2.3 Statistics of the synaptic response

It is standard to consider AP’s as point events because of their very short duration and

because, at a single neuron, they are almost replicas of each other. On the contrary post

synaptic potentials (PSP), at a single neuronal membrane, differ in sign, size and duration,

depending on the amount of transmitter that released, the number and type of post-synaptic

receptors activated and consequently the type and number of ion channels that produced

the PSP. However, if one focuses on one pair of neurons, it is sometimes the case that they

are connected by only one synaptic contact6. If we then neglect the fact that the amount

of transmitter vesicles release fluctuates, and assuming theuni-vesicular release hypothesis

(section1.2.4), then a certain train of spikes propagating along the pre-synaptic fiber, will

elicit a number of synaptic releases. These will produce in turn a train of PSP’s of equal

amplitude, which, regardless of the temporal extent, will be a decimated version of the input

train. In this sense, the synapse is making a stochastic transformation of the spike trains

arriving at the synaptic bouton into trains of, what we generally refer to as, synaptic re-

sponses. This description fails if one takes into account the fact that two neurons may be

6This is the case for instance in the previously discussed CA3-CA1 synapses in the hippocampus, where

it is been shown [Dobrunz, 2002, Hanse and Gustafsson, 2001a] that when the Shaffer collateral pathway is

stimulated with bursty patterns, the amplitudes of the elicited EPSCs measured in the CA1 pyramidal cell were,

excluding failures, equal in amplitude, reflecting the fact that one fiber makes a single synapse with only one

active zone

Page 37:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.2. Methods 23

connected by several contacts, which in the cortex are usually between 1-15 [Markram et al.,

1997b, Walmsley et al., 1998]. If this is the case, the PSP’s produced by the connection

of a single pair of neurons, do not have the same magnitude because they can be produced

by the release of several vesicles. Notice that this does not imply that we are rejecting the

uni-vesicular release hypothesis.

In other more formal words, we are modeling the synapse as a stochastic mapping

tini=0 −→ tjm

j=0 (2.2.10)

wheretini=0 are the times of the input spikes andtjm

j=0 are the times of the synaptic

responses, a subset of the previous spike-times set.

Since our final technical purpose is to analytically compute the information (both Fisher

and Shannon informations) transmitted by this synaptic channel, once the p.d.f of the input

ρisi(t) is defined, we need to calculate the p.d.f. of the output. We make use here of the fact

that the input is chosen to be a renewal process, so that the distribution of the traintini=0

is completely defined by the distribution of the ISI’sρisi(t). Thus, for a generalρisi(t) it

is feasible to compute the distribution of the intervals between synaptic responses/releases

ρiri(t). Nonetheless, this last quantity is not enough to define the distribution of the responses

because the output is not anymore a renewal process, except in the instance where the number

of docking sites,N0, is one. The explanation of this is the following:

Renewal character Let us suppose that at timet0 a release occurs. At timet+0 = t0 + δt

there are stilln vesicles in the RRP. The distribution of the time of the next release

ρiri(t − t0) depends on several random sources: firstly on the times at which new

spikes will reach the terminal. Since the input is renewal this dependency will be the

same every time there is a release. Second, on the recovery of new vesicles to the RRP

and in turn its size at the time of a spike arrival. Although the recruitment of a vesicle

for an empty site is again a renewal process (which even keeps no memory of the

time the site became empty) the distribution of thetotal number of vesiclesN in the

RRP at timet does keep memory of how many were there at times beforet0. In other

words, the distributionρiri(t − t0) would be different if we know there aren vesicles

in the RRP at timet+0 than if there aren + 1. If, on the contrary, there is only one

docking site (N0 = 1), the RRP size immediately after a release is zero. This happens

because the only vesicle site, that was occupied, just got depleted. Therefore, there is

no dependence on the number of vesicles in the RRP att+0 because it is always zero.

In conclusion, the resulting stochastic process that underlies the response generation is

renewal if and only ifN0 = 1. Only in this particular case, we will be able to compute

the correlation functionCr(t) of the responses using eq.2.2.5.

Page 38:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

24 Chapter 2: A model of synaptic depression and unreliability

We define the probability function of the sizeN of the RRP at timet as℘N(t). Since our

input is stationary, unless we make use of some recent information about the state of the

RRP, the function℘N(t) will not depend ont, hence we will write℘ssN , wheress stands for

the stationary state. This distribution can be computed for a general renewal inputρisi(t|ν)

and a general release probabilitypr(N) by following the next three steps.

• First we establish a system ofN0 + 1 equations for the℘N(t) (N = 0 . . . N0) at the

times of two consecutive spikes(tj−1, tj) that define an ISI of duration∆ = tj − tj−1:

℘N(tj) = ℘N(t+j−1) Prec(0, ∆|N0 −N) +

+ ℘N−1(t+j−1) Prec(1, ∆|N0 −N + 1) +

+ . . . + ℘0(t+j−1) Prec(N, ∆|N0) (2.2.11)

where℘N(t+j−1) = ℘N(tj−1 + δt) is the probability an instant after the (j-1)-th spike

arrival and relates to the probability just before the arrival through

℘N(t+j ) = ℘N(tj) (1− pr(N)) + ℘N+1(tj)pr(N + 1) (2.2.12)

The system of equations2.2.11comes from the dynamics that the probabilities℘N(t)

obey over the interval∆ between two spikes: at timetj there might beN vesicles in

the RRP if, 1) there wereN at timet+j−1 and during∆ no new sites were recovered, or

2) there were(N − 1) at t+j−1 and one site recovered, or 3)..., or N+1) there were no

vesicles in the RRP butN of them recovered in the interval∆.

• We average the system2.2.11over the stimulus ensemble∆ using the p.d.f.ρisi(∆|ν).

℘N(tj) = ℘N(t+j−1) 〈Prec(0, ∆|N0 −N)〉 +

+ ℘N−1(t+j−1) 〈Prec(1, ∆|N0 −N + 1)〉 +

+ . . . + ℘0(t+j−1) 〈Prec(N, ∆|N0)〉 (2.2.13)

where the brackets〈·〉 represent

〈Prec(N, ∆|N0 −N)〉 =∫ ∞

0d∆ρisi(∆|ν) Prec(N, ∆|N0 −N) (2.2.14)

• We then substitute eq.2.2.12into the system2.2.13and finally impose the stationary

condition which sets that in the steady state:℘N(tj−1) = ℘N(tj) = ℘ssN . Thus, we

reach a system ofN0 +1 linear equations for theN0 +1 probabilities℘ssN . This system

is degenerate (there are onlyN0 independent equations), but we can complete it by

adding the normalization condition∑

N ℘ssN = 1. Solving this system one obtains the

exact distribution of the size of the RRP,N , for an inputρisi(t|ν) and a release function

pr(N) in the stationary regime.

Page 39:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.2. Methods 25

We focus now on the main calculation of this chapter which describes how to obtain the

expression of the p.d.f. of the IRI’s,ρiri(t|ν), given a general renewal inputρisi(t|ν). We

start by making the following expansion

ρiri(∆|ν) = ρ(1)iri (∆|ν) + ρ

(2)iri (∆|ν) + ρ

(3)iri (∆|ν) + . . . (2.2.15)

whereρ(i)iri(t|ν) is the p.d.f. of having an IRI of length∆ composed ofi consecutive ISI’s, or

put differently, the probability that, given that there was a release att0, thei-th event arriving

at t0 + ∆ is the first one to trigger a new release. We give here the expression of the first of

these terms and leave the details of the calculation for the AppendixB.

ρ(1)iri (∆|ν) =

N0∑N=0

℘ssN︸ ︷︷ ︸

(i)

N0−N+1∑n1=0

Prec(n1, ∆|N0 −N + 1)︸ ︷︷ ︸(ii)

ρisi(∆|ν)︸ ︷︷ ︸(iii)

pr(N − 1 + n1)︸ ︷︷ ︸(iv)

(2.2.16)

The probability that a synaptic release occurs is a combination of all the stochastic dis-

tributions of the problem: the sizeN of the RRP (℘ssN ), the numbern1 of recovered vesicles

(Prec), the arrival of an input spike after a time-window∆ (ρisi(∆|ν)) and finally the release

probability upon arrival of the spike (pr). The factor(i) represents the weighted sum of all

possibles sizes of the RRP at an arbitraryt0 given that a spike succeeded to release a vesicle

at that time. The tilde over℘ssN accounts for this conditioning which could be expressed using

the Bayes rule:

℘ssN =

℘ssNpr(N)∑N0

k=0 ℘ssk pr(k)

(2.2.17)

The term(ii) corresponds to the probability that between the releases att0 and att0 + ∆, n1

new vesicles recover (wheren1 = 0 . . . N0 − N + 1, because att+0 there areN0 − (N − 1)

empty sites). The expression(iii) represents the probability that the next spike, after the one

at t0, arrives att0 + ∆ and the last term(iv) is the probability that it succeeds.

Subsequent termsρ(i)iri(t|ν) (i = 2, 3, . . .) of the expansion2.2.15are constructed like

the one described but including the convolution product ofi distributionsρisi in each of

the terms. This in turn means nesting more summations with the respective probability

functions in a systematic manner (see AppendixB). To perform the explicit calculation of

all the summations in each termρ(i)iri(t|ν) an expression for the functionpr(N) is needed.

Most choices here make the calculation non-feasible. It is viable, however, for the step-

like function of our model (eq.2.2.1). Still the convolution ofi distributionsρisi remains

unsolved in each of the terms.

Page 40:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

26 Chapter 2: A model of synaptic depression and unreliability

We shall now introduce an alternativedivide-and-conquerinspired approach to look at

our synaptic model, that will be very convenient for the calculation. Since the two stochastic

sources of the synapse namely, the recruitment of vesicles for the RRP and the unreliable

release, are independent, we will treat them as separated phases of a two-stage channel. This

can be done as a consequence of the following renormalization property which any synaptic

model where recovery and release are model by independent distributionsPrec(n|N) and

pr(N), posseses,

ρisi(∆|ν)Prec, pr−→ ρiri(∆|ν) ≡ ρisi(∆|ν)

U−→ ρdisi(∆|νd)

Prec, p∗r−→ ρiri(∆|ν)

The renormalized approach (r.h.s. of previous diagram) reads as follows: spikes come

through the first stage of the channel which is just a non-activity-dependent “filter” that

decimates the train with a constant probability(1 − U) or in other words, every spike has

a probabilityU of crossing. After this stage we have a diluted version of the original train

which we have calledρdisi(∆|νd) pointing that at least7 the rateν has a different valueνd,

which will be termeddiluted rate. The second stage of the synaptic channel is an activity-

dependent filter where depression takes place. It is equivalent to the original one-stage model

except for the fact that the release probability has been renormalized topr(N)∗ = pr(N)U

ensuring thatpr(N)∗ < 1 for all N . To understand why this two-stage model may be con-

venient let us apply the same transformation to the step-like model. In this simple case,

pr(N)∗ = Θ(N − Nth). Because of this, the stochasticity in the second stage is only due

to the recovery process, so that the unreliability in the release process has been transfered

to the first, non-activity-dependent stage. It is interesting to observe that in the limit where

the recovery of vesicles occurs very fast (τv → 0), the second stage becomes trivial, because

the RRP size is alwaysN0 and transmission at this phase is always successful. In this limit,

the global transformation of the spike train into a response train is reduced to that of the first

stage, i.e.ρiri(∆|ν) = ρdisi(∆|νd).

Coming back to the computation ofρiri(∆|ν) (eqs.2.2.15-2.2.16), we can apply it to the

computation of the second-stage transformation by just substitutingρisi → ρdisi andpr(N) →

p∗r(N) = Θ(N −Nth).

The last step then, is to sum the expansion2.2.15. To accomplish this, we apply to the se-

ries the Laplace transform, which has the well-known property to convert each convolution

into a simple product in Laplace space. With this trick the expansion becomes a geomet-

rical series and can be explicitly operated. Notice that so far in this calculation, the input

distributionρisi(∆|ν) was not specified. Therefore, we obtain an expression of the Laplace

7We will see later that for the exponentially correlated input we are using, also theCVisi is transformed

asCV disi =

√(CV 2

isi − 1) + 1. The third parameterτc is invariant to this transformation and the correlation

Cc(t) remains as in2.2.6but withCV disi instead ofCVisi.

Page 41:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.2. Methods 27

transform ofρiri(t|ν), for a general renewal input, which reads

ρiri(s) = ρdisi(s) − ℘ss

Nth

ρdisi(s + N+

τv)[1− ρd

isi(s)]

1− ρdisi(s + N+

τv)

(2.2.18)

which depends on the Laplace transformρdisi(s) of the diluted version of the input distribution

ρisi(∆|ν). The variableN+ = N0 − Nth + 1 is thex-size of the upper portion of the step

functionp∗r(N). We also introduced the conditional probability℘ssNth

, which was defined for

a generalN in eq.2.2.17, but that we rewrite here for this particular case

℘ssNth

=℘ss(Nth)

℘ss(Nth) + ℘ss(Nth + 1) + . . . + ℘ss(N0)(2.2.19)

that is, the probability that the RRP size isNth given that the size is equal or larger thanNth.

In the particular case where there is only one vesicle docking site (N0 = 1), both variables

℘ssNth

andN+ equal one.

Still the transformation of the first stage, i.e.ρisi(∆|ν) −→ ρdisi(∆|νd), must be estab-

lished. Expandingρdisi in a series like2.2.15, each of the terms does not depend on any

vesicle dynamics anymore. They only depend on the constant probabilitiesU and(1 − U)

of being or not transmitted, respectively, in the following way

ρdisi(∆|νd) = ρisi(∆|ν)U +

∫ ∆

0dxρisi(x|ν)ρisi(∆− x|ν)(1− U)U + (2.2.20)

+∫ ∆

0dx∫ ∆

0dyρisi(x|ν)ρisi(y|ν)ρisi(∆− x− y|ν)(1− U)2U + . . .

which Laplace reads

ρdisi(s) =

U

(1− U)

∞∑k=1

[ρisi(s)(1− U)]k =

=Uρisi(s)

1− (1− U)ρisi(s)(2.2.21)

Taking eq.2.2.21into eq.2.2.18one obtains an expression forρiri(s) in terms of the input

distributionρisi(s) and of the synaptic parameters [τv, U , N+]. Thus, no dependence onNth

nor onN0 appears. This occurs because, in the stationary state, the size of the RRP,N , never

goes belowNth−1 since no vesicle can be released ifN < Nth. For this reason, the absolute

size of the RRP is functionally irrelevant and all it matters is the relative distance from the

maximum size,N0, to the edge of the step,Nth. Since we have the freedom to giveNth any

value, we set it equal to one, makingN+ = N0. In conclusion, the only relevant synaptic

parameters are[τv, U, N0].

To obtain finally an explicit analytical expression forρiri(∆|ν), one must first define the

distributionρisi(∆|ν), and then apply the Laplace anti-transform to eq.2.2.18. In the Results

section we will show the particular easy case of Poisson input, while the derivation of the

formula for the correlated case is given in AppendixD.

Page 42:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

28 Chapter 2: A model of synaptic depression and unreliability

2.2.4 Population of synapses

As a step further in the analysis of the transmission properties of dynamical synapses,

we study a population ofM pre-synaptic neurons each of them making a single contact, like

in the previous model, onto the same post-synaptic cell (see fig.2.4). We will assume that

the M units fire with the same statistics, given byρisi(t|ν) (which in turn depends on the

input parametersCVisi andτc) and the p.d.f. of the ratef(ν). We will impose that given the

rateν, the spike trains arriving through each of the axons are independent, that is, there are

no spatial cross-correlations (apart from those arising because all the cells fire with the same

ν). We model the synaptic response to this multi-channel stimulus as a vector of IRI’s∆iwhich represents the set of intervals between responses produced in each of theM synapses.

. . .

∆ 1

3

2

Μ

In

pu

t:

Po

pu

lati

on

of

neu

ron

sC

Vc

[ν, τ

,

]

3

2

1

M

Figure 2.4: A population ofM neurons making single contacts onto a target cell. All cells in

the population share the same firing statistics given byCVisi, τc andf(ν). The parameters of

the synaptic contacts are, however, distributed according with the distributionD(U,N0, τv).

The input code is the input rateν. The output code is theM -dimensional vector of IRI’s

∆i.

Motivated by puzzling findings showing that, among a population of synapses of the

same class, the synaptic variables (RRP maximum size, probability of release in a stationary

state, recovery time,..) take a wide distribution of values [Allen and Stevens, 1994, Dobrunz

and Stevens, 1997, Murthy et al., 1997, 2001], we try to asses the functional relevance of

such a heterogeneity in terms of information transmission. We characterize the population

of synapses by a joint probability distributionD(U,N0, τv) which is partially determined by

experimental data. In particular we refer to results obtained from the CA3-CA1 hippocampal

synapses [Allen and Stevens, 1994, Murthy et al., 1997, Dobrunz and Stevens, 1997, Murthy

Page 43:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.2. Methods 29

et al., 2001]. Firstly, Murthy et al.[1997] found that the distribution of the probability of

release when the synapse is at rest,p1 (not stimulated for a long period of time), can be well

fitted by a gamma function of order2, namely

Γλ(p1) = λ2p1e−λp1 (2.2.22)

with λ = 7.9.

In our model We identify this probabilityp1 with the probability of release when the

RRP size is maximal, i.e.p1 = pr(N = N0) = U . As already mentioned at the beginning

of this chapter, several works [Dobrunz and Stevens, 1997, Murthy et al., 2001, Hanse and

Gustafsson, 2001a, Dobrunz, 2002] found that in a population of synapses,p1 correlates with

the RRP maximum sizeN0. In other words, the marginal joint probability ofU and the RRP

maximum sizeN0 do not factorize:D(U,N0) = f(N0)R(U |N0) whereR(U |N0) 6= R(U).

In particular,Murthy et al. [2001] found that when averaging the value ofU across the

population for a fixed value ofN0, the data can be fitted with the single contact model (see

section2.2.1), that is:

〈U〉N0 = 1− (1− pv)N0 (2.2.23)

with pv ' 0.05. Using a different frequency of stimulation,Hanse and Gustafsson[2001a]

find that the maximal size of the RRPN0 takes much lower values. Fitting their data with

the same release model (eq.2.2.23) yields values ofpv ' 0.3− 0.6.

Therefore, we use equations2.2.22and2.2.23as requirements which theD(U,N0, τv)

must fulfill. Thus, we write the marginal joint probability ofU and N0 as the product

D(U,N0) = f(N0)R(U |N0) and we model the conditional probabilitiesR(U |N0) with the

following ansatz8

Rq(U |N0) =qN0+1

N0!UN0e−qU (2.2.24)

which means that for eachN0, the distributionR(U |N0) is a gamma function of orderN0 +1

and parameterq. This new parameter has to be determined with the experimental constraints

(eqs.2.2.22-2.2.23).

Imposing first that the marginal ofU has to equal the function given in eq.2.2.22,

Γλ(U) =∞∑N0

f(N0)Rq(U |N0)

From this expression we obtain the coefficientsf(N0) by expanding both sides in powers of

U and identifying terms of the same order. In this way, we find that thef(N0) are polynomi-

als in λq

of gradeN0 + 1 (see AppendixC).

8The experimental works described [Dobrunz and Stevens, 1997, Murthy et al., 1997, 2001] did not collect

enough data to fit the distribution ofU for a fixedN0.

Page 44:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

30 Chapter 2: A model of synaptic depression and unreliability

To fix the parameterq, we make use of the second constraint of eq.2.2.23, making an

expansion up to first order in the parameterpv, that is

〈U〉N0 = 1− (1− pv)N0 ' pv N0 (2.2.25)

Performing the average〈U〉N0, our model gives:

〈U〉R(U |N0) ≡∫

U Rq(U |N0) dU =N0 + 1

q(2.2.26)

Thus, identifying the slopes of these two equations9, we find that1q' pv.

Summarizing, we first propose that the family of distributionsR(U |N0) (N0 = 1, 2 . . .)

is a gamma basis of parameterq; then we determine this parameter as a simple function

of the experimentally obtainedpv; finally we determine the probability functionf(N0) for

N0 = 1, 2 . . . Nmax0 (where the cutoffNmax

0 has been chosen so that the cumulative prob-

ability∑Nmax

0N0=1 f(N0) is close enough to one) as polynomials inλ

q, whereλ is the second

experimentally measured parameter. In fig.2.5 (bottom) we show the plane(N0, U) with

55 realizations from the final distributionD(U,N0) when the values ofpv andλ are taken

from [Murthy et al., 2001] and [Murthy et al., 1997], respectively. The theoretical curve

1 − (1 − pv)N0 along with a straight line representing the model average〈U〉R(U |N0) are su-

perimposed for comparison. The figure resembles the experimental one in ([Murthy et al.,

2001], fig.1-B). The top figure is the histogram of the release probability of 600 realizations

together with the experimental fit, equation2.2.22(see [Murthy et al., 1997] fig. 2-B).

Since there is no experimental data about howτv is distributed in the population and

whether it correlates or not withN0 andp1, we will apply an optimization criterion presented

in next section to determine the complete joint distributionD(U,N0, τv).

2.3 Results: synaptic response statistics

We will describe now the main statistical features of the synaptic responses, as a conse-

quence of the stochastic dynamics of the synaptic channel. We will depict these statistical

properties of the response train starting from particular simple cases (Poisson input,N0 = 1)

with which we may gain an intuition, and we will tackle later the harder general case.

9If the parameterpv takes a larger value, like in [Hanse and Gustafsson, 2001a] which is pv ∼ 0.5, the

expansion in equation2.2.25can be performed with respect to the parameterN0, which in those cases is small

(∼ 1). Thus, we identify1q ' −ln(1− pv).

Page 45:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.3. Results: synaptic response statistics 31

0 2 4 6 8 10 12 14 16 18 20N

0

0

0.2

0.4

0.6

U

0 0.2 0.4 0.6 0.8 1U

0

1

2

3

4

Γλ(U

)

Figure 2.5: Model for the population distributionD(U,N0). Top: Histogram of600 real-

izations of the marginal distribution of the release probabilityΓλ(U), which is fitted with

a gamma function of order 2 and parameterλ (this constraint has been taken from exper-

imental findings shown in [Murthy et al., 1997], where the valueλ = 7.9 has been taken

from). Bottom: the plane(N0, U) with 55 sampled points drawn from the population joint

distributionD(U,N0). The black line represents the experimental fit to the averaged release

probability〈U〉R(U |N0) which equals:1− (1− pv)N0 [Murthy et al., 2001]. The straight red

line represents the average〈U〉R(U |N0) = (N0 + 1)/q given by our model. We have taken

q ' pv, whose numerical value (taken from [Murthy et al., 2001]) readspv = 0.057.

2.3.1 Single docking site:N0 = 1

As we already mentioned, the case where there is at most one vesicle ready-for-release,

yields a renewal synaptic output process. This exhibits several advantages: first it permits to

calculate the connected conditional rate (that we will call correlation function for simplicity)

Ccr(t) of the responses and with it, compute the coefficient of variation of the IRI’s,CViri.

Moreover, in the next chapter, we will be able to compute the information conveyed in the

number of responsesn(T ) in a large time windowT (see section3.2).

Page 46:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

32 Chapter 2: A model of synaptic depression and unreliability

2.3.1.1 Poisson input

WhenCVisi = 1 the input statistics becomePoissonwith ρisi(∆) = νe−∆ν . In this

case the expressions for the IRI distributionρiri(∆|ν) and the correlation function of the

responsesCcr(t) read [de la Rocha et al., 2002] (see AppendixD)

ρiri(∆|ν) =νU

1− νUτv

(e−νU∆ − e−∆/τv) (2.3.1)

Ccr(t) = − νU

1 + νUτv

e−t/τr (2.3.2)

where we have defined a new time constantτr ≡ τv

1+νUτv, which corresponds to the time

scale of the correlations induced by the synaptic depression. It is now very helpful to recall

the interpretation of the synapse as a two-stage channel in which the afferent spikes are first

removed with a constant probabilityU , while in a second activity-dependent phase, they

reliably elicit a response if there is a vesicle at the RRP. Because now the input is Poisson,

the first stage becomes a trivial dilution of the train by a factorU which is equivalent to

renormalize the rateνd = Uν. The random synaptic response process defined by eq.2.3.1, is

acombinationof two exponentials arising from the two Poisson processes involved: the input

process with a diluted rateνd = Uν, and the replenishment of the docking site, which spends

a time drawn from an exponential distribution, and consequently it can be seen as a Poisson

process with rate1/τv. The ratio between the two rates,νUτv, determines which process is

faster and which is slower and hence dictates the behavior of the responses: whenνUτv 1,

the recovery of a vesicle is so fast that no depression can be observed and the release process

reproduces the renormalized Poisson input with rateνd. In the opposite limit, which will be

referred to hereafter as thesaturationregime,νUτv 1, almost every spike finds the site

empty, and when it is occupied a new afferent spike provokes its release straight away. In

this limit the output becomes again Poisson but with rate1/τv. As a consequence of this, in

the two limits the amplitude of the correlationCcr(t) (eq.2.3.2) vanishes to zero. However,

although both limits posses identical statistics (except for the rate) they differ enormously

in terms of information transmission. In the saturation limit, no information aboutν can be

transmitted because the post-synaptic neuron is onlylisteningthe input-independent refilling

of the RRP. This saturation regime can be better described by analyzing the rateνr and the

coefficientCViri of the responses, which for Poisson input read (AppendixD)

νr =νU

1 + νUτv

(2.3.3)

CViri = 1− 2τrνr =1 + (νUτv)

2

(1 + νUτv)2(2.3.4)

We can now analytically verify from these expressions that in the saturation regimeνUτv 1, νr approaches the asymptotic value1

τvandCViri → 1.

Page 47:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.3. Results: synaptic response statistics 33

0 0.5 1 1.5 2 2.5τ

v [s]

0

0.5

1

1.5

2

CV

2 iri

0 20 40 60 80 100ν [Hz]

0

0.5

1

1.5

2

ν r [H

z]

CV=1CV=1.5CV=2.5

Figure 2.6:(Top) Synaptic transfer function: the synaptic response rateνsr vs the input rate

ν. τv = 500 ms. Vertical arrows indicate the position of the corresponding saturating input

rateνsat(χ = 0.75) (see text). It can be observed that the correlated inputs saturate later, than

the Poisson input, to the same limit value1τv

. (Bottom) Squared coefficient of variation of

the IRI’s,CV 2iri, vs τv for several values ofCVisi. Top plot inset values apply to both plots.

Other parameters are:ν = 10 Hz(bottom),U = 0.5 andτc = 50 ms.

In order to formalize where does this saturation regime begin, we defineνsat(χ) as the

input rate at which the output rateνr has reached a fractionχ of its total dynamic range1τv

,

which means thatνr(νsat) = χτv

. Adopting this functional definition, we arrive atνsat(χ) =χ

(1−χ)Uτv. We realize now that we can arbitrarily fix the parameterχ, because the choice of a

different value, implies just a multiplication by a constant factor. Thus, we setχ = 0.5 and

finally obtain

νsat =1

Uτv

(2.3.5)

Thus, we recover the definition ofAbbott et al.[1997] of the limiting input rate beyond

which, no information can be transmitted in the stationary state about this spike rate. Since

the parameterU has been found to be plastic [Markram and Tsodyks, 1996a], the range

below saturation can be adjusted in an activity-dependent manner.

What happens between the two Poissonian limits? Fig.2.6(bottom, black line) illustrates

the behavior ofCViri as a function ofτv. In both limits, whenτv → 0 andτv → ∞, the co-

efficient of variationCViri → 1. In the whole intervalCViri < 1, which means that the train

Page 48:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

34 Chapter 2: A model of synaptic depression and unreliability

is more regular than Poisson. This can also be concluded by inspection of the correlation

functionCcr(t) (eq. 2.3.2): because it is an exponential with negative amplitude, the coeffi-

cient of variation must be negative (see eq.2.2.6). The fact that the correlation is negative,

implies that after a release occurs, and during a time window of size∼ τr, the probability

that the synapse triggers subsequent responses is reduced. Intuitively, the response train be-

comes more regular than the Poisson input because very short intervals have been removed,

that is, it is very unlikely that two spikes separated by a short ISI of size∆ (or more exactly

∆ τv) elicit, both releases. This occurs because a vesicle must be recruited to occupy

the empty docking site before the second spike comes, and this takes on average a timeτv.

Fig. 2.6also shows thatCViri has a minimum which is achieved precisely when this average

recovery timeτv equals the renormalized mean ISI,1Uν

. It is remarkable that, at this point,

the output release process becomes a gamma of order 2:ρisi(∆) = U2ν2 ∆ exp(−Uν∆)

whose coefficient of variation is exactly1√2.

2.3.1.2 Correlated input

Let us now study what happens when the input is correlated, that is we use theρisi(∆)

given by the two exponential distribution of eq.2.2.3. In this case,ρiri(∆) andCcr(t) result

in a combination of four and three exponentials, respectively (see AppendixD).

ρiri(∆|ν) =U

τv

(C1 e∆s1 + C2 e∆s2 + C3 e−

∆τv + C4 e

− ∆τ1

)(2.3.6)

Ccr(t) = K1 e−t( 1

τv−s1) + K2 e−t( 1

τv−s2) + K3 e−

tτc (2.3.7)

The coefficientsCi, Ki are fractional functions of both input and synaptic parameters. The

time constantτ1 = ( 1τv

+ 1τc

)−1 ands1 ands2 are the roots of a second order equation which

is shown in the AppendixD. From these expressions we can derive the output variablesνr

andCViri which read (see AppendixD)

νr =νU

1 + τvνU + τvUCV 2

isi−1

2(τv+τc)

(2.3.8)

CV 2iri =

H

(τc + τv + τv Uν τc + τv Uα + τv2Uν)2 (2.3.9)

where

H = τv4U2ν2 + 2 τv

3U2ν2τc + τv2U2α2 + 4 τv

2U2ν α τc + 2 Uα τv2 + τv

2 + τc2 +

+τv2U2ν2τc

2 + 2 τv U2α2τc + 4 τv Uα τc + 2 τv U2ν α τc2 + 2 τv τc + 2 Uα τc

2

andα = (CV 2isi − 1)/2.

Page 49:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.3. Results: synaptic response statistics 35

The first result that can be easily extracted from these formulas is that the saturation rate

νsat is altered by the input correlations in the following way

νsat =1

Uτv

+CV 2

isi − 1

2(τv + τc)(2.3.10)

which means that if the input stimulus is positively correlated the saturation regime is boosted

towards higher values of the input rate. In fig.2.6(top) we have plotted the synaptic transfer

functionνr(ν). Along with the Poisson input (black line) there are two more values of the

input CVisi (red and green lines). The three curves tend asymptotically to1τv

but the two

correlated ones approach slower than the Poisson case. Three small arrows have been placed

on the rate axis to indicate the corresponding valueνsat(χ = 0.75)10 in of each curve. It is

interesting to remark here that, since the dynamical range of the output rate does not change

with the input parameters, the increase of the the non-saturating range due to the positive

correlations, has the prize of a poorer encoding of this larger range. This can be seen by

comparing the slope of the functionνr(ν) of the Poisson and correlated examples showed in

fig.2.6for small values of the input rateν (below the first arrow).

In figure2.6(bottom) we have also plotted the coefficientCViri of the same three exam-

ples, with a fixedν = 10 Hz, as a function ofτv. For large values ofτv (saturation limit)

all curves tend to the same asymptote atCViri = 1. However, when the vesicle dynamics

are rapid (smallτv) the correlated cases produce more irregular trains than the Poisson input

case. This is the consequence of the larger variability of the ISI’s in the input (in particular

CVisi = 1.5 andCVisi = 2.5) and the fast recovery which permits short ISI’s to cross the

synapse. In the limitτv = 0 where there is no depression (the RRP gets instantaneously re-

plenished), the curves of the correlated cases do not hit theCViri-axis at their correspondent

input valuesCVisi because of the the unreliability represented by factorU . Particularly, the

effect on the variability (CVisi → CViri) of a non-dynamical stochastic filterU is to make

the output closer to a Poisson. This is expressed by:

CV 2iri − 1 = U(CV 2

isi − 1) (2.3.11)

As τv starts to increase, the response trains rapidly become more regular, as happened in

the Poisson case explained above. Now a larger value ofτv is needed to reach the minimum.

Since the minimumCViri reached is roughly the same than in the Poisson case, the reduction

of variability, measured as the distance from the maximum ofCViri (at τv = 0) to the

minimum, is greatly increased as the input correlations grow. Thus, we conclude that in

terms of regularizing the response train, short-term depression becomes more efficient the

10We have picked a different value ofχ for the sake of graphical clarity. Using the previously defined

νsat(χ = 0.5) would imply a simple rescaling of the arrowsx-position by a13 factor:

Page 50:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

36 Chapter 2: A model of synaptic depression and unreliability

ττ τ

τ τ τ

Poisson spikes

Bursty spikes

Figure 2.7: Why do correlated trains saturate for higher input rates than Poisson trains? The

picture shows two caricatured spike trains, namely a Poisson and a bursty correlated train,

with equalrateν. It is also shown how they would be transformed into a train of releases if

they arrive at a reliable (U = 1) synapse with a single docking site (N0 = 1). Vertical lines

represent spikes on a temporal axis:solid lines represent spikes transmitted whiledashed

lines represent unsuccessful ones, which found the RRP empty. Shadowed orange boxes

represent the time during which the release site is empty: at the left border of each box the

release of the vesicle occurs, while at the right border the docking of a new vesicle takes

place. We have set the recovery time equal at each release to illustrate the explanation more

clearly. Because of theclustered structureof the bursty train, there exists a larger interval

between the docking of a vesicle a the next release (blue arrows), which indicates that the

synapse saturates less than in the Poisson case. (Note:τ in the figure corresponds toτv in

the text).

more positively correlated is the stimulus. Perhaps a non-trivial comment is to indicate that,

at the point in which theCViri curve crosses the horizontal lineCViri = 1, the output is not

a Poisson process11.

Why do positively-correlated trains saturate for higher rates than Poisson trains (or than

negatively-correlated trains, data not shown)? The answer is illustrated in figure2.7 where

two spike trains transformed by the recovery dynamics have been caricatured, namely a

Poisson and a bursty input. Both spike trains have the same rateν. In the correlated input,

pulses tend to cluster so that, to keep the rate fixed, the interval between clusters (bursts)

11Recall that the constraintCV = 1 is a necessary but not sufficient condition for a renewal process to be

Poisson.

Page 51:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.3. Results: synaptic response statistics 37

is larger than the mean ISI. The orange boxes indicate the time in which the synapse is

empty so that any spike falling inside it, fails to elicit any response. The right end of each

window denotes the time at which a new vesicle has docked. For the bursty input, more

spikes are filtered out by the vesicle depletion (illustrated as an orange box), meaning that

νr is lower than for the Poisson input. However, a larger fraction of failures in transmission

does not imply more saturation. What is saturation then? Blue arrows indicate the time

interval between a docking event and a release. As the synapse is more and more saturated,

this temporal distance shrinks, eventually becoming zero (complete saturation). Thus, we

can observe that for the Poisson input the average of this interval is smaller than in the bursty

input. The key point is that when the same number of spikes (that is, the input rate kept fixed),

are clustered the distance between clusters increases as the burst gather more spikes (see also

figure2.3). Thus, at high rates, the same increase inν will produce a larger increase inνr if

the input is correlated because the “remaining” temporal range to codeν (blue intervals) is

larger in this case.

The last relevant effect of depression, related to the reduction of variability, is the trans-

formation of correlations. In figure2.8 (left) we represent the correlation function of the

response train,Ccr(t) (eq. 2.3.7) for different values ofτv, as well as the correlation of the

input spike trainCc(t) (eq. 2.2.6). The caseτv = 0 (no depression) shows that the output

correlation equals the input multiplied by a factorU . As soon asτv has a finite value, the

correlation for short intervals is rapidly reduced because of the effectivesynaptic refractory

timethat the vesicle replenishment imposes. For larger values ofτv, Ccr(t) becomes negative

for all t, and in the saturation limit, when the statistics of the responses is governed by the

docking of the vesicles, the correlation vanishes.

It is known [Atick, 1992, Dan et al., 1996, Nadal et al., 1998] that de-correlation, or

more generally, minimal redundancy, is directly related with the maximization of the infor-

mation transmission. However, it turns out that in our system this connection is clearly not so

straightforward, due to the fact that in the saturation limit, where the output does not convey

any information about the input rateν, the response train is indeed uncorrelated. Despite

this observation, we will compare the mutual information with the total magnitude of the

correlation function that will be quantified by means of the following measure

K =∫ ∞

0dt |Cc

r(t)| (2.3.12)

i.e. the integral of the absolute value (to avoid cancellations due to the sign) of the correlation.

This absolute measure of the correlation is plotted in fig.2.8 (right) for the same parameter

values chosen for the left side plot. It is interesting to realize that even if the general trend of

K asτv grows is to decrease, it draws a local minimum before it goes asymptotically to zero,

Page 52:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

38 Chapter 2: A model of synaptic depression and unreliability

0 0.1 0.2 0.3 0.4 0.5

∆ [s]

−2

0

2

4

6

Crc (∆

) [H

z]

inputτ

v= 0 ms

τv= 10 ms

τv= 50 ms

τv= 500 ms

τv= 2 s

0 0.2 0.4 0.6 0.8

τv [ms]

0.1

0.15

0.2

0.25

0.3

K

τc= 100 ms

τc= 500 ms

Figure 2.8:Left plot: Connected conditional rateCcr(∆) (also called correlation function in

the text) of the synaptic responses for several values ofτv. The input connected conditional

rate,Cc(t), is also shown for comparison (thick black line).Right plot: Integral of the

absolute value ofCcr(∆) (denoted asK in the text) as a function ofτv. From both plots it

can be deduced that in the limitτv →∞ the output is completely decorrelated. However,K

displays, for someτc values, a non-monotonic behavior. Parameters:ν = 10 Hz, τc = 100

ms.(left),CVisi = 1.5, U = 0.5, N0 = 1.

2.3.2 Multiple docking sites:N0 > 1

When the synaptic contact is endowed with several vesicle docking sites, upon arrival

of a spike, transmission occurs with probabilityU only if the ready-releasable pool (RRP)

contains at leastNth vesicles. As previously mentioned (section2.2.3), in the stationary

state, the RRP size never gets belowNth − 1, so that the RRP absolute size can be redefined

by settingNth = 1. Because of this, the synapse will release transmitter with probability

U if there is one or more vesicles ready for release. The main difference with respect to

the single docking site scenario is that now, if a long enough time interval elapses with no

spikes hitting the terminal, the RRP size,N , can grow untilN ∼ N0. At that point, several

consecutive releases must occur (approx.N0) before the pool gets depleted and depression

takes place. Thus, if the number of sitesN0 is large, even if the recruitment of vesicles is not

too fast, depression would be moderate, becauseN would usually be well over the edge of

the step-like function, that isN 1.

Page 53:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.3. Results: synaptic response statistics 39

0

0.5

1

1.5

2

ρ iri(∆

| ν)

[s

−1 ]

N0= 1, τ

v= 0.5 s

N0= 2, τ

v= 1 s

N0= 3, τ

v= 1.5 s

N0= 4, τ

v= 2 s

CV=1 (Poisson)

0

1

2

3

4

CV=2 (Bursts)

0 1 2 3∆ (s)

0

0.5

1

1.5

2

[s−1

]

ρ(∆ | ν; [N0=1,τ

v=0.5])

0 1 2 3∆ (s)

0

1

2

3

4

Figure 2.9: Distribution of the synaptic responsesρiri(∆|ν) for several values of the number

of docking sitesN0 and different recovery rates1/τv. Top plots: distribution ρiri(∆|ν)

for a Poisson input (top left) and a correlated input withCVisi = 2 (top right). Bottom

plots: Decomposition ofρiri(∆|ν) in the two terms shown in equation2.3.13, namely: i) the

leading termρiri(∆, [N0 = 1, τv ]N0

) which equals the distribution given by a single docking

site with a renormalized recovery rate; ii) a perturbative term which includes the qualitatively

different response effects obtained with several docking sites (e.g. the non-zero probability

of obtaining IRI’s of infinitely small size). It can be observed in the two right panels, that the

introduction of input short-range correlations changes severely the shape ofρiri only when

N0 > 1. Top plot inset values apply to all plots. Bottom plot inset values apply to bottom

plots. Other parameters are:ν = 10 Hz, τc = 50 ms.,U = 0.5, τv = 0.5 s.

Page 54:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

40 Chapter 2: A model of synaptic depression and unreliability

Because our aim is focused on the optimization of the information transmission through

the synapse, and since depression will turn out to be beneficial for this purpose, we now write

the distribution of IRI’s,ρiri(∆), of a multiple site synapse as a perturbative extension of a

single site model. The reason for this strategy is that, as will be shown in the next section, the

more moderate is the depression trace due to a largeN0, the worse will be the transmission

(at the optimal value ofτv). In this sense, the single docking site model performs better than

the multi-docking site model12, which under some choice of the parameters can resemble a

single docking site model. Thus, if we manipulate the expression of the Laplace transform

of the IRI distributionρdiri(s) eq.2.2.18we find thet it can be written like as

ρiri(s, [N0, τv]) = ρiri(s, [1,τv

N0

]) + (1− ℘ss1 )

ρdisi(s + N0

τv)(1− ρd

isi(s))

1− ρdisi(s + N0

τv)

(2.3.13)

which means that the distribution forN0 docking sites recovering withτv, ρiri(∆, [N0, τv]),

equals the distribution of the single site case with recovery constantτv

N0, plus a perturbative

term which is multiplied by the factor(1− ℘ss1 ) = ℘ss(N>1)

℘ss(N≥1)(see eq.2.2.19). This factor rep-

resents the probability that, known that the RRP is not empty, it holds more than one vesicle.

This expression however, could be indeed partially deduced: let us suppose that the synapse

is near the saturation regime, or in other words that a large fraction of the afferent spikes

found the RRP empty because the vesicle recovery machinery did not work fast enough. In

this case the probability that the RRP size exceeds one is small (℘ss(N > 1) 1) and

therefore the factor(1 − ℘ss1 ) ∼ 0. Then, if we sit at the synaptic contact and watch the

vesicles randomly dock at any of theN0 sites, they would arrive as a Poisson process of rateN0

τv, which results of the superposition of theN0 independent Poisson processes with rate1

τv

which govern the refilling of each individual site. Thus, the multi-site model becomes ef-

fectively a single-site contact with a recovery time constantN0 times smaller (i.e.N0 times

faster), andρiri(∆, [N0, τv]) is well approximated byρiri(∆, [1, τv

N0]). When the stationary

state of the synapse is far from the saturation regime, then the second term in eq.2.3.13

becomes comparable to the first one. In the limit whenN0 becomes very large (and the rest

of the parameters are fixed) the output distributionρiri(∆, [N0, τv]) → ρdisi(∆).

Figure2.9 shows the distributionρiri(∆|ν) for several values ofN0. As more docking

sites have been included, the recovery time has been changed in such a way that the ratioτv

N0is kept constant. This, as we just explained, leads to a comparison of different con-

tacts with the same the limit frequency (the maximum frequency at which releases can take

place). For this reason, the tails of the distribution coincide, independently ofN0 (top plots).

12Depending on the optimization criterion the optimal transmission will occur with a single docking site, if

the Fisher information is picked, while a two-docking site contact will be optimal if the Shannon information

is chosen.

Page 55:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.3. Results: synaptic response statistics 41

The bottom plots show the decomposition explained in equation2.3.13into a leading term

ρiri(∆|ν, [1, τv

N0]) (which is common to all values ofN0) and a perturbative term (second term

in the r.h.s. of eq.2.3.13), which includes the effect of several vesicles being docked at the

same time. In both Poisson and correlated cases, the introduction ofN0 > 1 provokes a

qualitativechange at the origin, namely, thatρiri(∆ = 0|ν) > 0. This change is particularly

prominent when the input is correlated (exhibited in the large peak at the origin reflect-

ing the effect of the input bursts). Thus, we can conclude that adding a new docking site

while renormalizing the recovery rate, only produces a substantial change when going from

N0 = 1 to N0 = 2. This will have important consequences when computing the information

for different values ofN0.

0 10 20 30 40 50 60 70ν [Hz]

2

4

6

ν r [H

z]

0

2

4

6

8

ν r [H

z]

N0=1

N0=2

N0=3

N0=4

Figure 2.10: Synaptic response rateνsr vs the input rateν for different values ofN0. CVisi =

1 (top) andCVisi = 2.5 (bottom). The straight dashed line is the limitN0 → ∞ where no

depression is observed. Vertical arrows indicate the position of the corresponding saturation

rateνsat(χ = 0.75) (see text). As theN0 increases,νsat also increases. Top plot inset values

apply to both plots. Parameters:τc = 50 ms.,τv = 500 ms. andU = 0.5.

In figure 2.10 the synaptic transfer functionνr(ν) is plotted for different values ofN0

while τv has been kept fixed. The behavior in all cases is qualitatively the same. It is clear

though, that the rate to whichνr saturates grows withN0 as N0

τv. It is also apparent, that the

greaterN0 the later in theν-axis depression takes place (seen as a deviation from the straight

line of slopeU ). The position ofνsat is represented by the vertical arrows on theν-axis.

Page 56:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

42 Chapter 2: A model of synaptic depression and unreliability

2.4 Tables of symbols

MODEL OF SYNAPTIC CONTACT

pv Single vesicle fusion probability , defined as the probability that an

individual docked vesicle undergoes exocytosis upon arrival

of a spike (see section2.1)

pr(N) Probability of auni-vesicularrelease as a function of the number

of docked vesicles,N

N0 Maximum number of vesicles in the

readily-releasable-pool (RRP)

Nth Threshold in the number of docked vesicles above which

pr(N > Nth) = U

U Release probability when the number of docked

vesicles isN > Nth

τv Mean recovery time of a single docking site

Prec(n, t|N0 −N) Probability thatn out ofN0 −N empty docking sites

are recovered in a time windowt.

℘N(tj) Probability ofN vesicles being docked upon

arrival of thej − th spike

℘ssN Probability ofN vesicles being docked upon

arrival of a spike in the stationary state

℘ssN Probability ofN vesicles being docked upon arrival of a spike

in the stationary state known that it will trigger a response

Table 2.1: Parameters and functions of the model of a pre-synaptic terminal

Page 57:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.4. Tables of symbols 43

MODEL POPULATION OF SYNAPSES

D(U,N0, τv) Population joint distribution of the synaptic

parameters[U,N0, τv]

D(U,N0) Population joint distribution of the synaptic

parameters[U,N0]

f(N0) Marginal probability of the population parameterN0

R(U |N0) Conditioned probability ofU givenN0, modeled

as Gamma functions of orderN0 + 1

q Parameter of the Gamma functionsR(U |N0)

Γλ(U) Marginal probability ofN0

λ Parameter of the Gamma functionΓλ(U)

Table 2.2: Parameters and functions of the population of synapses distribution

Page 58:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

44 Chapter 2: A model of synaptic depression and unreliability

INPUT STATISTICS

ν Pre-synaptic firing rate

CVisi Coefficient of variation of pre-synaptic inter-spike-intervals (ISI’s)

τc Temporal scale of the exponential auto-correlations of the input

ρisi(t|ν) Probability distribution function of the inter-spike-intervals (ISI’s)

given the input rateν

[β1, β2, ε] Parameters of the p.d.f.ρisi(t|ν) which are

complex functions of the physical parameters[ν, CVisi, τc]

C(t, t′) Two-point correlation function, defined as the probability

of finding spikes at timest andt′

Cc(t) Connected conditional rate, defined as the extra probability of

finding a spike at timet given that there wasanotherone at time zero

Nb Mean number of spikes within a burst

Tb Mean duration of a burst

f(ν) Probability distribution function of the input ratesν

ρisi(t) Probability distribution function of the inter-spike-intervals (ISI’s)

for the whole ensemble of inputs

νd Rate of thediluted input process, defined

as the result of decimating the input by a random factorU

ρdisi(t|νd) P.d.f. of the intervals of thedilutedspike train

ρisi(s) Laplace transform of the ISI’s p.d.f.ρisi(t|ν)

Table 2.3: Parameters and functions used to model the input spike statistics.

Page 59:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

2.4. Tables of symbols 45

SYNAPTIC RESPONSE STATISTICS

νr Rate of synaptic responses (or releases)

CViri Coefficient of variation of the inter-response-interval (IRI’s)

ρiri(t) Probability distribution function of inter-response-intervals (IRI’s)

Ccr(t) Connected conditional rate of responses, defined as the probability of a release

at timet given there was one at time zero

τr Time scale of the exponential correlations of the releases

when the input is Poisson

νsat Saturation frequency, defined as the input rate at which

the output rateνr has reached half of its maximal value

K Absolute measure of the response correlations, defined as

the integral of the absolute value ofCcr(t)

Table 2.4: Parameters and functions used to model the synaptic response statistics

Page 60:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 61:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Chapter 3

Information transmission through

synapses with STD

3.1 Introduction

In this chapter we will try to answer some of the fundamental issues addressed in this

thesis, which can be resumed in a simple general question: how do the biophysical properties

of the synapses, namely the fact that vesicle release is unreliable and the effect of short term

depression (STD), affect theinformation transmission from the pre to the post-synaptic

terminal?

To answer this question we will build upon the synaptic model presented in the previ-

ous chapter. In particular, we will compute analytically the information transmitted by one

synapse composed of a single contact (or synaptic specialization) but many vesicle docking

sites like the one described in section2.2.1. Moreover, we will make use of the statistical

properties of the synaptic responses computed in section2.3of the previous chapter. There,

we obtained an analytical expression for the distribution of the inter-response-intervals (IRIs)

ρiri(∆|ν) when the input is a renewal process with exponential temporal correlations be-

tween the spike times. Therefore, the distributionρiri(∆|ν), along with the statistical quan-

titiesνr andCViri, will be the starting point of our information calculation.

It is necessary to describe here the basic hypothesis and simplifications that we make to

address the problem of the quantification of the information.

• Code: rate, timing,... The first decision one takes when calculating information is

which are the input and output signal variables or, in other words, which is the code.

We will assume that the input signal is the pre-synaptic firing rateν, which means that

the pre-synaptic neuron uses arate codeto transmit information to the post-synaptic

47

Page 62:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

48 Chapter 3: Information transmission through synapses with STD

cell. In the output, we will consider two different choices: the precise timing of the

responses, conveyed in the size of the consecutive IRIs, and the number of responses

in a certain time window. It is obvious that the first of these codes always gives more

information than the second because, besides the number of responses, it provides the

particular temporal pattern. However it is not so trivial to state that, if one only wants

to determine the rate of the particular process, knowing the temporal pattern still gives

you an advantage over knowing just the number responses. In the general case, the

estimation of the rate is always better with a temporal code than with a counting code.

Only in certain particular cases, like the Poisson process, no information is lost when

the spike count is used to estimate the rate [van Vreeswijk, 2001]. For this reason, in

the cases that it is technically possible, we will compute the information contained in

the output by both codes.

• Information per response vs. information per unit time. Because we are inter-

ested in finding the values of the synaptic parameters that optimize the information

transmission, a pertinent question immediately arises: is it desirable to maximize the

information per synaptic response or the information per second, what is commonly

called information rate. The latter case would be relevant if the post-synaptic neuron

needed to decode the inputν as rapidly as possible without any metabolic constraints.

On the contrary, if the cell is trying to extract the information in an efficient way in

terms of energy consumption, no matter the velocity of the process, it would maximize

the information per response. It seems clear that these two scenarios are limit situa-

tions unlikely to occur in the brain. Perhaps the plausible optimization occurring in the

cortex, if any, takes into account both aspects in a complex manner, lying in between

the two simple limits. Nevertheless, if one takes care not to reach unreasonable results,

the analysis of these two approaches gives a valid approximation to the more realistic

problem. For instance, if the optimal value of the release probability becomes infini-

tively small, that would be an absurd result because it would lead to the meaningless

conclusion that it would take an infinitively large time to extract the information con-

tained in the output, since the output rate would be almost zero. In conclusion, we will

analyze both the information per response and per unit time, discussing every time the

significance of the optimization results.

• Stationarity vs transient. The third assumption is that the analysis is carried out in

a stationary situation where the input firing rateν is constant and the synapse is in a

stationary state. This stationarity does not mean that the synaptic vesicle dynamics

remain in equilibrium, but only that thestatisticalproperties of the synapse (such as

the probability that the docking sites are empty) do not depend on time. The reason

Page 63:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.1. Introduction 49

for this assumption is simply to start with the easiest non-trivial case in which the

calculations are analytically tractable. Several previous works [Abbott et al., 1997,

Tsodyks and Markram, 1997] have proposed that short-term depression might play

an important role in the transmission of information over non-stationary periods, in

particular during the transient following a sudden change of the input firing rate. The

analysis of transient aspects however, are left for future work.

• Temporal correlations. The last key-point we have introduced in our analysis is the

presence of temporal auto-correlations in the spike trains (see section2.2.2). As orig-

inally proposed byGoldman et al.[1999], Goldman[2000] and as we will show in

this chapter [de la Rocha et al., 2002], depressing synapses are specially well suited to

transmit information from positively correlated spike trains. This will turn out to be a

key-point when looking for the optimal value of the vesicle recovery timeτv: it will

be shown that, unless there are positive correlations in the input, the optimal recovery

time is zero,τopt = 0, which means that a non-depressing synapse is optimal (this

is the case for instance when the input is Poisson). As soon as positive correlations

are introduced, depressing synapses become advantageous over static ones, meaning

τopt > 0. Because the correlations considered are positive and exponentially shaped,

their existence implies an increase in the variability of the ISI’s with respect to the

Poisson case (see the expression for the conditional rate in eq.2.2.6). For this reason,

in this particular case increasing(decreasing) the variability of the input is equivalent

to increasing(decreasing) the positive autocorrelations. Nevertheless, this equivalence

breaks down when we reduce the variability belowCV = 1. In this case decreasing

the variability means introducingnegativecorrelations (see eq.2.2.6). From hereafter

we will make this correspondence (correlations = variability) when is pertinent without

further explanations.

The ability of STD to increase the information transmission about the precise input spike

pattern has been studied previously by different groups [Matveev and Wang, 2000a, Gold-

man, 2000, Fuhrmann et al., 2002]. Goldman[2000] computed the information about the

input spike times contained in an IRI. He found that STD is advantageous when compared

with an unreliable synapse in which the release probability is renormalized so that the frac-

tion of successful spikes in both cases is the same. He also showed that in the case in which

the input is embedded with positive correlations, the transmission capacity of a depressing

synapse is enhanced. However, from his results one concludes that both release unreliability

and short-term depression (produced for example by vesicle depletion) always decrease the

amount of information transmitted in comparison with a reliable synapse (U = 1) with no

depression (τv = 0). In other words, he shows that the existence of these two characteristics

Page 64:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

50 Chapter 3: Information transmission through synapses with STD

of the synapses in the brain cannot be justified in terms of the optimization of the informa-

tion transmission.Fuhrmann et al.[2002] studied the information about the input spike times

conveyed in a different alternative variable: the magnitude of a single EPSP. Because of the

existence of non-linearities due to the presence of short-term depression and facilitation,

they show that the amplitude of a single EPSP carries information about the times of several

preceding spikes. They also find an optimal vesicle recovery time constantτopt > 0, which

implies that depression is increasing the capacity of the channel, but the perfectly reliable

synapse always performs better than the unreliable one, i.e. they findUopt = 1.

3.2 Methods

We introduce now the different information measures we have computed and briefly show

how they are obtained and used. We are interested in establishing what the information

transmission capacity of synapses with short-term depression is, and to answer whether the

synaptic parameters can be tuned to optimize the transmission.

3.2.1 Fisher Information

The Fisher informationJ is a quantity widely used in the context of parameter estimation

[Blahut, 1988, Frieden, 1999]. It quantifies the encoding accuracy of a given particular value

of the input variables. In our case, this variable is the input rate,s = ν. J is related to the

mean squared stimulus reconstruction error,σ2s , through the Cramer-Rao inequality [Blahut,

1988]

σ2s ≥

1

J(3.2.1)

Therefore, it provides a measure of how important is the error in the estimation ofν.

When taking the IRIs temporal length as the output encoding variable, that is, the timing

code, we will refer toJsr(ν|∆) as the Fisher information about the rateν given one interval

between responses,∆, and we will compute it using the usual definition [Blahut, 1988]

Jsr(ν|∆) =∫

d∆ρiri(∆|ν)

(∂

∂νlog ρiri(∆|ν)

)2

(3.2.2)

When the output encoding is the number of responses within a time windowT , n(T ), one

needs to calculate the probabilityg(n(T )|ν) of n(T ). This is impracticable for a generalT ,

so we will use the expression valid for the largeT limit [ van Vreeswijk, 2001], in which

g(n(T )|ν) becomes a Gaussian distribution with meanνrT and varianceFT νrT . FT is the

Page 65:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.2. Methods 51

Fano factor of the responses defined (for any kind of train) as

FT =V ar[n(T )]

< n(T ) >(3.2.3)

where< n(T ) > andV ar[n(T )] are the mean and the variance of the count of events in a

time windowT . In the limit mentioned, the Fisher information of the rateν givenn(T ) is

up to first order [van Vreeswijk, 2001]

JT (ν|n) =T

FT νr

(∂νr

∂ν

)2

(3.2.4)

The first quotient of the r.h.s. of this equality represents the Fisher information about there-

sponses rateνr given the number of releases,JT (νr|n). Becauseνr is a continuous function

of the spike rateν (what we call the synaptic transfer function), the partial derivative trans-

forms JT (νr|n) → JT (ν|n). Nevertheless, the Fano factor is analytically tractable only if

there exists an expression for the response correlation function,Cr(t), which can be obtained

only if N0 = 1 (see section2.2.2). Because of this restriction, the information contained in

the countn(T ) will be calculated only in the particular instance where there is exactly one

vesicle release site (N0 = 1). In this case, the synaptic response is a renewal process, so for

large enough time windowsT the Fano factorFT equals the squared coefficient of variation

of the IRIs,FT ' CV 2iri (see e.g. [Gabbiani and Koch, 1998]). We will make use of this

property and writeJT (ν|n) as a function ofCV 2iri.

Finally, to obtain the Fisher informationper synaptic response, we have to take the ra-

tio betweenJT (ν|n) and the average number of releases in the windowT , which is νrT ,

resulting [de la Rocha et al., 2002]

Jsr(ν|n) ≡ JT (ν|n)

νrT=

1

CV 2iriν

2r

(∂νr

∂ν

)2

(3.2.5)

Besides the information per response, we will use the informationper unit time, which

is obtained, in the case of the count code, dividingJT (ν|n) by T . Thus, we simply drop the

dependence ofJT (ν|n) onT and rename it asJ(ν|n).

J(ν|n) ≡ JT (ν|n)

T=

1

CV 2iriνr

(∂νr

∂ν

)2

(3.2.6)

For the timing code, we had shown the expression for the Fisher information given the size

of an IRI,Jsr(ν|∆) eq. 3.2.2. To transform this expression into the information contained

in the response timesper second, all we have to do is multiply by the rate of the responses,

νr, obtaining:

J(ν|∆) = νrJsr(ν|∆) = νr

∫d∆ρiri(∆|ν)

(∂

∂νlog ρiri(∆|ν)

)2

(3.2.7)

Page 66:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

52 Chapter 3: Information transmission through synapses with STD

3.2.2 Mutual Information

The second measure of information content we will employ is the mutual information,

also called Shannon information [Shannon, 1948, Blahut, 1988, Cover and Thomas, 1991].

An important difference with the Fisher information is that it is not defined for a fixed value

of ν, but for a whole input ensemble, defined by its probability distributionf(ν). The mu-

tual information not only quantifies the ability of discerning which stimulus was presented

by observation of the output response, but it also accounts for thecapacityof the input-

output channel, which basically measures the number of distinguishable signals that can be

communicated through that channel. This means that in the hypothetical case of perfect

discrimination of the stimulus (case in which the Fisher Information would be infinity), the

information would be upper bounded by the entropy of the stimulus1. This upper bound is

directly related with an essential characteristic of the mutual information which states that it

is not additive, that is, the observation of independent realizations of the response does not,

in general, increase the information linearly. On the contrary the Fisher information does

grow linearly with the number of independent output observations [Nadal, 2000, Brunel and

Nadal, 1998].

We use the following definition for the information aboutν conveyed in an interval be-

tween responses∆ [Cover and Thomas, 1991]

I(∆; ν) = H(∆)− 〈H(∆|ν)〉ν (3.2.8)

where the angular brackets denote average over the input ensemblef(ν), and the entropies

H(∆) and〈H(ν|∆)〉ν are defined as

H(∆) = −∫ ∞

0d∆ ρiri(∆) log2 ρiri(∆) (3.2.9)

〈H(∆|ν)〉ν = − 〈∫ ∞

0d∆ ρiri(∆|ν) log2 ρiri(∆|ν) 〉ν (3.2.10)

and the IRI distributionρiri(∆) is obtained by averaging over the input distribution the con-

ditioned distribution:

ρiri(∆) = 〈 ρiri(∆|ν) 〉ν =∫

dνf(ν)ρiri(∆|ν) (3.2.11)

We will use an exponential function for the input rate distribution for two reasons: first

because it has been measured that neurons follow in some areas such a distribution [Treves

et al., 1999]. Second, because in the context of parameter estimation the prior distribution

is typically chosen as the maximal entropy distribution, taking into account all the known

1In the case that the input signal is a real random variableX with a certain distributionf(X), the entropy

H(X) is infinity, meaning that in the case where the noise in the channel is zero, the information is infinity.

Page 67:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.2. Methods 53

constraints on the parameter. In our case, the estimated parameter/stimulus is the input rateν,

and the only constraint we impose is that it has a given finite averageν. With this restriction

the function which maximizes the entropy is an exponential distribution:

f(ν) =e−ν/ν

ν(3.2.12)

The information per unit time, that is, the information rateR(∆; ν) , is defined as

R(∆; ν) = νr I(∆; ν) (3.2.13)

In both approaches, Fisher and mutual information, we will look for those values of the

synaptic parametersτv andU which maximize the information. We will denote these optimal

values byτopt andUopt. We will study first how these optimal variables depend on both the

input parameters[ν, CV, τc] and the other synaptic parameters, and then make a quantitative

comparison of their values with those in the experimental literature.

3.2.3 Mutual Information in a population of synapses

Besides computing the information transmitted by a single synaptic contact, we will

consider the information conveyed in the responses produced by a population of synapses.

This population of synapses are activated independently (their activity is not correlated) by

a population of pre-synaptic neurons which fire with the same statistics, that is, the sameν,

CV andτc. As illustrated in figure2.4in the previous chapter, we model the response to this

multi-synaptic stimulus as an array of IRIs,∆iMi=1. Since each synapse is stimulated by a

different spike train (though with the same statistics) given the input rateν the distribution

of ∆iMi=1 factorizes as follows

ρiri(∆i|ν) =M∏i=1

ρiri(∆i|ν) (3.2.14)

When this conditional independency holds, there exists an explicit relationship between the

Fisher information aboutν given∆iMi=1, J(∆i|ν), and the mutual information between

the variablesν and the M-dimensional∆iMi=1 which is valid in the largeM limit [ Rissanen,

1996, Brunel and Nadal, 1998]

I(∆i; ν) '∫

dν f(ν) log(1

f(ν)) −

∫dν f(ν) log(

√2πe

J(∆i|ν)) (3.2.15)

Due to the independence of the individual responses∆i given ν, J(∆i|ν) is just the

sum of the individual informations, which, for largeM , can be replaced by the average over

Page 68:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

54 Chapter 3: Information transmission through synapses with STD

the distributionD(U,N0, τv) of the synaptic parameters

J(∆i|ν) =∑M

i=1 Ji(∆i|ν) ' M∑N0

∫ ∫dUdτvD(U,N0, τv)J(∆|ν) (3.2.16)

Using the relation between Fisher information and Shannon information of a population,

we are able to compute the latter for a general input p.d.f.f(ν) and a general synaptic

distributionD(U,N0, τv).

If a neuron could modify the properties of the synapses impinging onto its dendrites, and

if its purpose was to read the rateν of a population of afferent stimuli, which follow a certain

distributionf(ν), the optimal way to achieve this goal would be to change in acoordinate

manner, the physiological values of the parameters[U,N0, τv] of all the pre-synaptic termi-

nals contacting upon its dendritic tree, according to the optimal distributionDopt(U,N0, τv).

This function is defined as the p.d.f. which maximizes the information about the input rate,

that the times of the responses, produced by each contact, convey.

We would like to think that CA3-CA1 synapses, from which we have partially modeled

this distribution, are optimized in this sense. Thus, the answer to the question of how the

values of the recovery time constantτv are distributed and how they correlate withN0 andU

would arise if we could optimizeD(U,N0, τv) under the constraints the experimental find-

ings impose [Dobrunz and Stevens, 1997, Murthy et al., 1997, 2001]. This is however a

non-practicable calculation if one tries to accomplish it in an analytical fashion. We will

therefore, compute an approximate solution to this problem by means of a different opti-

mization criteria. Instead of optimizing globally through the population, the optimization

will be performed locally. Thus, the relevant solution may be lost with this procedure. Let

us write the complete synaptic distribution as follows

D(U,N0, τv) = f(N0) R(U |N0) P (τv|N0, U) (3.2.17)

We have mentioned before that at a single contact, there exists an optimal valueτopt for the

recovery time of the synapse which maximizes the mutual informationI(ν ; ∆). This τopt

depends firstly on the input parametersCV andτc and on the rate distributionf(ν), but also

in the other two synaptic parametersU andN0. Thus, we hypothesize that every synaptic

bouton (defined by the values ofU andN0) individually adjusts the velocity of its synaptic

machinery by modifyingτv so that it matches the optimal valueτopt(U,N0). In this case, the

population distributionD(U,N0, τv) would take the form

D(U,N0, τv) = f(N0) R(U |N0) δ(τv − τopt(N0, U)) (3.2.18)

and we can reach the marginal distribution ofτv by simply summing and integrating overN0

andU

g(τv) =∑N0

∫dU D(U,N0, τv) (3.2.19)

Page 69:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.2. Methods 55

=∑N0

∫dUf(N0) R(U |N0) δ(τv − τopt(N0, U))

It is important to remark here that this procedure does not lead us in general, to the

optimal solutionDopt(U,N0, τv). By assuming that each contact tunes the parameterτv

individually, we prevent the ensemble of synapses from beating the problem in a cooperative

manner (which would be the case in the instance of optimizingD(U,N0, τv) as a whole).

3.2.4 Optimization with Metabolic considerations in the recovery rate

When optimizing the mutual information by adjusting a number of biophysical param-

eters, or a certain distribution of them, one must not ignore that the mechanism underlying

information transmission consumes energy. An organism must have an optimal representa-

tion of the sensory world, but at what prize? The final goal of achieving such an efficient

encoding is the organism survival. It would be a disadvantage for the animal to have the

most efficient encoding if, for instance, having a certain parameter equal to its optimal value

requires more energy than the extra metabolic gain (e.g. obtained by procuring food with the

optimized sensory mechanism). Another example would be the rates at which tha neurons

fire. If instead of emitting AP’s at 10-50 Hz they fired at 1000 Hz, perhaps the temporal

resolution of the visual system could be enhanced. However, firing at those rates our neu-

rons would consume many more ATP molecules, and the benefit of having such a super-fast

vision would not make the elevated rates worthwhile.

In conclusion, when making a particular optimization, one must consider the efficiency of

such a maximization in terms of metabolic consumption [Levy and Baxter, 1996, Baddeley

et al., 1997, Laughlin et al., 1998, Balasubramanian et al., 2001, de Polavieja, 2002]. The

simplest way to do it is to maximize the mutual information per unit energy,I which can be

defined asBalasubramanian et al.[2001], de Polavieja[2002]

I =I(X; Y )

E(~p)(3.2.20)

whereI(X; Y ) is the information conveyed byY aboutX, andE(~p) is the energy required

in the transmissionX → Y (which may include the production of the signalX, its transfor-

mation intoY , the reading of the responseY , etc.). One must propose, therefore, this energy

function, E(~p), which depends on the parameters~p to be optimized. If the details of the

metabolic consumption are known, the energy function can be derived from the biophysical

mechanisms taking place in the transmission. If, on the contrary, this dependence is un-

known, one needs to make anansatzaboutE(~p). The simplest one is that the energy scales

up linearly with energy-consuming variables such as the firing rate, the vesicle recovery rate,

Page 70:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

56 Chapter 3: Information transmission through synapses with STD

the release probability , etc.[Levy and Baxter, 1996, Baddeley et al., 1997, Balasubramanian

et al., 2001, de Polavieja, 2002]:

E(~p) = β r(~p) (3.2.21)

whereβ is a constant whose units are energy divided by the units of the variabler(~p), which

depends on the parameters~p. In this case, the value of~p which maximizesI does not depend

on theβ because it just represents a proportionality factor.

Let us introduce an example of efficient metabolic consumption for the synaptic channel

we are interested in. Here the input signalX is always the pre-synaptic rateν. The param-

eters to be optimized are the recovery time constantτv and the release probabilityU , i.e.

~p = τv, U. If we consider now that the output code is given by the times of the responses

in a given unit time, ∆t, we can think of an energy function which depends linearly on the

release rater = νr. In other words, it seems reasonable to state that the energy needed to

generate the output variable (the times of the responses in a time window) depends on the

mean number of responses per unit time,νr, which means that

E(τv, U) = β νr(τv, U) (3.2.22)

In this way the information per unit energy would read,

I(∆t; ν) =I(∆t; ν)

β νr

(3.2.23)

We have previously introduced the mutual information between the input rate and an inter-

response interval,I(∆; ν) , and the information rateR(∆; ν) = νr I(∆; ν) . Since, as men-

tioned in the previous section, the Shannon information is not additive [Cover and Thomas,

1991], the information conveyed by the responses lying in a time window does not equal,

in general, the information of a response times the number of them occurring in that time

window. When the responses are conditionally independent2, the information rateR(∆; ν)

is always larger or equal than the information of the responses occurring in a unit time,

I(∆t; ν) ≤ R(∆; ν) = νr I(∆; ν) (3.2.24)

which using eq.3.2.23leads to

I(∆t; ν) ≤ I(∆; ν) (3.2.25)

2This means that given the input signalν, the joint probability of many responses factorizes into the product

of individual probabilities. If this condition does not hold, the information given by several responses together

can be larger than the sum of the information each provides separately. This is known assynergy.

Page 71:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.2. Methods 57

Therefore, despite the inequality, one may consider that optimizing the informationper re-

sponse, I(∆; ν) , is approximately3 equivalent to maximizing the information of the re-

sponses produced in a small time window, considering the metabolic consumption due to the

generation of these responses, i.e.I(∆t; ν).

Another example refers to the cost of the recovery rate of the vesicles. If one assumes

that structural constraints make the synapses have only a small number of docking sites,

therefore, in order not to loose the information of incoming spikes, after the release of the

vesicles which were ready for release, the synapse should replenish those empty docking

sites in a fast manner. It might be the case that the optimal value of that recovery rate is

infinite, that is, the mean time elapsed between release and recovery equals zero (and thus

the synapse is never “empty”). However it seems reasonable to state that the number of

ATP molecules to refill an empty site at an infinite rate should be much larger then the case

of a slower recovery. This implies not only that the consumption rate (the number of ATP

molecules consumed per unit time) is higher but that each single recovery expends more the

faster it occurs. This can be formalized by assuming that the energy functionE(τv) grows

monotonically with the recovery rate1τv

. In particular we assume,

E(τv) =β

τv

(3.2.26)

In this case, the information about the input rate per unit energy and per response reads

I(∆; ν) =I(∆; ν)

E(τv)= β I(∆; ν) τv (3.2.27)

In the results section, it will be shown that, considering the metabolic consumption in this

way, primes larger values ofτv to be optimal, because of the higher cost of a fast replenish-

ment. In addition to the mutual information per unit energy we will also compute the Fisher

information per unit energy, defined as the ratioJE(~p)

.

3.2.5 Numerical methods

The specific way in which all the information measures have been computed varies from

one to another. As intended along most of this thesis, when tractable, the calculation has

been overcome analytically. Therefore, the Fisher information using the count code has been

computed analytically, because its calculation does not require to perform any integral and

all the necessary variables have been obtained in this way. However, in the cases of the Fisher

3Because the response rates are smallνr (∼ 1 − 5 Hz) and therelevanttime window is smaller than one

second, the number of responses within it will be just a few. This makes the additive approximation a better

estimation [Nadal, 2000]

Page 72:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

58 Chapter 3: Information transmission through synapses with STD

information based on the synaptic event timing, and of the mutual information, the integrals

present in their definitions do not have a primitive4, so they have been carried out numerically

using standard methods [Press et al., 2002]. To find the optimum values of the parameters

that maximize a certain quantity, except when explicitly indicated, numerical algorithms

have been used because analytics became rapidly very tedious. In this particular case, the

numerical approach was much faster giving almost the same insight into the problem.

To obtain the marginal of the optimal distribution of synaptic parameters defined in equa-

tion 3.2.20, a Monte Carlo simulation was performed generating values ofU andN0 using

the analytical expression of the marginalD(U,N0). For each data sample the optimalτv was

computed and an histogram of optimal values was obtained.

3.3 Results

In this section we will start analyzing how does the information depend on the input

frequencyν and on the input distribution of ratesf(ν). Afterwards, we will address the

question of how does the information depend on the biophysical parameters of the synapse

[τv, U, N0] and whether it is possible or not to find optimal values for them, which maximize

the information.

3.3.1 Rate dependence of information measures

3.3.1.1 Dependence of the Fisher information onν

We first study which input rates are best represented by the train of post-synaptic re-

sponses, that is, which rates would be reconstructed with the minimal error. This is expressed

by the Fisher information that, as we mentioned, sets a lower bound for the mean squared

error of the best estimator. We will assume, since now on, that the estimator that saturates

this bound exists so that we use the Cramer-Rao bound as an equality that states:σ = 1√J.

Because we are going to compare the errorσ for different values of the rateν, to make this

comparison adequate, we compute the relative errorε defined as the ratio between the error

and the signal:

ε =σ

ν=

1

ν√

J(3.3.1)

Depending whether we are referring to the Fisher information per response or per unit time,

we will express the relative error asεr andεt, respectively.

4Even in the simplest case in which the input is Poisson, the conditioned distribution of the responses

ρiri(∆|ν) is the sum of two exponentials [de la Rocha et al., 2002]. Thus, taking the logarithm of this sum of

exponentials immediately leads to an integrand with no primitive.

Page 73:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 59

What is the expected behavior ofεr andεt as a function ofν? Firstly, the two of them

have to be strongly influenced by the most obvious effect of synaptic depression: saturation.

As we discussed in the previous chapter, whenever the input rate is very high (ν 1Uτv

),

or the vesicle recovery process is very slow (τv 1Uν

), it happens that a large fraction of

incoming spikes find the synaptic bouton depleted of vesicles, and therefore do not succeed

to trigger any response detectable by the post-synaptic terminal5. In this saturating regime,

more accurately defined byντvU 1 (see section2.3), changes inν would hardly produce

any change in the response train statistics, and therefore the estimation ofν will be poor.

Thus, beyond some point, if we keep increasing the rateν, the relative errorε (no matter

whether it is per response or per unit time) has to increase and tend to infinity as we approach

the limit ν →∞.

Now, what happens when we approach the opposite limitν → 0? The first obvious

consequence is that the number of synaptic responses, within a fixed time window, tends to

zero, that is, we have less and less information in the post-synaptic cell about what occurs in

the afferent fiber. Thus, the Fisher information per second must tend to zero whenν → 0,

which means thatεt →∞. In conclusion, because the Fisher information is continuous inν,

there exists a non-trivial minimum forεt which is achieved for a particular value ofν, that

we shall denote byνmin.

In fig. 3.1 (bottom) the relative errorεt is plotted as a function ofν for several values

of the inputCVisi. Both theεt obtained fromJ(ν|∆) (timing code, solid lines) and from

J(ν|n) (counting code, dashed lines) are included. Several things must be noticed: i) In

all the cases there is a minimum at a value nearνmin ∼ 5Hz. ii) Comparing the codes,

timing vs. counting (for the same value ofCVisi), although the error is always smaller for

the timing case, for small values of the rate (ν < 10 Hz) both solid and dashed lines almost

superimposed, meaning that counting responses provides an estimation as good as observing

the timing. iii) Different inputCVisi’s have slightly differentνmin (see inset fig.3.1), or more

exactly, the bigger the inputCVisi the largerνmin. iv) However, in the case of timing code,

εt is very similar, for all the values ofCVisi explored, in theν range shown in the figure.

Unlike this insensitivity for the ISI variablity, the case of the counting code shows that, forν

beyondνmin, the larger theCVisi the worse is the estimation.

Because both codes, timing and counting, reach the minimum at the same valueνmin, we

can easily obtain its analytical formula using the simple expresion derived forJ(ν|n) (eq.

5To be precise, although closely related to the loss of a large fraction of spikes (which find the synapse

empty of vesicles), it has been shown in section2.2.3the level saturation is not equivalent to the fraction of

release failures. Certainly the latter it is a necessary condition to achieve saturation but is not sufficient (see the

example shown in figure2.7.

Page 74:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

60 Chapter 3: Information transmission through synapses with STD

0123456789

10111213

ε r (re

lativ

e er

ror

per

resp

onse

)

cv= 1cv= 1.5cv= 2

0 10 20 30 40 50

ν [hz]

0

1

2

3

4

5

6

7

8

9

10

ε t (re

lativ

e er

ror

per

sec.

)

0 5 10

Figure 3.1: Relative reconstruction errorε per response (top) and per second (bottom) versus

input rateν. The relativeε is derived from the Fisher Information (see eq.3.3.1) given the

number of responses (dashed lines) or the times of the responses (solid lines). Top plot inset

values apply for both plots. Inset, represents a magnification of the minima in bottom plot.

Parameters:τc = 50 ms,τv = 500 ms,U = 0.5 andN0 = 1.

3.2.6), that we write down again here:

J(ν|n) =1

CV 2iriνr

(∂νr

∂ν

)2

(3.3.2)

which yields to an expression forεt which reads

εt =CViri

√νr

ν

∂ν

∂νr

(3.3.3)

For a Poisson input, it takes the following expression as a function of the input and the

synaptic parameters

εt =

√(τv

2ν2U2 + 1) (1 + τv ν U)

Uν(3.3.4)

and its minimum occurs at

νmin =0.657

Uτv

(3.3.5)

Let us recall here, that the saturation frequencyνsat defined in the previous chapter, when

the input is Poisson (eq.2.3.5), is exactlyνsat = 1Uτv

. Thus, we can rewrite the rate where

the minimum is achieved asνmin = 0.657 νsat. The interpretation of this dependency is

straightforward: the reconstruction error per secondεt does not decrease monotonically as

Page 75:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 61

the input rate increases (due to the occurrence of more responses per second) because the

response rateνr rapidly saturates, (see the transfer function,νr(ν), in figure2.6). For that

reason, the later the transfer function saturates ( i.e. the biggerνsat), the laterεt will start

to increase because of this saturation, meaning a larger value forνmin. Furthermore, as we

saw in the previous chapter, introducing aCVisi > 1 in the input, boosts the saturation rate

νsat towards higher values (see eq.2.3.10). Consequently,νmin takes larger values the larger

is the inputCVisi (see inset of fig.3.1). We have checked numerically that the relation

νmin = 0.657 νsat also holds approximately whenCVisi > 1.

On the contrary, figure3.1(top) shows that the reconstruction relative errorper response,

εr, does not go to infinity asν goes to zero. This implies that if, in the limitν = 0, the

post-synaptic cell observes a response, it conveys some information to make an estimation.

Despite the reconstruction error in this limit is minimal, it is not a biophysically relevant

situation because there are barely any responses, and it would take a huge time to be able to

make any estimation.

Observing the quantitative aspect of the estimation, we find that for any values ofν the

relative error is always bigger than one. This means that the best we can do, is to estimate

the input rateν with an errorσ which is at least as large asν. This is obviously a very

poor estimation. Things get even worse beyond the saturation rateνsat where the error is

several times the signalν. We have to remind, however, that this is an estimation made by

the observation of one response in asinglesynapse. If we think that the post-synaptic neuron

receivesN afferent inputs providing the same signalν, assuming that the spike trains coming

along the pre-synaptic fibers are independent, the total Fisher information would be the sum

of N terms like the one we are analyzing. Hence, the errorε would be proportional to 1√N

,

and it can be made infinitively small by just pooling enough independent synaptic responses.

Concerning the value ofνmin, for the parameter values chosen in figure3.1, it ranges,

for different inputCVisi’s, between2 − 5 Hz. These low values ofνmin, apparently more

inside the interval of spontaneous cortical activity than in the range of rates associated with

the performance of a cognitive task, are due to the inverse relation with the recovery time

constantτv whose values in cortical pyramidal neurons varies between 0.4-1.5 s. [Markram,

1997, Varela et al., 1997, Finnerty et al., 1999, Varela et al., 1999, Petersen, 2002]. For

a Poisson input, since the release probability,U , runs from 0.1-0.95 [Rosenmund et al.,

1993, Hessler et al., 1993, Markram, 1997, Murthy et al., 1997, 2001], νmin ranges from0.657

0.95∗1.5∼ 0.5 Hz to 0.657

0.1∗0.5∼ 14Hz.

To conclude, in the case of the estimation error per second, there exists a finite non-zero

value of the input rateνmin for which a minimal reconstruction error is achieved. This value

is directly proportional to the saturation frequency defined in chapter2, so it is sensitive to

Page 76:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

62 Chapter 3: Information transmission through synapses with STD

0

0,05

0,1

0,15

0,2

Info

[bi

ts]

CV= 1CV= 1.5CV= 2CV= 2.5

Exponential input distribution: f(ν)= e(-ν/<ν>)/<ν>; τ

c= 20 ms., τ

v= 500 ms., U=0.5, N

0=1

0 10 20 30 40 50 60 70 80

<ν> [hz]

0

0,05

0,1

0,15

0,2

0,25

Info

Rat

e [b

its/s

]

Figure 3.2: InformationI(∆; ν) (top) and information rateR(∆; ν) νr I(∆; ν) (bottom) ver-

sus input rateν. Inset values apply for both plots. Parameters:τc = 20 ms, τv = 500 ms,

U = 0.5, N0 = 1.

the introduction of auto-correlations in the input by increasing the value ofCVisi. Besides,

both the counting and the timing code provide similar performance before we enter into the

saturating regime. When the estimation becomes severely affected by saturation, the timing

code seems to be less influenced than the counting code.

3.3.1.2 Dependence of the Mutual Information onf(ν)

We now study the question of what are the best input ensembles to be transmitted across

a unreliable depressing synapse. In other words, how does the Shannon information depend

on, for example, the mean input rateν? How does the choice of the input distributionf(ν),

affect the information transmitted? As opposed to the Fisher information, the computation

of the mutual information requires to define the probability distribution of the stimulus. We

take, to start with, an exponential functionf(ν) = e−ν/ν

νand later we will explore other

possible choices. A lower cut-offνinf was taken in the distributionf(ν) to account for the

experimental fact that neurons are seldom completely “quiet”, but in general they fire at least

with a low spontaneous rate which falls around 1-2 Hz [Sanseverino et al., 1973, Abeles,

1982, Legendy and Salcman, 1985]. Thus, the p.d.f. of the input rate reads:

f(ν) =e−

ν−νinfν

ν(3.3.6)

Page 77:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 63

if ν > νinf , and zero otherwise. This implies that the mean input rate is not anymore the

parameterν, but ν + νinf . However since we will take a rather smallνinf ∼ 1 Hz, we will

keep callingν the mean rate.

Figure3.2(top plot) responds to the question of which is the bestmeaninput rateν to op-

timally transmit information inonesynaptic IRI, when the input has fixed autocorrelations,

i.e. CVisi andτc are constant. The four values of the inputCVisi displayed, show a tuned

curve with a maximum at different rates, ranging between10 − 20 Hz. The maximum at a

non-trivial value of the mean rate can be explained by understanding why the information

drops to zero at both limitsν → 0 andν →∞. When the mean rate goes to zero, the distri-

bution converges, in the sense of distributions, to the Dirac functionδ(ν − νinf ). What does

this imply? A Dirac delta p.d.f. indicates that the only stimulus with non-zero probability is

ν = νinf . This in turn means, that the number of input messages isone, because the channel

transmits always the same frequency. This implies that the information is zero, because the

input rate is known previous to any observation.

The opposite limitν → ∞ approaches zero for a different reason: saturation. Whenν

is very large ( or in generalUντv >> 1, see section2.2.3) the synapse saturates and the

rate of responses is rather insensitive to changes in the input. In section3.3.2.2, it will be

shown that the two entropiesH(∆) and〈H(∆|ν)〉ν converge to the same function ofτv as

saturation becomes more and more prominent. Although increasing the mean input rateν

does not imply thatall input rates saturate, it makes the fraction of input messagesν which

are not well transmitted larger. As a consequence,I(∆; ν) tends to zero asν increases. As

a result of these behaviors at the limitsν → 0 and ν → ∞, I(∆; ν) exhibits a maximum.

In other words, given a synapse defined by the parametersU andτv , there exists an optimal

exponential input distribution, namely the one whose mean valueν maximizesI(∆; ν) in

the way just explained.

This optimization is possible regardless of the input correlations. It occurs for any values

of CVisi andτc . The effects of introducing correlations, namely the enhancement and shift of

the maximum, are due to the displacement of the saturation frequency. This can be observed

in figure3.2, where higher inputCVisi ’s produce a maximum at a higherν . The bottom plot

of this figure, shows that the information rate,R(∆; ν) , displays the same kind of behavior,

although now the mean rate at which the maximum is achieved, is larger (∼ 15 − 30 Hz.)

We want now to test whether this tuning inν can be reproduced for other choices of input

distributions. Figure3.3shows three different input p.d.f.f(ν), namely the exponential and

two Gamma distributions of order3 and10 defined as [Feller, 1950]

f(ν) =( p

ν )p

νp−1 e−pνν

(p−1)!(p = 1, 3, 10) (3.3.7)

where the order of the gamma is the parameterp. Notice that the exponential distribution is

Page 78:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

64 Chapter 3: Information transmission through synapses with STD

0 10 20 30 40 50

ν [Hz]

0

0,025

0,05

0,075

0,1

0,125

0,15

f(ν)

p=1p=3p=10

Gamma input distributions of order p and mean rate 10 Hz.

Figure 3.3: Three input Gamma distributionsf(ν) (orderp = 1, 3 and10) constructed in

such a way that all have meanν = 10 Hz (see expression3.3.7)

the particular casep = 1. These gamma functions have been constructed in such a way that

the mean rate does not depend onp, so that it is always equal toν .

Figure3.4 shows the informationI(∆; ν) as a function ofν for the three distributions,

p = 1, 3 and10. In the three cases the qualitative behavior is the same: all of them show

an optimal value ofν , although different in each case: the higher the order, the lower is the

optimal mean rate. Independently of the tuning, the overall scale decreases severely as the

order of the Gammap increases. This happens because the input entropyH(ν) associated

with each distribution decreases asp increases. An indirect proof is obtained by looking at

the input varianceV ar[ν], which is usually proportional to the entropy. For the particular

normalization chosen for the Gamma distributions (see eq.3.3.7) it reads6: V ar[ν] = ν2

p.

This means that asp increases, the range of possible input rates deacreases, that is, the

number of input distinguishable messages decreases makingI(∆; ν) become smaller.

To summarize, whatever the choice off(ν) within this family of distributions, the exis-

tence of an optimal mean rateν , seems to be a robust effect. The value of the optimalν

6If we had made the standard choice for the Gamma distributions, i.e.

f(ν) =ap e−aν νp−1

(p− 1)!(3.3.8)

the variance would be proportional to the orderp: V ar[ν] = pa2 . However in this case, the mean of the

distributions would not be the same for different ordersp, that is,< ν >= pa , and increasing the orderp would

produce a diminution ofI(∆; ν) because of the increasing of< ν >.

Page 79:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 65

0 10 20 30 40 50 60 70

ν [hz]

0

0,005

0,01

0,015

0,02

0,025

I(∆;

ν) [

bits

]

0

0,05

0,1

0,15

0,2I(

∆;ν)

[bi

ts]

τc= 50 ms., U=0.5, τ

v= 500 ms., N

0= 1

00,010,020,030,040,050,060,07

I(∆;

ν) [

bits

]

CV= 1CV= 1.3CV= 2

Gamma 1

Gamma 3

Gamma 10

Figure 3.4: InformationI(∆; ν) versus the mean input rateν , for three different input

distributionsf(ν), namely three Gamma functions of orderp =, 3 and10 (see eq.3.3.7).

Inset values apply to all plots. Parameters:τc = 50 ms,τv = 500 ms,U = 0.5, N0 = 1.

depends on the distribution chosen and on the correlations of the input, but it ranges between

5 − 20 Hz. When looking atR(∆; ν) , the tuning to an optimalν is even more prominent

and the maximum occurs at higher values ofν .

3.3.2 Optimization of the recovery time constantτv

We focus our attention now in determining how does the biophysical synaptic variable

τv influence the way the information is transmitted to the post-synaptic cell. Our aim is to

check the dependence of the different information measures onτv, and to compute the value

τopt that maximizes the observed quantity.

3.3.2.1 Optimizingτv with the Fisher Information

We start by analyzing the Fisher Information as a function ofτv. As done before, we will

distinguish the two output coding strategies (counting and timing), and the Fisher informa-

Page 80:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

66 Chapter 3: Information transmission through synapses with STD

tion per response and per unit time.

The dependance of the reconstruction error onτv, is dictated again by the two major

effects of depression in the response trains statistics (see section2.3):

i) Reduction of the variabilityCViri, by means of the elimination of very short IRIs,

resulting in a distribution for the IRIs closer to a Gamma function of order two (which

has aCViri = 1√2) than to an exponential (withCViri = 1).

ii) Saturation of the response rate.

The first, enhances the information because the estimation of the rate (or equivalently the

mean ISI) is easier in a regular train than in a very variable train. If we consider the limit

case in which the variability is zero, i.e. a periodic train withCViri = 0, the estimation error

committed by the observation of one IRI would be zero, because all the intervals equal the

mean IRI. In this case the Fisher information would be infinity. This is precisely expressed

by the dependence ofJsr(ν|n) andJ(ν|n) on the inverse ofCV 2iri, shown in equations3.2.5

and3.2.6.

Contrariwise, saturation diminishes the information due to the fact that eventually the

output rate becomes insensitive to changes in the input signalν. This is also shown explicitly

in the expressions ofJsr(ν|n) andJ(ν|n) by the squared derivative of the transfer function∂νr

∂ν(see equations.3.2.5and3.2.6) which, as the synapse starts saturating, decays rapidly to

zero.

The synaptic parameterτv tunes the strength of depression and consequently the level

of saturation of the system. In the case of Poisson input (where the saturation frequency is

defined asνsat = 1Uτv

, see section2.3), for a fixed frequencyν, a saturation recovery time

can be defined,τsat ≡ 1Uν

, as the value ofτv beyond which the system starts to saturate.

Qualitatively, if the replenish of vesicles takes a time much smaller than the typical diluted

ISI7, τv 1Uν

= τsat, the fraction of spikes that do not trigger any response due to vesicle

depletion is negligible so that changes inν may be observed inνr. In the limit τv → 0

the synapse does not show any depression at all. We will call the model at this limit, a

static synapse. On the other hand, whenτv 1Uν

= τsat, the synapse shows a strong

depression and saturation is prominent. Therefore, it seems like, regarding the consequences

of saturation on the information transfer alone, the optimalτv to estimate the rateν will be

zero, that is, a static synapse.

However, as illustrated in the previous chapter, the plot ofCViri versusτv shows that,

before saturation starts to play a dominant role, the variability of the output train of is severely

7As we described in section2.2.3 the complete 3-parameters synaptic model can be reduced to a 2-

parameters model by absorbing the parameterU into the input. The resulting input is basically adiluted

version of the original input. The diluted rate equalsνd = Uν.

Page 81:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 67

0,002

0,004

0,006

0,008

0,01

J pe

r re

spon

se [

s-2] CV=1

CV=1.5CV=2

0,1

1

(dν r/d

ν)2

0 0,2 0,4 0,6 0,8τ

v [s]

0,8

1

1,2

1,4

1,6

CV

iri

Figure 3.5: Optimization of the recovery time constantτv regarding the Fisher information

per responseJsr. Top: Solid linesrepresentJsr(ν|∆)(the Fisher information given the length

of an IRI) versus the recovery timeτv; dashed linesrepresentJsr(ν|n) (counting code);

dotted linesrepresent a counting code when the output is a “forced” Poisson (see text).

Middle: Squared derivate of the transfer functionνr(ν) in a logarithmic scale.Bottom:

Solid linesrepresent the coefficient of variation of the IRIs,CViri. Dotted linesrepresent

the squared inverse of the coefficient,1CV 2

iri(which is the term which appears in the Fisher

information). Colors of the inset applies to all plots and all lines types. Parameters:ν = 10

Hz, τc = 50 ms,U = 0.5, N0 = 1.

reduced with respect the input spike train. For a Poisson input,CViri reaches its minimum

valueCViri = 1√2

and then tends asymptotically to one, asτv is increased (2.6). The same

qualitative behavior occurs when the input has correlations of short range (τc ≤ τv), namely

CViri falls down to a minimum and then increases towards one (see figure2.6). Therefore,

as far as variability is concerned, the optimal estimation would be achieved at theτv where

CViri reaches its minimum, because there the output train is the most regular.

The question now is, what is the outcome of this sort ofpull and pushmechanism in

which saturation pulls theτopt towards zero, while regularization pushes it towards a finite

non-zero value? The answer depends critically on the amount and time scale of the input

correlations. The explanation of this dependency is depicted in figure3.5. In order to better

illustrate the action of both saturation and elimination of variability, we have isolated the

effect of the first by thinking about the following example:

Page 82:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

68 Chapter 3: Information transmission through synapses with STD

• Poissonresponsetrain. Let us suppose that theresponsestatistics are always Poisson

with rate νr, i.e. the same saturating function we have derived in eq.2.3.8. We

are going to plot the Fisher information contained in the train of synaptic responses

about the input rateν, but assumingthat the output is Poisson. In this way, we may

separate the effect of saturation, and later include the effect of the variability. The top

plot in fig. 3.5 displays the Fisher information per response (dotted lines). Different

colors represent again different values of theCVisi. Because this case is by hypothesis

Poisson, both codes, timing and counting, coincide. As expected for this case, thepull

and pushcompetition, in which just one player (saturation) is activated, is resolved

by an optimalτopt = 0. For any value of the inputCVisi the information decays

monotonically asτv increases. The factor∂νr

∂ν, responsible of the information loss with

τv, is shown in the middle plot of fig.3.5. It decreases monotonically fromU towards

zero. Summarizing, in this hypothetical case, any amount of depression, that isτv > 0,

increases the estimation error of the rate.

Including the effect ofCViri, yields a quite different result, as it showm in figure3.5 (top

plot). Different colors represent different magnitudes of the input correlations, which we

now examine separately.

• Poisson input.When the input is a Poisson train, the maximal information is achieved

at τopt = 0 for both codes. Still, the information is always larger than in the example

previously considered, where theresponseswere Poisson, becauseCViri is smaller

than one for any value ofτv (see black solid line in the bottom plot). This means

that the train is more regular than Poisson and consequently the estimation is better.

CViri is plotted againstτv in the bottom plot of the figure together with1CV 2

iri(dotted

lines). This is exactly the factor that appears in the information quantitiesJsr(ν|n)

andJ(ν|n). Thus, we conclude that in spite of the fact that the output train is maxi-

mally regular forτv 6= 0 (∼ 200 ms), saturation effects dominate andno compromise

between the two is obtained: the static synapse still performs better.

• Correlated input. When the input has aCVisi > 1, the Fisher information per re-

sponse, in both coding schemes, shows a maximum for a finiteτopt > 0. The explana-

tion of this can be read out from the behavior of theCViri (bottom plot). Forτv = 0

(static synapse),CV 2iri = (CV 2

isi − 1)U + 1 (eq. 2.3.11). Hence the variability of the

responses for smallτv is high, reflecting a correlated input withCVisi > 1. This makes

the Fisher information smaller than in the Poisson case, whenτv ∼ 0. As τv increases,

the response train becomes rapidly more regular untilCViri reaches a minimum, which

is very close to the minimum for the input Poisson case. As a consequence, thereduc-

tion of variability, defined as the difference between the maximum value atτv = 0 and

Page 83:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 69

the minimum, is now bigger. Because of this more severe elimination of variability, the

Fisher information presents a maximum for aτopt > 0. It is interesting to notice that

although the effect of the regularization of the response train is strong,τopt is still be-

low the position of the minimum ofCViri. The reason is that saturation ispulling τopt

down towards zero so that, at the end, a compromise between the two effects arises.

Finally, because for stronger correlations saturation occurs at higher values ofτv, the

caseCVisi = 2 (blue lines) shows a biggerτopt that the caseCVisi = 1.5 (red lines).

The discussion above holds for both codes. The reason is that, as we saw in the previous

section, both coding schemes behave surprisingly similar in the range where saturation is not

dominant, i.e.τvνU 1 (compare solid and dashed lines in the top plot of fig.3.5). Since

the maximization of the Fisher information takes place in a non-saturating regime (i.e.τopt

< τsat), both codes display quantitatively almost the same result, that is, both give the same

τopt .

After analyzing the Fisher information per response, we plot in figure3.6 (bottom plot)

the Fisher information per unit time. The ratio between both is simplyνr, namelyJ = νrJsr,

whereνr is a decreasing function ofτv (remember thatνr is constrained by0 < νr <1τv

). Thus, unless the maximum is very prominent in the information per responseJsr, the

information per unit timeJ is a monotonically decreasing function ofτv , meaningτopt = 0.

In conclusion, unless the stimulus is a very irregular spike train (CVisi ∼ 3 meaning a bursty

trains with many impulses per burst, data not shown), in what regards the reconstruction

error per time unit, synaptic depression is a disadvantage.

3.3.2.2 Optimizingτv with the Mutual Information

We now study the dependency of the Shannon information onτv . We take as the p.d.f of

the input stimulus, an exponential function, with a lower cut-offνinf = 1 Hz, which reads

f(ν) = e−ν/ν

ν. The mutual (middle plot) and Fisher (top and bottom plots) informations are

plotted together in figure3.6to compare their dependence onτv. This is done by settingν, in

the Shannon case, equal to the rateν in the Fisher case. Despite the absolute values (which

correspond obviously to different units),I(∆; ν) follows very well the behavior of the Fisher

information per response: the maximum occurs for a non-zeroτopt only when the input is

correlated and the values ofτopt are approximately the same.

Because of this analogous behavior of the Shannon and Fisher information per response,

when computing the information per unit time, the result is also qualitatively the same to

the Fisher information per time unit (data not shown): unless correlations are very large, the

information per time unit decreases monotonically, meaningτopt = 0.

Page 84:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

70 Chapter 3: Information transmission through synapses with STD

0

0,002

0,004

0,006

0,008

0,01

J pe

r re

spon

se [

s-2]

0

0,05

0,1

0,15

0,2

0,25

I(∆;

ν)

[bits

] CV= 1CV= 1.5CV= 2

0 0,1 0,2 0,3 0,4 0,5τ

v [s]

0

0,01

0,02

0,03

0,04

0,05

J pe

r se

cnod

[s-2

]

Figure 3.6: Optimization of the recovery time constantτv , and comparison of the Fisher

information per responseJsr, per unit timeJ and the mutual informationI(∆; ν) . Top:

Solid linesrepresentJsr(ν|∆) (the Fisher information given the length of an IRI).Dashed

lines representJsr(ν|n) (counting code) as a function ofτv . Middle: Mutual information

I(∆; ν) . Bottom: Fisher information per second:Solid linesrepresentJ(ν|∆) anddashed

linesrepresentJ(ν|n). Colors of the inset applies to all plots and all lines types. Parameters:

ν = 10 Hz, τc = 20 ms,U = 0.5, N0 = 1.

It is well known property [Rieke et al., 1996, Dayan and Abbot, 2001], that for a given

renewalprocess with a fixed mean rate, the inter-event-interval distribution that maximizes

the entropy of the train is an exponential, that is, a Poisson process. This result can also be

phrased in the following way: any correlations between the time of the events decrease the

entropy with respect to a Poisson process. However maximizing the mutual information is

not equivalent to maximize the entropy of the response. If we rewrite the expression of the

mutual information of eq.3.2.8

I(∆; ν) = H(∆)− 〈H(∆|ν)〉ν (3.3.9)

we see the the maximization ofI(∆; ν) requires a compromise between the maximization

of the differential entropy of the responseH(∆) but without allowing the noise entropy

〈H(∆|ν)〉ν to get too big. Whenever the synapse enters in the saturation regime, the distri-

bution of the responses begins to be dominated by the vesicle recovery dynamics, and thus

to be independent of the input rate. When this occurs, the noise entropy〈H(∆|ν)〉ν equals

Page 85:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 71

the response entropyH(∆) (i.e. when the input signal,ν, is fixed, the response contains

the same variability than when the whole ensemblef(ν) is considered), and the information

vanishes. In other words, no matter the magnitude of the response entropyH(∆): saturation

always makes the information fade away.

Figures3.7 and3.8 show, together with the mutual information (middle plot), the two

entropies,H(∆) and〈H(∆|ν)〉ν (top plot), and the efficiency of the transmission (bottom

plot) as a function ofτv for a correlated input (CVisi = 1.5). The efficiency is defined as

e =I(∆; ν)

H(∆)(3.3.10)

Again we have computed the hypothetical example in which the response output is “forced”

to be Poisson with a rateνr(ν).

The mutual information is defined as the distance between the two entropies. Hence,

we observe in fig.3.7 that although both entropies grow monotonically withτv the distance

between them varies. In the case of Poisson output, the only observable effect is that the

noise entropy slowly approaches the response entropy, leading to a monotonous decrease of

the mutual information and the efficacy (red dashed lines in middle and bottom plots). In the

un-constrained case, for largeτv both terms also approach each other, making the information

decrease to zero. For very smallτv, when there is barely any depression and the statistics of

the output are dominated by the input, the response entropyH(∆) becomes much smaller

than the equivalent term in the Poisson output case. This happens because, in this range,

the output has large positive correlations inherited from the input and falls far from its upper

Poissonian bound. Whenτv ∼ 70 ms,H(∆) saturates this upper bound because depression

introduces negative correlations that cancel the positive correlations of the input. It is in this

range where the mutual information reaches its maximum. In fig.3.8 the same plots have

been drawn with a larger scale inτv . In the top plot we can check that for very large values

of τv , H(∆) again saturates its upper bound. This is indeed due to de-correlation as well,

but in this case no maximization of the information is achieved because both entropiesH(∆)

and〈H(∆|ν)〉ν have converged to the same value.

3.3.2.3 Several vesicles:N0 ≥ 1

After proving that a synapse with a single vesicle can be tuned to recover with an optimal

rate 1τopt

, we will consider now a synapse with several docking sites,N0. As we described in

section2.3.2of chapter2, N0 represents the maximum number of vesicles that can be ready

for release, i.e. the size of the RRP is alwaysN ≤ N0. Since we are assuming that at most

one vesicle fuses the membrane upon arrival of a spike [Edwards et al., 1976, Triller and

Page 86:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

72 Chapter 3: Information transmission through synapses with STD

0 0,05 0,1 0,15 0,2τ

v [s]

0,015

0,02

0,025

0,03

effi

cacy

0,15

0,2

0,25

0,3

I(∆;

ν) [

bits

]

-0,8

-0,6

-0,4

-0,2

0

0,2

0,4

0,6

Ent

ropy

H(∆)<H(∆|ν)>νH(∆) ouput Poisson<H(∆|ν)>ν ouput Poisson

Figure 3.7: Analysis of the mutual informationI(∆; ν) and entropiesH(∆) and〈H(∆|ν)〉νas a function ofτv . In the three plots two examples with the same parameters are plotted:

dashedrepresent the “forced” Poisson (see text) whilesolid represent the same case but

considering the variability of the responses.Top: EntropyH(∆) and conditioned entropy<

〈H(∆|ν)〉ν > of the inter-response-interval.Middle: Mutual InformationI(∆; ν) . Bottom:

Efficacy of the transmission defined in eq.3.3.10. Parameters:ν = 10 Hz, τc = 20 ms,

CVisi = 1.5, U = 0.5, N0 = 1.

Page 87:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 73

0 0,2 0,4 0,6 0,8 1 1,2 1,4τ

v [s]

0

0,01

0,02

0,03

0,04

effi

cacy

0

0,1

0,2

0,3

0,4

I(∆;

ν) [

bits

]

0

1

2

3E

ntro

py

H(∆)<H(∆|ν)>νH(∆) ouput Poisson<H(∆|ν)>ν ouput Poisson

Figure 3.8: Same representation as in figure3.7but showing a different range ofτv.

Korn, 1982, Stevens and Wang, 1995], the probability of release is defined by a function of

the number of docked vesicles (eq.2.2.1):

pr(N) = UΘ(N − 1) N = 0, . . . , N0 (3.3.11)

We want to address now the question of wherther a synapse whereN0 > 1 transmits in-

formation better, worse or in the same manner as a single docking site synapse. The presence

of several vesicles ready for release has two important consequences:

1. There is a quantitatively different transfer functionνr(ν) which saturates latter (νsat

grows withN0) and to a larger value, the larger isN0. By analytic arguments we ar-

gued in section2.3.2that if the stationary probability thatN > 1 is small compared

with the probability thatN = 1, then the synaptic contact behaves effectively like

a one vesicle synapse but with recovery time constantT = τv

N0, that is, the effective

uniquevesicle recoversN0 times faster. For this reason, two scenarios for comparing

synapses with different number of docking sites can be considered. The first compari-

son is the straightforward one, where we just add docking sites to a synapse and keep

Page 88:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

74 Chapter 3: Information transmission through synapses with STD

the recovery time of each site,τv, constant. In this scenario, the number of releases

per second, that isνr, will increase as we introduce docking sites (see fig.2.10). The

second possible situation consists in a comparison where, as one includes more dock-

ing sites, the maximal response rate is kept constant. This is achieved by changingτv

each timeN0 is varied, so that

τv

N0

= T = constant (3.3.12)

whereT is a time constant whose inverse sets the maximal possible value of the re-

lease rate,νr < 1T

. In this way, we would compare a synapse with one docking site

replenished with rate1τv

with another with two sites each recovering with a slower rate1

2τv. Renormalizingτv in this manner does not make the response rates for different

N0’s equal, but just equals the range ofνr.

2. The elimination of variability does not occur with the same efficiency. ForN0 = 1,

the process of regularization of the train (i.e. the reduction ofCVout = CViri with

respect toCVin = CVisi) is based on the incapability of the single site synapse to

trigger consecutive responses when the spikes come very close together (e.g. in bursts).

Because there is only one release site, after a vesicle undergoes exocytosis, some finite

time is needed to refill the site with a new vesicle. During this time, the rest of the

spikes in the burst find the terminal empty and are filtered away. If insteadN0 > 1,

there is a non-zero probability that two or more spikes in a burst elicit a response. This

makes the regularization of the input train, or phrased in another way, the reduction of

correlations present in the incoming spikes, less efficient. Thus, those IRIs produced

by spikes within a burst will increase the variability of the train of responses making

the estimation worse. Comparing the shape of the distribution of IRIs whenN0 is one,

and when is larger (figure2.9), it can be observed that, only in this latter case,ρiri(∆|ν)

shows a skewed high peak at the origin, which reflects the bursty nature of the input,

and accounts for the probability that two spikes within a burst elicit a response.

Comparison of different N0’s at τv fixed

In this case the range of the output rateνr increases withN0 as0 < νr < N0

τv. This

implies that saturation is pushed upwards to larger values of the inputν (or τv). ForN0 > 1,

more than one vesicle can be ready for release and theshowerof afferent spikes has to be

more “intense” (higher rate) to keep the synapse depleted, that is, in saturation. This effect

was shown in fig.2.10, where the arrows in thex-axis indicate the value of the saturation

rate,νsat. In terms of information transmission, boosting the position ofνsatincreases the

Page 89:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 75

0,002

0,004

0,006

0,008J sr

(∆|ν

) [s

²]N

0=1

N0=2

N0= 3

N0= 4

N0= 5

0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1τ

v [s]

0

2

4

6

8

10

ν e [hz]

Figure 3.9: Optimization ofτv regarding the Fisher information per response, for different

values of the readily releaseable pool sizeN0. Top: Fisher Information per response for a

time output code,Jsr(ν|∆), as a function ofτv , for several values ofN0. Bottom: Response

rateνr versusτv . Inset values apply to both plots. Parameters:ν = 10 Hz, CVisi = 1.8,

τc = 10 ms,U = 1.

ability of the channel to transmit higher input rates. In figure3.9 (top plot) we have plotted

theFisher information Jsr(ν|∆) as a function ofτv, for several values ofN0, when the input

is correlated (CVisi = 1.8). For high values ofτv, the cases whereN0 is small, i.e. one or

two, feel the saturation and the information falls to zero, whereas for largeN0, i.e. four or

five, the information remains almost constant (in the range ofτv shown). Moreover, in these

latter cases a maximum is hardly detected, because the bursts are poorly filtered away and

no regularization of the train occurs. On the other hand, whenN0 = 1, 2, if τv goes from0

to 100 ms, firstly, the response rate drops down reflecting that many spikes are being filtered

(bottom plot); secondly, because of this filtering, the variability is severely reduced and the

information shows a maximum (top plot).

What happens to themutual information ? The result, displayed in the top plot of figure

3.11, is qualitatively the same. However, the caseN0 = 1 seems to stand apart from the cases

N0 > 1. More specifically, the systematic behavior (observed in the Fisher information),

Page 90:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

76 Chapter 3: Information transmission through synapses with STD

0,002

0,004

0,006

0,008

J(∆|

ν) [s

²]

N0=1

N0=2

N0= 3

N0= 4

N0= 5

0 0,05 0,1 0,15 0,2 0,25 0,3 0,35 0,4 0,45 0,5Τ = τ

v/N

0 [s]

0

2

4

6

8

10

ν e [hz]

Figure 3.10: Optimization ofT = τv

N0regarding the Fisher information per response, for

different values of the readily releaseable pool sizeN0 (whereτv in each case varies such

thatτv = N0 T .) Top: Fisher Information per response for a time output code,Jsr(ν|∆), as

a function ofT , for several values ofN0. Bottom: Response rateνr versusT . Inset values

apply to both plots. Parameters:ν = 10 Hz, CVisi = 1.8, τc = 10 ms,U = 1.

namely the displacement of the maximum towards higher values ofτv, and the decline of the

information at the maximum, occurs only forN0 ≥ 1. Moreover, the optimization of the

Shannon information is achievable for higher values ofN0 than for the Fisher information.

This can be checked by comparing the instancesN0 = 5 of figures3.9 and3.11: while the

Fisher information shows a negligible maximum, in the mutual information the bump is still

prominent.

In conclusion, if the synapse is able to optimize the biophysical parameterτv, then it

would be more convenient to set just one (or two for the mutual information) docking site,

than to provide the synapse with many. If on the contrary,τv is fixed at a large value (e.g.

∼ 1 s.), then including more docking sites would increase the estimation performance and

the information transfer, particularly for high input rates.

Page 91:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 77

0.05 0.1 0.15 0.2

T [s]

0.1

0.15

0.2

0.25

0.3

I [b

its]

0 0.1 0.2 0.3 0.4 0.5τ

v [s]

0.1

0.15

0.2

0.25

0.3

I [b

its]

N0= 1

N0= 2

N0= 3

N0= 4

N0= 5

Figure 3.11: Optimization ofτv andT = τv

N0regarding the mutual informationI(∆; ν) .

Top: InformationI(∆; ν) versusτv for different values of the readily releaseable pool size

N0. Bottom: InformationI(∆; ν) as a function ofT , for several values ofN0 (whereτv in

each case varies such thatτv = N0 T .). Inset values apply to both plots. Parameters:ν = 10

Hz, CVisi = 1.8, τc = 10 ms,U = 1.

Comparison of different N0’s at T = τv

N0fixed

Figure3.10(top plot) shows theFisher information as a function ofT , in this different

comparative scenario. Indeed, this plot was obtained by rescaling thex-axis of fig. 3.9by a

factor 1N0

(for eachN0 case). The addition of more docking sites now gives little advantage

in terms of diminishing the saturation effects (the bottom plot illustrates the response rate

νr for different values ofN0). In all cases,νr converges to the same limit frequency,1T

, as

T is increased. The top figure shows that all the maxima arealignedand become less and

less prominent asN0 is increased. This is consistent with what we just saw: the reduction

of variability is less efficient as more vesicles can be dockedat the same time. Thus the

beneficial effect of depression, as a correlation filter, is not performed in such a proficient way

anymore. The fact that all the maxima are aligned can be traced back to the decomposition

of the statistics of the IRI made in previous chapter (see eq.2.3.13): the multiple sites case

Page 92:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

78 Chapter 3: Information transmission through synapses with STD

0 10 20 30 40 50

ν [hz]

0

0,2

0,4

0,6

0,8

1

1,2

1,4

τ opt [

s]

τc= 100ms

τc= 50ms

τc= 20ms

τc= 5ms

cv=1.8, U=0.5, N0= 1

0 10 20 30 40 50

ν [hz]

cv= 1.2cv= 1.5cv= 2cv= 2.5

τc= 50ms, U=0.5, N

0= 1

Figure 3.12: Optimal recovery timeτopt as a function of the input rateν, for different magni-

tudes and temporal scales of the input correlations. Parameters:CVisi = 1.8 (left), τc = 50

ms (right),U = 0.5, N0 = 1.

with a renormalized recovery time constant, behaves like the single release site model plus a

perturbative term (which vanishes in the saturation regime), that accounts for the probability

of more than one vesicle being ready at the same time.

The bottom plot in figure3.11, shows theShannon information as a function ofT .

The outcome again follows a behavior similar to that of the Fisher information, that is, all

maxima are aligned and less prominent asN0 is increased. However, the caseN0 = 1 aligns

out of this general trend: the optimalT is larger andI(∆; ν) at the maximum is smaller than

for N0 = 2, 3. In this sense, the mutual information seems to be more sensitive than the

Fisher information to this qualitative difference, namely, that forN0 > 1, there is a non-zero

probability of having an infinitively small IRI.

3.3.2.4 Dependence ofτopt on the other parameters

In this section we analyze the dependence of the optimal recovery time constantτopt on

the rest of the parameters, and we will try to bound it quantitatively taking into consideration

Page 93:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 79

0

0.1

0.2

0.3

0.4

0.5

τ opt (

s)

N0= 1

N0= 2

N0= 3

N0= 4

ν= 10 Hz; ; U= 1

0

0.1

0.2

0.3

0.4

0.5τ op

t (s)

1 1.25 1.5 1.75 2 2.25 2.5

CV

0

0.1

0.2

0.3

0.4

0.5

τ opt (

s)

τc= 10 ms

τc= 25 ms

τc= 50 ms

Figure 3.13: Optimal recovery timeτopt of the Fisher information per response, as a function

of the inputCVisi, for different number of docking sitesN0. Colors in the inset apply for all

plots. Parameters not indicated in the figure:ν = 10 Hz, U = 1

the natural ranges of the other parameters.

In a first approximation, what sets the scale ofτopt ? The order of magnitude ofτopt is

determined by the renormalized input rate,νd = Uν. The rest of the parameters also affect

the value ofτopt but the changes are small. Figure3.12 shows theτopt which maximizes

Jsr(ν|n) as a function ofν, for several values of the inputCVisi andτc. The dependence is

close to an hyperbolic relation. When a non-linear regression was performed, the relation in

all cases was

τopt =C

να(3.3.13)

whereC is a constant that depends in the other input parametersCVisi and τc, whereas

α ∼ 0.75 for all the examples. Because of this dependence, if one assumes that the firing

rate range in which information is transmitted in the brain lays from around10 to 100 Hz,

and takeU from 0.1 to 0.95 [Markram, 1997, Murthy et al., 1997], the range ofτopt obtained

is: 0 < τopt < 200 ms. In order to obtain higher values ofτopt lower input rates must

be considered [de la Rocha et al., 2002]. It is interesting to mention here thatFuhrmann

et al. [2002] find by optimization of the information (using different input-output codes) a

relationτopt = 1Uν

. In the last section of the chapter we will discuss the compatibility of this

result with the experimental data. The dependence ofτopt on the input correlation magnitude

(that is, the inputCVisi) is depicted in figure3.13. As we know, the correlations pushed the

Page 94:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

80 Chapter 3: Information transmission through synapses with STD

0

0,002

0,004

0,006

0,008

0,01

J pe

r re

spon

se [

s-2]

0

0,05

0,1

0,15

0,2

0,25

I(∆;

ν)

[bits

] CV= 1CV= 1.5CV= 2

0 0,1 0,2 0,3 0,4 0,5τ

v [s]

0

0,01

0,02

0,03

0,04

0,05

J pe

r se

cnod

[s-2

]

Figure 3.14: Optimal recovery timeτopt of the Fisher information per response, as a function

of the time scaleτc of the input correlations, for different values of theCVisi (left), and

different number of docking sitesN0 (right). Parameters left:ν = 10 Hz, U = 0.5, N0 = 1.

Right: ν = 10 Hz, U = 1, CV = 1.8.

saturation up towards higher values. At the same time, increasingCVisi makes the response

coefficient of variationCViri reach its minimum for a higherτv. Both effects cooperate to

increase the value ofτopt asCVisi increases. In the same figure, it can be observed that, if

N0 > 1, having a correlated input is not a sufficient condition to findτopt > 0 (what we will

call to besubject to optimization). TheCVisi-threshold, over which the synapse is subject to

optimization, grows withN0 and decreases with the correlation timeτc. Thus, if the synapse

has many docking sites, it can only be optimized if the input is a very bursty train, which

implies short correlation range (τc 1ν) and large positive correlations (CVisi ∼ 2). The

variability (or equivalently the auto-correlation) present in the input is filtered in a more

efficient manner if the temporal scale of the correlations is small. This points in the direction

that short term depression is best suited to transmit information conveyed in spike trains

made up of bursts. In figure3.14the relation betweenτopt andτc is explicity analyzed: the

top plots representτopt as a function ofτc , while the bottom plots show the ratio between

the informationτopt and atτv = 0. This ratio gives a quantitative measure of how much

advantageous is to have depression tuned to its optimal value (τv =τopt ), as opposed to a

static synapse with no depression (τv = 0). Solid lines in the four plots represent theτopt

when the Fisher information was considered, whereas dashed lines apply for theτopt of the

Page 95:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 81

0 1 2 3 4 5 6 7N0

0

0,05

0,1

0,15

0,2

0,25

τ opt (

s)

τc= 100 ms

τc= 50 ms

τc= 25 ms

τc= 10 ms

τc= 4 ms

ν= 10 Hz; CV= 1.8; U= 1

0 1 2 3 4 5 6 7N

0

1

1,2

1,4

1,6

1,8

2

J(τ v=

τ opt)

/ J(τ

v=0)

Figure 3.15:Top: Optimal recovery timeτopt of the Fisher information per response, as

a function of the number of docking sitesN0, for different values ofτc . Bottom: Ratio

between the Fisher information at the maximum and at the origin, i.e.Jsr(τv=τopt)Jsr(τv=0)

. Colors in

the inset apply for both plots. Parameters:ν = 10 Hz, CV = 1.8, U = 1

.

mutual information. In the first place, we notice that the range ofτc over whichτv can

be optimized is clearly upper bounded. This happens because the height of the maximum

decreases continuously withτc . Roughly speaking, the filter time windowτopt must grow

with the temporal extent of the correlations which need to be removed. But ifτc keeps

increasing, we need to increaseτopt beyondτsat, so that the information drops to zero. This

is shown in the left top plot of fig.3.14. For this reason, when the input has a largerCVisi, the

upper bound ofτc , where saturation starts, is larger due to the dependence of the saturation

regime on the inputCVisi . On the contrary, adding docking sites has the opposite effect, that

is, it shrinks the range ofτc where the input is subject to optimization.

Finally, we have plotted the dependence ofτopt on the parameterN0. The τopt which

optimizes the Fisher information is represented in figure3.15while the case of the mutual

information is drawn in fig.3.16. In both figures, the bottom plot represents the ratio of

the information atτopt and atτv =0. As it has been previously discussed, increasingN0

decreases the capability of optimizingτv , as well as the advantage of having depression at

its optimal tuning over a static synapse. Eventually, aN0 is reached where depression does

not constitute any advantage anymore. Ifτc is large (∼ 100 ms) this maximalN0 is reached,

for the parameters chosen in fig.3.15, already atN0 = 1. As we temporally bound the

Page 96:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

82 Chapter 3: Information transmission through synapses with STD

0

0,05

0,1

0,15

0,2

τ opt [

s] cv=1.5cv= 2

ν=10, τc= 20 ms, U=1

0 1 2 3 4 5 6 7

N0

1

2

3

4

I(τ v=

τ opt)/

I(τ v=

0)

Figure 3.16: Top: Optimal recovery timeτopt of the mutual informationI(∆; ν) , as a

function of the number of docking sitesN0, for different values of the inputCV . Bottom:

Ratio between the mutual information at the maximum and at the origin, i.e.I(τv=τopt)I(τv=0)

.

Colors in the inset apply for both plots. Parameters:ν = 10 Hz, τc = 20 ms,U = 1

.

correlations, i.e. decreaseτc , the maximalN0 becomes bigger. In the case of the mutual

information thislimiting effect occurs for higher values ofN0.8

3.3.2.5 Metabolic considerations

As described in section3.2.4, we study now the effects of considering the metabolic

consumption in the optimization of the information transmitted. To do this, we use the energy

function proposed in equation3.2.26to account for the metabolic expenditure produced in

the recovery of vesicles. This energy function depends linearly on the recovery rate as

E(τv) =β

τv

(3.3.14)

The information about the input rate per unit energy and per response reads (eq.3.2.27)

I(∆; ν) =I(∆; ν)

E(τv)= β I(∆; ν) τv (3.3.15)

8Solving our model requires to work out analytically theN0 × N0 linear system of eq.2.2.13. Due to

computational constraints (basically, the calculation exceeded the RAM memory of the computer), we could

not compute the solution of the model for values ofN0 beyond 7, but one can tell by continuity arguments what

the result for higherN0’s would be.

Page 97:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 83

0

0.1

0.2

0.3

Info

. per

Ene

rgy

unit

CV= 1CV= 1.5CV= 2

0 0.2 0.4 0.6 0.8 1 1.2 1.4

τv [s]

0

0.5

1

1.5

Info

. per

tim

e un

it pe

r E

nerg

y un

it

Figure 3.17: Optimization of the recovery time constantτv regarding the mutual informa-

tion per unit energyI(∆; ν) . Top: Dashed linesrepresent the mutual informationI(∆; ν)

(without metabolic considerations) whilesolid linesrepresent the information per unit en-

ergyI(∆; ν) (see eq.3.3.15). Bottom: Dashed linesrepresent the information rateR(∆; ν) ,

while solid linesrepresent information per unit time per unit energy,R(∆; ν) τv. Both quan-

tities in both plots are shown in arbitrary units and have been normalized to have similar

overall scale. Inset values apply for both both plots and for solid and dashed lines. Parame-

ters: ν = 10 Hz, τc = 20 ms,U = 0.5, N0 = 1.

We have plottedI(∆; ν) versusτv in figure3.17(top plot, solid lines) along with the in-

formationI(∆; ν) (dashed lines), for the same values chosen in the previous figure3.6. Both

quantities are illustrated in arbitrary units and have been normalized so that their magnitudes

are similar (in these arbitrary units).

The first thing which must be noticed is thatI(∆; ν) is zero whenτv = 0, for any value of

the inputCV . Besides, the optimalτopt which maximizesI(∆; ν) is not optimal forI(∆; ν)

in any case. On the contrary,I(∆; ν) reaches its maximum value for a largerτv, and seems

to saturate at that maximum for higherτv. However, it is hard to determine numerically the

limit value ofI(∆; ν) asτv goes to infinity, because it is the product of a vanishing integral,

which is computed numerically, multiplied byτv which tends to infinity. Therefore, we have

to conclude that this particular result is not veryrobust to the election ofE(τv): any other

power of the rate1τv

smaller than one would have lead to a vanishingI(∆; ν) at the limit

τv →∞

In the bottom plot of figure3.17 (solid lines) the information rate per unit energy, i.e.

Page 98:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

84 Chapter 3: Information transmission through synapses with STD

0

0.2

0.4

0.6

0.8

1

J pe

r re

spon

se p

er E

nerg

y un

it

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

τv [s]

0

0.2

0.4

0.6

0.8

1

J pe

r tim

e un

it pe

r E

nerg

y un

it

CV= 1CV= 1.5CV= 2

Figure 3.18: Optimization of the recovery time constantτv regarding the Fisher informa-

tion per unit energyI(∆; ν) . Top: Dashed linesrepresent the Fishermutual information

Jsr(ν|∆) (without metabolic considerations) whilesolid linesrepresent the ratioJsr(ν|∆)E(τv)

,

meaning the reconstruction error per response and per energy unit.Bottom: Dashed lines

represent the Fisher information per secondJ(ν|∆) while solid linesrepresentJ(ν|∆)E(τv)

, mean-

ing the reconstruction error per second and per energy unit. Both quantities in both plots

are shown in arbitrary units and have been normalized to have similar overall scale. Inset

values apply for both both plots and for solid and dashed lines. Parameters are as in fig.3.17:

ν = 10 Hz, τc = 20 ms,U = 0.5, N0 = 1.

R(∆;ν)E(τv)

, is depicted as a function ofτv, along withR(∆; ν) , for comparison purposes (both

quantities have been normalized in arbitrary units). As opposed to the behavior ofR(∆; ν)

which decreases monotonically, the information rate per unit energy shows a maximum for

values ofτv between200 and400 ms, depending on the correlations of the input, that is, the

CV . Regardless of which quantity, namelyI(∆; ν) or R(∆;ν)E(τv)

, we are optimizing, it is clear

that considering the energetic cost of the recovery process, by computing the information

per unit energy, punishes small values ofτv, which are costly, and pushes the optimalτopt

towards higher values in the range of several hundreds. Other choices of the dependence of

E(τv) on τv (e.g. logarithmic, quadratic,...) would have produced different curves but, as

long as they are monotonically increasing functions ofτv, their qualitative result would be

the same: priming higherτv values over low ones. Notice that, with the choice made here,

no extra parameter was required to obtain this higherτopt .

In figure 3.18 we have plotted the Fisher information per response (top plot) and per

Page 99:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 85

unit time (bottom plot) divided by the energy functionE(τv). In both cases, the information

per unit energy shows a maximum for a value ofτv higher than the one obtained with no

metabolic cost constraint. In particular, a non-zeroτopt is now obtained even for the case

in which the input is Poisson (something that never happened when no energy considera-

tions were taken into account). The new values ofτopt range, for the parameters chosen in

this example, from150 ms to400 ms, again in agreement with the result from the mutual

information.

The maximization ofI(∆; ν) performed in previous sections resulted in an optimalτopt

ranging in50 − 150 ms. Now, the introduction of a metabolic cost constraint changes this

result quantitatively makingτopt equal to200−600 ms. As mentioned above, these last values

are indeed already in the range of the values observed in neocortical synapses [Markram,

1997, Varela et al., 1997, Markram et al., 1998a, Finnerty et al., 1999, Varela et al., 1999,

Petersen, 2002].

3.3.3 Optimization of the release probabilityU

In this section we will analyze the relevance of the synaptic parameterU , which rep-

resents the probability of release when the RRP is not empty. We will again address the

question of whether it is possible to maximize the information by tuning the value ofU . The

optimal value which maximizes the information will be denoted byUopt . We will start by

exploring the Fisher information, which in some cases can be analytically maximized, and

latter we will focus in the Shannon information.

3.3.3.1 OptimizingU with the Fisher Information

Let us start studying the Fisher informationper response, Jsr. A naive approach tells us

that whenU approaches zero, the information should decrease to zero because in that limit

synaptic channel is “closed”. In these circumstances, it seems that no estimation of the input

rateν can be accomplished by the post-synaptic cell. However, as it will be derived now, the

limit of Jsr whenU → 0 is not zero but a finite value.

As already shown in previous sections, both coding strategies, counting and timing, re-

sult in almost the same Fisher information (only in the saturation regime the differences are

appreciable). Therefore, we will use the analytical expression forJsr(ν|n), the Fisher infor-

mation per response given the total number of responses, to explore the dependence onU .

Let us study first a Poisson input and later, we will consider the correlated situation. Thus,

whenCVisi = 1, Jsr(ν|n) equals (see eq.3.2.5),

Jsr(ν|n) = 1CV 2

iriν2r

(∂νr

∂ν

)2=

[(1 + νUτv)

ν2U2[1 + (νUτv)]

] [U2

(1 + νUτv)4

]

Page 100:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

86 Chapter 3: Information transmission through synapses with STD

0

0,5

1

1,5

2

ν r [hz

]

U=1, CV=1U=0.1, CV=1U=1, CV=2U=0.2, CV=2

CV=1, τ= 500 ms

0 10 20 30 40 50

ν [hz]

0

0,05

0,1

0,15

0,2

0,25

0,3

dνr/d

ν

Figure 3.19: Analysis of the transfer functionνr(ν) for two values of the release probability

U . Top.- Black lines: response rateνr as a function of the input rateν when the input is

Poisson andU = 1 (solid) andU = 0.1 (dashed).Redlines: the same but for a correlated

input (CV = 2). Brown straight lines: transfer function when the recovery time constant

τv equals zero (νr = Uν), for two values of the release probability :U = 1 andU = 0.1.

Bottom.- Derivative of the transfer function,νr(ν)′, for the cases depicted above (except

whenτv = 0). Greenvertical lines mark the value ofν at whichνr(ν; U = 1)′ = νr(ν; U =

0.1)′. Whenν is beyond the green line the changeU = 1 → U = 0.1 makes the slope of

the transfer function increase, while ifν is below the green line, the change makes the slope

decrease. Inset values apply for both plots. Parameters:τc = 20 ms,τv = 500 ms,N0 = 1.

= 1ν2[1+(νUτv)2]

(3.3.16)

One might think that the Fisher information should tend to zero asU → 0 because the

transfer functionνr(ν) becomes completely flat and equal to zero. However, when asU → 0

the decreasing behavior of the derivative∂νr

∂ν, which is the factor that captures this intuitive

idea, is compensated by the variance factorCV 2iriν

2r because both go to zero asU2. For this

reason, we consider that the optimization ofJsr(ν|n) by tuningU is not well defined because

it shows a paradoxical behavior asU → 0. Moreover,U = 0 turns out to be the optimal

value to estimateν (data not shown). To claim that the cell would optimizeU following

the maximization of the quantityJsr(ν|n) and thus making the synapse completely useless,

seems not very reasonable, so we will disregard this criterion. What happens if we consider

the Fisher informationper secondJ(ν|n)? The expression ofJ(ν|n), when the code is the

Page 101:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 87

counting of responses and the input is Poisson, reads (eq.3.2.6)

J(ν|n) = Jsr(ν|n)νr =U

ν[1 + (νUτv)2](1 + νUτv)(3.3.17)

By inspection, it is clear thatJ(ν|n) does converge to zero asU → 0. The question now is,

why could it be advantageous that the release were unreliable, i.e.U < 1? If the synapse is

static (τv = 0 and therefore no depression at all) the Fisher information for a Poisson input

equalsJ(ν|n) = U/ν. Thus,J(ν|n) grows linearly withU , a behavior that can be traced

back to the transfer function, which in the static case is simplyνr = Uν. U would be playing,

therefore, the role of the gain. As a consequence, the larger the release probability, the larger

the gain and the better represented is the inputν in the outputνr, making the estimation more

accurate.

Nevertheless, whenτv 6= 0, the transfer function is not linear anymore but it saturates

to 1τv

. Figure3.19illustrates the functionνr(ν) (top plot) and its derivate (bottom plot) for

two values ofU , for Poisson input (solid lines) and for a bursty input (CVisi = 2, dashed

lines). The release probabilityU still controls the gain but only at the origin, that is, when

ν = 0 (this is shown in the figure by the brown straight lines which are exactlyνr = Uν).

Then, when the input is Poisson, decreasing the reliability fromU = 1 to U = 0.1 (see

figure) results in a decrease of the slope of the gain functionat the originand at low values

of ν. However, for higher values ofν (beyond the vertical green dotted line) the gain has

increased! This reflects the dependence of the saturating frequencyνsat = 1Uτv

(eq.2.3.5) on

the parameterU : decreasingU pushes upνsat so that for those inputsν that saturate when

U = 1, with U = 0.1 they will be better represented because the transfer function is steeper

at that value ofν.

The range of rates for which the change[U = 1 → U = 0.1] is advantageous is de-

termined by the intersection of the derivatives of each transfer function (bottom plot). If

the input is correlated, the result is qualitatively the same, except that now, forU = 1 the

non-saturating range has been expanded (see eq.2.3.10). Therefore, one has to go to larger

values ofν to make the change[U = 1 → U = 0.1] worthwhile9. To summarize, if the

input rateν is high enough, it might be beneficial to make the synapse unreliable. However,

if U gets close to zero the informationJ(ν|n) vanishes, because no responses are observed

within a unit time (eq.3.3.17). Thus, there must be an optimal valueUopt to estimate the

input rateν, lying between a reliable synapse withU = 1 and adeadsynapse withU = 0.

Furthermore, input correlations are not needed to obtainUopt > 0, i.e. this optimization can

9The fact that whenU = 0.1 the transfer function does not exhibit much difference between Poisson and a

bursty input withCVisi = 2 is the following: the enlargement of the non-saturating regime scales withU (see

eq.2.3.10), so for small values ofU the output is less sensitive to the value ofCVisi .

Page 102:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

88 Chapter 3: Information transmission through synapses with STD

0

0,005

0,01

0,015

0,02

0,025

0,03

J p

er s

econ

d [s

-2]

CV=1CV=1.5CV=2

ν=10 hz, τc= 50 ms, N

0= 1

0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1

U

0,0005

0,001

0,0015

0,002

0,0025

0,003

J p

er s

econ

d [s

-2]

0

0,001

0,002

0,003

0,004

0,005

0,006

J p

er s

econ

d [s

-2]

τv = 100 ms

τv = 500 ms

τv = 1000 ms

Figure 3.20: Optimization of the release probability ,U , regarding the Fisher information per

second, when the output code is the number of responses. The three plots representJ(ν|n)

versusU for three values of the inputCV (colors in the inset apply to all plots). The larger

the recovery time constantτv , the “stronger” are the depression effects, and the lower is the

optimalUopt .

be achieved for both Poisson and correlated inputs. The only requirement is thatντv 6= 0,

that is, that depression is present.

Figure3.20displays the Fisher information per second, as a function ofU , for different

values ofτv. Only in the top plot both output codes were plotted, showing that they are very

similar. In the other two plots only the timing code is exhibited for clarity. The cases depicted

try to represent three different levels of depression namely: weak (τv = 100 ms), moderate

(τv = 500 ms) and strong depression (τv = 1 s.). These three levels of depression could

also be obtained by fixing the value ofτv and increasing the input rateν. For the Poisson

input (black lines), in all three cases there exists a non-trivialUopt , i.e. 0 < Uopt < 1. As

τv increases depression becomes stronger and, for a fixed rateν, saturation is increasingly

important. Hence, the optimalUopt becomes smaller as the vesicles turn slower. The strategy,

that the model synapse seems to adopt to estimateν, is simple: “If you are not fast enough

to get ready toattend everyone, just reduce the number ofvisitsby randomly getting rid of a

fractionU . Only in that case you are still sensitive to changes in thevisitsrate”.

Input correlations change little the result only in the case whereUopt falls near one: as

Page 103:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 89

0

0.2

0.4

0.6

0.8

1

τ

0

5

10

15

20

ν

0

0.2

0.4

0.6

0.8

1

Figure 3.21: 3-D plot of the optimal release probability ,Uopt , as a function ofν andτv,

when the input is Poisson (see eq.3.3.19). It can be observed that only for lowντv, the

reliable synapse, i.e.Uopt = 1, is optimal. (Note:τ in the axis label stands for theτv of the

text.)

CVisi increases, the optimalUopt boosts slightly to higher values.

If the input is Poisson, an analytical expression forUopt can be easily derived. Taking the

derivative ofJ(ν|n), the equation∂J(ν|n)∂U

= 0 reads,

2τ 3v ν3U3

opt + τ 2v ν2U2

opt − 1 = 0 (3.3.18)

whose solution reads

Uopt = min(1,0.657

ντv

) (3.3.19)

Figure23 shows the 3-D hyperboloid defined by the functionUopt(ν, τv). It can be ob-

served that only when the productντv is very small (more exactly whenντv < 0.657), the

completely reliable synapseU = 1 is optimal. As soon asτv increases to plausible values

andν ranges in frequencies around10− 20 Hz, the optimalUopt drops very fast to values in

the interval0.1− 0.2.

When the input is correlatedUopt does not change too much. As fig.3.22shows, unless

τc is very small (∼ 10 ms), it barely changes with respect to the Poisson input, ifCVisi is

increased. In that case, the input correlations pushUopt closer to one. For example, if for

Poisson inputUopt is not too small (which only happens if the productντv is not too big), a

bursty input (CVisi ∼ 2, τc ∼ 5 ms) with the sameν would have a largerUopt . Hence, the

Page 104:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

90 Chapter 3: Information transmission through synapses with STD

0 10 20 30 40 50

ν (Hz)

0

0.2

0.4

0.6

0.8

1

Uop

t

CVisi

= 1

CVisi

= 1.5

CVisi

= 2

0

0.2

0.4

0.6

0.8

1

Uop

t

0 10 20 30 40 50

ν (Hz)

Figure 3.22: Optimal release probability ,Uopt , as a function ofν, when the input is corre-

lated. Common parameters:U = 1, N0 = 1, CV = 1 (black lines),1.5 (red lines) and2

(blue lines). Top left:τv = 0.1 s.,τc = 50 ms. Top right:τv = 0.1 s.,τc = 5 ms. Bottom left:

τv = 0.5 s.,τc = 50 ms. Bottom right:τv = 0.5 s.,τc = 5 ms.

range where the reliable synapse is optimal for Poisson is enlarge for correlated inputs (see

top left plot in fig.3.22). In conclusion, for correlated inputs is more likely the case that the

reliable synapse performs optimally.

3.3.3.2 OptimizingU with the Mutual Information

We now focus on the question of finding the optimal values of the release probabilityU ,

that maximize the mutual informationI(∆; ν). We will make use of an exponential p.d.f of

input rates defined in previous sections (see eq.3.3.6): fν(ν) = e−(ν−νinf )/ν

ν.

In the first place, in the limitU → 0, I(∆; ν) converges to zero.10 A precise formal

10 As we saw in previous section, that is not the case ofJsr(ν|∆) (neither ofJsr(ν|n)) which showed a

paradoxical behavior in this limit. However we must emphasized here the fundamental differences between the

Fisher and Shannon information. Although related (and behave similar in many cases) they are qualitatively

distinct: the Fisher information in this limitU = 0 only provides an lower bound for the reconstruction error.

It does not say anything about the existence of an estimator which saturates this bound. This paradoxical limit,

seems to be one of those cases in which there is not such an estimator (simply because there are no responses to

estimate with). On the contrary the information gives a much accurate quantification about how much can the

input ensemble be constrained, by observing the output. In this caseU = 0, the observation ofno responses

does not restrict the number of possible input messages and thusI(∆; ν) equals zero.

Page 105:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 91

explanation of this limit is the following:

Limit U → 0 Let us divide the synaptic model into the equivalent two-stage channel defined

it in section2.2.3of chapter2. The first stage is a non-activity dependent random filter

which decimates the input spike train with a probabilityU , while in a second step we

incorporate the vesicle dynamics. The out-coming process of the first stage, is what

was defined as the diluted (or decimated) inputρdisi(∆|ν). If the input is Poisson, this

dilution is equivalent to renormalizing the input rate by a factorU , i.e. the output is

the same input process but with a lower rate:

ρdisi(∆|ν) = ρisi(∆|Uν) (3.3.20)

Let us compute the non-conditioned distributionρdisi(∆) of the diluted train:

ρdisi(∆) =

∫∞0 dνρd

isi(∆|ν)fν(ν) =∫∞0 dνρisi(∆|Uν)fν(ν)

and let us assume, for simplicity, thatνinf = 0 and change the integration variable as

ν ′ = Uν obtaining

ρdisi(∆) =

∫∞0 dνρisi(∆|Uν)

exp(−νν

)

ν=

∫ ∞

0dν ′ρisi(∆|ν ′)

exp(−ν′

Uν)

Uν=

=∫∞0 dν ′ρisi(∆|ν ′)fUν(ν

′)

The final expression ofρdisi(∆) is simply the input distributionρisi(∆) when the mean

of the input ensemble has been renormalized toUν. Thus, as we takeU → 0, the

p.d.f. fUν(ν′) tends to the Dirac delta distributionδ(ν ′). When this occurs, the entropy

H(∆d) associated to the distributionρdisi(∆) becomes zero11 and thetotal informa-

tion I(∆; ν) also vanishes (because at the intermediate variable∆d, the entropy has

vanished).

DoesI(∆; ν) exhibits a non-monotonic behavior as the probabilityU is increased from

zero to one? The answer depends on the amount of depression. An unreliable synapse (U <

1) still has some advantages over the reliable one (U = 1) in terms of transforming the gain

of the transfer function. Figure3.23(top plots) shows the mutual information as a function

of U for the same three values ofτv chosen in fig.3.20(for the Fisher information), and the

mean rateν = 10 Hz Although the same values for the parametersτv and ν were chosen,

the three situations do not represent as before weak (τv = 100 ms), moderate (τv = 500

11Thedifferentialentropy[Cover and Thomas, 1991] does indeed converge to minus infinity. If we discretize

the signals by for instance binning the time, the entropyH(∆d) would converge exactly to zero asU → 0.

Page 106:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

92 Chapter 3: Information transmission through synapses with STD

0

0.02

0.04

0.06

0.08

0.1

τv= 1 s

0 0.2 0.4 0.6 0.8 1

U

0

0.05

0.1

0.15

0.2

0.25

0

0.05

0.1

0.15

0.2

τv= 500 ms

0 0.2 0.4 0.6 0.8 1

U

0

0.02

0.04

0.06

0.080

0.05

0.1

0.15

0.2

0.25In

fo [

bits

]

CV=1CV=1.5CV=2

τv= 100 ms

0 0.2 0.4 0.6 0.8 1

U

0

0.5

1

1.5

Info

Rat

e [b

its/s

]

Figure 3.23: Optimization of the release probability regarding the mutual information

I(∆; ν) for different values of the input correlations magnitudeCV and the recovery time

constantτv. Top: Shannon informationI(∆; ν) versusU . Bottom: Information Rate versus

U . Left plots: τv = 100 ms. Center plots:τv = 500 ms. Right plots:τv = 1 s. Inset values

apply for all plots. Common parameters:ν = 10 Hz, τc = 50 ms,N0 = 1.

ms) and strong depression (τv = 1 s.). Now there is an exponential ensemblefν(ν) of input

rates, implying that the three instances represent negligible, low and moderate depression.

Therefore, for the Poisson case (black lines), only in the last two examples it comes out

thatUopt < 1. Again, the larger is the productντv (the stronger the depression) the smaller

resultsUopt . Because in these examplesUopt lies near one for Poisson, the introduction of

correlations disrupts the maxima and makesUopt = 1 (see explanation in previous section).

When we consider the information rateR(∆; ν) = νrI(∆; ν) , sinceνr is a monotonically

increasing function ofU , we obtained qualitatively the same result except that the optimal

Uopt is slightly pushed up towards one.

In conclusion, when maximizing the mutual informationI(∆; ν) if depression is strong

and the saturation effects are present, an unreliable synapse transmits more information than

a reliable one.

Page 107:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 93

3.3.4 Optimization of the distribution of synaptic parameters

In this section we will optimize the population distribution of synaptic parameters,D(U,N0, τv)

introduced in section2.2.4. After optimizing, we will compute the mutual information that

the array of responses generated by the population of synapses,∆iMi=1, conveys about the

population rateν (see fig.2.4).

As explained in previous chapter2, in the population model proposed, the distribution

can be expressed as

D(U,N0, τv) = f(N0) R(U |N0) P (τv|N0, U) (3.3.21)

wheref(N0) andR(U |N0) are determined with two requirements derived from experimen-

tal data (see section2.2.4). On the contrary,P (τv|N0, U) will be determined by thelocal

optimization of individual synapses as (section3.2.3)

P (τv|N0, U) = δ(τv − τopt(U,N0)) (3.3.22)

By this procedure we can obtain the marginal distribution ofτv, g(τv), by just integrating

D(U,N0, τv) overU andN0, after the optimization (eq.3.2.20).

To study the effect of the distribution in the transfer of information, we have picked three

values of the parameterq, of the Gamma functionsR(U |N0), which areq = 13.2, 11.2 and

9.3. The first choice give rise to a distribution more similar to the ones reported in [Dobrunz

and Stevens, 1997] and [Murthy et al., 2001], where the mean number of docking sites12

< N0 >∼ 5. On the contrary, the choiceq = 9.3 is closer to a the distribution reported by

Hanse and Gustafsson[2001a] where the mean number ofprimedvesicles when the synapse

is at rest, is∼ 1. We would like to test which of the of distributions results in a larger

information, when the optimization described is applied.

The second parameter of the distributions,λ, has been chosen to be the same in the three

cases and equal to the value reported byMurthy et al. [1997], i.e. λ = 7.9. Thus, after

computing the joint distributionsD(U,N0) for the differentq’s, we have generated three

populations, each one composed ofM = 745 synapses. At each synapse (determined by

the valuesU,N0) the value ofτv is determined by maximizing the informationI(∆; ν) at

that particular synapse. The populations of synapses built in this way are, on the one hand

optimally tuned to transmit information, and on the other hand, are qualitatively in agreement

with the experimental data found in hippocampal CA3-CA1 synapses [Murthy et al., 1997,

2001, Hanse and Gustafsson, 2001a].

12WhatDobrunz and Stevens[1997] andMurthy et al.[2001] really measured was the size of the RRP when

the synapse was at rest. Under our perspective, this is equivalent to the maximal size of the RRP which isN0.

Page 108:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

94 Chapter 3: Information transmission through synapses with STD

0 1 2 3 4 5 6 70

0,2

0,4

0,6

0,8

0 0,05 0,1 0,15 0,20

20

40

60

80

0 0,05 0,1 0,15 0,20

20

40

60

80

100

120

0 0,05 0,1 0,15 0,2

τv

(ms)

0

50

100

150

0 1 2 3 4 5 6 70

0,2

0,4

0,6

0,8

0 1 2 3 4 5 6 70

0,2

0,4

0,6

0,8

0 0,2 0,4 0,6 0,8 10

1

2

3

4

0 0,2 0,4 0,6 0,8 10

1

2

3

4

0 0,2 0,4 0,6 0,8 10

1

2

3

4

q= 13.2

q= 11.2

q= 9.3

< N0 >= 2.25

< N0 >= 1.89

< N0 >= 1.25

Figure 3.24: Three examples of optimization of the population distributionD(U,N0, τv).

Each example represents a different value ofq (indicated in the right plots), an is illustrated

by three plots of the same color (top panel black; middle panel red and bottom panel blue).

Within each panel:(Top left plot) Histogram ofU generated by745 synapses (which can

be fitted with the Gamma functionΓλ(U) of eq. 2.2.25, with λ = 7.9. (Bottom left plot)

Histogram ofN0 generated by the same synapses.(Right plot) Colored bars represent the

histogram ofτv (which is an estimation of the marginalg(τv)) after the local optimization

was performed (see text). Blue line represent the histogram ofT = τv

N0. Other parameters

used:f(ν) =exp(ν−νinf )/ν

ν, ν = 20 Hz, νinf = 1 Hz, τc = 20 ms,CVisi = 2.

Page 109:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.3. Results 95

Figure3.24shows the results of the construction of these three populations. In each of

the three panels (vertically aligned), the bottom left plot shows the histogram of the release

probabilityU (which, in the three cases can be fitted with the Gamma functionΓλ(U) of eq.

2.2.25). The top left plot shows the histogram ofN0, which differs from one panel to the

other due to the different choice ofq. The mean value< N0 > is shown inside the plot. In

the top panel (black), whereq = 13.2, the histogram ofN0 decays slowly up to the value

N0 ∼ 5 beyond which the probabilityf(N0) is very small. Therefore, the mean number of

sites is< N0 >= 2.25. On the other hand, in the bottom panel, whereq = 9.3, the sameN0-

histogram shows a rapid decay, and only the probabilitiesf(1) andf(2) are clearly different

from zero. In this case< N0 >= 1.25.

In the right plot of each panel, the colored bars illustrate the histogram ofτv . The shape

of the marginalg(τv) can be estimated from this histogram. In the three examples, it has

a prominent peak at∼ 60 ms. The top and the middle ones, show a significant tail, which

in the first reaches almostτv ∼ 200 ms. The superimposed blue line shows the histogram

of the variableT = τv

N0. As previously shown (section3.3.2.3), it represents the recovery

time of theprobability. This is indeed the variable which can be directly measured in the

experiments. Its bimodal shape can be traced back to the dependence ofτopt on the number of

vesicles (section3.3.2.3). There, it was shown that there is a qualitative difference between

the optimization of a single site synapse and a multi-site synapse.τopt depends linearly on

N0 onlywhenN0 ≥ 2. WhenN0 = 1, τopt takes a value “apart” of the linear trend it follows

whenN0 ≥ 2 (see fig.3.16). As a result, the small an wide bump centered atT ∼ 25 ms is

produced by the synapses withN0 ≥ 2, while the tall and narrow peak centered atT ∼ 60

ms is the outcome of single-site synapses. With this explanation, it is clear why the small

bump is less prominent in the bottom panel: the number of synapses withN0 ≥ 2 is smaller.

Once the three populations have been generated, we have computed the information that

the array of IRI’s produced in each synapse,∆iMi=1, conveys aboutν. Because the mutual

information is not additive [Cover and Thomas, 1991], the information conveyed by the

sum of responses is not the sum of the informations. We will use eq.3.2.15to compute

I(∆i; ν). To computeJ(∆i|ν) first, we simply sum the individual contributions of each

contact (eq.3.2.16):

J(∆i|ν) =M∑i=1

Ji(∆i|ν) (3.3.23)

The result in the three cases reads:

I(∆i; ν) =

7.0 bits, if q = 13.2

6.9 bits, if q = 11.2

6.9 bits, if q = 9.3

(3.3.24)

Page 110:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

96 Chapter 3: Information transmission through synapses with STD

Therefore, we conclude that the differences between these populations are not manifested in

the value ofI(∆i; ν), since a variation of0.1 bits is within the error.

3.4 Conclusions and Discussion

A couple of previous works have addressed the question of whether short term depression

may increase the information transmitted from the pre-synaptic terminal, to the post-synaptic

cell [Goldman, 2000, Fuhrmann et al., 2002, de la Rocha et al., 2002]. All of them conclude

that, under certain assumptions, depression may be an advantageous strategy to transmit

information. In the last two chapters, we have studied the problem in detail, going further in

several aspects: i) A wide family of correlated inputs was considered which consists in all the

renewal processes with exponential correlations. ii) A synaptic model which includes a RRP

with several vesicles was used. iii) Most of the calculations were carried out analytically,

providing a deeper understanding of the problem.

Furthermore, we have adopted different encoding variables in the input and output: in our

case, the inputsignal is the spike rate,ν, whereas in the output, we consider two encoding

schemes, timing and counting.

The main results obtained here are the result of the interaction between two effects

which take place at depressing synapses, namely saturation [Abbott et al., 1997, Tsodyks

and Markram, 1997] and the transformation of auto-correlations13 [Goldman et al., 1999,

2002, de la Rocha et al., 2002], what, when they are exponentially shaped and positive, can

be viewed as a reduction of variability.

We briefly enumerate the main results:

1. If the rest of the parameters are fixed, increasing the input rateν always increases the

reconstruction errorper responsemade in the estimation ofν (see top plot fig.3.1). In

the estimation performedper unit time, there exists an optimalν, which for realistic

values of the parameters, is around∼ 5− 15 Hz (see bottom plot fig.3.1).

2. On the other hand, for a wide family of rate distributionsf(ν) (i.e. Gamma distribu-

tions) there exists an optimal mean input frequencyνopt (∼ 5 − 20 Hz) at which the

mutual information is maximal (see fig.3.4).

3. If the input is an un-correlatedPoissonprocess (or has negative correlations), a static

synapse (τv = 0) is more advantageous than a depressing synapse (τv > 0). Moreover,

13When the input is embedded with positive temporal autocorrelations between the spikes, depression acts as

“filter” of those correlations. However, if the input is un-correlated, depression introducesnegativecorrelations

[Goldman et al., 1999]

Page 111:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.4. Conclusions and Discussion 97

both Fisher and mutual informations are monotonically decreasing functions ofτv (see

top plot fig.3.5and middle plot in fig.3.6).

4. For the caseN0 = 1, if the input has short-range positive correlations, there always

exists a non zero vesicle recovery time,τopt > 0, which gives a larger information

(both Fisher and Shannon) than the static situation, i.e.τv = 0 (see top plot fig.3.5

and middle plot in fig.3.6). However, for realistic values of the parameters,τopt falls

in the interval50− 200 ms, smaller than the values reported by the experiments.

5. WhenN0 > 1, having input positive correlations is not a sufficient condition to obtain

τopt > 0. However, there exists a minimal value ofCVisi above whichτopt > 0. This

CVisi “threshold” is larger the bigger isN0, and also increases withτc (fig. 3.13). We

can summarize all these dependencies by a heuristic inequality which must hold to

haveτopt > 0:

ντcN0 < constant (CV 2isi − 1) (3.4.1)

6. The optimal value of the vesicle recovery time is inversely proportional to the diluted

rateνd = Uν (fig. 3.12):

τopt ∝ 1(Uν)α , 0 < α < 1

7. When considering the metabolic consumption as a constraint to obtain high recovery

rates 1τv

, the optimization ofτv results in values which range from200 to 1000 ms

depending on other parameters. These values are in accordance with the experimental

data.

8. For τv =τopt , a synapse with one (if we consider the Fisher information, fig.3.9) or

two (for the Mutual information, fig.3.11) docking sites (N0 = 1, 2) performs as well

(Poisson input) or better (correlated input) as a multiple sites terminal (N0 > 2).

9. If the synapse does not show short term depression (τv = 0), the optimal release

probability is one in all the cases, i.e.Uopt < 1 (figs. 23and3.22).

10. But, if the synapse shows depression, the Fisher information per unit time, the mutual

information and the information rate may have a non-trivial optimal release probabil-

ity, Uopt < 1. In other words, unreliability becomes an advantage to transmit informa-

tion (eq. 3.3.19). As opposed to the optimization ofτv , input correlations are not

necessary to obtainUopt < 1. Indeed, the larger the correlations the more unlikely that

Uopt = 1.

Page 112:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

98 Chapter 3: Information transmission through synapses with STD

11. The more depression is present (the largerUντv >>) the lowerUopt . For the Poisson

case (when considering the Fisher information per unit time), the following relation

holds (eq.3.3.19)

Uopt ∝const.

ντv

(3.4.2)

The capacity of synaptic activity-dependent processes to increase the computational power

of neurons has been largely discussed in the last years: as a gain control mechanism [Ab-

bott et al., 1997]; as non-linear temporal filters [Natschlager et al., 2001, Maass and Zador,

1999]; as a mechanism for reading a synchrony code [Senn et al., 1998]; as a mechanism to

decode information conveyed in bursts [Lisman, 1997, Matveev and Wang, 2000a]; as mem-

ory buffers [Maass and Markram, 2002]; as redundancy removing filters [Goldman et al.,

1999, 2002, de la Rocha et al., 2002]; as a mechanism which affects the efficacy of different

codes [Tsodyks and Markram, 1997, Fuhrmann et al., 2002, de la Rocha et al., 2002] and

as a mechanism responsible of adaptation [Chance et al., 1998a, Adorjan and Obermayer,

1999]. The study of its implications at the network level has only started but has already

yielded interesting results [Tsodyks et al., 1998, 2000, Reutimann et al., 2001, Loebel and

Tsodyks, 2002, Pantic et al., 2002].

In the present chapter, we have comprehensively studied the way activity-dependent de-

pression transforms the temporal auto-correlations among the input spikes, leaving the train

of synaptic responses with a different temporal structure. In particular, we have shown that,

if the input auto-correlations are positive and are confined in a short range (e.g. bursts), that

transformation can filter out spikes in an informatively efficient way. Although the informa-

tion conveyed in a certain time window decreases (basically because there are less responses

per unit time) each synaptic response conveys more information about the input rate than if

no spikes where filtered. If on the other hand, we consider the information rate but imposing

a metabolic constraint in the recovery rate of the vesicles (which simply states that higher

recovery rates cost more energy than low ones), again a non-zero recovery timeτv is advan-

tageous over a non-depressing synapse. In both cases, the transmission is efficient in terms

of metabolic consumption, that is, we have optimized the informationper unit energy.

In addition, we have shown that, if depression is prominent, the existence of unreliability

in the synaptic transmission can be explained in terms of optimization of the information.

Unreliable synapses result to transmit a larger information per response. What seemed to be

a “bug” in the circuity has turned to be an advantageous strategy to overcome the saturation

produced by short term depression.

Quantitatively, our prediction of the optimal recovery time constantτv , when no metabolic

constraint in the recovery rate is imposed, seems to be lower than what has been measured

in the experiments: if the inputν ∼ 10− 20 Hz, τopt ranges from50− 150 ms, while exper-

Page 113:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.4. Conclusions and Discussion 99

iments in slice report values from400 ms [Markram et al., 1998a, Petersen, 2002] to 2000

ms [Dobrunz and Stevens, 1997]. This inconsistency could be due to a least three reasons: i)

The experimental conditions do not resemblein-vivoconditions, that is, that synapses in the

slice do not behave like in an intact brain (something imputable to many different factors14

[Steriade, 2001]) ii) Small rates (2 − 5 Hz) might carry information so that they are not, as

usually considered, spontaneous rates. iii) The metabolic consumption of the recovery must

be considered in the optimization.

On the other hand, if the recovery metabolic cost is considered in a simple manner (by

means of an energy function linear on the recovery rate), the values of the optimalτopt are

in the range200 − 1000 ms, and again bursty inputs give larger values. It seems then like,

perhaps in synapses with STD, the recovery time value has arisen as a trade off between

energy efficiency and maximization of the information transfer.

The third synaptic parameter which was analyzed in the present chapter is the size of the

RRP,N0, i.e. the maximum number of ready vesicles. The result of its optimization depends

on the ability of the synapses to adjust theirτv to its optimal value: ifτv = τopt the optimal

N0 is one or two depending on whether we maximize the Fisher or the Shannon information,

respectively. If on the contrary,τv is fixed to a large value (e.g.∼ 1 s.) larger values

(N0 > 7) are optimal. If the first scenario occurs, then our findings are in agreement with

the recent experimental findings [Hanse and Gustafsson, 2001a,b] which observed that the

most immediately releasable pool (what they call the pre-primed pool) has on average one or

two vesicles [Hanse and Gustafsson, 2001a]. These experiments also show that the recovery

of the pre-primed pool takes less than a hundred milliseconds. Besides the pre-primed pool

they suggest there exists a docked pool with slower recovery kinetics and a larger size. In this

way, our result of an optimal pool with one or two vesicles and fast recovery rates (τv = 50

ms), could represent the pre-primed pool which has a strong an immediate impact on the

release probability change, and therefore could eliminate the correlations due, for instance,

to bursts. A second and larger pool, harder to deplete, could have a different role, and it was

not modeled in this work. It is necessary to mentioned, that this perspective of two pools

with different sizes was proposed byMatveev and Wang[2000b] to explain the largepaired

pulsedepression observed in neocortical slices [Markram and Tsodyks, 1996a, Varela et al.,

1997] which seems at odds with a large unique RRP. At this point, ifN0 = 1, 2, our model

would be able to reproduce such a high paired pulse depression.

We would like now to mention a parameter relationship, which has appeared in some of

our results, which is in agreement with previous works [Fuhrmann et al., 2002], and to make

14We have notice for instance that slice recordings are often performed at room temperature (∼ 22o C.)

which, as explained by [Dobrunz and Stevens, 1999], may contribute to slow down the kinetics of neurons up

to a factor of3

Page 114:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

100 Chapter 3: Information transmission through synapses with STD

an experimental prediction with it. The two synaptic parametersU andτv , and the input rate

ν have revealed, in different partial results, to be related, when one of them is optimized by

Uopt ∝ 1

τv ν(3.4.3)

τopt ∝ 1

U ν(3.4.4)

Experiments, however, have shown little correlation betweenτv andU [Petersen, 2002]. Our

prediction is that those neurons which code information at high rates (highν) would have,

either very unreliable synapses (lowU ) or they would recover fast (lowτv ). On the other

hand, if the synapses of a pre-synaptic neuron have very slow recovery (largeτv ) or are very

reliable (U ∼ 1), it should fire at low rates.

Future work

The work presented in this chapter is a first step towards understanding the role of the

different biophysical properties of cortical synapses in information processing. To start with,

we have chosen an elementary input-output system: a model of a pre-synaptic terminal which

captures the effects of unreliability and short-term depression. Other mechanisms, such as

facilitation, post-synaptic dynamics like desensitization of transmitter receptors or synaptic

conductance kinetics, should be added in further work.

To test whether the results obtained here, are the same when the input is a natural stimulus

with positive auto-correlations, is also important to validate the conclusions. Thus, it is our

purpose to compute numerically the information transmitted by bursty stimuli recorded in

hippocampal pyramidal cells [Dobrunz and Stevens, 1999, Fenton and Muller, 1998].

The analysis carried out about the information transmitted by a population of synapses,

should be continued. The results obtained so far, do not explain the benefits of having an

heterogeneous distribution of parameters neither what sort of distribution would provide a

better representation of the input. Obtaining, by numerical methods, the optimal distribu-

tion Dopt(U,N0, τv) and comparing the information transfer then with that of the locally

optimized distribution, would clearly show whether or not such heterogeneity is relevant to

transmit an ensemble of inputs.

Information theory provides a mathematical framework for quantifying how well we can

establish the stimulus identity by observing the responses. However, it does not tell us how a

neuronextractsand processes the information present in the responses, neither what sort of

computations can be performed by using this information. For example, it has been shown

[Fuhrmann et al., 2002] that the size of an EPSP conveys information about the time of the

previous spikes. In this thesis we have worked under the hypothesis that information in the

Page 115:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

3.4. Conclusions and Discussion 101

input was encoded in the firing rate. Whether neurons use one or the other (or both) coding

strategies, depends on their capacity to extract (decode) the information of such a signal, and

to encode their output response in the same manner (assuming that from one stage to the next,

the code does not change). Therefore, we consider that an interesting line of investigation

that complements the work presented in this chapter, is the analysis of the integration and

response of the whole cell, to complex temporal structured inputs. This analysis would

provide hints of the mechanism that neurons might use to process the information conveyed

in their inputs. The next two chapters point in this direction.

Page 116:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 117:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Chapter 4

Synaptic current produced by a

population of synchronized neurons

4.1 Introduction

A major challenge in neuroscience is to understand how a cortical neuron, receiving a

time-varying signal composed of small current pulses, from around103 − 104 [Braitenberg

and Schuz, 1991] synapses distributed along its dendritic tree, is capable to perform any

computation. If we had to find a simile, it would be like reading the information encoded

in the outcome of a shower, where thousands of drops (representing spikes) are hitting our

skin in a continuous manner, and all our work would be to make a computation out of that

barrage of water and encode the solution, in the best case, in a series of simple binary sounds

coming out our throat. Moreover, several facts contribute to make such a fabulous task

even more complex: a) the spike trains impinging the dendrites are highly irregular [Softky

and Koch, 1993], seemingly random, which sounds incompatible with he simplest view of

cortical processing [Softky and Koch, 1993, Softky, 1995]; b) these spike trains have to

transmit its information to the post-synaptic neuron across cortical synapses which, as posed

in the previous chapter, are often highly unreliable [Allen and Stevens, 1994, Dobrunz et al.,

1997], meaning that a large fraction of afferent spikesfail in reaching the post-synaptic cell;

c) these synapses display a large number of different activity-dependent processes such as

short-term depression, facilitation, augmentation, post-tetanic potentiation, etc. (see [Zucker

and Regehr, 2002] for a review), which make that post-synaptic responses wax and wane

as the pre-synaptic cell dictates with its activity. In addition, these dynamical processes are

often widely heterogenous in their quantitative features from one synapse to another, even

along synaptic boutons of the same axon [Murthy et al., 1997, Markram et al., 1998a, Gupta

et al., 2000].

103

Page 118:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

104 Chapter 4: Synaptic current produced by synchronized neurons

Since many years, it has been generally agreed that the neurons output firing rate contains

information about its inputs [Adrian, 1926, Werner and Mountcastle, 1965, Tolhurst et al.,

1983, Tolhurst, 1989, Britten et al., 1992, Tovee et al., 1993]. Recently there has been grow-

ing evidence that the precise times of the spikes also conveyed information [Panzeri et al.,

2001, Rieke et al., 1997, Bialek et al., 1991, Softky, 1995]. Hence, there exists a debate in

which the two poles are represented by two limit coding hypothesis, namely: i) a pure rate

code in which the signal is represented by just the mean number of spikes in a certain time

window, and where the variance of that number is only noise; ii) the simple idea that the

spike patterns produced by a single neuron encode information as if it speaks binary mean-

ingful words [Rieke et al., 1997], that is, the exact position of each spike is not noise but

signal.

However, there are experimental indications that other forms of coding lying in be-

tween these two coding schemes are used by the central nervous system ([deCharms and

Merzenich, 1996, Bergman et al., 1995, Murthy and Fetz, 1996, Vaadia et al., 1995, Prut

et al., 1998] see also [deCharms and Zador, 2000] for a review). In particular, the gener-

ally calledCoordinated-coding hypothesissuggests that the temporal relation among signals

from multiple neurons plays a crucial role in the processing of messages in the brain [Carr,

1993, Hopfield, 1995, 1996]. To decode the information content of a spike train one must

compare its temporal pattern with the output of other neurons. Information cannot be ex-

tracted by independent analysis of the spike trains neither by pooling the independent votes

of individual neurons.

The synchronous firing of action potential (AP) by neurons from a given population has

been hypothesized as a particular instance of coordinated coding ([Gray and McCormick,

1996, Singer, 1999], see also [Salinas and Sejnowski, 2001] for a recent review). General-

izing the idea of synchrony to longer delays in the firing of coordinated neurons and more

complex temporal relations,correlatedneural activity is believed to play an important func-

tion is cortical processing such as attention, gain modulation, and sensory encoding [Salinas

and Sejnowski, 2001].

The impact of correlated input in the response of a neuron has been studied by many dif-

ferent groups [Bernander et al., 1994, Murthy and Fetz, 1994, Stevens and Zador, 1998, Feng

and Brown, 2000, Bohte et al., 2000, Salinas and Sejnowski, 2000, Rudolph and Destexhe,

2001, Kuhn et al., 2002, Moreno et al., 2002] .

The response of a neuron with STP has also been studied in previous works [Tsodyks

and Markram, 1997, Markram et al., 1998a,b, Abbott et al., 1997]. The most important

result found by both groups is that no information about an constant input rate can be trans-

mitted, when operating in the saturation regime (section2.3). In this regime the response

Page 119:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.2. Parameterization of the afferent current 105

of the output neuron reflects only changes of the rate during a brief transient [Tsodyks and

Markram, 1997]. If the input is a time varying signal, the outcome of the neuron, working

in saturation, is sensitive to proportional changes of the input rather than to the absolute

magnitude of those changes [Abbott et al., 1997].

In chapter2 the statistics of the output of a model of vesicle depletion were analyzed. The

implications in the information transmission through that stochastic channel was studied in

chapter3. In the present chapter we will examine how those synaptic releases are integrated

by the post-synaptic cell. In previous chapters, however, we only considered positive tempo-

ral auto-correlations in the input spike train, because the analysis was performed in asingle

synaptic contact. Since the system now will be a spiking neuron receiving afferents from

approximately104 synaptic contacts, our aim here will be to quantify the relevance of the

spatial cross-correlations between the impinging trains of AP’s. In this way, two examples

of cross-correlations will be considered: i) first, the correlations existing between synaptic

contacts which “belong” to the same pre-synaptic neuron. In other words, we will extend the

model of connection introduced in previous chapters, from a single contact to an arbitrary

number of synaptic contacts between the pre and the post-synaptic neuron. ii) The case in

which a pre-synaptic population of neurons fire with a certain grade of synchrony will be

considered, i.e. zero-lag cross-correlations will be included in the pre-synaptic activity.

The goal in this last two chapters is twofold: first, to obtain an analytical description of

the statistics of the current as well as the first moment of the output rate, as a function of the

synaptic parametersτv andU (see section2.2.1), and the cross-correlations of the incoming

spike trains. Secondly, to evaluate the results in terms of the ability of synchrony and STD

to increase the capacity to process and communicate information between neurons.

In the present chapter we will model and parameterize the afferent current and we will

obtain the expressions of its first and second order statistics. In the next chapter, the response

of a leaky integrate-and-fire neuron (LIF) receiving this current will be analyzed, and the re-

sults and conclusions of these two chapters will be laid out. It will be shown that the presence

of STD results in a wide range of effects which could not be obtained with synchrony itself.

4.2 Parameterization of the afferent current

We will first specify the parameterization of the current, which will be integrated by

the output neuron in the following chapter. Here is a brief description, which follows the

explanations in chapter1 of the way vesicle release takes place and how the synaptic current

is then generated. We will also explain the several simplifications introduced in order to

obtain a tractable model.

Page 120:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

106 Chapter 4: Synaptic current produced by synchronized neurons

How is the synaptic current generated? When an action potential invades the pre-synaptic

bouton, one vesicle, which isready-for-release, may fuse the cell membrane, releasing its

transmitter content to the synaptic cleft. Then, the transmitter interacts through a chemical

reaction in which a certain number of molecules bind to a closed ionic channel receptor and

open it. The opening of a synaptic channel can be modeled with a simple gating variable,

0 < PS(t) < 1, which measures the probability for the channel to be open. Because the time

it takes to open the ionic channel is much shorter than the time it takes to close, the dynamics

of this gating variablePs(t) can be modeled as (see e.g. [Dayan and Abbot, 2001])

dPs(t)

dt=−Ps(t)

τs

+ (1− Ps(t))Pmax

∑rel

δ(t− ti) (4.2.1)

wherePmax is the maximum fraction of open channels when a release occurs and all chan-

nels were previously closed. The sum∑

rel δ(t − tk) represents the series of releases which

take place at timesti (i = 1, 2, . . .) The synaptic time constantτs is a characteristic of

the type of receptor associated with that synaptic conductance. For excitatory conductances,

channels mediated by AMPA receptors have a time constantτs ∼ 5 ms. while those me-

diated by NMDA receptors is usually much largerτs ∼ 150 ms. In the case of inhibitory

synapses, GABAA time constant isτs ∼ 10 ms.. The conductance of a specific synapsei,

is obtained by multiplying the gating variablePs(t) by a constant parameter,gi (which rep-

resents the maximum conductance in that particular synapse achieved when all channels are

open),gi(t) = giPs(t). Finally, the current produced by that synaptic contacti is defined as

Ii(t) = gi(t)(V (t)− Es) = giPs(t)(V (t)− Es) (4.2.2)

whereEs is the reversal potential of the synaptic channel, andV (t) is the membrane potential

of the neuron. If we neglect the fluctuations of the membrane potentialV (t) around its

mean value, we can substitute this dependence in equation4.2.2by writing the mean voltage

V instead ofV (t). Therefore, following this model, if for example one AMPA mediated

conductance is activated by spontaneous release of vesicles at a very low rate (∼ 1 Hz), if

we patch the post-synaptic neuron with an electrode and record the incoming currents at the

soma, we would observe a series of excitatory post-synaptic currents (EPSC’s) which would

be described by the following expression

I(t) = giPmax(V − Es)(e−t−t1

τs Θ(t− t1) + e−t−t2

τs Θ(t− t2) + . . .)

(4.2.3)

= giPmax(V − Es)∑k

e−t−tk

τs Θ(t− tk)

The sum overk accounts for the consecutive releases taking place at contacti. Let us

include now the synaptic activity of the rest of the synaptic contacts. Although the particular

Page 121:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.3. Afferent spike trains 107

location of each synaptic contact in the dendritic arbor affects the way EPSC’s are integrated

[Rall and Segev, 1987, Segev and Rall, 1998, Schutter, 1999, Magee, 2000], we will neglect

the spatial dimension by assuming that the EPSC’s from each synapse add linearly at the

soma. Therefore, the total synaptic current coming fromN contacts takes the simple form

I(t) =N∑i

giPmax(V − Es)∑k

e−t−ti

kτs Θ(t− tik) (4.2.4)

The net charge entering the cell at each release is the integral of the EPSC, that is,Q =

giPmax(V − Es)τs. We will make a final simplification of this scheme, which is to assume

that the synaptic time constantτs is much shorter than the rest of the time constants of the

problem, namely the membrane time constantτm (which will be defined in section5.1) and

the vesicle recovery time constantτv. This is a reasonable approximation in the case of the

AMPA mediated currents, not so good for GABAA, and is definitely wrong in the case of the

long time courses given by the NMDA-mediated conductance. Assuming that the EPSC’s

(and IPSC’s) are very brief pulses of current entering the cell, we can finally express the total

synaptic current as

I(t) 'N∑i

Cm Ji

∑k

δ(t− tik) (4.2.5)

We have introduced the quantityJi = giPmax(V−Es)τs

Cmwhich has voltage units and exactly

represents the magnitude of the instantaneous jump of the membrane voltage upon the in-

jection of an EPSC. In other words,Ji is the size of the excitatory post-synaptic potential

(EPSP) produced by a release at the synaptic contacti. Cm is the capacitance of the cell

membrane.

Now, what determines the statistics of the afferent currentI(t)? As can be seen in eq.

4.2.5, the timestik, wherei is the contact indexi = 1, 2, . . . N , andk the releaseindex

k = 1, 2, . . ., will dictate the temporal structure of current. In addition, the times at which

the pre-synaptic AP’s reach the pre-synaptic terminals, will regulate in time the occurrence

of releases. Since in the next chapter the afferent current will be simplified, by means of the

diffusion approximation (section5.2.1), and described as a Gaussian process, we will only

compute its statistics up to second order. Before starting the calculation of the correlation

function of the total current, we will describe the correlations between the spikes and we will

compute the correlations between the releases.

4.3 Afferent spike trains

As we said in the Introduction, we would like to study the interaction between short

time scale positive spatial cross-correlations, that is, synchrony among pre-synaptic neurons,

Page 122:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

108 Chapter 4: Synaptic current produced by synchronized neurons

and unreliable short-term depressing synapses. Thus, we define the AP’s coming from pre-

synaptic neuroni as a series of delta functions centered at the spike times:

Si(t) =∑

l

δ(t− tli) (4.3.1)

Superimposing the activity of all pre-synaptic cells, one obtains a compound train made up

of the contributions ofC pre-synaptic cells

S(t) =C∑i

∑l

δ(t− tli) (4.3.2)

The second order statistics of this compound train is completely defined by the two point

connected correlation function [Bialek et al., 1991] defined as

C(t, t′) ≡< (S(t)− < S(t) >) (S(t′)− < S(t′) >) > (4.3.3)

where the angles< · > denote average over the ensemble of input trains. The function

C(t, t′) represents the excess of probability of finding a spike (independently of the neuron

it is coming from) at timet given the occurrence of one spike at timet′. If there were no

correlation between the spikes, this function would equal a Dirac delta function centered at

zero. This would be the case, if the afferent spikes were independent Poisson process. We

model each afferent individual fiber spiking activity as an stationary Poisson processes with

identical rateν. However, the individual trains are not independent of each other, but are

partially synchronized, that is, they show cross-correlations with zero time scale. This is

formally imposed by rearranging the correlation function as

C(t− t′) = < S(t) S(t′) > − < S(t) >< S(t′) > (4.3.4)

= <∑C

i,j

∑l,m δ(t′ − tli) δ(t− tmj ) > − (

∑Ci <

∑l δ(t− tli) >)2

=C∑i

<∑l,m

δ(t′ − tli) δ(t− tmi ) >

︸ ︷︷ ︸same neuron

+C∑

i6=j

<∑l,m

δ(t′ − tli) δ(t− tmj ) >

︸ ︷︷ ︸diff. neurons

−C2ν2

Since all the Poisson processes are stationary (ν 6= ν(t)) the correlation function can not

depend on any absolute timet or t′ but only on the difference between both. Moreover, we

have replaced<∑

l δ(t− tli) > by the rateν because it represents the probability of finding a

spike at timet. We have reached an expression where two terms appear: the autocorrelation

of each spike train arriving from thesame neuron, and a second term, which represents the

cross-correlations fromdifferent pre-synapticneurons. Because individual trains are Poisson

processes, the probability of finding two spikes at timest andt′ is just the squared rateν2,

Page 123:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.4. Statistics of the synaptic releases 109

except ift = t′ and the two spikes are the same. This sort of trivial correlation is formulated

with a Dirac delta function centered att− t′, so that the first term in eq.4.3.4reads,C∑i

<∑l,m

δ(t′ − tli) δ(t− tmi ) >

︸ ︷︷ ︸same neuron

= (4.3.5)

C∑i

<∑

l

δ(t′ − tli) δ(t− tli) >︸ ︷︷ ︸same spike

+C∑i

<∑l 6=m

δ(t′ − tli) δ(t− tmi ) >

︸ ︷︷ ︸diff. spike

=

= Cνδ(t− t′) + Cν2

The second term, accounting for the the cross-correlation, reflects the fact that given a

spike emitted by thei-th neuron at timet, there is a probabilityρ that thej-th neuron spikes

at timet as wellC∑

i6=j

<∑l,m

δ(t′ − tli) δ(t− tmj ) >

︸ ︷︷ ︸diff. neurons

= C(C − 1) ν ρ δ(t− t′) + C(C − 1)ν2 (4.3.6)

We take this expression as a definition of the correlation parameterρ, which grades the

strength of the cross-correlations: ifρ = 0 all pre-synaptic neurons fire independently, while

if ρ = 1 theC individual trains are all copies of the same train, or in other words, the spikes

from different neurons are all aligned in time. This way of establishing correlations has been

previously used by many authors [Rudolph and Destexhe, 2001, Kuhn et al., 2002]. Now

putting equations4.3.5and4.3.6together in eq.4.3.4we reach

C(t− t′) = C ν [1 + ρ (C − 1)] δ(t− t′) (4.3.7)

4.4 Statistics of the synaptic releases

We turn now to the calculation of the correlation function of the synaptic releases pro-

duced by the afferent spike statistics defined above. We will make use here of the correlation

function computed for a single contact in chapter2. First we will define the dynamics of

the probability of transmission in a single contact. Later, we introduce a generalization of

the single-contact model between two neurons in which the number of synaptic contacts

established by two neurons is a new parameterM .

4.4.1 Dynamics of one synaptic contact between two neurons

In this section we will solve the dynamics of the system, composed of a synapse with

a single docking site (N0 = 1) receiving a Poisson input. This simple synaptic model is

Page 124:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

110 Chapter 4: Synaptic current produced by synchronized neurons

UτPoisson

Figure 4.1: Model of a single synaptic contact. Spikes are treated as point events as well as

synaptic responses. A release may occur with probabilityU whenever a spikes arrives at the

synapse and a vesicle is ready-for-release. After a release, a new vesicle gets prepared within

a stochastic time taken from an exponential distribution of parameterτv.

illustrated in figure4.1. The probability of transmission upon arrival of a spike equals the

probability that a vesicle is ready-for-release, times the probability of release when the vesi-

cle is docked,U . We will only define the dynamics of the former, because the parameterU

is constant1.

We can now determine the dynamics of the probability for the vesicle being ready-for-

release at two different levels: i) for a particular realization of the input spike trains. ii) for

the whole input ensemble, i.e. the dynamics without the knowledge of which is the particular

Poisson realization of the input train. We shall denote the first probability aspv(t), and its

dynamics will depend on the precise times of the spikes. We will refer to the second one as

< pv(t) >, where the brackets mean average over all the input trains, and its dynamics will

not depend on the precise spike times but on the firing rateν.

Dynamics ofpv(t)

Let us consider first the case where we have a given input spike train, described by the

functionS(t) =∑

l δ(t − tl), and we would like to know how does our initial knowledge

about the state of the vesicle site evolves in time. The diagram of figure4.2describes a time

stepdt of the dynamics of the system. Let us suppose that our vast wisdom makes us detect

the presence of the vesicle at timet′, but with some uncertainty, i.e. weknow the value of

pv(t′). During a very short time stepdt, two processes contribute to change the probability

for the vesicle to be ready at timet′ + dt:

• a) If a spike arrives between(t′, t′+dt), then, if the vesicle was prepared, with proba-

bility U it is released. This is expressed by the negative transition rate:−pv(t′) S(t) U

(red line in diagram4.2).

1If we had considered facilitation,U would have its own dynamics in which every afferent spike will make

it increase towards one, and in the interval between spikes it would decrease exponentially to zero [Tsodyks

and Markram, 1997].

Page 125:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.4. Statistics of the synaptic releases 111

1- Release pv(t+dt)

1-pv(t)=P

0(t)

pv(t)=P

1(t)

1-pv(t+dt)

Release

Recov

er

Figure 4.2: Diagram of the temporal evolution of a system composed of a single synaptic

contact. At timet the contact has a vesicle prepared for release with probabilitypv(t) or

the contact is empty with probability1 − pv(t). A time stepdt later, a transition may have

occurred: a release may occur if the vesicle is ready-for-release , a spike arrives , and release

is successful. The recovery of the contact may occur if the docking site was empty and a

new vesicles suddenly occupies it.

• b) If the vesicle is not avaliable, there is a probability that in the next time step it gets

recovered. This contributes with the positive transition rate:(1−pv(t′))τv

(green line in

diagram4.2)

Thus the dynamics ofpv(t) read:

dpv(t + dt)

dt=

1− pv(t)

τv

− pv(t) U∑k

δ(t− tk) (4.4.1)

This stochastic model has been previously used in other works [Senn et al., 2001, Fuhrmann

et al., 2002]. The only indirect experimental validation, so far, has been that, when averaging

the synaptic response over trials with the same stimulus, the mean response can be fitted with

the time evolution of the product2 Upv(t) [Tsodyks and Markram, 1997]. Because response

recordings of single trials in neocortical neurons are very noise (see fig.1 C-D of [Markram

and Tsodyks, 1996a]), it is not easy to directly validate this stochastic model. However,

it would be interesting to compare other moments of the response, like the mean squared

response for instance, in order to confirm whether this stochastic model is valid or not.

2BecauseU pv(t) represents the probability of an all-or-none process, if we compute the mean of this

process, assessing the value 1 when there is a release and 0 when there is not, the evolution of the first moment

coincides with that ofU pv(t).

Page 126:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

112 Chapter 4: Synaptic current produced by synchronized neurons

The deterministic mean-response model ofTsodyks and Markram[1997], and other

equivalent versions [Varela et al., 1997], have been extensively used in the literature [Tsodyks

and Markram, 1997, Fuhrmann et al., 2002, Tsodyks et al., 1998, 2000]. As we will see later,

it provides the same description of the mean response, and subsequently of the mean current,

as the stochastic model. However, they differ in the description of the second-order statistics

(like the current variance). Nevertheless, it is an approximate correct description of higher

order statistics of the stochastic model, if the number of synaptic contacts is very large (of

order 100). This will be shown in the next chapter.

Dynamics of〈pv(t)〉

How does the probability of the state of the vesicle evolves in time when all we know

about the input is that is a Poisson process of rateν? All we need to do is to take the

average of eq.4.4.1over the input ensemble of trains [Tsodyks et al., 1998]. We denote this

averaging by angular brackets〈·〉. The only non-trivial term is the product〈∑k pv(t)δ(t −tk)〉. However, since the arrival of Poisson spikes does not keep memory of any of the

previous events, and sincepv(t) depends only of the activitybeforetime t, each term of the

previous sum factorizes obtaining〈pv(t)〉〈∑

k δ(t−tk)〉 = 〈pv(t)〉ν(t) [Tsodyks et al., 1998].

If the input is not Poisson, this factorization would not be correct because the arrival time of

a spike would depend, aspv(t) does, on the previous spikes. Thus, the dynamics of〈pv(t)〉are governed by [Tsodyks et al., 1998]

d〈pv(t + dt)〉dt

=1− 〈pv(t)〉

τv

− 〈pv(t)〉 U ν(t) (4.4.2)

If the input rate does not vary in time, the solution of this equation reads,

〈pv(t)〉 = 〈pv(0)〉 e−t/τc + 〈pssv 〉 (1− e−t/τc) (4.4.3)

〈pssv 〉 =

1

1 + Uντv

, τc ≡τv

1 + Uντv

(4.4.4)

where〈pv(0)〉 is the probability at time zero, andτc is the characteristic time constant of the

dynamics3. The stationary state of this probability ,〈pssv 〉, gives us the responses rateνr (eq.

2.3.3in section2.3) for the Poisson case: all we need to do is multiply it byU , times the

probability of arrival of a spikeν:

νr = 〈pssv 〉 Uν =

1 + Uντv

(4.4.5)

3Do not confuse thisτc with the correlation time scale of the input trains introduced in chapter2.

Page 127:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.4. Statistics of the synaptic releases 113

Figure 4.3: Illustration of the different ways two neurons may be connected. Picture(a)

depicts a mono-synaptic connection, this is the case in which a given axon makes only one

contact onto the post-synaptic cell.(b) In this case the connection is multi-synaptic: the same

axon establishes up to three synaptic boutons with the same post-synaptic neuron.(c) This

case illustrates the instance in which within a single synaptic bouton, there exists several

synaptic specializations(three in this case) where vesicle release occurs. In our modeling

situation(b) and(c) will be treated as an equivalent connection: both connections are made

up of threefunctionalsynaptic contacts (taken from [Natschlager, 1999]).

4.4.2 Several synaptic contacts between two neurons

Cortical neurons are seldom connected by only one contact. A single axon may some-

times make multiple contacts onto the same post-synaptic cell. Moreover, an individual

synapse may be have several active zones or release sites (see figs.4.3and1.2). Since we are

neglecting the spatial dimension of the output neuron, these two distinct ways of increasing

the connectivity between two neurons are, according to our approach, equivalent. We will

use the termfunctional contact[Zador, 1998] to refer to both situations. Trying to quantify

how many functional contacts exist between two cortical neurons, seems to be an experi-

mental arduous task. Therefore, the literature is extensive and some times contradictory.

Numbers in the neocortex range from one up to twenty or thirty in some cases[Abeles, 1991,

Markram et al., 1997a,b, Larkman et al., 1997]. In somatosensory cortex of developing rats,

Markramet al revealed a range between four and eight, with an average of 5.5 [Markram

et al., 1997a,b]. GABAergic interneurons tend, nevertheless, to establish more contacts, from

10 to 20 [Gupta et al., 2000]. In the hippocampus different groups report different numbers:

while some groups communicate to find a single contact between CA3 and CA1 pyramidal

cells [Stevens and Wang, 1995, Hanse and Gustafsson, 2001a] others report values between

3 and 18 [Stricker et al., 1996, Larkman et al., 1997]. In addition, it is interesting to mention

Page 128:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

114 Chapter 4: Synaptic current produced by synchronized neurons

Figure 4.4: Schematic picture of the calyx of Held synapse, a giant excitatory synapse in the

mammalian auditory pathway. The pre-synaptic terminal contains hundreds of individual

synaptic specializations in which transmitter release occurs (taken from [Walmsley et al.,

1998].

particular examples in the nervous system in which the numberM of functional contacts is

much higher. The first classical example of a synapse with hundreds of contacts are neuro-

muscular junctions [del Castillo and Katz, 1954a, Katz and Miledi, 1968], which due to their

size, were the foremost accessible synapses studied, giving the earliest understandings about

the quantal nature of synaptic release [Katz, 1996]. Other extensively studied example are

the giant synapses in the auditory system known as the end-bulbs and calces of Held [Held,

1893, Ramon y Cajal, 1995] which are excitatory central synapses in the mammalian brain-

stem [von Gersdor et al., 1997, Weis et al., 1999, von Gersdorff and Borst, 2002]. A calyceal

synaptic bouton, contains hundreds of excitatory synaptic specializations, as depicted in fig-

ure4.4, and shows short-term depression induced by the depletion of pre-synaptic resources

[von Gersdor et al., 1997, Weis et al., 1999, Trussell, 2002, von Gersdorff and Borst, 2002].

To summarize, we will consider the situation in which the pre-synaptic neurons makeM

functionalcontacts4 with the target neuron, and we will investigate the impact in the response

of variations inM . The M contacts, that each of theC pre-synaptic neurons establish,

receive the same spike train, but each of them is provided with its own independent vesicle

dynamics (in particular a single docking site model, i.e.N0 = 1, see2.2.1)5. This causes,

4We will index the contacts sharing the same input spike train with the Greek lettersα, β and will leave the

Latin indexesi, j to denote contacts from different afferent neurons, andk, l to index the spike time arrivals.5It is interesting, however, to establish here an equivalence between this model withM functional contacts

(which so far are identified with the active zones, regardless of the synaptic bouton they belong to) where at

most one vesicle is released, to the model that one obtains if the univesicular release hypothesis is ignored. Let

us take our model of a single synaptic contact (described in section2.2.1of chapter2) with N0 docking sites

Page 129:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.4. Statistics of the synaptic releases 115

that releases in each of theM contacts are highly correlated. This means that if we observe

a release in one contact att′, it is very likely that more releases occur in the rest of theM −1

contacts at the same time than by chance. However, because of the independent vesicle and

release dynamics, when the common spike train is known (the times of the spikes are given),

the response in each of theM contacts is an independent process. This property is usually

denote as conditional independence, since the joint probabilities of release in each of the

M contacts only factorize if all previous spikes in the train are given. Before addressing

the calculation of the correlation function of the synaptic releases, we will first formalize

this conditional independence and derive some quantities, that will be necessary later in this

chapter.

The first question we would like to prompt is, can we model the release of theM contacts

as a binomial distribution of parametersM andpr, wherepr is defined as the probability that

a spike triggers a response in one contact? This has been the way in which several authors

have modeled this simple system (see [Fuhrmann et al., 2002] for instance), but as we will

see now it is not a strictly correct description. The answer then is no, the response of the

M contacts is not a binomial, simply because the processes are not independent, but only

independent when all the previous spike times are given. If for instance, the vesicle recovery

time is very long (very slow recovery) and release occurs when the vesicle is docked with

probability close to one (U ∼ 1), if we observe that a spike arriving att′ elicits a release in

contactα, it will be more likely to find a release at contactβ than if we had no knowledge

about what happened at the other. The explanation is easy: the release at theα-th contact

indicates that the vesicle there, was ready yet, probably because the previous spike came a

long time ago (since we set a slow replenishment of vesicles). Because both contacts share

the same incoming spikes, then the last spike which arrived at theβ-th contact did it a long

time beforet′, making the recovery of the vesicle more likely to happen. This dependency

occurs for arbitrary values ofU andτv .

In order to quantify this dependency, we need to compute the conditional probability

that a certain spike att′ finds a ready-for-release vesicle given that a common contact did

find one at the arrival of the same spike. These probabilities have to be averaged over the

input ensemble, i.e.〈·〉. We will denote this conditional probability as〈pv(β|α)〉. Although

and now assume that docked vesicles may fuse the membrane independently in such a way that, at a single

active zone, it can be produced as many simultaneous releases as docking sites are located in the area of the

active zone. In this case we do not identify a release site with an active zone, but with a docking site (see

comment in section1.2.4). In this description, the model of synaptic connection would be formalized exactly

in the same way as we have modeled a connection made of many functional contacts with a single docking site.

Thus, our model ofM contacts may be reinterpreted as a model with a single active zone withM docking sites

(release sites) where no univesicular release paradigm is adopted.

Page 130:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

116 Chapter 4: Synaptic current produced by synchronized neurons

Poisson

Figure 4.5: Schematic description of a model connection with two synaptic contacts (M =

2). Input spikes arrive along the axon and aredeliveredto each of the contacts. Thus, the

spikes that arrive at both contacts are completely synchronous (see dotted lines). At the same

time, each contact is embedded by its own vesicle machinery. Both recovery of vesicles and

release upon arrival of a spike, occur at each contact in an independent manner. Nevertheless,

this does not mean that the probability of a spike triggering a release at one of the contacts

is independent of what happens in the other contact (see text)

not explicity written, this quantity is computed in the stationary state. The derivation of

〈pv(α, β)〉 is neither trivial nor straightforward. To achieve it, we need to solve the stationary

state of a system like the one illustrated in figure4.5. This is a system composed of two

contacts which share the same input but have independent vesicle refilling dynamics. Since

we are only concerned about the second order correlations, it is enough to solve a system

with two contacts instead of including all theM contacts. The calculation is described in

detail in AppendixE. The result is the following:

〈pv(β|α)〉 =1

[1 + Uντv (1− U/2)](4.4.6)

As we see, this conditioned probability (eq.4.4.6) differs of the non-conditioned (〈pssv 〉 =

11+Uντv

) demonstrating that the two processes are not independent. However, if we would

only care about themeannumber of releases upon arrival of a spike, using a binomial distri-

bution of parametersM and〈pssv 〉U , will lead us to the right solution (which isM〈pss

v 〉U ).

This happens because the mean is independent of the second order correlations. Never-

theless, we need to calculate the two-point correlation of the current so that the binomial

hypothesis fails. We can add here that, if at the spike arrival we know that there areN sites

Page 131:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.4. Statistics of the synaptic releases 117

ready, the total number of releases would actually be a binomial of parametersN andU

[Matveev and Wang, 2000b].

By inspection of〈pv(β|α)〉 in eq. 4.4.6, we can state that it is always larger than the

non-conditional one, and that the difference among them is maximal the larger the product

ντv (when depression is strong and the response saturates) and when the release is reliable.

4.4.3 Release correlations among two synapses from different neurons

As we established in section4.3, spike trains coming from different neurons display

spatial cross-correlations of temporal range zero (synchrony) which are quantified by the

parameterρ, defined in expression4.3.6. It has been named asρ because it coincides with

the correlation coefficient of the spike count defined as:

ρ ≡ 〈ni(T )nj(T )〉 − 〈ni(T )〉〈nj(T )〉√〈(ni(T )− 〈ni(T )〉)2〉 〈(nj(T )− 〈nj(T )〉)2〉

(4.4.7)

The quantityni(T ) refers to the number of spikes produced by neuroni in a time windowT ,

and the angular brackets represent an average over repetitions of the stimulus. For the zero-

lag correlations defined (eq.4.3.6), this expression does not depend on the time window

T .

We now turn to investigate how these spatial correlations are passed on to the synaptic

releases. We start by computing the conditional probability〈pv(j|i)〉 of having a release at

thej-th synapse given there is been one, at the same time, at thei-th synapse. The calcula-

tion resembles the one of〈pv(β|α)〉, except that now the trains arriving at each contact are

different and only a fraction of the spikes are synchronized. This is caricatured in figure4.6

where the two synapses are stimulated with spikes from two distinct Poisson processes, and

a fraction of them fall temporally aligned (see vertical blue dashed lines in fig.4.6).

The calculation ofpv(j|i) is described in AppendixF. The final expression reads

〈pv(j|i)〉 =1[

1 + Uντv (1− Uρ2

)] (4.4.8)

A few limits can be verified by simple inspection: i) When we makeρ = 1 theC afferent

fibers provide the same spike train, and we recover the conditioned probability〈pv(β|α)〉of eq. 4.4.6. ii) In the limit ρ = 0, the conditional probability becomes equal to the non-

conditioned probability〈pssv 〉 = 1

1+Uντv, indicating that, since the inputs are now indepen-

dent, the docking of vesicles becomes also independent.

We are ready to go now through the main calculation of this chapter which consists in

deriving the first and second order statistics of the total afferent current.

Page 132:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

118 Chapter 4: Synaptic current produced by synchronized neurons

ρ

Poisson 2

Poisson 1

Figure 4.6: Schematic description of two synaptic contacts belonging to different pre-

synaptic neurons whose activity is correlated. A fraction of the input spikes coming along

each axon arrives at the contacts at the same time, they are synchronized. Each contact has

its own independent synaptic machinery which produces unreliability and short-term depres-

sion. However, this does not imply that the probability of a spike triggering a release at one

of the contacts is independent of what happens in the other contact unless the coefficient of

correlation equals zeroρ = 0 (see text)

4.5 Statistics of the total afferent current

At this point, we have gathered all the elements needed to compute the statistics of the

afferent current, which has been parameterized as described in section4.2. We include in eq.

4.2.5the generalization that describes the connection between two neurons asM functional

contacts, leading to the following expression,

I(t) =C∑i

M∑α

Cm Ji,α

∑k

δ(t− tki,α) (4.5.1)

Consequently, we have adopted a further simplification: we assume that the number of

contacts that theC pre-synaptic cells establish is the same. We also assume that EPSP’s sizes

Ji,α are distributed according to a probability distributionf(J) with meanJ and variance

J2∆2 (i.e. the new parameter∆ is the coefficient of variation of theJ ’s). This distribution of

synaptic efficacyJ , may account for different aspects which were not explicitely modeled:

the heterogeneity of the response amplitudes across synapses due to variable size of synaptic

vesicles [Atwood and Karunanithi, 2002], the variable number of post-synaptic receptors

[Walmsley et al., 1998]; the spatial distribution of synaptic contacts across the dendritic

arbor which strongly influences the depolarization at the soma [Segev and Rall, 1998]. The

Page 133:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.5. Statistics of the total afferent current 119

variability observed in the quantal response amplitudes of single cortical synapses [Korn

et al., 1984, Larkman et al., 1997] may not be included in this list since this is a variability

from release to release, whereas in this model the singleJi,α are constants extracted from the

population distributionf(J). In order to avoid confusion between the membrane capacitance

Cm and the number of pre-synaptic neuronsC, we drop the dependence on the first from the

current, and, later on, when considering the equation of the integrate-and-fire neuron we will

put it back.

4.5.1 The mean of the afferent current

We are interested now in computing the mean of the afferent current, that is

µ = < I(t) > =C∑i

M∑α

Ji,α <∑k

δ(t− tki,α) > (4.5.2)

The series of deltas between brackets in the r.h.s. equals the average number of releases

found at timet. This quantity equals to the average number of spikes arriving at timet,

times the fraction of them which elicited a release. This fraction is exactly the probability of

success of a spike in triggering the release, which in turn equals the probability〈pssv 〈, times

the probability of releaseU . Putting all together

µ = =∑C

i

∑Mα Ji,α ν 〈pss

v 〈 U =C∑i

∑αM

Ji,αν U

1 + Uντv

(4.5.3)

Since the number of post-synaptic contacts a cortical neuron makes is of the order of104,

i.e. N ≡ CM = 104, the sum of synaptic efficacies self-average leading to

µ = C M JUν

1 + Uντv

(4.5.4)

This expression can be interpreted as the mean current ofCM inputs bombarding events of

strengthJ at a rateνr = Uν1+Uντv

. Two things must be remarked: concerning the first order

statistics i) there is no difference between contacts coming from the same neuron or from

different neurons, in other words, the meanµ depend only on the productCM = N which

is the total number of contacts. ii) there is no trace of the spatial correlation coefficientρ6.

6In chapter2 we derived the release rate in a single contact when the spike train had auto-correlations. That

rateνr depended on the input auto-correlations parametersCVisi andτc (eq.2.3.8). On the contrary, the release

rate now, does not show any trace of the cross-correlations. This can be explained because there, the temporal

structure of a single afferent affected the recovery dynamics of synapse, while now, individual contacts have

no information about cross-correlations and only see Poisson trains.

Page 134:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

120 Chapter 4: Synaptic current produced by synchronized neurons

Once again, we underline the most characteristic signature of short-term depression

present in the mean current: saturation. The mean current is upper bounded by what we

will call the limit mean current,µlim, which reads

µlim =CMJ

τv

(4.5.5)

The discussion about the consequences of saturation in terms of information transmission

from the pre to the post-synaptic terminal has been extensively covered in chapter2. At the

end of the next chapter new important consequences, at the level of the response of the output

neuron, will be presented.

4.5.2 Correlations of the current

We define the correlation function of the current as

C(t, t′) ≡ < (I(t)− < I(t) >) (I(t′)− < I(t′) >) >= < I(t) I(t′) > −µ2(4.5.6)

Now let us work out the first term of the r.h.s. of this equation. Substituting the definition of

the current eq.4.5.1, we obtain

< I(t) I(t′) > = <C∑i,j

M∑α,β

Ji,αJj,β

∑k,l

δ(t− tki,α)δ(t′ − tlj,β) > (4.5.7)

We break this expression into three terms, namely i) one that represents theauto-correlations

at the single contacts (termA(t′− t)), ii) one that represents thecross-correlationsbetween

pairs of contacts with common pre-synaptic cell (termB), iii) and the last that constitutes

thecross-correlationsbetween pairs of contacts which receive spikes coming from different

pre-synaptic neurons (termC(t′ − t)).

< I(t) I(t′) > =C∑i

M∑α

J2i,α <

∑k,l

δ(t− tki,α)δ(t′ − tli,α) >

︸ ︷︷ ︸same contact:A(t′−t)

+ (4.5.8)

+C∑i

M∑α 6=β

Ji,αJi,β <∑k,l

δ(t− tki,α)δ(t′ − tli,β) >

︸ ︷︷ ︸same pre−neuron:B(t′−t)

+ (4.5.9)

+C∑

i6=j

M∑α,β

Ji,αJj,β <∑k,l

δ(t− tki,α)δ(t′ − tlj,β) >

︸ ︷︷ ︸diff. pre−neurons: C(t′−t)

(4.5.10)

We will treat each term separately so that we can carry out the operations step by step.

To be more concise, we will make use of two previously defined variables that we now write

Page 135:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.5. Statistics of the total afferent current 121

again: i)νr which represents the probability of finding a synaptic response in a single contact,

at any time:

νr =Uν

1 + Uντv

(4.5.11)

ii) The time constant of the dynamics of〈pv(t)〉 (see eq.4.4.3):

τc ≡τv

1 + Uντv

(4.5.12)

4.5.2.1 Auto-correlation in single contacts

The sum of deltas within the brackets in the first termA(t′ − t), equals simply the non-

connected auto-correlation function of the releases in one contact for Poisson input, com-

puted in chapter2 (see eq.2.3.2) In that chapter we defined a related quantity: the condi-

tional rate. This function is obtained from the correlation by connecting it, dropping the delta

function and normalizing by the rate (see eq.2.2.5):

νcr(t) = νr (1− e−t/τc) t > 0 (4.5.13)

We will use it here for the sake of clarity.

Now, the termA(t′ − t) can be operated obtaining

A(t′ − t) =C∑i

M∑α

J2i,α [νr δ(t′ − t) + νrν

cr(|t′ − t|)] = (4.5.14)

=C∑i

M∑α

J2i,α

[νr δ(t′ − t) + ν2

r (1− e−|t′−t|/τc)

](4.5.15)

Now applying the self-averaging argument explained above one obtains

A(t′ − t) = C M J2[νr δ(t′ − t) + ν2

r (1− e−|t′−t|/τc)

]= (4.5.16)

= C M J2 (1 + ∆2)[νr δ(t′ − t) + ν2

r (1− e−|t′−t|/τc)

]where we have used the relation between the variance, and the first and the second moments

of J :

V ar[J ] = ∆2 J2 = J2 − J2 (4.5.17)

4.5.2.2 Cross-correlation between pairs of contacts with the same input train

The second termB(t′ − t) representing the correlation between responses generated in

contacts sharing the same pre-synaptic neuron is more complicated:

B(t′ − t) =C∑i

M∑α 6=β

Ji,αJi,β [νr δ(t′ − t) 〈pv(β|α)〉U+ (4.5.18)

Page 136:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

122 Chapter 4: Synaptic current produced by synchronized neurons

+ νr 〈pv(β|α)〉 U νcr(|t′ − t|) + (4.5.19)

+ νr 〈pv(β|α)〉 (1− U) ν 〈p∗v(|t′ − t|)〉 U + (4.5.20)

+ νr (1 − 〈pv(β|α)〉) νcr(|t′ − t|) ] (4.5.21)

The four terms inside the square brackets represent theprobability of occurrence of four

situations which cover the whole event’s space, namely

4.5.18 There is a release in contactα (νr); therefore, there is a spike hittingβ at the same

time (δ(t′− t)); the vesicle atβ is ready, known that the one atα was prepared as well

(〈pv(β|α)〉); the vesicle atβ is also released (U ).

4.5.19 There is a release in contactα at t (νr); the vesicle atβ was ready att, known the

one atα was prepared as well (〈pv(β|α)〉); the vesicle atβ was then released att (U );

there is a new release atβ after |t′ − t| time, known that att+ the contact was empty

(νcr(|t′ − t|)).

4.5.20 There is a release in contactα at t (νr); the vesicle atβ was ready att, known the one

at α was prepared as well (〈pv(β|α)〉); the release atβ fails to succeed att (1 − U );

there is a new spike arriving at timet′ (ν); the vesicle is ready att′ known it was ready

at t (〈p∗v(|t′ − t|)〉); the recovered vesicle atβ is released att′.

4.5.21 There is a release in contactα at t (νr); the vesicle atβ wasnot ready, however, at that

time ((1 − 〈pv(β|α)〉)); there is a release atβ at timet′ known that att+ the contact

was empty (νcr(|t′ − t|)).

The only new term we have introduced is the evolution of the probability of the vesicle

to be ready〈p∗v(|t′− t|)〉. This is just a particular case of the general solution of〈pv(t)〉 given

by equation4.4.3with initial condition equal to one, i.e.〈pv(0)〉 = 1:

〈p∗v(|t′ − t|)〉 = 〈pv(0)〉 e−|t′−t|/τc + 〈pss

v 〉 (1− e−|t′−t|/τc) =

= e−|t′−t|/τc +

1

1 + Uντv

(1− e−|t′−t|/τc) (4.5.22)

Of the four terms listed, the first one is a positive correlation of time lag zero, caused

by the fact that the contactsα andβ are stimulated with the same input train. The other

three will turn out to be a negative exponential due to depression,. If we just substitute, the

functions〈pv(β|α)〉, 〈p∗v(|t′ − t|)〉 andνcr(t), and consider that the two summations of the

product of synaptic efficacies inB(t′ − t) also self-average, after some simple algebra one

Page 137:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.5. Statistics of the total afferent current 123

obtains

B(t′ − t) = C M (M − 1)J2

[νr δ(t′ − t)

U (M − 1)

1 + Uντv

− (4.5.23)

− ν2r e−

|t′−t|τc

U (M − 1) (1 + Uντv/2)

1 + Uντv

+ ν2r

](4.5.24)

It is interesting to detect that depression, which introduces negative auto-correlations in a

single contact, spreads out its effects to the cross-correlations of two contacts receiving the

same spike train. In other words, if one observes a release inα at timet, during some period

(of the orderτc) there will be less chances to have a release inβ due to depression, even if we

have no information whether there was a release inβ at that timet! Later it will be shown

that this also happens when the contacts do not receive the same trains, but only correlated

trains. The stronger this correlation, the larger will be the influence of depression in the

cross-correlograms.

4.5.2.3 Cross-correlation between pairs of contacts with correlated input trains

We finally focus on the third termC(t′−t) of the correlation function of the current, which

account for the cross-correlations between synaptic contacts from different pre-synaptic neu-

rons. Its structure is very much the same as the termB(t′−t), except that now the conditioned

probabilities are〈pv(j|i)〉 instead of〈pv(β|α)〉.

C(t′ − t) =C∑

i6=j

M∑α,β

Ji,αJj,β [νr ρ δ(t′ − t) 〈pv(j|i)〉U+ (4.5.25)

+ νr ρ 〈pv(j|i)〉 U νcr(|t′ − t|) + (4.5.26)

+ νr ρ 〈pv(j|i)〉 (1− U) ν 〈p∗v(|t′ − t|)〉 U + (4.5.27)

+ νr ρ (1 − 〈pv(j|i)〉) νcr(|t′ − t|) + (4.5.28)

+ νr (1 − ρ) ν 〈p∗∗v (|t′ − t|)〉 U ] (4.5.29)

The first four terms inside the squared brackets are equivalent to the four terms in the ex-

pression ofB(t′ − t) and do not need further comment, except the presence ofρ which is

needed to consider the situations in which two spikes arrive synchronously at contactsi and

j. The fifth term within the squared brackets was not present in the expression ofB(t′ − t).

It describes the situation in which a spike arrives at terminali at t and evokes a response

(νr); there is not a synchronouspartner arriving at j (1 − ρ); then, another spike arrives

at j (ν); the vesicle is now present (〈p∗∗v (|t′ − t|)〉) and is finally released. The probability

〈p∗∗v (|t′ − t|)〉 of finding j occupied, when all we know is that the vesicle ini was prepared

at timet, is obtained by the temporal evolution of〈pv(t′)〉 from an initial condition att equal

to 〈pv(j|i)〉 (see eq.4.4.3):

Page 138:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

124 Chapter 4: Synaptic current produced by synchronized neurons

〈p∗∗v (|t′ − t|)〉 = 〈pv(j|i)〉 e−|t′−t|/τc + 〈pss

v 〉 (1− e−|t′−t|/τc) = (4.5.30)

=e−|t

′−t|/τc

1 + Uντv(1− Uρ/2)+

(1− e−|t′−t|/τc)

1 + Uντv

Now, substituting the expressions of〈pv(j|i)〉 (eq. 4.4.8), 〈p∗v(|t′ − t|)〉 (eq. 4.5.22),

〈p∗∗v (|t′ − t|)〉 (eq. 4.5.30) and the the conditional release rateνcr(|t′ − t|) (eq. 4.5.13) into

eqs.4.5.25-4.5.29and self-averaging the summation of the synaptic efficacies one obtains,

C(t′ − t) = C(C − 1)M2 J2

[νr δ(t′ − t)

U ρ(C − 1)M

1 + Uντv(1− Uρ/2)− (4.5.31)

− ν2r e−

|t′−t|τc

Uρ(C − 1)M(1 + Uντv/2)

1 + Uντv(1− Uρ/2)+ ν2

r

](4.5.32)

4.5.2.4 Total current correlation

We have expressed the correlation function of the current as

C(t, t′) = < I(t) I(t′) > −µ2 = A(t′ − t) + B(t′ − t) + C(t′ − t) − µ2 (4.5.33)

We now need to introduce the last two variables which will parameterize this correlation

function as done byMoreno et al.[2002]. The delta varianceσ2w and theexponential vari-

anceΣ2, will be the coefficients of each of the two correlation terms in the following way:

C(t, t′) = σ2w δ(t′ − t) +

Σ2

2τc

e−|t′−t|

τc (4.5.34)

If we operate equation4.5.33above and identify these new coefficients, their expression

after some algebra read

σ2w = C M J2 νr

[(1 + ∆2) +

U(M − 1)

1 + Uντv(1− U/2)+

Uρ(C − 1)M

1 + Uντv(1− Uρ/2)

]

=C M J2 Uν

1 + Uντv

[(1 + ∆2) +

U(M − 1)

1 + Uντv(1− U/2)+

Uρ(C − 1)M

1 + Uντv(1− Uρ/2)

](4.5.35)

and

Σ2 = −2C M J2 τc νr

[(1 + ∆2) +

U(M − 1)(1 + Uντv/2)

1 + Uντv(1− U/2)+

Uρ(C − 1)M(1 + Uντv/2)

1 + Uντv(1− Uρ/2)

]

=−2C M J2 U2ν2τv

(1 + Uντv)3

[(1 + ∆2) +

U(M − 1)(1 + Uντv/2)

1 + Uντv(1− U/2)+

Uρ(C − 1)M(1 + Uντv/2)

1 + Uντv(1− Uρ/2)

](4.5.36)

Page 139:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.5. Statistics of the total afferent current 125

Each of the three terms of the sum within the squared brackets, come from the auto-

correlation at a single contact (A(t′ − t)), the cross-correlation between contacts sharing

the same pre-synaptic neuron (B(t′ − t)), and from the cross-correlation between contacts

with synchronized inputs (C(t′ − t)). The most interesting aspect of the last two terms is

the effect of the interaction between the positive correlations, regardless whether they come

from the several contacts per neuron (in the first) or fromad hocsynchrony (in the second),

and STD. This interaction is expressed in the dependence of their denominators on the input

rate. When the productUντv becomes large enough, the trace of the correlation vanishes.

In other words, the effect of the positive correlations will be less and less detectable as the

synapses start to saturate.

We will define here the quantityσ2lim, which will be helpful to describe the behavior of

the current as a function of the input rate, and is defined as follows:

σ2lim ≡ lim

ν→∞σ2

w =C M J2 (1 + ∆2)

τv

(4.5.37)

We do not define the limit ofΣ2 because it is simply zero.

The second interesting feature is the symmetry which exists between the two terms of the

cross-correlations. More exactly, for every choice of the parameters not including synchrony

(ρ = 0) but including several contacts per connection (M > 1), there is an “equivalent”

election of parameters withM = 1 andρ > 0, which produce approximately the same

statistics of the current. The transformation is the following

C

M > 1

ρ = 0

−→

C ′ = CM

M ′ = 1

ρ′ ' 1C

(4.5.38)

This transformation leads to a new current with the same meanµ and with variancesσ2w

andΣ2 which areapproximatelythe same. Although the numerators of the cross-correlation

terms are invariant under this transformation, the denominators are not, so this is just an

approximate invariant transformation. However, the qualitative behavior is very similar and

the limiting value of the delta varianceσ2lim is just the same. Therefore, from hereafter we

may talk about the positive correlations present in the input referring to the two somehow

equivalentsources of positive correlations we have implemented in the input:M > 1 and

ρ > 0.

Now, what is the intuitive meaning of the delta varianceσ2w and the exponential variance

Σ2 and how do they affect the response of the neuron if they do at all? To answer this question

Page 140:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

126 Chapter 4: Synaptic current produced by synchronized neurons

we will first give a more qualitative argument in terms of the familiar quantity of the variance

of the event count, and in the next chapter introduce the mathematical formalism to quantify

the effects of the correlations on the response of a model neuron.

The variance of the synaptic release count

Since the current has been finally simplified to a summation of point events which are the

releases (eq.4.5.1) the noise around the mean valueµ that the output neuron will perceive,

is related to the variance of the number of releases falling in a time window equal to the

integration timeτm. We can easily defined the stochastic train of releases as

R(t) =C∑i

M∑α

∑k

δ(t− tki,α) (4.5.39)

By inspection we can realize thatR(t) equals the currentI(t) if we assume thatJi,α = 1 for

all i andα. Thus, we can infer that the connected cross-correlation function of the releases,

CR(t, t′), equals that of the current, when we makeJ = 1 and∆ = 0.

Now if we define the random variableN(T ) as the number of releases that occur within a

time window of lengthT , its varianceσN(T ), which depends onT , is related to the connected

cross-correlation of the releasesCR(t, t′) through (see e.g. [Rieke et al., 1997])

σ2N(T )(t) =

∫ t+T

t

∫ t+T

tds ds′ CR(s, s′) (4.5.40)

Since our input is stationary, the result must be independent of the absolute timet. We

now take equation4.5.34and integrate it,obtaining

σ2N(T ) = σ2

w T + Σ2

[T − τc(1− e−T/τc)

](4.5.41)

It must be remembered, that bothσ2w and Σ2 appearing in this equation are obtained

from the original ones (eqs.4.5.35and4.5.36) by makingJ = 1 and∆ = 0. Finally, we

can check the dependence ofσ2N(T ) on the length of the time windowT . If the time window

T << τc we expand the exponential up to first order inTτc

and the dependence onΣ2 vanishes

resulting inσ2N(T ) ' σ2

w T . On the opposite limitT >> τc we neglect the term proportional

to τc obtainingσ2N(T ) ' (σ2

w + Σ2) T . So in both cases the variance of the count grows

linearly withT (which is a general property of renewal processes [Cox, 1962]), but the slope

depends on the ratioT/τc. It is easy to observe thatfor the same pricewe have computed

the variance of the input current integrated in a window of durationT defined as:

IT (t) =∫ t+T

tdsI(s) (4.5.42)

Page 141:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

4.5. Statistics of the total afferent current 127

and whose variance reads

σ2IT

(t) =∫ t+T

t

∫ t+T

tds ds′ C(s, s′) (4.5.43)

= σ2w T + Σ2

[T − τc(1− e−T/τc)

]where the expression is identical to that ofσ2

N(T ) but now the coefficientsσ2w andΣ2 are those

defined in eqs.4.5.35and4.5.36. Now a pertinent question to prompt is: what is the noise

magnitude perceived by the neuron when integrates this current? or in other words, which

is the relevant time window size in which we should measure the variance of the current?

This is a profound and difficult question, to which we will give an heuristic answer based

on the empirical analysis of the response function of the neuron (see [Moreno et al., 2002]

for details). As it will be shown later, the current fluctuations detected by the neuron are

those observed at the temporal scale given by the membrane time constantτm (see section

5.1). This implies that the variance of the current integrated over a time windowτm, σ2Iτm

, is

the relevant noise measure. As a consequence, if the time scale of the negative exponential

correlations induced by synaptic depression is much longer than the membrane time constant,

τc >> τm, thenσ2Iτm

= σ2w τm and depression will have barely any effect on the input noise.

On the contrary ifτc << τm, thenσ2Iτm

= (σ2w + Σ2) τm and the input variance is then

affected by the negative correlations introduced by depression, which, due toΣ2 < 0, will

diminish the magnitude of the fluctuations. If we define an effective varianceσ2eff , in this

last situation it would equalσ2eff = σ2

w + Σ2 whereas in the case of long range correlations

it will equal σ2w. This addition of the two variancesσ2

w andΣ2, whenτc is small, can be

predicted by inspection of the correlation function of the currentC(t′ − t) of eq. 4.5.34. If

we take the limit of this function asτc goes to zero, we obtain that the exponential becomes

a Dirac delta function centered int′− t, so that that the two deltas add. The result is a single

delta function with a coefficient equal toσ2w +Σ2. The next question we have to address then

is, does it ever happen thatτc << τm? If we recall the expression ofτc it reads

τc =τv

1 + Uντv

(4.5.44)

Sinceτv ∼ 500 − 1500 ms andτm ∼ 10 − 20 ms, we immediately realize thatτc << τm

will only be possible ifUντv > 100 which means that more than a hundredU -diluted spikes

would reach the synapse during the time it takes, on average, to a new vesicle to get ready,τv

. In this saturating regime the release statistics are mainly dictated by the docking dynamics

of the vesicles, which were set to be Poisson. The negative correlations then arise because

it is not exactly the case that when a vesicle docks a release occurs (which would imply a

de-correlated train of releases). After the vesicle docks, a short time elapses until a new

spike makes it fuse the membrane. The resulting output process seems like a Poisson with a

Page 142:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

128 Chapter 4: Synaptic current produced by synchronized neurons

variable refractory period of mean length1Uν

(the mean time it waits a prepared vesicle until

a spike makes it undergo exocytosis). Thus, in this regime, there are negative correlations

of time scale 1Uν

in the release process. The quantitative effect of these short-range negative

correlations will be illustrated in the next chapter.

Page 143:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Chapter 5

The response of a LIF neuron with

depressing synapses

5.1 The Leaky Integrate and fire (LIF) neuron

The model of neuron we have chosen to investigate the impact of synchrony and synaptic

depression on the output response, is the well-known integrate-and-fire (IF) neuron (see e.g.

[Ricciardi, 1977, Tuckwell, 1988]). Due to its simplicity, this model has allowed neurosci-

entists to extract, by means of analytical calculations and simulations, correct basic ideas

about the function of single neurons and neural networks [Amit and Tsodyks, 1991, Shadlen

and Newsome, 1994, Amit and Brunel, 1997, Wang, 2001]. Its first assumption consists in

reducing the neuron to a point in space, neglecting the spatial distribution of the dendrites

and assuming that the potential inside the neuronV , which actually should be a space and

time functionV (x, y, z, t), depends only on time,V (t). Besides, the mechanism underlying

the generation of an action potential, are quite stereotyped: when the membrane potential of

a neuron reaches a value around−55 mV it approximately follows a stereotyped trajectory.

Therefore the IF model avoids to explicitely model its generation, and focuses on how the

neuron integrates the current when the membrane potential is below threshold. In its sim-

plest version, the integrate-and-fire model ignores the active conductances of the membrane

and reduces the properties of all membrane channels to a single passive leakage conductance

gL, which produces a leak current which pushes the membrane potential towards the resting

valuesEL. Thus, in the Leaky integrate-and-fire model (LIF), the membrane acts as a capac-

itor with a resistor set in parallel. The synaptic current loads the capacitor with positive or

negative charge (depending on whether the inputs are excitatory or inhibitory) and the leak

current allows for the capacitor to unload. This is expressed by the first order differential

129

Page 144:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

130 Chapter 5: The response of a LIF neuron with depressing synapses

equation

CmdV (t)

dt= gL(EL − V (t)) + I(t) , if V (t) < θ (5.1.1)

whereCm is the capacitance of the membrane andθ is the voltage threshold. Whenever the

potential reaches this threshold, a spike is emitted and the potential is artificially placed at a

reset value,V (t+) = H, where it remains constant during a refractory timeτref . Thus, the

generation of an AP and refractoriness, are explicitely modeledad hoc. The equation can

be rewritten in more common and simpler manner, if one multiplies both sides by1gL

and

changes the origin of the potential settingEL = 0:

τmdV (t)

dt= −V (t) + RmI(t) , V (t) < θ (5.1.2)

whereτm = CmRm is what is called the membrane time constant, and the membrane resis-

tanceRm is just the inverse of the leak conductance. Thus, the four parameters of the model,

namelyτm, Rm, θ andτref are free parameters which have to be phenomenologically fitted

from single intra-cellular recordings.

In a recent work,Rauch et al.[2002] have explored the validity of the LIF neuron model

by comparing the theoretical output rate prediction, with the response of a real neuron stim-

ulated in vitro through a dynamic clamp with an artificially generated current resembling

synaptic activity. The results seem to demonstrate how in almost all cases the model param-

eters could be fitted so that the model neuron imitates the response in the experiment with

great accuracy.

The choice of the LIF neuronal model turns out to be twofold advantageous: i) it is very

simple to be programmed and computationally inexpensive, allowing us to run simulations

of long time periods without consuming too much CPU time; and ii) what is more important,

we will make use of a well developed set of analytical tools with which we will obtain a

theoretical expression for the output rate of a LIF neuron integrating a current produced by

the activity of depressing synapses receiving synchronous inputs.

5.2 The analytical calculation of the output rate of a LIF

neuron

In this section we will give a brief overview of the mathematical tools coming from the

theory of stochastic processes, which will allow us to obtain a formula for the output rate of

a LIF neuron excited by the current described in previous sections. Therefore our starting

point will be the equations describing the statistics of the current, eq.4.5.4and eq.4.5.34for

Page 145:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.2. The analytical calculation of the output rate of a LIF neuron 131

the meanµ and the correlation functionC(t′ − t), and the equation describing the evolution

of the membrane potentialV (t) (eq.5.1.2).

Our aim is to describe the dynamics of the potential of the neuronV (t) as it follows from

the differential equation of the LIF model (eq.5.1.2) when the current termI(t) is replaced

by the synaptic current given in eq.4.5.1:

τmdV (t)

dt= −V (t) + Rm Cm

C∑i

M∑α

Ji,α

∑k

δ(t− tki,α) (5.2.1)

Using the definition of the membrane time constantτm ≡ RmCm this can be written as:

dV (t)

dt= −V (t)

τm

+C∑i

M∑α

Ji,α

∑k

δ(t− tki,α) (5.2.2)

The trajectory followed byV (t) can be described easily: the arrival of a release at timetki,αproduces a discontinuity in its trajectory by a jump of sizeJi,α. This is an EPSP. After the

jump, V (t) smoothly relaxes to zero in an exponential manner. Because the generation of

releases is a stochastic process, the potential itself becomes also a stochastic process, which

is continuous in time (the time is not discretized), but discontinuous in space, that is, the

trajectory ofV (t) in the voltage axis is not continuous.

The problem of quantifying the output rate of a LIF neuron given the statistics of the

input current, is equivalent to solve themean first-passage timeproblem (see e.g. [Ricciardi,

1977, Tuckwell, 1988]). This refers to compute the mean of the random variableT defined

as the first time the potential hits thresholdV (T ) = θ when the neuron has emitted a spike at

time zero and has then been reseted toH . In other words,T is the inter-spike-interval (ISI),

and its mean< T > gives us the output rate through

νout ≡1

< T >(5.2.3)

However, solving this problem whenV (t) is discontinuous is a rather complex job, that

can be simplified by what is called the diffusion approximation.

5.2.1 The diffusion approximation

In order to characterize the trajectories of the potential as continuous, an extra assumption

must be made on the input current. Let us suppose that we replace the input current, which is

a stochasticpoint process, by aGaussianprocessG(t) (see e.g. [Ricciardi, 1977]) with the

same mean< G(t) >= µ and correlation function< G(t)G(t′) >= C(t− t′). Because this

Page 146:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

132 Chapter 5: The response of a LIF neuron with depressing synapses

process is continuous in time, the potentialV (t) becomes a continuous variable too. Now the

question is, when is this substitution a good approximation? The diffusion approximation

turns out to begood (it roughly gives the same results as the point process current) when

two conditions are fulfilled: i) the sizes of the EPSP are small compared with the distance

from the reset potential to the threshold, i.e. the number of simultaneous jumps required for

the potential to reachθ starting fromH must be large. ii) The number of synaptic releases

arriving in a time window of the order ofτm must be large. The second condition is naturally

satisfied since the number of impinging afferent synapses received by a cortical neuron is of

the order of103 − 104 [Braitenberg and Schuz, 1991], even in the case where they fire at

low spontaneous rates of about∼ 2 Hz, the number of events occurring in a time window

of length τm is of the order of 100 [Shadlen and Newsome, 1998]. The first assumption

depends critically on the size of the EPSC’s (and IPSC’s) entering the neuron when a vesicle

release is triggered, but it could also be violated if for instance the synchronous arrival of

n releases occur frequently, resulting (in our simple linear model of integration of EPSC’s)

in a EPSP which would ben times larger than the individual ones. Experiments have re-

ported a wide range of EPSP’s and EPSC’s sizes, which in many cases is hard to discern if

the depolarization was monosynaptic (produced by the release of one transmitter vesicle) or

was the superimposed print of the synchronous releases occurring in multiple contacts. An

interesting way to measure the amplitude of the EPSP which leads to no confusion about

whether is a monosynaptic or multi-synaptic depolarization, consists in measuring the am-

plitudes of the spontaneous EPSC’s evoked by the spontaneous release of vesicles. Because

in this case release is not triggered by spiking activity, synchronous releases are rather rare.

Typical numbers to the size of EPSC’s range from5 to 30 pA [Berretta and Jones, 1996,

Stevens and Zador, 1998], which taking a membrane resistanceRm ∼ 10 − 100 MΩ gives

an EPSP range ofJ ∼ 0.01 − 0.3 mV. However, double whole-cell patch clamp recordings

in cortical pyramidal neurons, in which the pre-synaptic neuron is stimulated to fire, and the

EPSP’s are recorded at the soma of the post-synaptic cell, give EPSP’s amplitudes which can

be bigger than1 mV. ([Markram and Tsodyks, 1996a], see also [Larkman et al., 1997] who

used the method ofminimal extracellular stimulationin hippocampal CA1 synapses). This

prominent depolarization is likely to occur because of multiple contacts existing between

cells.

If in addition, we add an extra degree of synchronizationρ in the activity of the afferent

cells, the probability that the membrane potential undergoes large discontinuities, due to the

arrival APs at the same time and the subsequent release of vesicles in multiple contacts,

increases and spoils the diffusion approximation. Nevertheless, in a wide range of values of

J andM this approximation will turn out to give an excellent fit of the simulation data.

Page 147:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.2. The analytical calculation of the output rate of a LIF neuron 133

5.2.2 The solution ofνout for a white noiseinput

Let us first consider the case where the input current can be approximated by a stationary

Gaussian processG(t) with mean< G(t) >= µ and connected correlation function of the

form

C(t′ − t) = < G(t)G(t′) > −µ2 = σ2wδ(t′ − t) (5.2.4)

Because the correlation function has only the delta term, its Fourier transform is just a con-

stant, expressing that all frequencies are equally represented. For this reason, this type of

stochastic process is commonly called awhite noise. If in the case of our input, depression

were negligible (τv ∼ 0), thenΣ2 ∼ 0 (see eq.4.5.36) and the correlation function would be

a white noise. This does not mean that all correlations in the input have vanished. The cross-

correlations due to synchrony in the pre-synaptic population (ρ) and the cross-correlations

due to multiple-contacts are still there and its effect is contained inσ2w (see eq.4.5.35). How-

ever, these two sources of cross-correlations produce zero-lag correlations, which do not

change the white noise nature of the input, that is based on the fact that at two different in-

stants of time, the process is uncorrelated. Therefore, including zero-lag correlations (like

perfect synchrony or multiple contacts) has the effective result of a renormalizing the input

varianceσ2w [Salinas and Sejnowski, 2000, Moreno et al., 2002].

If now we make a LIF neuron integrate the white noise defined by the two parametersµ

andσ2w, the solution of the mean first passage time reads [Ricciardi, 1977, Tuckwell, 1988]

ν−10 =< T >0= τref +

√πτ∫ Θ

Hdtet2(1 + erf(t)) (5.2.5)

whereerf(t) is the error function defined as

erf(t) =2√π

∫ t

0dx e−x2

, (5.2.6)

the limits of the integral are functions ofµ andσ2w defined as

H =H − µτ

σw√

τm

(5.2.7)

Θ =θ − µτ

σw√

τm

(5.2.8)

and we have denoted the output rate byν0, to distinguish it from its final expressionνout

when correlations, due to depression, are present.

Page 148:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

134 Chapter 5: The response of a LIF neuron with depressing synapses

5.2.3 Perturbative solution ofνout for a correlated input

The calculation outlined here is described in detail in [Moreno et al., 2002]. If we assume

that the input current is a stationary Gaussian processG(t) with correlation function given

by (see eq.4.5.34)

C(t′ − t) = < G(t)G(t′) > −µ2 = σ2wδ(t′ − t) +

Σ2

2τc

e−|t′−t|/τc (5.2.9)

then the current is completely determined by the following four variables:

• Mean current:µ

• Delta variance:σ2w

• The temporal scale of the correlation, redefined by the parameter:k ≡√

τc

τm

• The exponential varianceΣ2, redefined as thecorrelation magnitude: α ≡ Σ2

σ2w

Now, as we previously mentioned, in the limit whereτc → 0 the exponential becomes a

Dirac delta function and a white noise current with an effective varianceσ2eff = σ2

w + Σ2

is recovered. The solution in this case would be the same asν0 but with the new effective

varianceσ2eff :

ν−1eff = τref +

√πτ∫ Θeff

Heff

dtet2(1 + erf(t)) (5.2.10)

Heff =H − µτm

σeff√

τm

(5.2.11)

Θeff =θ − µτm

σeff√

τm

(5.2.12)

On the other hand whenτc → ∞ the exponential vanishes and we recover exactlyν0.

Nonetheless, an exact analytical solution for a generalτc cannot be computed. However if

the parameterα is small andτc << τm or τc >> τm a perturbative solution aroundνeff and

ν0 respectively can be obtained [Moreno et al., 2002].

Whenτc << τm , the quantitiesk andα are treated as perturbative parameters. The

solutionνout is analytically found by expanding the Fokker-Planck equation associated with

the diffusion process [Moreno et al., 2002] in powers ofk =√

τc/τ , and calculating the

terms exactly for allα = Σ2/σ2w for the zero order, and perturbatively inα ≥ 0 up to the first

non trivial correction for the first order. The obtained firing rate can be written as

νout = νeff − α√

τcτν20R(Θ) (5.2.13)

Page 149:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.2. The analytical calculation of the output rate of a LIF neuron 135

whereR(t) is defined as

R(t) =

√π

2et2(1 + erf(t)) (5.2.14)

In the opposite limitτc >> τm, the perturbative parameter isk−1. The solutionνout up

to the orderk−2, and exact for allα reads

νout = ν0 +D

τc

(5.2.15)

where the auxiliary variableD reads

D ≡ ατ 2mν2

0 [τmν0(R(Θ)−R(H))2

1− ν0τref

− ΘR(Θ)− HR(H)√2

]

In the next section we will compare the output rateνout of a simulated LIF neuron with

the analytical predictions given by equations5.2.13and5.2.15, as a function of the input rate

ν. Because the correlation timeτc depends on the input rate (see eq.4.5.44), we will go from

the regimeτc >> τm to the regimeτc << τm passing through the intermediate regime in

which τc ∼ τm. In order to cover the whole range ofτc we have performed an interpolation

of νout between the two regimes. The interpolating curves have been determined by setting

the firing rate in the smallτc range (τc < τm) asνout = νeff +A1√

τc+A2τc whereA1 andA2

are unknown functions ofα and of the neuron and input parameters. Although the function

A1 has been analytically calculated (eq.5.2.13) for smallα, this procedure takes into account

higher order corrections which match more accurately the observed data for larger values

of α. In any case, the interpolation and the analytical prediction are very similar for small

α. In the large correlation time limit (τc > τ ) the analytically computed expression given

in eq.(5.2.15), νout = ν0 + C/τc, was used. The functionsA1 andA2 were determined by

interpolating these two expressions, with conditions of continuity of its rate and its derivative,

at a convenient interpolation pointτc,inter ∼ τm, which is indeed the only parameter which

has to be fitted.

5.2.4 Several input populations

We have considered so far that theC pre-synaptic neurons have all the same properties

(the same firing rateν, same correlation coefficientρ, same number of contactsM , etc).

However, we will often need to put together cells with common physiological properties

(e.g. excitatory and inhibitory cells) or cells with similar selectivity properties.

In the simulations we have performed we have set a excitatory pre-synaptic population

made up ofCe cells with a mean synaptic efficacyJe > 0, and an inhibitory population of

Ci neurons with mean synaptic efficacyJi < 0. For simplicity we have assumed that neu-

rons from different population are not cross-correlated and the synaptic properties were set

Page 150:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

136 Chapter 5: The response of a LIF neuron with depressing synapses

to be the same, i.e. the synaptic parameters were equal for both types of synapses. Although

these simplifications may seem to over-simplify the natural situation where, for instance, the

dynamical properties of the synapses depend strongly on whether they connect interneurons,

pyramidal neurons or pairs of these two types of cells [Gupta et al., 2000], our purpose id

to understand the combined role of synchrony and depression alone. The more complex sit-

uation in which different pre-synaptic populations make connections with the target neuron

with different properties, such as different magnitudes of depression (U, τv), different num-

ber of contactsM or even display other activity dependent mechanism such as facilitation

[Zucker and Regehr, 2002], is left for future work.

Therefore, in this simplified scenario the expressions of the mean current and of the

variances are modified by simply adding the contributions from each population:

µ = µe + µi (5.2.16)

σ2w = σ2

w,e + σ2w,i (5.2.17)

Σ2 = Σ2,e + Σ2,i (5.2.18)

5.3 Results

In this section we will analyze the implications of synchronization of input spike trains

plus short-term synaptic depression on the response of a neuron which integrates this synap-

tic activity. The results obtained are independent of the neuron model chosen, a LIF neuron,

because, as we will discuss latter, they can be deduced from the properties of the input current

itself. In other words, the most exciting result, which will be a non-trivial non-monotonic

behavior of the output of the neuron, is already present in the input current so that any other

more sophisticated spike generator model (e.g. a Hodgkin-Huxley model) would lead, pre-

sumably, to the same qualitative behavior.

In order to compare different instances in which the input connections are made of dif-

ferent number of contacts, that is examples of currents with differentMs, we need to adopt

a certain criterion to compare things which are equivalent in some sense. ChangingM while

keeping the rest of the parameters constant would imply a new current with a different mean

µ, delta varianceσ2w and exponential varianceΣ2. The change in the output would be due to

the composition of several effects and it would be hard to discriminate what really caused a

certain change. We have adopted, therefore, two criteria to study these situations, in which

the mean currentµ = CMJνr remains constant while varyingM , so that the changes in the

response are entirely due to the change in the current variance:

1. To change the number of pre-synaptic neuronsC so thatCM remains constant. In this

Page 151:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 137

case less input cells would provide the same mean current because they would establish

more synaptic contacts with the target neuron. Besides, with this criterion also the

limit varianceσ2lim = C M J2 (1+∆2)

τv(see eq.4.5.37) remains invariant so that in the

limit where the input rate is very big, all choices ofC andM , such thatCM = const.,

have to converge to the same response statistics. Another advantage of this comparison

criterion is that (as mentioned in section4.5.2.4) it will turn out that setting the number

of contacts larger than oneM > 1 while renormalizing the number of pre-synaptic

cells, is approximately equivalent to introducing a certain amount of synchronyρ > 0

while keepingM = 1 andC constant (see eq.4.5.38). Because of this property, when

studying different values ofM (and renormalizingC) we will also be studying the case

in which while all the other parameters are kept fixed, the synchronization is tuned by

changingρ.

2. The second reasonable way to absorb the change inM keepingµ invariant is to renor-

malize the synaptic efficaciesJ so thatMJ = const. This can be thought of as the case

in which a synaptic bouton creates, from a single active zone, several synaptic special-

izations where transmitter release occurs. This genesis of new contacts would occur,

however, with no increase of the synaptic resources such as transmitter molecules or

post-synaptic receptors but only with a redistribution of synaptic efficacies between

the newly created active zones. This hypothetical redistribution would imply that the

efficacyJi,α at each of theM contacts would decrease because the same amount of

resources (transmitter molecules, transmitter receptors) are now spread and used by

M contacts. This situation may result a little artificial because, both the probability

of releaseU and the recovery rate1/τv would remain constant at each contact while

vesicles would be filled with less glutamate, for instance1. In this second compara-

tive scenario, only the mean rateµ remains constant to variations ofM , and the limit

varianceσ2lim (eq.4.5.37) decreases as1/M . This is in agreement with a naive applica-

tion of the Central Limit Theorem: increasing the number of independent realizations

of the random release, the renormalized summed response approximates the mean re-

sponse with an error that vanishes as1/√

M . In this situation, the limitM →∞ leads

us to the phenomenological model of short-term depression of Tsodyks and Markram

(see e.g. [Tsodyks and Markram, 1997]). In this model the noise coming from the

stochastic recovery of the vesicle, has been washed out by pooling the response from

1Redistribution of the post-synaptic receptors into several zones would imply a rescaling of the synaptic

efficacy parameterJ only in the case where the released transmitter saturates the receptors in such a way that

the amount of synaptic current entering the cell is not bounded by the quantity of released transmitter but by

the number of receptors.

Page 152:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

138 Chapter 5: The response of a LIF neuron with depressing synapses

many independent synaptic contacts (or by averaging over many experimental trials,

as they did to obtain the nice response traces which are accurately fitted by their deter-

ministic model). The consequences of assuming such anaveragedmodel of STD over

an stochastic one would be revealed by the comparison of the result obtained for finite

M ’s with the caseM = ∞.

Now we turn to analyze the implications of the major constraint imposed by synaptic

depression: the saturation of the mean current.

5.3.1 The saturation ofµ and the subthreshold regime

Before going into the analysis of the current saturation, we will define the sub- and supra-

threshold regimes. By inspection of the equation of the LIF neuron, eq.5.2.2, we can in-

tuitively see that the stochastic process defined by the temporal evolution of the membrane

potentialV (t) of the cell, has a mean value2 µ[V ] = µτm whereµ is the mean of the Gaus-

sian coloredcurrent by which we have replaced the source term of eq.5.2.2. This can be

rigorously proven [Tuckwell, 1988]. This Gaussian process has no current units because it

is not exactly the currentI(t) defined in eq.4.5.1. Indeed, it isRm/τm times the currentI(t)

(see how we have transformed eq.5.2.1into 5.2.2). Therefore, this Gaussian source term

(which will be called current for simplicity) has units of voltage divided by time. Now, if

µ[V ] = µτm < θ this implies that the mean membrane potential, after it has been reset toH

and stayed there during a periodτref , starts integrating the current and approaches the value

µ[V ]. Once it is there, it fluctuates around it with a varianceσ2[V ] proportional to the current

varianceσ2w (for details see [Tuckwell, 1988]). Because this mean valueµ[V ] falls belowθ

, the potential reaches threshold because of the input fluctuations, i.e. the input noise. In

the limit case in which the input has no noise, ifµτm < θ the neuron would never fire. In

conclusion, in the sub-threshold regime3 the input fluctuations are crucial to make the neu-

ron fire [Tsodyks and Sejnowski, 1995, vanVreeswijk and Sompolinsky, 1996, Shadlen and

Newsome, 1998, Salinas and Sejnowski, 2000]. If on the contraryµτm > θ the evolution

of V (t) would be driven mostly by the mean currentµ and not so much by the fluctuations

σ2w: on its way fromH to the mean current the potential will cross threshold and a moderate

amount of noise would jitter a little the precise crossing time. Figure5.1shows an example

2To be precise, this is the mean voltage of an IF neuronwithout threshold, that is, of the variableV (t)governed by equation5.2.2alone. If the effects of the threshold are taken into account the meanµ[ν] would be

smaller than the given value.3We will characterize the sub-threshold regime by the inequalityµ < θ instead of specifying thatµ refers

to the mean voltage. When plotting the source current, in order to visualize if it determines a sub- or supra-

threshold regime we will usually plot the current timesτm so that it has the proper voltage units.

Page 153:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 139

of the evolution of the membrane potential in two situations, one in which the current is

sub-threshold (top) and one in which it is supra-threshold (bottom). Another major conse-

quence of these two different modes of activity is the regularity of the output spike pattern

measured by the coefficient of variation of the IRIs,CVisi [Shadlen and Newsome, 1998,

Feng and Brown, 2000, Salinas and Sejnowski, 2000, Renart, 2000]. As can be seen in the

figure, the variability of ISIs in the case where the firing activity is dominated by the fluctu-

ations is much larger (CV = 1.06) than in the case in which the potential is mainly driven

by the mean current (CV = 0.68). Because cortical neurons seem to fire in a very irregular

fashion [Softky and Koch, 1993], close to a Poisson process, the subthreshold regime has

been postulated as a potentialmechanismto generate such a high variability [Shadlen and

Newsome, 1998, Stevens and Zador, 1998]. Under this hypothesis the subthreshold regime

was obtained by abalancebetween the excitatory and inhibitory currents (that is why this

regime is also known as the balanced regime).

Now, does short-term depression play any role in determining if the working regime of

the neuron is sub or supra-threshold? Let us take a closer look at the saturation value of the

mean currentµlim, defined as the value of the mean current when the pre-synaptic cells fire

with infinite rateν = ∞ (see eq.4.5.5in section4.5.1):

µlim =CMJ

τv

(5.3.1)

If we put into this expression biologically plausible values for the parametersτv ∼ 500−1500

ms andJ ∼ 0.01 − 0.5 mV, we can compute the number of active synapsesCM needed to

obtain a mean currentµ above threshold, i.e. µlim > θ. Assuming thatθ = 20 mV and that

the membrane time constantτm = 10 ms, τv = 1000 ms andJ = 0.2 mV the inequality

reads

CMJ

τv

τm

(5.3.2)

CM > τvθJτm

' 10000 (5.3.3)

In the lower limit of our intervals, that is,τv = 500 ms andJ = 0.5 mV, this number comes

down to500. This means that in other to enter in the supra-threshold regime (µτm > θ)

at least500 strong synaptic contacts with fast short-term depression need to be activated

with infinite spike rate, or up to10000 contacts in the case that their efficacies are moderate

(J = 0.2 mV) and depression is not so fast recovering (τv = 1000 ms).

These figures give an idea of how large a population of pre-synaptic neurons has to

be in the case itwantsto put the target neuron in a supra-threshold state (notice that if the

connections are not mono-synaptic,M > 1, the numbers given above do not correspond with

the number of pre-synaptic cells). They motivate us to state that depressionmay imposethe

Page 154:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

140 Chapter 5: The response of a LIF neuron with depressing synapses

1200 1250 1300

t (ms)

-10

0

10

20

30

40

50

V (

mV

)

1000 1200 1400 1600 1800 2000

t (ms)

-30

-15

0

15

30

45

V (

mV

)

CV= 0.68

CV= 1.06

Figure 5.1: Two examples showing the simulation of the membrane potential evolution of

a leaky integrate-and-fire neuron.Top plot: Sub-threshold balanced regime: the synaptic

efficacies were chosen so that the mean current is zero, i.e.µτm = 0 (dashed line). Thus, the

potential fluctuates around zero, and eventually crosses the threshold (dotted line) driven by

a positive fluctuation of the current (σwτ 1/2m = 14 mV). As a result the firing pattern is very

irregular (CV = 1.06). Bottom: Supra-threshold regime: a different choice of the synaptic

efficacies leads to a mean currentµτm = 33.75 > θ (dashed line) larger than the threshold

(dotted line). The potential evolves from the reset potentialH towardsµ, crossingθ in its

way. This results in a regular firing pattern (the inter-spike-intervals have little variability,

CV = 0.68) which is quite insensitive to the fluctuations of the current (which are about the

same size as in the top plot,σwτ 1/2m = 13.66 mV). Neuron parameters:θ = 20 mV, H = 10

mV,τm = 20 ms,τref = 2 ms. Current parameters:Ce = 4000, Ci = 1000, νe = νi = 25

Hz. Top: Je = 0.14 mV,Ji = 0.56 mV Bottom:Je = 0.14 mV,Ji = 0.53 mV

Page 155:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 141

working regime to be sub-threshold in those circumstances in which the pre-synaptic signal

is not transmitted by a huge number of cells, but by an amount of the order of a thousand

neurons. This number seems to be of the order of a cortical functional population size, or

perhaps a little bit above. For all these reasons we have chosen the sub-threshold regime

as the relevant case to study, in which in addition, the current noise plays a critical role.

Therefore, since our aim was to study the effect of synchrony plus STD on the response of a

neuron, we observe that depression naturally sets a working regime in which the modulation

of noise by the input synchrony is maximally expressed.

The particular situation in which we will analyze the output of the neuron is the follow-

ing: Let us suppose that the target neuron is receiving inputs from many cells, pyramidal cells

and interneurons at low rates (∼ 2 Hz) If this excitatory and inhibitory activity is approxi-

mately balanced, it could be modeled as a white noise with zero mean an a certain variance

σ2bg. Now, over this background we focus on a specific population trying to transmit a certain

signal to the target neuron through depressing synapses. This population consists ofC neu-

rons whose activity can be correlated with zero time-lag, that is, some of their spikes might

be synchronous. Because the number of contacts that this population makes onto the output

neuron will not surpass the critical number needed to produce a supra-threshold current, the

modulation of the input noise is crucial to make the output neuron fire. In the following,

we will show how critical is the number of contacts in each connectionM and the level of

synchronyρ in the modulation of the input fluctuations and the way this modulation is read

out by the target neuron.

5.3.2 The modulation of the variance

We will first show the effect of multiple-contact connections between the cells on the pre-

synaptic population and the target neuron. Afterwards, we will investigate the effects due to

synchrony. As explained above, we will vary the number of contactsM while keeping the

mean rateµ constant in two ways:

5.3.2.1 VaryingM , with CM constant

Varying M in this way, is equivalent to compare the currents produced by populations

with different sizesC, such that each one establishes the same number of total contacts,CM .

Thus, the mean current,µ, produced by each of these populations is equal, but the variance,

σ2w, is not. What is then the value ofσ2

w, as a function of the input rateν, whenM is varied?

Figure5.2 exhibits a three dimensional plot in whichσ2w is plotted as a function ofM and

ν, for CM = 1. The first striking feature is that as soon asM > 1, the varianceσ2w as a

Page 156:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

142 Chapter 5: The response of a LIF neuron with depressing synapses

0 10 20 30 40 50

Rate

5

10

15

M

0

1

2

3

4

5

6

7

Figure 5.2: Current varianceσ2w (per contact) as a function of the number of contactsM and

the input rateν whenCM is kept fixed and equal to one.σ2w is given in units ofJ2 andν is

given in Hz. Other parameters are:∆ = 0, U = 0.7, τv = 500 ms andρ = 0.

function ofν shows anon-monotonicbehavior with a maximum. This maximum becomes

more and more prominent asM increases. Although it is not completely clear in the figure

(because of the smallν range shown), in the limitν → ∞ the variance converges toσ2lim

independently of the number of contactsM . As discussed before in this section, this happens

becauseσ2lim depends on the productCM which is held constant. On the other hand, we also

have the correlation magnitudeα = Σ2

σ2w

, which measures the importance of the exponential

correlations in the input and whose value4 −1 < α < 0 tells us how much the input current

deviates from an uncorrelated white noise (see section5.2.3). This correlation magnitudeα

does not depend neither on the number of pre-synaptic cellsC nor on the synaptic efficacy

J , but it does depend onM (see eqs.4.5.35and eq.4.5.36). Therefore the way it varies

with M is the same in both comparison schemes. Figure5.3 illustratesα as a function ofM

andν. First we observe that in the limitν → ∞, α vanishes. This happens because while

σ2w tends to a finite value,Σ2 goes to zero asν increases. The second observation is that,

although qualitatively similar, for differentMsα exhibits a behavior which is scaled up, i.e.

the magnitude of the minimum grows with the number of contactsM and the convergence

4The correlation magnitude is bounded by−1 and zero because the exponential varianceΣ2 is negative due

to intrinsic properties of the correlations introduced by depression whichsubtractprobability for a subsequent

release after we observe one. Besides,Σ2 < σ2w because as we showed, the variance of the release count in

a time windowT , which obeysT >> τC , equals(σ2w + Σ2)T . Because the variance is defined non-negative

|Σ2| cannot be larger thanσ2w

Page 157:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 143

10

20

30

40

50

M0

1020

3040

50nu

-1

-0.8

-0.6

-0.4

-0.2

0

Figure 5.3: Correlation magnitudeα as a function of the number of contactsM and the input

rateν. Parameter values were chosen as figure5.2: ∆ = 0, U = 0.7, τv = 500 ms,ρ = 0.

to zero is also slower (asM increases, the rate needed to make the correlation magnitude

α negligible, is bigger). In figure5.3 all the current properties (µ, τc , σ2w, Σ2 andα ) are

plotted together as a function of the input rateν for several values of the number of contacts

(M = 2, 10, 50, 150) when besides the population current, a background zero mean white

noise has been included in the input. In the top plot the mean currentµ(ν) and the correlation

time scaleτc are common to allM values. We can check that the mean currentµ, which is

exhibited in voltage units (that is, what is shown is the productµτm), saturates below the

threshold value20 mV. The time constantτc displays a similar though inverse behavior: it

saturates to zero with the same convergence rate. The first consequence can be extracted

already at this point: since the exponential correlations will be detectable by the output

neuron only ifτc <τm (see section5.2.3), the input current can be taken as a white noise

with varianceσ2w until the mean current starts to saturate, because as it saturatesτc becomes

smaller thanτm and corrections can be observed in the output. The middle plot showsσ2w and

Σ2 for severalMs. Here,σ2w contains the contribution of the input population under analysis

plus a constant termσbg due to background activity. Allσ2w lines show a maximum (second

plot), although for smallM it is not detectable in the plot. In the same plot we observe that

Σ2 displays a similar behavior asσ2w but for negative values. However, the ratio between

the two variances, that is the correlation magnitudeα shown in the bottom plot, grows as

M increases. As a consequence we can assert that the effect of the exponential negative

correlations parameterized byα andτc , will only be detectable whenM is large. The reason

Page 158:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

144 Chapter 5: The response of a LIF neuron with depressing synapses

0

5

10

15

20

µ [

mV

]

Effects on the input current caused by multiple-contacts M>1

-100

-50

0

50

100

150

σ w

2 , Σ2 [

mV

]

M= 2M= 10M= 50M= 150M=1, ρ= 0.05

0 20 40 60 80 100ν [Hz]

-0,8

-0,6

-0,4

-0,2

0

α

0

100

200

300

400

500

τ c [ms]

Figure 5.4: Current parameters as a function ofν for several values of(C, M) whereCM

is invariant. The current is composed of: i) a background created by an excitatory popula-

tion of 2000 neurons with synaptic efficacyJ = 0.05 and an inhibitory population of500

interneurons withJ = −0.2, both firing at a constant rate of2 Hz. ii) The population under

focus is composed ofC excitatory neurons withJ = 0.19, each one makingM contacts so

that the total number of contacts is alwaysCM = 3750. In all cases except when mentioned,

ρ = 0. Top plot: Superposition of two graphs, namely the mean currentµτm (left axis) and

the correlation timeτc (right axis). Both apply for all combinations of(C, M) becauseτc

does not depend on either of them, andµτm depends on the product. It is important to notice

that µlimτm falls below the thresholdθ, located at20 mV. Middle plot: Current variance

σ2wτm (solid lines) and exponential varianceΣ2τm (dashed lines) for four different values of

(C, M). The dotted line represents a mono-synaptic configuration with synchronyρ = 0.055

to illustrate the equivalence betweenM > 1 andρ > 0 (see text).Bottom plot: Correlation

magnitudeα = Σ2/σ2w. Inset values apply for middle and bottom plots.

Page 159:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 145

is that, although for anyM α exhibits a minimum (more pronounced the bigger isM ) and

there is aν range in which it is not negligible, for low number of contacts this range is small

and, what is more important, in this rangeτc >τm . Then, asM is increased, the range where

α is not negligible scales up whileτc remains insensitive to the variation ofM . Thus, for

largeM , τc becomes smaller thanτm whenα is still noticeable and therefore the correction

from the white noise will be important. This will be illustrated in the coming figures of the

output rate.

The position of the maxima inσ2w also differs from oneM to another. Actually an

expression for the maximum positionνmax, can be easily obtained from the formula ofσ2w,

eq.4.5.35. When one assumes thatρ = 0 (no synchronization) and that∆ = 0 (all efficacies

are equal) the maximum is achieved at

νmax = 2−2 + U −

√(1− 3 M + 2 M2) U2 (2− U)

τ U (4− U2 − 4 UM + 2 U2M)(5.3.4)

which only depends on the parametersM andU . In fact, from this expression we can deduce

for which pairs ofM andU there exists a maximum. This happens, whenever the previous

expression ofνmax is non-negative, which occurs if the number of contacts exceeds

M >U + 2

2U(5.3.5)

This implies that the smaller isU , the more contacts are needed to produce a non-monotonic

behavior ofσ2w. If release is completely reliableU = 1, two contacts are enough to produce

a maximum, though it will not be very prominent. The value ofνmax is plotted as a function

of the two parametersU andM in figure 5.5 for τv = 500 ms. Although at a first glance

it seems to be widely modulated by these parameters in a sort of hyperbolic behavior, we

will show that effectively it always falls in a low rate value. To measure how prominent

is the bump in the variance we have plotted the ratio between theσ2w at the maximum and

the limit valueσ2lim. This ratio is shown in fig.5.5 (bottom plot) for the sameτv as before.

Comparing both plots, we now see that the bump becomes prominent when bothM andU

are large, and in this regionνmax shows a large plateau at low rates. Changing the value of

τv within a plausible range is of little help. Therefore, we conclude that this non-monotonic

behavior becomes more relevant the bigger the number of contacts and the closer the release

probability is to one. Whenever these two things happen, the maximum is located at low

rates (∼ 5− 10 Hz).

What is the impact of the synchronization of the incoming spikes (ρ > 1) on the mod-

ulation of the varianceσ2w and the correlation magnitudeα? As it was commented when

the expression forσ2w was derived, the implications ofρ > 1 seemed to be equivalent to

settingM > 1 while renormalizingC to keepCM constant. This is shown in figure5.4

Page 160:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

146 Chapter 5: The response of a LIF neuron with depressing synapses

24

68

1012

1416

1820

M

0.2

0.4

0.6

0.8

1

U

0

10

20

30

40

50

[hz]

2 4 6 8 10 12 14 16 18 20

M0.2

0.4

0.6

0.8

1

U

1

2

3

4

5

ratio

Figure 5.5: Positionνmax of the variance maximum (top) and the ratioσ2w(νmax)σ2

lim(bottom) as

a function of the number of contactsM and the input rateν, when the productCM is held

constant. Notice that the region in the(M, U) plane where the ratio is significantly above

one, is the plateau of theνmax plot where the rate always falls below10 Hz. The white corner

in the top plot represents a region where there is no maximum, and corresponds to the red

closer corner in the bottom plot. Other parameters:τv = 500 ms,∆ = 0

Page 161:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 147

where, together with the case in which(C = 25, M = 150, ρ = 0), we have included the

case(C = 3750, M = 1, ρ = 0.05, dotted line). Although lines do not overlap, the behavior

of σ2w, Σ2 and thereforeα is very similar. The meanµ and the correlation timeτc do not

depend onρ. Thus, we have demonstrated that the synchronization of the Poisson activity

within a given population ofC cells which connect mono-synaptically (M = 1) to a given

target neuron, is qualitatively equivalent to a decrease inC together with an increase inM

in such a way that the product remains invariant. This is interesting for practical applica-

tions: first, it simplifies the situation from an input in which the trains are cross-correlated

to a situation in which fewer afferent trains are independent but make several contacts onto

the post-synaptic cell. Dealing with independent trains is always simpler than dealing with

cross-correlated ones. Secondly, simulations of correlated trains are always harder to im-

plement than those with independent afferent trains. Thus, we have performed numerical

simulations only for the case of independent Poisson5 spike trains impinging on connections

made up of several contacts (by means of this equivalence, these simulations cover the cases

with cross-correlated inputs).

Figure5.6confirms this point by showing an example in which a population with a fixed

number of cells (C = constant) establishing mono-synaptic contacts (M = 1) increases

the zero-lag cross-correlations among its neurons. The outcoming current defined by all its

parametersµ, σ2w, Σ2, α andτc resembles very much the previous situation in whichM was

above one. The values ofρ used here are all lower than0.05 which falls in the range of the

experimental data [Zohary et al., 1994].

5.3.2.2 VaryingM , with MJ constant

We turn now to the second case which consists in determining the effect of an increase

in the number of contactsM while renormalizing the synaptic efficaciesJ , so that the mean

current remains invariant. A common argument among neuroscientists has been that, since

synaptic release is unreliable, i.e.U < 1, neurons create several contacts among them to

average out the intrinsic unreliability by pooling independent responses. A few works have

shown that this idea indeed works: Zador (1998) computes the information conveyed in the

output spike pattern of a neuron about the input spike times when synapses are unreliable.

This work shows that this information grows monotonically as the number of contacts be-

tween the cells increase.Fuhrmann et al.[2002] show that the information contained in the

5Without going into details, we must also bound this equivalence to the case in which the inputs are Poisson.

In the case in which, for instance, spikes come in bursts, this equivalence would not hold. In other words,

depression would make different the cases in which the groups of contacts which are correlated are quenched

(M > 1, ρ = 0) from the situation in which they are all time-varying (M = 1, ρ > 0) (see [Senn et al., 1998]

for an example where this subtle point makes the difference)

Page 162:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

148 Chapter 5: The response of a LIF neuron with depressing synapses

0

5

10

15

20

µ [m

V]Synchrony effects on the input current caused by ρ>0

-100

-50

0

50

100

σ w2 , Σ2 [

mV]

ρ= 0 ρ= 0.01ρ= 0.02ρ= 0.05

0 10 20 30 40 50 60 70 80 90 100ν [Hz]

-0,8

-0,6

-0,4

-0,2

0

α

0

100

200

300

400

500

τ c [ms]

Figure 5.6: Current parameters as a function ofν for several values of the correlationρ

where all theC = 3750 neurons make a mono-synaptic connectionM = 1. The current is

composed of a background generated as in fig5.4and the component coming from the focus

population.Top plot: Superposition of two graphs, namely the mean currentµτm (left axis)

and the correlation timeτc (right axis). Both apply for all values ofρ because neither of them

depend onρ. The mean currentµ saturates again below thresholdµlimτm < θ. Middle plot:

Current varianceσ2wτm (solid lines) and exponential varianceΣ2τm (dashed lines) for four

different values ofρ (see inset).Bottom plot: Correlation magnitudeα = Σ2/σ2w. Inset

values apply for middle and bottom plots. Notice that this figure resembles the previous

figure5.4

Page 163:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 149

5

10

15

M

0 10 20 30 40 50nu

0

0.2

0.4

0.6

0.8

1

Figure 5.7: Current varianceσ2w (per contact) as a function of the number of contactsM and

the input rateν whenMJ = J ′ is kept fixed.σ2w is given in units ofJ ′2. Other parameters

are:∆ = 0, U = 0.7, τv = 500 ms,ρ = 0 andC = 1 so that it represents the variance per

contact.

amplitude of an EPSP about the time at which previous spikes arrived at the synaptic ter-

minal also increases monotonically withM , because many contacts eliminate noise in this

output signal. However the argument of establishing multiple contacts as a mechanism for

pooling out the unreliability, seems at odds with the existence of cortical synapses with a

high probability of release [Paulsen and Heggelund, 1994, 1996, Bellingham et al., 1998,

Stratford et al., 1996]. If unreliability may be overcome by tuning the biophysical mecha-

nism at the single synaptic bouton so that release probability increases, why would neurons

not try to beat this limitation by establishing more and more contacts? The question seems

to be still opened, suggesting that multiple contacts between neurons may be useful for other

purposes. In this section we study the transformation of the current second order statistics

when a single contact spreads up into severalsmalleractive zones6. The net effect of this

increase is a strong decrease of the varianceσ2w. This is shown in figure5.7. In the limit

M →∞, atMJ fixed,σ2w becomes

limM→∞

σ2w = σ2

dm

C J ′2 Uν

1 + Uντv

[UM

1 + Uντv(1− U/2)

](5.3.6)

6As it is done in the whole thesis, we do not distinguish whether this multiplication offunctionalcontacts

occur within a synaptic bouton or if the pre-synaptic axon creates new synapses.

Page 164:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

150 Chapter 5: The response of a LIF neuron with depressing synapses

whereJ ′ is the constant product ofMJ . This limit varianceσ2dm displays again a non-

monotonic behavior with a maximum at

νmax = 1/21−

√9− 4 U

(−2 + U) Uτv

(5.3.7)

which again, for realistic values ofτv , is always lower than10 Hz. In this limit, the cur-

rent coincides with the current derived from the deterministic model (dm) of Tsodyks and

Markram[1997]. Although not shown in this thesis, we have analytically proven this iden-

tity by computing the correlation function of the deterministic model, obtaining the same

function one obtains in the limitM →∞ andJ → 0, while MJ =constant, of the stochas-

tic multiple contact model. In this phenomenological model the synaptic response is not

stochastic. Thus the varianceσ2dm is due only to the pre-synaptic stochasticity, and, there-

fore, its magnitude is smaller thanσ2w at any finite value ofM for all ν. A remarkable feature

of this model arises when the input rateν is very large and depression makes the synapse

saturate: in this regime the fluctuations in the current vanish. This occurs because the size of

the synaptic responses becomes smaller asν increases. In theν →∞ limit, the magnitude of

each EPSC becomes zero, as the number of them per unit time is infinite: the current is not a

point process any more but a constant flux of ions into the cell. Nevertheless,one should bear

in mind that this is an averaged model which essentially violates the quantal description of

the synaptic response because it assumes a continuum of EPSC sizes. It is useful for many

purposes but, when modeling the input current to a neuron up to second order statistics,

important differences between the deterministic and a stochastic model appear.

Figure5.8 shows several parameters of the synaptic current as a function of the input

rate forM = 1, 2, 10, 100. The last of thisM ’s is qualitatively similar to the deterministic

model. This time we have taken a population of both excitatory and inhibitory neurons, both

firing at the same rateν. Although inhibition is present, the net mean current is not balanced.

This happens because the number of excitatory neurons was taken bigger than the number of

inhibitory ones. Even though the meanµ is not zero, it saturates under threshold,µlimτn < θ.

For the chosen values ofU and∆ the exponential variance,Σ2, is the same for allM , but this

is not true in general. Like in the previous section, increasingM results in an enhancement

of the correlation magnitudeα with no change in the correlation timeτc . For this reason,

when the number of contacts is small, the effect of the exponential negative correlations is

weak, whereas for largeM or in the deterministic model, this correlation cannot be ignored

in order to quantify the neuron response rate, as we will see in next section.

Page 165:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 151

0

5

10

15

20µτ

(mV)

Effects on the input current cuased by multilpe contacts (M*J=const.)

-20

0

20

40

60

80

σ w2 τ, Σ 2τ

(mV)

M=1M=2M=10M=100 (= det. mod.)Σ

2

0 20 40 60 80 100 ν [Hz]

-1

-0,8

-0,6

-0,4

-0,2

0

α

0

100

200

300

400

500

τ c (ms)

Figure 5.8: Current parameters as a function of the input rateν for several values of(M, J)

whereMJ = J ′ is constant. The current is composed of: i) a excitatory population made

up ofCE = 2000 units with efficacyJ ′e = 0.6 mV and an inhibitory population ofCi = 500

interneurons withJ ′i = −1.6 mV, both firing at the same rateν. In each population neurons

are makingM contacts so that the total number of contacts grows as(Ce + Ci)M . In all

casesρ = 0. Top plot Superposition of two graphs, namely the mean currentµτm (left axis)

and the correlation timeτc (right axis). Both apply for all combinations of(M, J). Notice

thatµlimτm falls below thresholdθ = 20 mV. Middle plot Current varianceσ2wτm (colored

solid lines) and exponential varianceΣ2τm (dashed line) for four different values of(M, J)

(dashed line corresponds to all four cases).Bottom plot Correlation magnitudeα = Σ2/σ2w.

Inset values apply for middle and bottom plots. Parameters:U=1, τv = 500 ms.

Page 166:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

152 Chapter 5: The response of a LIF neuron with depressing synapses

5.3.3 The output rate of a LIF neuron

5.3.3.1 Output rate atCM fixed

Finally, we are going to analyze how the modulation of the meanµ, the varianceσ2w

and the correlation magnitudeα, affect the firing properties of an integrate-and -fire neu-

ron. We have done it both analytically and numerically through simulations. The analytical

procedure used to compute the output rateνout has been described in section5.2.3. It takes

into account corrections, arising from exponential correlations in the input current, to the

standard expression for the rate of a LIF neuron driven by a white noise [Ricciardi, 1977].

We start by plotting the output rateνout andCVout for the currents shown in figure5.4

which illustrated the case withCM fixed. Besides the input from this population, we added a

balanced current component with zero mean and varianceσ2bg, wherebg refers to background.

This background current was produced by the activity of an excitatory and an inhibitory pop-

ulation composed of2000 and500 neurons respectively, connected mono-synaptically, and

firing at 2 Hz. To distinguish between the two components of the current, namely the back-

ground and the current from the population under analysis, we will refer to the neurons

belonging to the latter as neurons frompop. As discussed previously in the Results section,

these examples for the current are qualitatively equivalent to the case in which a given pop-

ulation correlates its activity with different degrees of synchronizationρ. Therefore, there

exits a matching between theMs picked for the figure and the values ofρ which would pro-

duce the same response:M = 2, 10, 50 and150 correspond toρ = 0.00025, 0.0022, 0.013

and0.04 respectively. Figure5.9 shows: the simulation results; the analytical prediction

taking into account the corrections due to the exponential negative correlations (solid lines

top plots) along with the prediction neglecting this correction (dashed lines top plots); the

current parametersµ andσ2w (second from the bottom plot),α andτc (bottom plot) for com-

parison. The first thing to notice is that the output rate is strongly modulated by the variance

σ2w, whereas the mean current plays a less important role. This, of course, occurs because

all these currents are subthreshold. Thus, the output rate is much higher whenM is big,

because in that case the fluctuations of the current are larger and enable the potential to fluc-

tuate and hit threshold more often. The non-monotonic behavior exhibited by the current

varianceσ2w for largeM , is now inherited by the output rate. This is only evident in the

two cases with biggerM . In those cases (M = 50, 150 or ρ = 0.013, 0.04) , νout increases

very fast reaching a maximum value at aboutν ' 10 Hz. From thereafter, the rate decreases

as a negative power (fit not shown) to a non zero value independent ofM . This saturation

value coincides with the rate which would result from a Gaussian white noise current with

meanµlim (eq. 4.5.5) and varianceσ2lim (eq. 4.5.37). The maximum does not coincide with

the maximum ofσ2w. The reason is that whenσ2

w reaches its maximum, the meanµ is still

Page 167:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 153

0

5

10

15

20

µ [

mV

]

0

50

100

150

σ w

2 [mV

]M= 2M= 10M= 50M= 150

0,6

0,7

0,8

0,9

1

1,1

CV

out

0

100

200

300

400

τ c [ms]

0

10

20

30ν ou

t [Hz]

Response of a LIF neuron for different Ms

0 50 100 150 200

νout

[Hz]

-0,8

-0,6

-0,4

-0,2

0

α

Figure 5.9: Numerical results and theoretical prediction of thenon-monotonicresponse of a

LIF neuron for input current examples with differentMs, whileCM is held constant. Input

current is taken from fig.5.4. Top plot: Output rateνout as a function of the input rateν.

Solid lines are the theoretical prediction with the correction due to the exponential correlation

introduced by depression. Dashed lines are the theoretical predictions when neglecting the

effect of the exponential correlation, and therefore assuming the input to be a white noise.

Symbols (in a slightly different color for a better visualization) are the simulation results.

Orange squares mark the position of the simulation examples shown in figs.5.10-5.13.

Second plot: Output coefficient of variation of the ISI’s,CVout, for the three cases with

higher output rates.Third plot: (The same as in fig.5.4 has been plotted here to visualize

the input together with the output). Superposition of the mean currentµ (magenta line, right

axis) and the current varianceσ2w (left axis). Bottom plot: Superposition of the correlation

magnitudeα (left axis) and the correlation time constantτc (magenta line, right axis). Colors

in the top plot inset apply to all plots.

Page 168:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

154 Chapter 5: The response of a LIF neuron with depressing synapses

growing significantly so that it still contributes to the increase ofνout.

What about the correction to the white noise? As can be seen in the analytical predictions

of the output rate, significant differences only arise whenM is either50 or 150 (see fig.

5.9). For less contacts, the exponential correlations produce a small deviation from the white

noise case. Besides, forM = 50, 150, although the correlation magnitude (bottom plot) for

an input rate of about∼ 10 Hz is prominent,α ∼ −0.8, sinceτc τv , the correction is

small. Only for higher input rates, whenτc < τv, the correction becomes important. The

modification produced by these negative correlations is todecreasethe output rate because,

as theτc becomes smaller, the effective current variance that the LIF neuron starts to observe

is the sum of the varianceσ2w and the exponential varianceΣ2 (which is negative).

The comparison between simulation and theory is excellent except near to the maxima

in the casesM = 50, 150. The reason of this discrepancy is the violation of the diffusion

approximation (see section5.2.1). First, whenM is large, upon arrival of a spike along a

pre-synaptic fiber, up toM EPSC’s might be produced at the same time. This provokes a

substantial jump of the membrane potential and the diffusion approximation no longer holds:

the size of the discontinuities ofV (t) is too large to be ignored. Why do these cases with

largeM give a good fit for lager input rates then? The answer is once again saturation: as the

synapse starts to saturate a larger fraction of theM contacts (from the same fiber) are empty

when the spikes arrives so that only the rest have a chance to release transmitter. Therefore.

the synchronous arrival ofM EPSC’s to the output neuron no longer occurs and the diffusion

approximation becomes valid.

5.3.3.2 Saturations eliminates synchrony

The influence of depression on the spatial structure of the inputs, is illustrated in four

snap-shots of the neuronal response and its input current at four different input ratesν (fig-

ures5.10,5.11,5.12 and 5.13). We have chosen the most extreme of the examples(C =

25, M = 150) to clearly capture this transformation. The snap-shots have been taken at the

input rates indicated by the orange squares in the top plot of figure5.9.

These four figures (5.10,5.11,5.12,5.13) are composed of the following plots: i)Top

plot: A rastergram of the input spikes, where the the population under focus represents only

the first25 neurons, while indexes from25 up to 100 represent a fraction of the neurons

firing in the background at2 Hz. ii) Second plot:A rastergram of the synaptic releases. The

population under focus represents the first3750 contacts (because it has25 neurons making

150 contacts each), although just660 belonging to four pre-synaptic neurons are shown.

The rest of the contacts are mono-synaptic synapses established by background neurons. iii)

Third plot: Instantaneous current,Iτm(t), defined as the total current received in a time

Page 169:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 155

window of lengthτm . The blue line is the balanced background current, while the black

line represents the total currentIτm(t), that is, the background plus the current from the

population under focus. The red dashed line represents the averaged ofIτm(t) which equals

µτm iv)Bottom plot: Membrane potentialV (t). The spikes are addeda posteriorifor clarity.

The dotted line represents the threshold (θ = 20 mV) and the dashed red line represents the

mean potential< V (t) >= µτm. The four figures5.10,5.11,5.12 and5.13, differ in the

input rateν which is2, 10, 80 and200 Hz, respectively. Depression was parameterized with

τv = 500 ms andU = 1. This choice of reliable synapses was done to make the effect of

desynchronization more obvious. A different value ofU would only attenuate the modulation

of the fluctuations.

• In the first figure5.10, the population under focus is firing at the background level, so

it cannot be distinguished from the rest of the neurons in the spike rastergram. In the

release raster gram, the difference cannot be more obvious: while a background spike

produces one release, a spike from our population, produces up to150 releases at the

same time (although it looks like a vertical line, is in fact the composition of150 dots).

We should make clear that this rastergram of releases, only shows the EPSC’s produced

by four neurons of the input population (and about600 more from the background).

The rateν is small enough to leave time between spikes for the vesicles to recover, so

most of the times all the150 EPSC’s occur at the same time. Therefore, the arrival

of a single spike from this population, produces a big fluctuation in the input current

(third plot), which after the spike goes back to zero and then up again with the next

synchronous arrival of EPSC’s. In this way,Iτm(t) displays a large variance,σwτ 1/2m =

12 mV, but a small meanµτm = 6.8 mV. Meanwhile, the membrane potential keeps

going up and down driven by these large fluctuations. However, not all of them make

V (t) reach threshold because the baseline where it iswaiting between fluctuations is

too low (the average potential isµτm = 6.8 mV). Because of these big jumps in the

potential, the diffusion approximation gives a rather rough description of the process,

and the estimation ofνout is not so good.

• In the next figure5.11, the input rate has been increased toν = 10 Hz, so that now

the input spikes frompop arrive at a higher rate than the background. Still, the syn-

chronous spatial structure of the releases produced by neurons frompopis maintained:

the smallbarsare not so solid now but arrive much more often. This makes the fluctu-

ations of the current to be still very big and to occur more often. The result is that the

potential keeps making sudden jumps towards threshold. Because input fluctuations

occur in an almost continuous fashion, the mean potential has come up to almost12

mV making the distance to threshold shorter and much easier for a fluctuation to make

Page 170:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

156 Chapter 5: The response of a LIF neuron with depressing synapses

the neuron fire. At this point, although the squared input variance is smaller than before

σwτ 1/2m = 10.3 mV, the maximal rateνout is achieved because of a higher mean current

µτm = 11.8 mV. The difference between the actualνout and the analytical prediction

is due to the synchronous arrival of EPSC’s, which make the diffusion approximation

a coarse estimation. However, it still gives a qualitative correct behavior.

• In the third figure5.12, things have changed qualitatively. The input rate is well above

saturation,ν = 80 Hz so that more than97% of the contacts frompophave not ready-

for-release vesicles, so that the incoming spikes only trigger a response in a small

fraction of the contacts theyvisit. For this reason, the spatial structure in the raster-

gram of releases has been blurred, and the vertical bars are harder to detect. As a

result, the fluctuations of the current have dropped toσwτ 1/2m = 5.4 mV, while the

average line is almost at its highest valueµlimτm = 14.2 mV. The potentialV (t) lives

now closer toθ than before but it crosses it much fewer times because there are no

more big abrupt jumps. Now its trajectory looks smoother, although still shows some

sudden appreciable jumps. In any case, the diffusion approximation gives an accurate

prediction of the output rate which now isνout = 12.7 Hz (fig. 5.12).

• In figure5.13, we have set the input rate to a very large valueν = 200 Hz to emphasize

this idea: the spatially synchronized input pattern, which was present in the release

rastergram when the input rate was low, has now almost disappeared, and looks very

much the same as the pattern produced by the Poisson background activity (second

plot). Fluctuations in the current are now as low as those of a de-correlated white

noise.

The point we have tried to stress by this series of pictures can be summarized as follows:

if a given population of neurons fires with a certain degree of synchronization, short-term

synaptic depression would de-correlate those inputs if its firing rate is high enough (well

beyond the saturation frequencyνsat of eq. 2.3.5). We will make a brief discussion of this

result in the next section.

5.3.3.3 Output rate atMJ fixed

If we maintainMJ constant as we increaseM , the main effect is a substantial decrease

of the current varianceσ2w. Thus, since we have imposed that the population under focus

puts the output neuron in a subthreshold regime (µlimτm < θ), a decrease inσ2w results in a

decrement of the output rate. This is exactly what figure5.14shows. We have taken exactly

the same example as in figure5.8where a population composed of inhibitory and excitatory

neurons fires with rateν in such a way that the saturation current falls below threshold. We

Page 171:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 157

0

20

40

60

80

100

pre-

syna

ptic

neur

on

Input current: ν= 2 Hz, µ= 6.8 mV, σw

= 12 mV

3250

3500

3750

4000

4250

syna

ptic

cont

act

1000 1500 2000 2500t [ms]

0

10

20

30

40

50

V [m

V]

0

20

40

60

I τ m(t)

[mV]

νout

= 20 Hz, CVout

= 0.98

Figure 5.10: Evolution of the afferent spikes, synaptic releases, total afferent current and

membrane potential (I). The current is the same as in fig.5.4: a background component

of excitatory and inhibitory neurons (indexed in top plot from25 to 2524) making a single

contact (contacts indexed in second plot from3750 − 6249) onto the target neuron, and

firing as Poisson processes at a constant rate of2 Hz. The populationpophas25 excitatory

neurons (indexed in top plot from0 − 24) making M = 150 contacts each (indexed in

second plot as:0 − 149 neuron0, 150 − 299 neuron1, ..., 3600 − 3749 neuron24). The

statistics are Poisson with firing rate of2 Hz (in this figure). Top plot: Rastergram of the

incoming spikes labeled by input neuron index (each dot represent the arrival of a spike).

Second plot:Rastergram of the synaptic releases labeled by contact index. Vertical lines are

the superposition of synchronous releases by contacts belonging to the same neuron.Third

plot: Instantaneous currentIτm(t) defined as the afferent current in a time window of length

τm. Black line, represents the total current, while the blue line represents the background.

Red line represents the temporal average of the total current.Bottom plot: V (t) of the LIF

neuron integrating the current depicted in the plots above. Action potentials are addeda

posteriori for visualization. Brown line representsθ = 20 mV while re line the temporal

average ofV (t). Parameters:τm = 10 ms;τref = 2 ms;H = 10 mV.

Page 172:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

158 Chapter 5: The response of a LIF neuron with depressing synapses

0

20

40

60

80

100

pre-

syna

ptic

neur

on

Input current: ν= 10 Hz, µ= 11.8 mV, σw

= 10.3 mV

3250

3500

3750

4000

4250

syna

ptic

cont

act

1000 1500 2000 2500t [ms]

0

10

20

30

40

50

V [m

V]

0

20

40

60

I τ m(t)

[mV]

νout

= 27 Hz, CVout

= 0.88

Figure 5.11: Evolution of the afferent spikes, synaptic releases, total afferent current and

membrane potential (II). The same as in figure5.10but with an input rateν = 10 Hz.

Page 173:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 159

0

20

40

60

80

100

pre-

syna

ptic

neur

on

Input current: ν= 80 Hz, µ= 13.9 mV, σw

= 5.4 mV

3250

3500

3750

4000

4250

syna

ptic

cont

act

1000 1500 2000 2500t [ms]

0

10

20

30

40

50

V [m

V]

0

20

40

60

I τ m(t)

[mV]

νout

= 12, 7 Hz., CVout

= 0.72

Figure 5.12: Evolution of the afferent spikes, synaptic releases, total afferent current and

membrane potential (III). The same as in figure5.10but with an input rateν = 80 Hz.

Page 174:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

160 Chapter 5: The response of a LIF neuron with depressing synapses

0

20

40

60

80

100

pre-

syna

ptic

neur

on

Input current: ν= 200 Hz, µ= 14.1 mV, σw

= 4.3 mV

3250

3500

3750

4000

4250

syna

ptic

cont

act

1000 1500 2000 2500t [ms]

0

10

20

30

40

50

V [m

V]

0

20

40

60

I τ m(t)

[mV]

νout

= 6, 5 Hz., CVout

= 0.95

Figure 5.13: Evolution of the afferent spikes, synaptic releases, total afferent current and

membrane potential (IV). The same as in figure5.10but with an input rateν = 200 Hz.

Page 175:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.3. Results 161

considered four cases in which the population made mono-synaptic contacts with the target

neuron, or madeM = 2, 10, 100 contacts, while the efficacyJ was properly adjusted. This

last example whereM = 100 produced almost the same results as the deterministic model

(data not shown).

The output rate of the neuron (top plot fig.5.14) is severely affected by the redistribution

of contacts. Even when the change is fromM = 1 to M = 2 produces an important change

in νout. As in the previous scenario, whenM becomes large enough, the non-monotonic

behavior ofνout appears. However, now the absolute rates whenM is large have become

very small (νout ∼ 5 Hz). The reason is that since,σ2w decreases rapidly asM increases

(see middle plot), the distance from the mean currentµ to the thresholdθ, which remains

invariant to changes inM , becomes too big compared withσ2w. We could have done fine

tuning to setθ−µlim very small, so that even whenM is 100 νout would not be so small, but

in that case the differences between the differentM ’s would not be very clear.

As before, the effect of the negative correlation is significant only whenM is large, and

the correction setsνout below the prediction obtained with a white noise input.

The comparison between simulation and theory, becomes this time better asM increases,

because of the associated reduction ofJ which, in turn, makes the discontinuities of the

potential smaller.

5.3.4 Information beyond saturation of the mean current

One of the most important implications of saturation is that it restricts the range in which

informationabout the input ratecan be transmitted to the regime of low input firing rates

[Tsodyks and Markram, 1997]. This low-rate range has been bounded by the saturation

frequencyνsat (see chapter2, [Tsodyks and Markram, 1997, Abbott et al., 1997]) which sets

the magnitude of the input rate above which there is little chance to transmit information.

However, under the sub-threshold hypothesis the activity of the output neuron is not only

governed by the mean currentµ but also by the varianceσ2w and the correlation magnitude

α. So a question immediately arises: what is the saturation frequency of the varianceσ2w and

the correlation magnitudeα? If one takes the expression ofσ2w from eq.4.5.35and computes

the saturation frequencyν ′sat the result is7

7The computation ofν′sat was performed in the following way: an expansion ofσ2w around its asymptotic

valueσ2lim was performed up to first order in1ν . Thus,ν′sat is defined as the frequency at which the first order

correction equals the zero order term. Ifν > ν′sat then the leading term is larger than any other term. Although

the definition of the saturation frequency of the mean,νsat, derived in chapter2 (eqs. 2.3.5and2.3.10) was

defined in a different more simple way, both definitions coincide.

Page 176:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

162 Chapter 5: The response of a LIF neuron with depressing synapses

0

5

10

15

20

µτ (m

V)

0

20

40

60

80

σ w2 τ (m

V)

M=1M=2M=10M=100 (= det. mod.)

0 20 40 60 80 100

ν [Hz]

-1

-0,8

-0,6

-0,4

-0,2

0

α

0

100

200

300

400

500

τ c (ms)

0

5

10

15

20

25

ν out [H

z]Response of a LIF neuron for different Ms (M*J=const.)

Figure 5.14: Numerical results and theoretical prediction of thenon-monotonicresponse of

a LIF neuron for examples of input current with differentMs, whileMJ is held constant.

Input current is taken from fig.5.8. Top plot: Output rateνout as a function of the input rate

ν. Solid lines are the theoretical prediction with the correction due to the exponential correla-

tion introduced by depression. Dashed lines are the theoretical predictions when neglecting

the effect of the exponential correlation. Symbols are the simulation results.Second plot:

(The same as in fig.5.8has been plotted here to visualize the input together with the output).

Superposition of the mean currentµ (magenta line, right axis) and the current varianceσ2w

(left axis). Bottom plot: Superposition of the correlation magnitudeα (left axis) and the

correlation time constantτc (magenta line, right axis). Colors in the top plot inset apply to

all plots.

Page 177:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.4. An interesting experiment 163

ν ′sat =1

Uτv

(1 +

U(M − 1)

1− U/2+

Uρ(C − 1)M

1− Uρ/2

)(5.3.8)

This formula shows that if the input is de-correlated (ρ = 0) and the connections are

mono-synaptic (M = 1) both the saturation frequency of the mean currentνsat and the

saturation frequency of the varianceν ′sat coincide at 1Uτv

. However if the number of contacts

is bigger than one or the inputs are synchronizedν ′sat>νsat. This inequality only tells us that

the rate at which the varianceσ2w approaches its asymptotic value is slower than that of the

mean. Of course, this is not enough to guarantee that the response of the neuron depends on

the inputν whenν >> νsat. However, figure5.9shows that this is the case for some values

of M (and consequently ofρ), e.g. M = 150. Comparing the variation ofνout with that of

µ, σ2w andα for high rates (∼ 50 − 100 Hz) we can extract the conclusion that a significant

variation of the output is occurring while the mean currentµ is almost constant: information

is being transmitted: i) first, by means of the variance which is still varying as we can see

in the third plot of the figure. ii) Second, by means of a change in the negative exponential

correlationα whose convergence rate is much slower than those ofµ andσ2w (see bottom

plot). Therefore, the second order statistics of the input may enlarge the frequency range in

which information about the rate can be transmitted through depressing synapses.

5.4 An interesting experiment

One of the main results of this chapter is that synaptic short-term depression can filter

out cross-correlations if the input rateν is large enough to provoke the saturation of the

recovery machinery. This result makes a simple and direct prediction which can be tested

in a simple experiment. First, we need to choose a neural system in which synaptic short-

term depression has been observed. The experiment has to be performed in a slice, where

we need to record from an individual neuron under voltage clamp. With this configuration

we could register the current entering the cell. At the same time we should stimulate

extracellularly a bundle of fibers which are known to project onto the patched cell (this

is particularly easy in the hippocampus, for example, where one can oblate the CA3 area

from the slice and stimulate the Schaffer collateral fibers and record from a neuron in CA1).

The stimulation would be a train of Poisson electrical pulses which would excite a large

pool of proximal fibers creating a set of Poisson spike trains impinging the cell at a very

synchronous manner. Thus, synchrony is present because of the nature of the stimulation,

that is, because all the afferent activity was generated by the same external pulses. Then we

would record the afferent current for different stimulation frequencies,R. With that data,

Page 178:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

164 Chapter 5: The response of a LIF neuron with depressing synapses

we would compute the mean and the variance of the incoming current as a function ofR.

The prediction then would be the following: the mean of the current has to be a monotonic

function ofR that saturates to a certain value. On the contrary, thevariance of the current

will show a non-monotonic behavior with the stimulation frequencyR. In particular, for

high enough rates, when the mean current is saturating its upper bound, the variance should

decrease asR is increased. We could go a step further and, in the case were this afferent

current had an average below threshold (this could be adjusted by diminishing the amplitude

of the extracellular stimulus so that a minor number of fibers are excited), register the spiking

activity of this neuron. In that case, we should observe a non-monotonic behavior of the firing

rateνout as the stimulation frequencyR is increased, demonstrating that the signalR is being

transmitted by the variance of the current which decreases as a function ofR. Because high

stimulation frequencies are necessary to reach the saturating regime, one should take care of

possible adaptation occurring in the afferent axons or in the target cell.

5.5 Conclusions

In this chapter we have explored the impact of short-term synaptic depression on the

output rate of a neuron when the EPSC’s created by the afferent spikes have certain degree of

temporal synchrony. This spatial correlation among the release times was introduced by two

equivalent means: i) By the introduction of a generalization of the mono-synaptic connection

in which a pre-synaptic neuron makes only a single synaptic contact with the target neuron,

to a model in which the pre-synaptic neurons make a numberM of contacts, whereM can

be any integer number. ii) By means of a cross-correlation of zero time-lag among the spike

times of different cells. Since the study was performed first analytically, this allowed us

to extract deeper conclusions and to understand better the implications of the model. For

example we concluded that, two different features of the input, synchrony and multiple-

contacts, happened to give a qualitatively equivalent description of the afferent current, if

the input spike had Poisson statistics: they both effectively increase the magnitude of the

fluctuations of the currentσ2w and the magnitude of the negative exponential correlations

introduced by synaptic recovery of vesicles,α. This enhancement of the current varianceσ2w

was in agreement with previous works [Salinas and Sejnowski, 2000, Moreno et al., 2002].

The new effects exposed here are the result of a non-trivial inter-action between the positive

correlations and synaptic depression. They can be summarized as follows:

1. Saturation. The most noticeable effect of short-term depression is that it imposes

a saturation of the mean currentµ, i.e. the mean current tends, as the input rate in-

creases, asymptotically to a finite valueµlim. Many implications about this effect have

Page 179:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

5.5. Conclusions 165

been previously discussed [Tsodyks and Markram, 1997, Abbott et al., 1997]. In the

present work we have been interested in the idea that this saturation might lead to a

sub-threshold mean current (eq.5.3.1). Thus, given a population of neurons, if their

size is not very big, or their connections are not very strong, the mean current they

project to other neurons, will be sub-threshold, regardless the pre-synaptic firing rate.

A subthreshold regime has been commonly proposed under the excitation-inhibition

balance hypothesis [Shadlen and Newsome, 1994, vanVreeswijk and Sompolinsky,

1996, Tsodyks and Sejnowski, 1995, Amit and Brunel, 1997, Renart, 2000]. In this

scenario, both the excitatory and the inhibitory activity of the network are closely

balanced so that the fluctuations of the current acquire maximal importance. We are

not proposing that synaptic depression may result as an equivalent mechanism to the

balanced current, because both have very different properties. We would like just to

remark that both share the important common feature that the mean current remains

sub-threshold allowing the variance to play a central role in driving the neurons re-

sponse. The interesting issue of how this important constraint of saturation alters the

dynamics of a network, for instance during persistent activity [Wang, 2001], is the aim

of our future work.

2. Depression, a filter of spatial correlationsSince we have explored examples in which

the current was confined to be sub-threshold, the scaling up of the fluctuations due to

positive input correlations had a big impact, increasing the output rate. However, the

most interesting feature found, was that this positive correlation was wiped out by the

vesicle dynamics as soon as the input rate became large enough. In the first chapter

we repeated many times the idea that the release statistics becomes Poisson when

the input rate goes beyond saturation, no matter what the input statistics were. This

happened because in saturation, the synaptic response statistics were dictated more by

the recovery of the vesicle than by the input statistics. In the present chapter we have

simply taken that idea a little further to realize that, since the recovery of vesicles is

an independent process at each synaptic bouton, saturation has to impose also a spatial

de-correlation. And this is exactly what happens when a cross-correlated input crosses

a population of depressing synapses at a very high firing rate: the cross-correlations

are filtered away, implying that the enhancement of the fluctuations due to synchrony

are eliminated, resulting in anon-monotonicbehavior of the current varianceσ2w.

3. Information transmission at high rates We have shown how the output in the sub-

threshold regime can be tuned by the input fluctuations. Signaling by means of the

current variance is not a new idea and it has been shown to have important advantages

over the standard transmission by means of the mean current, like, for instance, a much

Page 180:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

166 Chapter 5: The response of a LIF neuron with depressing synapses

faster response by the output neuron [Silberberg et al., 2002, Moreno et al., 2002]. In

this work we have proven that the variablesσ2w andα converge to their asymptotic

values in a much slower way than the mean currentµ. This result suggests the idea

that transmission of information may occur for high rates at whichµ does not change,

butα andσ2w are still modulated by the input rate.

4. Non-monotonic response functionThe response function of a LIF neuron when the

input is positively correlated and it has depressing synapses can be non-monotonic

if µlimτm < θ, that is, the mean current is always subthreshold. The modulation of

σ2w can be detected by the input neuron even whenµ > θ but the decrease is always

much more subtle than in the subthreshold regime (data not shown). We have shown

however that, whenever the maximum is prominent, it is always located at low input

rates (< 10 Hz). This implies that this simple effect cannot be used in a straightforward

manner to build a tuned response function to apreferred input rate. Nevertheless,

the combination of a decreasing transfer function with an increasing one, may give

rise to a peaked response function in a subsequent stage of processing. Moreover,

this non-typical decreasing behavior prompts the question of whether single neurons

with the help of dynamical synapses may perform more complex computations than

simple integration. Here we have demonstrated that a simple combination of positive

correlations plus depression give, in a stationary situation, a complex behavior of the

output, but the addition of more realistic elements like auto-correlations among spikes

within a single spike train, facilitation, or the more interesting question of the response

to a time-varying input, may permit a single neuron to perform non-trivial tasks, which

should be studied in future work.

5. Cross-correlations vs auto-correlationsWhen synaptic dynamics are not considered,

the effect of positive cross- or auto-correlations in the input is qualitatively the same:

they rescale the fluctuations of the afferent current [Moreno et al., 2002]. However,

in this thesis we have shown how this is not the case when STD is considered. In the

first two chapters we worked out the idea that STD can transform the auto-correlations

by elimination of redundant spikes. Choosing the adequate values of the parameters,

this could be achieved in an efficient way in terms of information trasnmission. Be-

sides, auto-correlations modify the synaptic transfer function by boosting the saturat-

ing frequency. Cross-correlations, on the other hand, do not alter this function because

their effect does not appear in the first order statistics of the current. Moreover, cross-

correlations are filtered by the depressive synapses when the spike train saturates them.

Although considering STD when integrating a current with cross- or auto-correlation

can lead to de-correlation in both cases, the mechanism used in each case is different.

Page 181:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Appendix A

The exponentially correlated input

In this appendix we will show that any renewal process with exponential correlations has

an ISI distributionρisi(t) equal to the linear sum of two exponentials (see eq.2.2.3).

Our starting point is the assumption that the input spikes are modeled as renewal pro-

cesses withexponentialauto-correlation among the spikes times. This is formalized by set-

ting the correlation functionC(t) equal to

C(t) = νδ(t) +A

τc

e−t

τc + ν2 (A.0.1)

Notice that the un-connected version of the two point correlation was used. The parameter

A measures the area under the exponential, whileτc is the decaying time constant of the

exponential.ν is the input spike rate defined as the inverse of the mean inter-spike-interval:

ν ≡ 1<ISI>

. Now, what is the physical meaning of the area under the exponential? does it

depend on the input parametersν andτc or is it an independent degree of freedom? This

has a simple answer if one recalls the relation between the correlation function and the the

variability of the spike count (see section4.5.2.4). If we define the the random variableN(T )

as the number of spikes in a time windowT , its varianceV ar[N(t)] is then related to the

correlation function (without the Dirac delta term, i.e.C∗(t)) by means of (see e.g. [Gabbiani

and Koch, 1998])

V ar[N(T )] =∫ T

0

∫ T

0(C∗(t)− ν2) dt dt′ (A.0.2)

At the same time we know that the mean ofN(T ) is νT . The Fano factor of the spike count

is defined as the ratio between its variance and its mean

FT ≡V ar[N(t)]

< N(T ) >(A.0.3)

Thus, computing the integrals in eq.A.0.2, substituting them in the expression ofFT and

taking the limit of a large time windowT →∞, one obtains

F = 2A + 1 (A.0.4)

167

Page 182:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

168 Appendix A: The exponentially correlated input

In the largeT limit, for a renewal process, the Fano factor is related to the coefficient of

variation of the ISICVisi, defined as the ratio between the standard deviation and the mean

of the ISI’s:

CVisi ≡√

< ISI2 > − < ISI >2

< ISI >. (A.0.5)

This relation reads [Gabbiani and Koch, 1998]

CV 2isi ' FT (A.0.6)

Finally, we can express the areaA under the exponential of the correlation function as a

function of theCVisi which is simply,

A =CV 2

isi − 1

2(A.0.7)

A direct relationship can be established between the correlation functionC(t) and the

p.d.f. of the ISIs,ρisi(t) . If we assume thatρisi(t) is defined to be zero fort < 0, the relation

can be written as the following:

C(t) = ν(δ(t) + ρisi(t) +

∫ t

0dxρisi(x)ρisi(t− x)+ (A.0.8)

+∫ t

0dx∫ t

0dyρisi(x)ρisi(y)ρisi(t− x− y) + . . .

)This infinite series represents all the possible ways of finding two spikes at times zero and

t, namely:νδ(t) is the density joint probability of finding a spike at time zero and thesame

spike at timet. That is why the distribution functionδ(t) is centered at time zero, showing

that if the second timet is zero then the probability will be one, and otherwise it will be

null. The second term,νρisi(t), represents the joint probability density of finding a spike at

time zero and another one at timet, when there arezerospikes among them. The third term

stands for the joint p.d.f of having a spike at time zero and another one at timet, when there

was only one spike among them. Etc. Each of these terms is indeed the convolution ofn

functionsρisi(t) (n = 0, 1, 2, . . .), so that the previous formula can be expressed

C(t) = νδ(t) + ν∞∑

n=1

(∗ρisi(t))n (A.0.9)

where the asterisk∗ denotes the convolution operation, and(∗ρisi(t))n the convolution ofn

functionsρisi(t) . To perform this calculation explicity we resort to a well known property

of the convolution operation, namely that the Laplace transform [Gradshtein et al., 1980] of

the convolution of functions is the product of the Laplace transformed of the functions. We

therefore take the Laplace transform of the previous equation obtaining:

Page 183:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

169

C(s) = ν + ν∑∞

n=1(ρisi(s))n = ν

∞∑n=0

(ρisi(s))n (A.0.10)

But, not by chance, the last sum is simply a geometrical series which can be explicity

summed resulting

C(s) =ν

1− ρisi(s)(A.0.11)

Now, independently of this, we can compute the Laplace transform ofC(t) (defined in

equationA.0.1), resulting in

C(s) = ν +A

1 + τcs+

ν2

s(A.0.12)

We can then substitute this expression into eq.A.0.11 and after some algebra we obtain an

expression for the Laplace transform of ISI p.d.f.:

ρisi(s) =s(A + ν2τc) + ν2

s2ντc + s(A + ν2τc + ν) + ν2(A.0.13)

Denoting byβ1 andβ2 the roots of the denominator of this fraction we can rewriteρisi(s) as

ρisi(s) =β1(1− ε)

β1 + s+

β2ε

β2 + s(A.0.14)

where the three parameters[β1, β2, ε] are complex functions of the input parameters[ν, τc, CVisi],

namely

ε = 1/2

(2 τc ν + CV 2 + 1−

√λ) (

2 τc ν + CV 2 − 3−√

λ)

4 τ 2c ν2 + 4 τc ν (CV 2 − 3)− 2 τc ν

√λ + (CV 2 + 1)2 −

√λ(CV 2 + 1)

(A.0.15)

β1 = 4ν

2 τc ν + CV 2 + 1−√

λ(A.0.16)

β2 = 1/42 τc ν + CV 2 + 1−

√λ

τc

(A.0.17)

where the auxiliary variableλ is

λ ≡ 4 τc2ν2 + 4 τc ν CV 2 − 12 τc ν + (CV 2 + 1)2 (A.0.18)

The final step is to take the Laplace anti-transform ofρisi(s) given in eq. A.0.14. We

obtain the announced result:

ρisi(t) = (1− ε)β1e−β1t + εβ2e

−β2t , t > 0 (A.0.19)

Page 184:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

170 Appendix A: The exponentially correlated input

This p.d.f. displays a symmetry: it remains invariant under the transformationβ1 ↔β2, β2 ↔ β1, ε ↔ 1 − ε. Because of this, the mapping eq.A.0.15-A.0.18 is one of the two

possible variable choices. The inverse mapping, however, does not have any ambiguous

freedom, and it reads

ν =β1β2

(1− ε)β2 + εβ1

(A.0.20)

CV 2 =2((1− ε)β2

2 + εβ21

)((1− ε)β2 + εβ1)2

− 1 (A.0.21)

τc =1

(1− ε)β2 + εβ1

(A.0.22)

But are the three parameters[ν, τc, CVisi] independent? or in other words, doesρisi(t)

represent a well-defined probability distribution function for any choice of[ν, τc, CVisi]? The

answer is no. The renewal character of the process, constraints the space of input parame-

ters because the functionρisi(t) must satisfy a few conditions in order to be a well-defined

p.d.f. Thus, we must find the restrictions on[ν, τc, CVisi] by mapping the allowed values

of [β1, β2, ε], which in turn are obtained by formulating the conditionsρisi(t) must hold.

These conditions are simply: i) to be integrable, ii) to have norm one and iii) to be defined

as non-negative in the ranget > 0, that is,ρisi(t) ≥ 0 for all t > 0. The first and second

conditions hold by construction by just ensuring that the coefficients of the exponentials are

non-negative:β1, β2 ≥ 0. The condition which ensures thatρisi(t) is always positive or zero

can be broken up in two equivalent statements: a) thatρisi(t) is non-negative at the origin:

ρisi(t = 0) ≥ 0 and that the tail of the function is also positive1. This second statement is

equivalent to say that the coefficient of theslowestexponential (the one with the smallestβ)

is positive. Thus, let us assume without any lost of generality (due to the symmetry inρisi(t)

) thatβ2 > β1. Then, these two conditions are formalized by the following inequalities

(1− ε)β1 + εβ2 ≥ 0 (A.0.23)

(1− ε) > 0 (A.0.24)

After some manipulations, both inequalities can be summarized in

−β1

β2 − β1

< ε < 1 if β2 > β1 (A.0.25)

If we now take these inequalities, using the transformations given by equationsA.0.20-

A.0.22we arrive at the inequalities

ν, τc, CVisi ≥ 0 (A.0.26)

1It can be rigorously demonstrated that if a combination of two exponentials at the origin with a non-

negative value and converges to zero above the x-axis, then it is positive for all positivex values

Page 185:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

171

τc >1− CV 2

2νif CV < 1 (A.0.27)

Thus, for positively correlated inputs (CVisi > 1) any positive value ofν andτc is allowed,

whereas for negative correlation (CVisi < 1), the correlation time constant must exceed a

certain fraction of the mean ISI.

Page 186:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 187:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Appendix B

The computation ofρiri(t|ν) for a general

renewal input process

In this appendix we will show how to obtain the expression of the p.d.f. of the inter-

response intervals (IRI)ρiri(t|ν), given a general renewal inputρisi(t|ν), for the model with

a single synaptic contact described in section2.2.1.

This synaptic model is completely described by two functions:

i) The release probability (we adopt themono-vesicle release hypothesisso only one

vesicle can be released at a time, see section1.2.4) when the ready-for-release pool

(RRP) hasN vesicles:

pr(N) = UΘ(N − 1) N = 0 . . . N0 (B.0.1)

which is the step-like function introduced in section2.2.1, whereN0 denotes the max-

imum size of the RRP. Notice as well, that we have already substituted the initial

parameterNth (see2.2.1) by one. As explained in section2.2.3, consideringNth dif-

ferent from one is only relevant during transient states. In this appendix we consider

the stationary situation and therefore we can setNth = 1.

ii) The probability of refillingn docking sites during a time-window∆, whenN were

empty:

Prec(n, ∆|N) =

N

n

(1− e−∆/τv)n (e−∆/τv)(N−n) (B.0.2)

This expression relies on the assumption that the refilling of each empty docking site

occurs independently at a constant rate1τv

. This is a binomial distribution of parameters

N (the number of avaliable sites) and(1− e−∆/τv) the refilling probability of a single

docking site.

173

Page 188:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

174 Appendix B: Computing ρiri(t|ν) for any renewal input

sr

sr

sr sr

sr

sr

∆spk spk spk spk

spkspkspk

spk spkρ (∆)(1)

sr

ρ (∆)sr

ρ (∆)sr

(2)

(3)

+

+

+

. . .

. . .

n n

n n n

n 1

1

1 2

2

3

N N−1 N−1+n1ves. ves.ves.

t t

t 1

1 2

Figure B.1: Diagram showing the nomenclature and logic of the terms in whichρiri(t) is

expanded

We need compute now the probabilities℘N(t) of havingN vesicles ready in the RRP

at time t. We have already described in chapter2 section2.2.3, a method to obtain these

quantities for ageneralrenewal process in the stationary state. We will then assume that

℘N(t) (N = 0 . . . N0) are known in the stationary state (℘ssN ).

To computeρiri(∆|ν), we expand it in the following way,

ρiri(∆|ν) = ρ(1)iri (∆|ν) + ρ

(2)iri (∆|ν) + ρ

(3)iri (∆|ν) + . . . (B.0.3)

whereρ(i)iri(t|ν) is the p.d.f. of having an IRI of length∆ composed ofi consecutive ISIs.

FigureB.1 shows a diagram illustrating the meaning of the first three terms and the variables

used in their calculation. The first one reads

ρ(1)iri (∆|ν) =

N0∑N=0

℘ssN

N0−N+1∑n1=0

Prec(n1, ∆|N0 −N + 1) ρisi(∆|ν) pr(N − 1 + n1)

The structure of this first term may give us some intuition about the others. It can be de-

scribed as the probability to find a second release at time∆ given that there was one at time

zero and that no spikes failed to trigger a response within this time window∆. Term by term

it can be explained as:

i)∑N0

N=0 ℘ssN : The summed probabilities of havingN vesicles upon arrival of a spike at

time zero given that it elicits a release. Thehat above℘ssN stands for the conditioning,

Page 189:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

175

and it can be solved using the Bayes rule:

℘ssN =

℘ssNpr(N)∑N0

k=0 ℘ssk pr(k)

(B.0.4)

ii)∑N0−N+1

n1=0 Prec(n1, ∆|N0 −N + 1): The summed probabilities thatn1 empty sites are

recovered in the time window∆, whenN0 −N + 1 were avaliable at time0+.

iii) ρisi(∆|ν): The probability that a new spike, after the one which arrived at time zero,

comes at time∆.

iv) pr(N − 1 + n1): The probability that a release occurs, given that at time∆ there were

N − 1 + n1 vesicles in the RRP (there wereN at time0−, N − 1 at time0+ and

N − 1 + n1 at time∆.

Let us write now the second term of the expansion to see how these summations are

nested one in another:

ρ(2)iri (∆|ν) =

N0∑N=0

℘ssN

∫ ∆

0dt1ρisi(t1|ν)

N0−N+1∑n1=0

Prec(n1, t1|N0 −N + 1)×

× (1− pr(N − 1 + n1)) ρisi(∆− t1|ν) ×

×N0−N+1−n1∑

n2=0

Prec(n2, ∆− t1|N0 −N + 1− n1) pr(N − 1 + n1 + n2)

The structure of this term is similar to the termρ(1)iri (∆|ν). However we now need to consider

the occurrence of an unsuccessful spike at timet1 < ∆, and we must to integrate over this

time. Thus, we have added the probability that the spike att1 does not trigger any release:

(1− pr(N − 1 + n1). The recovery of vesicles has to be done now in two stages namely,n1

empty sites are occupied beforet1, andn2 are refilled betweent1 and∆.

These terms get complicated very soon as the number of ISIs in∆ is increased. Therefore

it is not feasible to sum up the complete series before substituting the functionspr(N) and

Prec(n, ∆|N) defined by the synaptic model. Other simple models of releasepr(N) have

been tested at this point. For instance, a linear release functionpr(N) = U N allows the

calculation of the two or three further terms, but we could not achieve the addition of the

complete series. An exponential release function, makes things even harder. Thus, we intro-

duce our synaptic model by substituting the functionspr(N) andPrec(n, ∆|N) of equations

B.0.1andB.0.2. We need to recall here the strategy adopted to perform the sum of expansion

B.0.3 (as introduced in section2.2.3): we will divide the synaptic channel into two stages,

namely i) a purely random channel which decimates the input with a probabilityU giving a

Page 190:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

176 Appendix B: Computing ρiri(t|ν) for any renewal input

decimated version of it,ρisi(t)→ρdisi(t) ; and ii) the vesicle dynamics which takes the diluted

input and gives the synaptic responses:ρdisi(t) → ρiri(t). This division of the problem into

two steps simplifies enormously the calculation since now, a spike may fail to provoke the

release of a vesicle only if it finds the RRP empty. This important point, allows us to reduce

the number of possibletrajectoriesthe system may follow from time0 to ∆.

Let us start with the termρ(1)iri (∆|ν). Substitutingpr(N) andPrec(n, ∆|N) in equation

B.0.4

ρ(1)iri (∆|ν) = ρd

isi(∆|ν)N0∑

N=0

℘ssN ×

×N0−N+1∑

n1=0

N0 −N + 1

n1

(1− e−∆/τv)n1 (e−∆/τv)(N0−N+1−n1) ×

× Θ(N + n1 − 1) (B.0.5)

In first place, the probability℘ssN=0 equals zero because, if at time zero the RRP is empty,

there is no chance to produced the first release (and the conditioning states that it happened).

The second thing to notice is that, if at time zeroN > 1, then the Heaviside function is

simply one for alln1 cases, and we can sum up to one all the recovery probabilities. When at

time zeroN = 1, only the case in which no vesicles are recovered (n1 = 0), is zero because

of Θ(N +n1−2), so we can sum up the recovery probabilities to one, and at the end subtract

the probability thatn1 = 0:

ρ(1)iri (∆|ν) = ρd

isi(∆|ν)

N0∑N=2

℘ssN + ℘ss

1 (1− e−N0∆/τv)

= ρd

isi(∆|ν)(1− ℘ss

1 e−N0∆/τv

)(B.0.6)

In the rest of the termsρ(i)iri (i > 1) only the caseN = 1 survives because, since there

have to be failures, we need the RRP to be empty at the intermediate timesti. After some

manipulation the second term reads

ρ(2)iri (∆|ν) =

∫ ∆

0dt1ρ

disi(t1|ν)ρd

isi(∆− t1|ν)℘ss1 e−N0t1/τv

(1− e−N0(∆−t1)/τv

)Following the same methodology and denoting the initial time ast0, we can write the

expression of thej − th term

ρ(j)iri(∆|ν) =

∫ ∆

0

j−1∏i=1

(dti ρd

isi(ti − ti−1|ν) e−N0(ti−ti−1)/τv

× ρdisi(∆− tj−1|ν) ℘ss

1

(1− e−N0(∆−tj−1)/τv

)(B.0.7)

Page 191:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

177

In order to perform the integrals we need to observe that each term is composed by the

convolutionof j functionsρdisi(t), and make use of the well known property of the Laplace

transform [Gradshtein et al., 1980], which says that the transform of the convolution of two

functions is the product of the transform functions, that is

L[∫ ∆

0ρ(t) ρ(t−∆) dt

]= L [ρ(t)]L [ρ(t)] = ρ(s)2 (B.0.8)

We also need the property

L[ρ(t) e−t/τ

]= ρ(s +

1

τ) (B.0.9)

which can be easily proven from the definition of the Laplace transform. With these two

properties we can now transform each termρ(j)iri(∆|ν) obtaining

ρ(1)iri (s) = ρd

isi(s) − ℘ss1 ρd

isi(s + N0/τv) (B.0.10)

ρ(2)iri (s) = ℘ss

1

[ρd

isi(s) ρdisi(s + N0/τv) − (ρd

isi(s + N0/τv))2]

(B.0.11)

...

ρ(j)iri(s) = ℘ss

1

[ρd

isi(s) (ρdisi(s + N0/τv))

j−1 − (ρdisi(s + N0/τv))

j]

(B.0.12)

Finally, we can sum the Laplace transform of the expansion:

ρiri =∞∑

j=1

ρ(j)iri(s) =

= ρdisi(s)− ℘ss

1

(1− ρd

isi(s)) ∞∑

j=1

(ρd

isi(s + N0/τv))j

=

= ρdisi(s)− ℘ss

1

(1− ρd

isi(s))

ρdisi(s + N0/τv)

1− ρdisi(s + N0/τv)

(B.0.13)

The last step has been performed observing that the last sum is simply a geometric series.

Now we can express the conditioned℘ss1 in terms of the non-conditioned probabilities (see

eq.2.2.19)

℘ssN=1 ≡

℘ss(1)

℘ss(1) + ℘ss(2) + . . . + ℘ss(N0)(B.0.14)

We have then reached the expression given in equation2.2.18. It gives the Laplace transform

of the p.d.f. of the IRIsρiri for a generalrenewal input, given by the transform of the p.d.f.

of thedilutedinput versionρdisi(s). The link between this decimated function and the original

ρisi(t) , was derived in section2.2.3and is given by equation2.2.21.

Page 192:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 193:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Appendix C

The calculation of the population

distribution D(U,N0)

In this section we will derive the computation of the distribution of the release probability

U and the number of docking sitesN0 across a population of synapses. This distribution ful-

fills two constraints derived from measurements done in hippocampal CA3-CA1 synapses,

namely: i) the marginal distribution ofU follows a Gamma function of order twoΓλ(U)

(whereMurthy et al.[1997] reportλ = 7.4). ii) the population averaged ofU for a fixedN0,

〈U〉R(U |N0), follows an exponential dependence inN0 〈U〉R(U |N0) = 1 − (1 − pv)N0 ' aN0

(whereMurthy et al. [2001] report pv = 0.055 andHanse and Gustafsson[2001a] report

pv ' 0.3− 0.7).

First, the joint distributionD(U,N0) is expressed in the following way:

D(U,N0) = f(N0) Rq(U |N0) (C.0.1)

Now we make theansatzthat the conditional distributionRq(U |N0) is a Gamma function of

orderN0 + 1 and parameterq:

Rq(U |N0) =qN0+1

N0!UN0e−qU (C.0.2)

Thus, the first constraint implies:

Γλ(U) ≡ λ2

2Ue−λU =

∞∑N0=1

f(N0)Rq(U |N0) (C.0.3)

and expanding both sides of the equality in powers ofU we obtain:

∞∑k=0

[(−λ)kλ2

2 k!

]Uk+1 =

∞∑j=0

∞∑N0=1

f(N0)Rq(U |N0)

(j)∣∣∣U=0

j!

U j (C.0.4)

179

Page 194:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

180 Appendix C: The calculation of the population distribution D(U,N0)

Identifying coefficients order by order (makingj = k + 1) we have

(−λ)k+2

2 k!=

∞∑N0=1

f(N0)Rq(U |N0)

(k+1)∣∣∣U=0

(k + 1)!k = 0, 1 . . .∞ (C.0.5)

But thenth derivative of the functionRq(U |N0) reads,

dnRq(U |N0)

dUn

∣∣∣∣∣U=0

=qN0+1

N0!

dn(UN0 e−qU)

dUn

∣∣∣∣∣U=0

=

=

qN0+1

N0!n!

(n−N0)!(−q)n−N0 if N0 ≤ n

0 otherwise(C.0.6)

Inserting this into eq.C.0.5, we obtain for each value ofk = 0, 1 . . .∞ a simple first

order equation withk+1 distribution coefficientsf(N0). These can be obtained in a recurrent

manner from:

(−λ)k+2

2 k!=

k+1∑N0=1

f(N0)(−)k+1−N0qk+2

N0! (k + 1−N0)!k = 0, 1 . . .∞ (C.0.7)

which after some manipulation reads

1

2

q

)k+2

=k+1∑

N0=1

[(−)N0−1k!

N0! (k + 1−N0)!

]f(N0) k = 0, 1 . . .∞ (C.0.8)

that can be written using the new variablej = k + 1(λ

q

)j+1

=j∑

N0=1

j

N0

(−)N0−1

jf(N0) j = 1 . . .∞ (C.0.9)

We show the first three values ofj:

f(1) =

q

)2

(j=1)

f(1)− f(2)

2=

q

)3

(j=2)

f(1)− f(2) +1

3f(3) =

q

)4

(j=3)

...

These equations form an infinite diagonal system of linear equations. We can easily

obtain the values off(N0) for N0 = 1, 2, . . . Nmax0 , where the cutoffNmax

0 is chosen in such

a way that the probability∑∞

N=Nmax0

f(N) is smaller than a given tolerance value. As can

Page 195:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

181

also be deduced from the system, the probabilitiesf(N0) are polynomials in(

λq

)of grade

N0 + 1.

To obtain the value ofq we need to make use of the second constraint, which yields,

〈U〉R(U |N0) =∫ ∞

0dU U Rq(U |N0) =

=N0 + 1

q(C.0.10)

All is left to do, is to take the fit to the data〈U〉R(U |N0) = 1− (1− pv)N0 [Murthy et al.,

2001], and set the following equality

N0 + 1

q' 1− (1− pv)

N0 (C.0.11)

Although this equality seems hard to fulfill, sincepv has a very small value (∼ 0.05) the

r.h.d. can be expanded up to first order, and then the slopes of both sides can be identified.

Figure2.5, shows how for the values of the experiments found byMurthy et al.[1997, 2001],

the approximation is rather good. Once we have set the value ofq, taking a value forλ, we

can obtain the probabilitiesf(N0) for all N0’s up to the upper cutoffNmax0 .

Page 196:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 197:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Appendix D

Computation of the output statistics for

the exponentially correlated input when

N0 = 1

D.1 Computation of the p.d.f. of the IRIsρiri(t)

In this appendix we will compute the final exact expression of the probability distribution

function of the inter-response-intervals (IRIs) for the particular cases: i) one single docking

site, i.e. the maximum size of the ready-releasable-pool (RRP) is only oneN0 = 1, and ii)

the case where the input is the renewal process with exponential auto-correlation function

defined in appendixA. Thus our starting point is the expression of the Laplace transform of

the p.d.f. of the IRIs (ρiri) as a function of the distribution of the decimated version of the

input ρdisi(s), given by equationB.0.13derived in appendixB that we now rewrite

ρiri = ρdisi(s)− ℘ss

1

(1− ρd

isi(s))

ρdisi(s + N0/τv)

1− ρdisi(s + N0/τv)

(D.1.1)

We also need equation2.2.21, which givesρdisi(s) in terms of the original inputρisi(t) , which

we also rewrite here,

ρdisi(s) =

Uρisi(s)

1− (1− U)ρisi(s)

We now take eq.D.1.1and makeN0 = 1. The probability℘ss1 becomes naturally one (see

eq. 2.2.19). We then take the expression ofρdisi(s) given by the second of these equations

(eq. D.1.2) and introduce it in the first. After some simple algebra we obtain an expression

for ρiri in terms of the original distribution of input ISIs:

ρiri =U [ρisi(s)− ρisi(s + 1/τv)]

[1− (1− U)ρisi(s)] [1− ρisi(s + 1/τv)](D.1.2)

183

Page 198:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

184 Appendix D: Computation of the output when N0 = 1

All we need to do now, is to take the expression ofρisi(s) derived in appendixA

ρisi(s) =s(A + ν2τc) + ν2

s2ντc + s(A + ν2τc + ν) + ν2(D.1.3)

and to introduce it inD.1.2. After some manipulations, it reads

ρiri(s) = Us2(Aτcτv + ντ 2

c τv) + s(Aτc + ντc(τv + τc)) + ν(τc + τv)

(sτv + 1)(sτcτv + τc + τv)(s2τc + s(1 + UA + ντcU) + νU)(D.1.4)

We must now anti-transform this equation. This is done by decomposing it in four frac-

tions with a single pole in the variables. Each of the four poles represents the decaying time

of an exponential inρisi(t) . After some algebra we arrive at the desired expression, a sum

of 4 exponentials (which we showed in eq.2.3.6):

ρiri(∆) =U

τv

(C1 e∆s1 + C2 e∆s2 + C3 e−

∆τv + C4 e

− ∆τ1

)(D.1.5)

where the coefficientsCi are:

Ci =si

(A

τcτ2+ ν

τ1− UE2

2

)+ ν

τcτ1− E2νU

τc

Di

(D.1.6)

Di = si

(−E1

(1

τv

+1

τ1

− E1

)+

2

τvτ1

− 2Uν

τc

)+

UνE1

τc

− 2

(1τv

+ 1τ1

)Uν

τc

+E1

τvτ1

for i = 1, 2. The other two read

C3 = A τvτ−1c

(1− Uν τ1 −

Uτv A

τc

)−1

(D.1.7)

C4 = −ν τv

(1− Uν τv −

Uτ2

τc

)−1

(D.1.8)

The decaying time constants of the four exponentials, come from the four poles of equa-

tion D.1.4. They are the vesicle recovery timeτv, the new time constant

τ1 ≡ τv τc

τc + τv

(D.1.9)

and the inverse of the two solutionss1, s2 of equation

s2 + s(1/τc + UA/τc + νU) + νU/τc = 0 (D.1.10)

which read

s1 = −1/2E1 + 1/2

√E1

2 − 4Uν

τc

(D.1.11)

s2 = −1/2E1 − 1/2

√E1

2 − 4Uν

τc

(D.1.12)

Page 199:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

D.2. Computation of the correlation function of the IRIs 185

Besides, we have defined the following auxiliary variables

τ2 ≡ τv τc

τc − τ(D.1.13)

E1 ≡ τ−1c + U

(A

τc

+ ν)

(D.1.14)

E2 ≡ A

τc

+ ν (D.1.15)

Finally, to show the dependence of all variables on the physical parameters, we write here

again the expression of the areaA in terms of the coefficient of variationCVisi (see appendix

A),

A =CV 2

isi − 1

2(D.1.16)

D.2 Computation of the correlation function of the IRIs

We will now obtain the output autocorrelation function of the synaptic responses. Since

for N0 = 1 the train of releases is a renewal process (see section2.2.3), we can relate the

p.d.f. of the IRIs,ρiri(t) , with the auto-correlation function of the responses,Cr(t), by means

of eq.A.0.11, originally derived for the input spike train, but now applied to the output train:

Cr(s) =νr

1− ρiri(s)(D.2.1)

However, to simplify the calculation, we will not compute the correlation,Cr(t) , but a the

conditional rate of responses,Cr(t) (see eq.2.2.5), which equals the correlation function

withoutthe Dirac function, and normalized by the rateνr. Its Laplace transform is

Cr(s) = Cr(s)/νr − 1 =ρiri(s)

1− ρiri(s)(D.2.2)

All we need to do now,substitute here equationD.1.4. After some algebra, we reach

Cr(s) = UQ1

Q2Q3

(D.2.3)

where

Q1 = 2 ν τc + 2 ν s2τc2τv + s22Aτc τv + 4 ν τc sτv + 2 ν τv + 2sAτc + 2 ν sτc

2

Q2 = s−1 (sτc + 1)−1

Q3 = 2 τc s2τv2 + 4 τc sτv + 2 sτv

2ν τc U + 2sτv2AU +

+ 2 sτv2 + 2 τv + 2 τv ν τc U + 2 τv

2ν U + 2 τc + 2τv AU

Page 200:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

186 Appendix D: Computation of the output when N0 = 1

The denominatorQ2 Q3 has four roots, one of thems = 0. This means that, when anti-

transforming we get a linear combination of three exponentials plus a constant term, namely

Cr(t) = K1 e−t( 1τv−s1) + K2 e−t( 1

τv−s2) + K3 e−

tτc + K4 (D.2.4)

where the coefficients of the exponentials are

K1 = − (s1 τc + 1) s1 (s1 τv − s2 τv − 1)

(s1 − s2) (s1 τv − 1) (s1 τc τv − τc + τv)

K2 = − (s2 τc + 1) s2 (s2 τv − s1 τv − 1)

(s2 − s1) (s2 τv − 1) (s2 τc τv − τc + τv)

K3 =(τv − τc) (s1 τc + 1) (s2 τc + 1)

(−τc + τv + s2 τc τv) (s1 τc τv − τc + τv)

K4 =(τc + τv) s2 s1

(−1 + s2 τv) (s1 τv − 1)

ands1 ares2 are the same as in eqs.D.1.11-D.1.12. Thus we have obtained the expression

for the conditional rate of the responses given in eq.2.3.7. With this expression we are able

to compute the rate of releasesνr and the coefficient of variation of the IRI’sCViri.

D.3 Release firing rateνr and coefficient of variationCViri

Once we have the IRI distribution and the autocorrelation function of the output train

we can easily calculate the firing rate by merely taking the limitt → ∞ of the conditional

rateCr(t). This is because it has to converge to the non-conditioned rate if things are well-

defined. Thus, the response rateνr equals the constant coefficientK4 of Cr(t). Using the

expressions ofs1 ands2 in terms of the input parameters (eqs.D.1.11-D.1.12) and simplify-

ing the result, we obtain

νr =νU

1 + τvνU + τvUCV 2−12(τv+τc)

(D.3.1)

which is the expression given in eq.2.3.8introduced in section2.2.3.

To calculate the coefficient of variation of the postsynaptic responses we use the relation-

ship established in appendixA, between the correlation function of a input renewal process

and itsCV , eqs. A.0.2, A.0.3 andA.0.6. There, it was expressed as [Gabbiani and Koch,

1998]:

Page 201:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

D.3. Release firing rateνr and coefficient of variationCViri 187

CV 2iri − 1

2=

∫∞0

∫∞0 (C∗r (t)− ν2

r ) dt dt′

νr T

=∫ ∞

0Cc

r(t)dt =

= K1 (1

τv

− s1)−1 + K2 (

1

τv

− s2)−1 + K3 τc (D.3.2)

All we need to do now, is to replace the expressions of the coefficientsKi and of the roots

si and do lots of algebra. The final expression of the coefficient of variation of the responses

reads

CV 2iri =

H

(τc + τv + τv Uν τc + τv UA + τv2Uν)2 (D.3.3)

where

H = τv4U2ν2 + 2 τv

3U2ν2τc + τv2U2A2 + 4 τv

2U2ν A τc + 2 UA τv2 + τv

2 + τc2 +

+τv2U2ν2τc

2 + 2 τv U2A2τc + 4 τv UA τc + 2 τv U2ν A τc2 + 2 τv τc + 2 UAτc

2

whereA is the area under the input exponential autocorrelation which equalsA = (CV 2isi −

1)/2.

Page 202:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments
Page 203:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Appendix E

Computation of the conditioned

probability 〈pv(β|α)〉

In this appendix we will compute the probability that a spike reaching theβ-th synaptic

terminal finds a vesicle ready for release, known that at theα-th contact, which belongs to the

samepre-synaptic neuron, there is one. This probability is averaged over the input ensemble

(see angle brackets) so that it does not depend on the times of the afferent spikes.

By direct application of the Bayes rule we find that

〈pv(β|α)〉 =〈pv(α, β)〉〈pv(β)〉

(E.0.1)

where〈pv(α, β)〉 is the joint probability that both contacts have their docking sites occu-

pied, and〈pv(β)〉 is the probability that a single contact (regardless of the index) is recovered

(computed in section4.4.1). So, in order to obtain the conditional probability〈pv(β|α)〉, we

need to calculate the joint probability〈pv(α, β)〉 of finding the two contacts filled.

At any time t, its state is well defined by the three probabilities℘2(t) ≡ 〈pv(α, β)〉,℘1(t) and℘0(t). These three functions stand for the probabilities of finding both, only one

or none of the two vesicles ready upon the arrival of a spike at timet. Note, that℘1(t) is

not equal to〈pv(β)〉, because the former represents the system with one contact recovered

and one empty, but the later considers only a single contact. To derive the expression of all

these probabilities, we must establish the system of differential equations which govern the

dynamics of the three probabilities.

FigureE.1 shows the probability flow diagram of the system of two contacts, like the

one shown before illustrating the single contact case (fig.4.2). The state of the system at

time t is defined by the number of vesicles avaliable for release, which can be two, one or

none1 We define the transition probabilitiesT [m → n; t] (n, m = 0, 1, 2) as the probabilities

1Do not confuse this system with the synaptic model of a single contact withN0 vesicle docking sites,

189

Page 204:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

190 Appendix E: Computation of the conditioned probability 〈pv(β|α)〉

P2(t)

P1(t+dt)

P0(t)

P1(t)

P2(t+dt)

P0(t+dt)

T[(2-->1]

T[0-->1]

T[(2-->1]

T[(2-->2]

Figure E.1: Diagram of the temporal evolution of a system composed of two synaptic con-

tacts belonging to the same pre-synaptic neuron. At timet the state of the system can be2

contacts ready for release with probability℘2(t), 1 contact ready with prob.℘1(t), or zero

contacts ready with prob.℘0(t). A time stepdt later, a transition may have occurred: a

release of one or two vesicles or the recovery of one vesicle, depending on the state of the

system at timet. (Note:Pi(t) the figure represent℘2(t) in the text.)

that the system switches its state, fromn vesicles at timet, to a new state withm vesicles at

time t + dt. Transitions between states occur by means of two processes: arrival of a spike

and subsequent release, or recovery of vesicles at any of the contacts. The probability that

a spikes arrives at both contacts equalsνdt, while the probability that a vesicle is recovered

depends on the number of avaliable ones: if both boutons are empty we havetwo recovery

Poisson processes acting at the same time (solid green line in fig.E.1), so the probability is

twice as much as if one contact is already occupied (dashed green line in fig.E.1). Thus, the

described in section2.2.1. Now, we are considering a system made up of a group of contacts (two in this

calculus) each of them having one or cero vesicles avaliable. The dynamics of both systems is different mainly

because now, more than one vesicle can be released at the same time, while in the single contact model, even if

there areN > 1 vesicles ready-for-release, at most one finally fuses the membrane.

Page 205:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

191

recovery transitions read

T [0 → 1; t] = ℘0(t)dt

τv

(E.0.2)

T [1 → 2; t] = ℘1(t)2 dt

τv

(E.0.3)

T [0 → 2; t] = ℘0(t)

(dt

τv

)2

(E.0.4)

On the other hand the release transitions (red lines in fig.E.1) depend on the arrival of a

spike and on the probability of releaseU :

T [2 → 1; t] = ℘2(t) ν dt 2 U (1− U) (E.0.5)

T [2 → 0; t] = ℘2(t) ν dt U2 (E.0.6)

T [1 → 0; t] = ℘1(t) ν dt U (E.0.7)

Naturally, the transitions in which the system state does not change (remaining transi-

tions, blue lines in fig.E.1) are obtained by applying the normalization principle which sets

that when the system is at any state, a transition must occur:

T [2 → 2; t] = ℘2(t)− T [2 → 1; t]− T [2 → 0; t] (E.0.8)

T [1 → 1; t] = ℘1(t)− T [1 → 0; t]− T [1 → 2; t] (E.0.9)

Thus, the probabilities at timet + dt can be obtained through

℘(n; t + dt) =∑m

T [m → n; t] = (E.0.10)

= T [n → n; t] +∑m6=n

T [m → n; t] =

= ℘(n; t)−∑m6=n

T [n → m; t] +∑m6=n

T [m → n; t]

n,m = 0, 1, 2

Taking the term℘(n; t) to the left side, dividing bydt and taking the limitdt → 0 one

reaches a three dimensional differential system. But using the normalization condition

℘(0; t) + ℘(1; t) + ℘(2; t) = 1 , ∀t (E.0.11)

we can get read of one of the three probabilities reaching the 2-dim. differential system

Page 206:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

192 Appendix E: Computation of the conditioned probability 〈pv(β|α)〉

d℘1(t)

dt= −℘1(t) (νU +

3

τv

) + 2[Uν(1− U)− 1

τv

]+

2

τv

(E.0.12)

d℘2(t)

dt= ℘1(t)

1

τv

− ℘2(t) Uν(2− U)

Making the derivatives equal to zero and solving the linear system we obtain the stationary

state solution which reads

℘ss1 =

U ν τv (2− U)

(1 + Uντv)[1 + Uντv (1− U/2)](E.0.13)

℘ss2 =

1

(1 + Uντv)[1 + Uντv (1− U/2)](E.0.14)

where the super indexss stands for stationary state. Finally, we can compute the condi-

tioned probability〈pv(β|α)〉 of finding a vesicle in theβ-th contact when a vesicle has been

observed in theα-th contact at any time in the stationary state2

〈pv(β|α)〉 =〈pv(α, β)〉〈pv(β)〉

=℘ss

2

〈pv(β)〉 =1

[1 + Uντv (1− U/2)](E.0.15)

2The lector must note how we indistinctly use observations atanytime or at the time of the arrival of a spike.

This ambiguous use is not accidental but due to the fact that when using Poisson spike trains, the probability

for the vesicle to be docked at an arbitrary timet is the same that at the arrival of a spike. This happens because

the well known property of the Poisson process of having no memory, so that an event can occur with the same

constant probability at any time.

Page 207:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Appendix F

Computation of the conditioned

probability 〈pv(j|i)〉

In this appendix we will compute the probability that a spike reaching thej-th synaptic

terminal finds a vesicle ready for release, known that at thei-th contact, which receives

spikes from a different neuron whose activity is cross-correlated with thej-th neuron, there

is already one. The calculations follows closely what was done in AppendixE, where the

same probability was derived for contacts belonging to the same neuron. In this sense, this

calculation is a generalization of the other one, since if we set the synchronization parameter

ρ = 1, it would be equivalent to the situations in which both contacts belong to the same

neuron.

The system is again described by the three probabilities℘2(t) , ℘1(t) and℘0(t), and the

flux diagram which describes its evolution is the same as in the previous case, fig.E.1. Be-

cause the recovery dynamics have not been altered, the three recovery transitionsT [0 → 1; t],

T [1 → 2; t] andT [0 → 2; t] are the same as in equationsE.0.2,E.0.3andE.0.4, respectively.

The release transitions, however, have changed because of the new configuration of the in-

coming fibers (compare figs.4.6and4.5), reading

T [2 → 1; t] = ℘2(t) [2νdt (1− ρ) U + νdt ρ2 U (1− U)] (F.0.1)

T [2 → 0; t] = ℘2(t) ν dt ρ U2 (F.0.2)

T [1 → 0; t] = ℘1(t) ν dt U (F.0.3)

The most significant change occurs in transitionT [2 → 1; t], in which the first term on

the right hand side equals the probability of having only one vesicle reaching one terminal

in the interval(t, t + dt) and succeeding in the release, while the second term constitutes

193

Page 208:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

194 Appendix F: Computation of the conditioned probability 〈pv(j|i)〉

the probability of observing two vesicles hitting the synapses at the same time but only one

triggering a release.

Now the probabilities at timet+dt are obtained from the probabilities a time step earlier

by means of

℘(n; t + dt) = T [n → n; t] +∑m6=n

T [m → n; t] = (F.0.4)

= ℘(n; t)−∑m6=n

T [n → m; t] +∑m6=n

T [m → n; t]

and reorganizing this expressions, taking the limitdt → 0 and making use of the normaliza-

tion condition eq.E.0.11, one obtains the following system of differential equations

d℘1(t)

dt= −℘1(t) (νU +

3

τv

) + 2℘2(t)[Uν(1− Uρ)− 1

τv

]+

2

τv

(F.0.5)

d℘2(t)

dt= ℘1(t)

1

τv

− ℘2(t) Uν(2− Uρ)

We take the system to the stationary state (℘i(t) = 0), and the solution reads

℘ss1 =

U ν τv (2− Uρ)

(1 + Uντv)[1 + Uντv (1− Uρ2

)](F.0.6)

℘ss2 =

1

(1 + Uντv)[1 + Uντv (1− Uρ2

)](F.0.7)

Finally, we compute the conditioned probability〈pv(j|i)〉 of finding a vesicle in thej-th

synapse when a vesicle has been observed in thei-th synapse, in the stationary state

〈pv(j|i)〉 =℘ss

2

〈pv(j)〉 =1[

1 + Uντv (1− Uρ2

)] (F.0.8)

On can now check, that ifρ = 1 we recover expression4.4.6concerning two contacts

from the same neuron.

Page 209:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Bibliography

E. D. Adrian. The impulses produced by sensory nerve endings: Part I.J. Physiol. (Lond.),

61:49–72, 1926.

G. Werner and V. B. Mountcastle. Neural activity in mechanoreceptive cutaneous afferents:

Stimulus-response relations, Weber functions, and information transmission.J. Neurophys-

iol., 28:359–397, 1965.

D.J. Tolhurst, J.A. Movshon, and A.F. Dean. The statistical reliability of signals in single

neurons in cat and monkey visual cortex.Vision Res., 23:775–785, 1983.

D.J. Tolhurst. The amount of information transmitted about contrast by neurons in the cats

visual cortex.Vis. Neurosci., 2:409–413, 1989.

K H Britten, M N Shadlen, W T Newsome, and J A Movshon. The analysis of visual

motion: a comparison of neuronal and psychophysical performance.J. Neurosci., 12:4745–

4765, 1992.

M. J. Tovee, E. T. Rolls, A. Treves, and R. P. Bellis. Information encoding and the response

of single neurons in the primate temporal visual cortex.J. Neurophysiol., 70:640–654,

1993.

C. Koch and I. Segev. The role of single neurons in information processing.Nature Neuro-

science, 3:1171–1177, 2000.

H. B. Barlow. Possible principles underlying the transformation of sensory messages. In

W. Rosenblith, editor,Sensory Communication, page 217. M.I.T. Press, Cambridge MA,

1961.

J. J. Atick. Could information theory provide an ecological theory of sensory processing?

Network: Comput. Neural Syst., 3:213–251, 1992.

J.-P. Nadal and N. Parga. Nonlinear neurons in the low-noise limit: a factorial code maxi-

mizes information transfer.Network: Computation in Neural Systems, 5:565–581, 1994.

195

Page 210:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

196 Bibliography

Y. Dan, J. J. Atick, and R. C. Reid. Efficient coding of natural scenes in the lateral geniculate

nucleus: experimental test of a computational theory.J Neurosci, 16:3351–3362, 1996.

J.-P. Nadal, N. Brunel, and N. Parga. Nonlinear feedforward networks with stochastic

outputs: infomax implies redundancy reduction.Network: Computation in Neural Systems,

9:1–11, 1998.

M. S. Goldman, P. Maldonado, and L. F. Abbott. Redundancy reduction and sustained

firing with stochastic depressing synapses.J Neurosci., 22:584–91, 2002.

C. E. Shannon. A mathematical theory of communication.AT&T Bell Labs.

Tech. J., 27:379–423, 1948. URLhttp://cm.bell-labs.com/cm/ms/what/

shannonday/paper.htm .

A. Borst and F. E. Theunissen. Information theory and neural coding.Nature Neuroscience,

2:947–958, 1999.

L. F. Abbott, J. A. Varela, K. Sen, and S. B. Nelson. Synaptic depression and cortical gain

control. Science, 275:179–80, 1997.

M. V. Tsodyks and H. Markram. The neural code between neocortical pyramidal neurons

depends on neurotransmitter release probability.Proc Natl Acad Sci USA, 94:719–723,

1997.

John E. Lisman. Bursts as a unit of neural information: making unreliable synapses reliable.

TINS, 20:38–43, 1997.

W. Senn, I. Segev, and M. Tsodyks. Reading neuronal synchrony with depressing synapses.

Neural Computation, 10:815–819, 1998.

F. S. Chance, S. B. Nelson, and L. F. Abbott. Synaptic depression and the temporal response

characteristics of v1 cells.J. Neurosci., 12:4785–99, 1998a.

W. Maass and A. M. Zador. Dynamic stochastic synapses as computational units.Neural

Computation, 11(4):903–917, 1999.

V Matveev and X-J Wang. Differential short-term synaptic plasticity and transmission of

complex spike trains: to depress or to facilitate?Cereb Cortex., 10:1143–53, 2000a.

Thomas Natschlager and Wolfgang Maass. Computing the optimally fitted spike train for a

synapse.Neural Comp., 13:2477–2494, 2001.

Page 211:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Bibliography 197

W. Maass and H. Markram. Synapses as dynamic memory buffers.Neural Networks, 15:

155–161, March 2002.

G. Fuhrmann, I. Segev, H. Markram, and M. Tsodyks. Coding of temporal information by

activity-dependent synapses.J. Neurophysiol, 87:140–148, 2002.

J. de la Rocha, A. Nevado, and N. Parga. Information transmission by stochastic synapses

with short-term depression: neural coding and optimization.Neucomputing, 44:85–90,

2002.

D. J. Amit and M. V. Tsodyks. Quantitative study of attractor neural network retrieving at

low spike rates: I. substrate–spikes, rates and neuronal gain.Network, 2:259–273, 1991.

M. N. Shadlen and W. T. Newsome. Noise, neural codes and cortical organization.Curr.

Op. Neurobiology, 4:569–579, 1994.

M. N. Shadlen and W. T. Newsome. The variable discharge of cortical neurons: implica-

tions for connectivity, computation, and information coding.J. Neurosci., 18:3870–3896,

1998.

M. F. Bear, B. W. Connors, and M. A. Paradiso.Neuroscience: Exploring the Brain. Wil-

iams & Wilkins, 1996.

KE Sorra and KM Harris. Occurrence and three-dimensional structure of multiple synapses

between individual radiatum axons and their target pyramidal cells in hippocampal area ca1.

J Neurosci, 13(9):3736–48, 1993.

Thomas Schikorski and Charles F. Stevens. Quantitative Ultrastructural Analysis of Hip-

pocampal Excitatory Synapses.J. Neurosci., 17(15):5858–5867, 1997. URLhttp:

//www.jneurosci.org/cgi/content/abstract/17/15/5858 .

B. Walmsley.Prog. Neurobiol., 36:391–423, 1991.

J. del Castillo and B. Katz. Quantal components of the end-plate potential.J. Physiol., 124:

560–573, 1954a.

B. Katz and R. Miledi. The role of calcium in neuromuscular facilitation.J. Physiol., 195:

481–492, 1968.

H. Held. Die centrale gehorleitung.Arch. Anat. Physiol., pages 201–248, 1893.

D.K. Ryugo, M.M. Wu, and T. Pongstaporn. Activity-related features of synapse morphol-

ogy: a study of endbulbs of held.J. Comp. Neurol., 365(1):141–158, 1996.

Page 212:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

198 Bibliography

Bruce Walmsley, Francisco J. Alvarez, and Robert E. W. Fyffe. Diversity of structure and

function at mammalian central synapses.Trends in Neurosciences, 21(2):81–88, 1998.

N.A. Hessler, A.M. Shirke, and R. Malinow. The probability of transmitter release at a

mammalian central synapse.Nature, 366:569–572, 1993.

C. Rosenmund, J.D. Clements, and G.L. Westbrook. Nonuniform probability of glutamate

release at a hippocampal synapse.Science, 262:754–757, 1993.

C. Allen and C. F. Stevens. An evaluation of causes for unreliability of synaptic transmis-

sion. Proc. Natl. Acad. Sci. USA, 91:10380–10383, 1994.

C. F. Stevens and Y. Wang. Changes in reliability of synaptic function as a mechanism for

plasticity. Nature, 371:704–707, 1994.

I. A. Boyd and A. R. Martin. The end-plate potential in mammalian muscle.J. Physiol.

Lond., 132:74–91, 1956.

B. Katz. The Release of Neural Transmitter Substances. Liverpool University Press, 1969.

H.C. Tuckwell. Introduction to Theoretical Neurobiology II. Cambridge University Press,

Cambridge, UK., 1988.

D.K. Smetters and A. Zador. Synaptic transmission: noisy synapses and noisy neurons.

Current Biology, 6(10):1217–1218, 1996.

O Paulsen and P Heggelund. The quantal size at retinogeniculate synapses determined

from spontaneous and evoked epscs in guinea-pig thalamic slices.J Physiol (Lond), 480:

505–511, 1994.

O Paulsen and P Heggelund. Quantal properties of spontaneous epscs in neurones of the

guinea-pig dorsal lateral geniculate nucleus.J Physiol (Lond), 496:759–772, 1996.

MC Bellingham, R Lim, and B Walmsley. Developmental changes in epsc quantal size and

quantal content at a central glutamatergic synapse in rat.J Physiol (Lond), 511:861–869,

1998.

KJ Stratford, K Tarczy-Hornoch, KA Martin, NJ Bannister, and JJ Jack. Excitatory synaptic

inputs to spiny stellate cells in cat visual cortex.Nature, 382:258–261, 1996.

William B Levy and Robert A. Baxter. Energy-Efficient Neuronal Computation via Quan-

tal Synaptic Failures. J. Neurosci., 22(11):4746–4755, 2002. URLhttp://www.

jneurosci.org/cgi/content/abstract/22/11/4746 .

Page 213:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Bibliography 199

A. Zador. The impact of synaptic unreliability on the information transmitted by spiking

neurons.J. Neurophysiol., 79:1219–1229, 1998.

H. Markram and M. Tsodyks. Redistribution of synaptic efficacy between neocortical pyra-

midal neurons.Nature, 382:807–810, 1996a.

Harold L. Atwood and Shanker Karunanithi. Diversification of synaptic strength: presy-

naptic elements.Nature Reviews Neuroscience, 3:497–516, 2002.

K. L. Magelby. Short-term changes in synaptic efficacy. In G.M. Edelman, W.E. Gall, and

W.M. Cowan, editors,Synaptic Function, pages 21–56. John Wiley and Sons, New York,

1987.

S.A. Fisher, T.M. Fischer, and T.J. Carew. Multiple overlapping processes underlying short-

term synaptic enhancement.Trends Neurosc., 20:170–177, 1997.

Robert S. Zucker. Short term synaptic plasticity.Annu. Rev. Neuroscience., 12:13–31, 1989.

Anthony M. Zador and Lynn E. Dobrunz. Dynamic synapses in the cortex.Neuron, 19:

1–4, 1997.

Henrique von Gersdorff and J. Gerard G. Borst. Short-term plasticity at the calyx of held.

Nature Reviews Neuroscience, 3:53–64, 2002.

Robert S. Zucker and Wade G. Regehr. Short-term synaptic plasticity.Annu. Rev. Physiol.,

64:355–405, 2002.

A.W. Liley and K.A.K. North. An electrical investigation of e ects of repetitive stimulation

on mammlian neuromuscular junction.J. Neurophysiol., 16:509–527, 1953.

J.I. Hubbard. Repetitive stimulation of the mammalian neuromuscular junction, and the

mobilization of transmitter.Journal of Physiology, 169:641–662, 1963.

C. F. Stevens and Y. Wang. Facilitation and depression at single central synapses.Neuron,

14:795–802, 1995.

Robert S. Zucker. Exocytosis: A molecular and physiological perspective.Neuron, 17:

1049, 1996.

E Neher. Vesicle pools and ca2+ microdomains: New tools for understanding their roles in

neurotransmitter release.Neuron, 20:389, 1998.

Page 214:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

200 Bibliography

R. Schneggenburger, T. Sakaba, and E. Neher. Vesicle pools and short-term synaptic de-

pression: lessons from a large synapse.Trends in Neurosciences, 25(4):206–212, 2002.

L. E. Dobrunz and C. F. Stevens. Heterogeneity of release probability, facilitation, and

depletion at central synapses.Neuron, 18:995–1008, 1997.

Sibylle Weis, Ralf Schneggenburger, and Erwin Neher. Properties of a Model of Ca++-

Dependent Vesicle Pool Dynamics and Short Term Synaptic Depression.Biophys.

J., 77(5):2418–2429, 1999. URLhttp://www.biophysj.org/cgi/content/

abstract/77/5/2418 .

V Matveev and X-J Wang. Implications of all-or-none synaptic transmission and short-term

depression beyond vesicle depletion: A computational study.J. Neurosci., 20:1575–1588,

2000b.

Julia Trommershauser. A semi-microscopic model of synaptic transmission and plastic-

ity. Master’s thesis, Georg-August-Universit at zu Gottingen, 2000. URLhttp://www.

theorie.physik.uni-goettingen.de/˜trommer/written.html .

Christian Rosenmund and Charles F. Stevens. Definition of the readily releasable pool of

vesicles at hippocampal synapses.Neuron, 16:1197, 1996.

T. Schikorski and CF. Stevens. Morphological correlates of functionally defined synaptic

vesicle populations.Nat Neurosci., 4(4):391–395, 2001.

D. Zenisek, J. A. Steyer, and W. Almers. Transport, capture and exocytosis of single synap-

tic vesicles at active zones.Nature, 406:849–854, 2000.

L. E. Dobrunz. Release probability is regulated by the size of the readily releasable vesicle

pool at excitatory synapses in hippocampus.Int. J. Dev. Neuroscience, 730:1–12, 2002.

V N Murthy and C F Stevens. Synaptic vesicles retain their identity through the endocytic

cycle. Nature, 392(6675):497–501, 1998.

Jian-Yuan Sun, Xin-Sheng Wu, and Ling-Gang Wu. Single and multiple vesicle fusion

induce different rates of endocytosis at a central synapse.Nature, 417:555–559, 2002.

C. F. Stevens and T. Tsujimoto. Estimates for the pool size of releasable quanta at a single

central synapse and for the time required to refill the pool.PNAS, 92:846–849, 1994.

Henry Markram, Yun Wang, , and Misha Tsodyks. Differential signaling via the same axon

of neocortical pyramidal neurons.PNAS, 95:977–979, 1998a.

Page 215:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Bibliography 201

Lu-Yang Wang and Leonard K. Kaczmarek. High-frequency firing helps replenish the

readily releasable pool of synaptic vesicles.Nature, 394:384–388, July 1998.

Gerald T. Finnerty, Langdon S. E. Roberts, and Barry W. Connors. Sensory experience

modifies the short-term dynamics of neocortical synapses.Nature, 400:367–371, 1999.

Juan A. Varela, Sen Song, Gina G. Turrigiano, and Sacha B. Nelson. Differential Depres-

sion at Excitatory and Inhibitory Synapses in Visual Cortex.J. Neurosci., 19(11):4293–

4304, 1999. URLhttp://www.jneurosci.org/cgi/content/abstract/

19/11/4293 .

Carl C.H. Petersen. Short-Term Dynamics of Synaptic Transmission Within the Excitatory

Neuronal Network of Rat Layer 4 Barrel Cortex.J Neurophysiol, 87(6):2904–2914, 2002.

URL http://jn.physiology.org/cgi/content/abstract/87/6/2904 .

J. Varela, K. Sen, J. Gibson, J Fost, LF Abbott, and SB Nelson. A quantitative description

of short-term plasticity at excitatory synapses in layer 2/3 of rat primary visual cortex.J

Neuroscience, 17:7926–7940, 1997.

Jeremy S. Dittman and Wade G. Regehr. Calcium dependence and recovery kinetics of

presynaptic depression at the climbing fiber to purkinje cell synapse.J. Neurosci., 18:

6147–6162, 1998.

CF Stevens and JF Wesseling. Activity-dependent modulation of the rate at which synaptic

vesicles become available to undergo exocytosis.Neuron, 20:1243–1253, 1998.

E Neher and T Sakaba. Estimating transmitter release rates from postsynaptic current fluc-

tuations.J Neurosci., 21:9638–54, 2001.

Jeremy S. Dittman, Anatol C. Kreitzer, and Wade G. Regehr. Interplay between Facili-

tation, Depression, and Residual Calcium at Three Presynaptic Terminals.J. Neurosci.,

20(4):1374–1385, 2000. URLhttp://www.jneurosci.org/cgi/content/

abstract/20/4/1374 .

L. E. Dobrunz, E. P. Huang, and C. F. Stevens. Very short-term plasticity in hippocampal

synapses.Proc. Natl. Acad. Sci. USA, 94:14843–14847, 1997.

Eric Hanse and Bengt Gustafsson. Release Dependence to a Paired Stimulus at a Synaptic

Release Site with a Small Variable Pool of Immediately Releasable Vesicles.J. Neurosci.,

22(11):4381–4387, 2002. URLhttp://www.jneurosci.org/cgi/content/

abstract/22/11/4381 .

Page 216:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

202 Bibliography

SF Hsu, GJ Augustine, and MB Jackson. Adaptation of ca(2+)-triggered exocytosis in

presynaptic terminals.Neuron, 17(3):501–12, 1996.

Mathew V. Jones and Gary L. Westbrook. The impact of receptor desensitization on fast

synaptic transmission.Trends in Neurosciences, 19(3):96–101, March 1996.

J. del Castillo and B. Katz. Statistical factors involved in neuromuscular facilitation and

depression.J. Physiol., 124:574–85, 1954b.

Anirudh Gupta, Yun Wang, , and Henry Markram. Organizing principles for a diversity of

gabaergic interneurons and synapses in the neocortex.Science, 287:273–278, 2000.

CM Hempel, KH Hartman, X-J Wang, GG Turrigiano, and SB Nelson. Multiple forms of

short-term plasticity at excitatory synapses in rat medial prefrontal cortex.J. Neurophysiol.,

83:3031–3041, 2000.

FR Edwards, SJ Redman, and B Walmsley. Statistical fluctuations in charge transfer at Ia

synapses on spinal motoneurones.J Physiol (Lond), 259(3):665–688, 1976. URLhttp:

//www.jphysiol.org/cgi/content/abstract/259/3/665 .

A Triller and H Korn. Transmission at a central inhibitory synapse. iii. ultrastructure of

physiologically identified and stained terminals.J Neurophysiol, 48:708–736, 1982.

S. J. Redman. Quantal analysis of synaptic potentials in neurons of the central nervous

system.Physiological Reviews, 70:165–198, 1990.

G Tong and CE Jahr. Multivesicular release from excitatory synapses of cultured hippocam-

pal neurons.Neuron, 12:51–59, 1994.

C. Auger, S. Kondo, and A. Marty. Multivesicular release at single f unctional synaptic

sites in cerebellar stellate and basket cells.J Neurosci, 18:4532–4547, 1998.

T.G. Oertner, B.L. Sabatini, E.A. Nimchinsky, and K. Svoboda. Facilitation at single

synapses probed with optical quantal analysis.Nat Neurosci., 5(7):657–664, 2002.

H. Markram, J. Lubke, M. Protscher, A. Roth, and B. Sakmann. Physiology and anatomy

of synaptic connections between thick tufted pyramidal neurones in the developing rat neo-

cortex.J. Physiol., 500:409–440, 1997a.

V Murthy, T Sejnowski, and C F Stevens. Heterogeneous release properties of visualized

individual hippocampal synapses.Neuron, 18:599–612, 1997.

Page 217:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Bibliography 203

W. Senn, M. Schneider, and B. Ruf. Activity-dependent development of axonal and den-

dritic delays or, why synaptic transmission should be unreliable.Neural Computation, 14

(3):583–620, 2002a.

Michael J. O’Donovan and John Rinzel. Synaptic depression: a dynamic regulator of synap-

tic communication with varied functional roles.Trends in Neurosciences, 20:431–433,

1997.

Frances S. Chance, Sacha B. Nelson, and L. F. Abbott. Synaptic depression and the tem-

poral response characteristics of v1 cells.Journal of Neuroscience, 18(12):4785–4799,

1998b.

P. Adorjan and K. Obermayer. Contrast adaptation in simple cells by changing the transmit-

ter release probability. In M. S. Kearns, S. A. Solla, and D. A. Cohn, editors,Advances in

Neural Information Processing Systems NIPS, volume 11, pages 76–82. MIT Press, 1999.

T. Natschlager, W. Maass, and A. Zador. Efficient temporal processing with biologically

realistic dynamic synapses.Network: Computation in Neural Systems, 12:75–87, 2001.

W Senn, M Schneider, and B. Ruf. Activity-dependent development of axonal and dendritic

delays, or, why synaptic transmission should be unreliable.Neural Comput., 14(3):583–

619, 2002b.

H. Markram and M. V. Tsodyks. Redistribution of synaptic efficacy: A mechanism to gen-

erate infinite synaptic input diversity from a homogeneous population of neurons without

changing absolute synaptic efficacies.J. Physiology, 90:229–232, 1996b.

Henry Markram, Dimitri Pikus, Anirudh Gupta, and Misha Tsodyks. Potential for multiple

mechanisms, phenomena and algorithms for synaptic plasticity at single synapses.Neu-

ropharmacology, 37:489–500, 1998b.

D.J. Jr. Hagler and Y. Goda. Properties of synchronous and asynchronous release during

pulse train depression in cultured hippocampal neurons.J. Neurophysiol., 85:2324–2334,

2001.

Eric Hanse and Bengt Gustafsson. Vesicle release probability and pre-primed pool at gluta-

matergic synapses in area ca1 of the rat neonatal hippocampus.The Journal of Physiology,

531(2):481–493, 2001a.

Eric Hanse and Bengt Gustafsson. Factors explaining heterogeneity in short-term synap-

tic dynamics of hippocampal glutamatergic synapses in the neonatal rat.The Journal of

Physiology, 531(1):141–149, 2001b.

Page 218:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

204 Bibliography

M. Raastad, J.F. Storm, and P. Anderson. Putative single quantum and single fiber excita-

tory postsynaptic currents show similar amplitude range and variability in rat hippocampal

slices.Eur. J. Neursci., 4:113–117, 1992.

AM Thomson and J Deuchars. Synaptic interactions in neocortical local circuits: dual

intracellular recordings in vitro. Cereb. Cortex, 7(6):510–522, 1997. URLhttp:

//cercor.oupjournals.org/cgi/content/abstract/7/6/510 .

V Murthy, T Schikorski, C F Stevens, and Zhu Yongling. Inactivity produces increases in

neurotransmitter release and synapse size.Neuron, 32:673–682, 2001.

Ling-Gang Wu and J. Gerard G. Borst. The reduced release probability of releasable vesi-

cles during recovery from short-term synaptic depression.Neuron, 23:821, 1999.

AM Thomson. Activity-dependent properties of synaptic transmission at two classes of

connections made by rat neocortical pyramidal axons in vitro.J Physiol (Lond), 502(1):

131–147, 1997. URLhttp://www.jphysiol.org/cgi/content/abstract/

502/1/131 .

T.C. Sudholf. The synaptic vesicle cycle revised.Neuron, 28:317–320, 2000.

Walter Senn, Henry Markram, and Misha Tsodyks. An algorithm for modifying neuro-

transmitter release probability based on pre- and postsynaptic spike timing.Neural Comp.,

13:35–67, 2001.

W. Softky and C. Koch. The highly irregular firing of cortical cells is incosistent with

temporal integration of random epsp’s.J. Neurosci., 13:334–350, 1993.

C. van Vreeswijk. Using renewal neurons to transmit information.European Biophysics

Journal, 29:245, 2000.

D.R. Cox.Renewal theory. Jonh Wiley, New York, 1962.

G Mandl. Coding for stimulus velocity by temporal patterning of spike discharges in visual

cells of cat superior colliculus.Vis. Res., 33:1451–1475, 1993.

E. Zohary, M. N. Shadlen, and W. T. Newsome. Correlated neuronal discharge rate and its

implication for psychophysical performance.Nature, 370:140–143, 1994.

W Bair, C Koch, W Newsome, and K Britten. Power spectrum analysis of bursting cells of

area mt in the behaving monkey.J. Neurosci., 14:2870–2892, 1994.

Page 219:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Bibliography 205

Andre A. Fenton and Robert U. Muller. Place cell discharge is extremely variable during

individual passes of the rat through the firing field.PNAS, 95(6):3182–3187, 1998. URL

http://www.pnas.org/cgi/content/abstract/95/6/3182 .

E.R. Kandel and W.A. Spencer.J. Neurophysiol., 24:243–259, 1991.

J.B. Ranck.Exp. Neurol., 41:462–531, 1973.

R. J. Baddeley, L. F. Abbott, M. Booth, F. Sengpiel, T. Freeman, E. A. Wakeman, and E. T.

Rolls. Responses of neurons in primary and inferior temporal visual cortices to natural

scenes.Proc. R. Soc. Lond. Ser. B, 264(1389):1775–1783, 1997.

H. Markram, J. Lubke, M. Frotscher, and B. Sakmann. Regulation of synaptic efficacy by

concidence of postsynaptic aps and epsps.Science, 275:213–215, 1997b.

C. van Vreeswijk. Information transmission with renewal neurons.Neurocomputing, 39:

417–422, 2001.

M. S. Goldman, S. B. Nelson, and L. F. Abbott. Decorrelation of spike trains by synaptic

depression.Neurocomputing, 26:147–153, 1999.

M. S. Goldman.Computational Implications of Activity-Dependent Neuronal Processes.

Ph. D. Thesis. Harvard University., 2000.

R. E. Blahut.Principles and Practice of Information Theory. Addison-Wesley, Cambridge,

MA, 1988.

B. Roy Frieden.Physics from Fisher Information: A Unification. Cambridge University

Press, 1999.

F. Gabbiani and C. Koch.Principles of spike train analysis, in ”Methods in neuronal

modeling: From ions to networks”, C. Koch and I. Segev ed.MIT Press, 1998.

T. M. Cover and J. A. Thomas.Elements of information theory. John Wiley, New York,

1991.

J.P. Nadal. Information theoretic approach to neural coding and parameter estimation: a

perspective. In B. Olshausena R. Rao and M. Lewicki, editors,Statistical Theories of the

Brain. MIT Press, 2000.

N. Brunel and J. P. Nadal. Mutual information, fisher information and population coding.

Neural Comp., 10:1731–57, 1998.

Page 220:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

206 Bibliography

A. Treves, S. Panzeri, E. T. Rolls, M. Booth, and E. A. Wakeman. Firing rate distributions

and efficiency of information transmission of inferior temporal cortex neurons to natural

visual stimuli.Neural Computation, 11:601–632, 1999.

J.J. Rissanen. Fisher information and stochastic complexity.IEEE Transactions on Infor-

mation Theory, 42(1):40 –47, January 1996.

WB Levy and RA Baxter. Energy efficient neural codes.Neural Comp., 8(3):531–

543, 1996. URLhttp://neco.mitpress.org/cgi/content/abstract/8/

3/531 .

L B Laughlin, R R de Ruyter van Steveninck, and J C Anderson. The metabolic cost of

neural information.Nature Neuroscience, 1:36 – 41, 1998.

Vijay Balasubramanian, Don Kimber, and II Michael J. Berry. Metabolically Efficient

Information Processing.Neural Comp., 13(4):799–815, 2001. URLhttp://neco.

mitpress.org/cgi/content/abstract/13/4/799 .

Gonzalo Garcia de Polavieja. Errors drive the evolution of biological signalling to costly

codes.J- theor. Biol., 214:657–664, 2002.

William H. Press, Saul A. Teukolsky, William T. Vetterling, and Brian P. Flannery, editors.

”Numerical Recipes in C: The Art of Scientific Computing ”. Cambridge University Press.,

2nd edition, February 2002.

H. Markram. A network of tufted layer 5 pyramidal neurons.Cereb. Cortex, 7(6):523–533,

1997. URL http://cercor.oupjournals.org/cgi/content/abstract/

7/6/523 .

E. Riva Sanseverino, L. F. Agnati, M. G. Maioli, and C. Galletti. Maintained activity of

single neurons in striate and non-striate areas of the cat visual cortex.Brain Res., 54:225,

1973.

M. Abeles.Local Cortical Circuits. An Electrophysiological Study. Springer-Verlag, 1982.

C. R. Legendy and M. Salcman. Bursts and recurrences of bursts in the spike trains of

spontaneously active striate cortex neurons.J. Neurophysiol., 53(926), 1985.

William Feller. An introduction to Probability Theory and Its Applications, volume 1. John

Wiley and Sons, 3 edition, 1950.

Page 221:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Bibliography 207

F. Rieke, D. Warland, R. R. de Ruyter van Steveninck, and W. Bialek.Spikes: exploring

the neural code. MIT Press, Cambridge, MA, 1996.

Peter Dayan and Larry F. Abbot.Theoretical Neuroscience: Computational and Mathe-

matical Modeling of Neural Systems. MIT Press, 1st edition, 2001.

Misha V. Tsodyks, Klaus Pawelzik, and Henry Markram. Neural networks with dynamic

synapses.Neural Comp., 10:821–835, 1998.

Misha Tsodyks, Asher Uziel, and Henry Markram. Synchrony generation in recurrent

networks with frequency-dependent synapses.J. Neurosci., 20:RC50, 2000.

J Reutimann, S Fusi, W Senn, V Yakovlev, and E Zohary. A model of expectation effects

in inferior temporal cortex.Neurocomputing, 2001.

Alex Loebel and Misha Tsodyks. Computation by ensemble synchronization in recurrent

networks with synaptic depression.ournal of Computational Neuroscience, 13(2):111–124,

2002.

Lovorka Pantic, Joaquin J. Torres, Hilbert J. Kappen, and Stan C.A.M. Gielen. Associative

memory with dynamic synapses.Neural Computation, 14(12), 2002.

L. E. Dobrunz and C. F. Stevens. Response os hippocampal synapses to natural stimulation

patterns.Neuron, 22:157–166, 1999.

Mircea Steriade.The Intact and Sliced Brain. MIT Press, 2001.

V. Braitenberg and A. Schuz. Anatomy of the cortex: Statistics and geometry. Spriger-

Verlag, Berlin, 1991.

W. Softky. Simple codes versus efficient codes.Curr. Opin. Neurobiol., 5:239–245, 1995.

Stefano Panzeri, Rasmus S. Petersen, Simon R. Schultz, Michael Lebedev, and Mathew E.

Diamond. The role of spike timing in the coding of stimulus location in rat somatosensory

cortex.Neuron, 29:769–777, March 2001.

F. Rieke, D. Warland, R. de Ruyter van Steveninck, and W. Bialek.Spikes: exploring the

neural code. MIT Press, Cambridge, MA, USA, 1997.

W. Bialek, F. Rieke, R. R. de Ruyter van Steveninck, and D. Warland. Reading a neural

code.Science, 252:1854–1857, 1991.

Page 222:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

208 Bibliography

R. C. deCharms and M. M. Merzenich. Primary cortical representation of sounds by the

coordination of action potentials.Nature, 381:610–613, 1996.

M. Abeles H. Bergman, I. Gat, I. Meilijson, E. Seidemann, N. Tishby, and E. Vaadia. Cor-

tical activity flips among quasi-stationary states.Proc. Nat. Acad. Sci. USA, 92:8616–8620,

1995.

V. N. Murthy and E. E. Fetz. Synchronization of neurons during local field potential os-

cillations in sensorimotor cortex of awake monkeys.J. Neurophysiol., 76(6):3968–3982,

1996.

E. Vaadia, I. Haalman, M. Abeles, H. Bergman, Y. Prut, H. Slovin, and A. Aertsen. Dy-

namics of neuronal interactions in monkey cortex in relation to behavioural events.Nature,

373:515–518, 1995.

Y Prut, E Vaadia, H Bergman, I Haalman, H Slovin, and M Abeles. Spatiotemporal struc-

ture of cortical activity: properties and behavioral relevance.J. Neurophysiol., 79:2857–74,

1998.

C deCharms and A Zador. Neural representations and the cortical code.Ann. Review of

Neurosci., 23:613–847, 2000.

CE. Carr. Processing of temporal information in the brain.Annu. Rev. Neurosci., 16:223–

243, 1993.

JJ. Hopfield. Pattern recognition com-putation using action potential timing for stimulus

representation.Nature, 376:33–36, 1995.

JJ. Hopfield. Transforming neural com-putations and representing time.Proc. Natl. Acad.

Sci. USA, 93:15440–44, 1996.

C. M. Gray and D. A. McCormick. Chattering cells: Superficial pyramidal neurons con-

tributing to the generation of synchronous oscillations in the visual cortex.Science, 274:

109–113, 1996.

W. Singer. Neuronal synchrony: A versatile code for the definition of relations?Neuron,

24:49–65, 1999.

E. Salinas and T. J. Sejnowski. Correlated neuronal activity and the flow of neural informa-

tion. Nature Reviews Neuroscience, 2:539–550, 2001.

Page 223:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

Bibliography 209

O. Bernander, C. Koch, and M. Usher. The effect of synchronized inputs at the single

neuron level.Neural. Comput, 6:622–641, 1994.

V. N. Murthy and E. E. Fetz. Effects of input synchrony on the firing rate of a three-

conductance cortical neuron model.Neural. Comput, 6:1111–11126, 1994.

C.F. Stevens and A. Zador. Input synchrony and the irregular firing of cortical neurons.

Nature Neuroscience, 3:210–217, 1998.

Jianfeng Feng and David Brown. Impact of correlated inputs on the output of the integrate-

and-fire model.Neural Computation, 12:671–692, 2000.

S. M. Bohte, H. Spekreijse, and P. R. Roelfsema. The effects of pair-wise and higher-order

correlations on the ring rate of a postsynaptic neuron.Neural. Comput, 12:153–179, 2000.

E. Salinas and T. J. Sejnowski. Impact of correlated synaptic input on output firing rate and

variability in simple neuronal models.J. Neurosci., 20:6193–6209, 2000.

Michael Rudolph and Alain Destexhe. Correlation detection and resonance in neural sys-

tems with distributed noise sources.Physical Review Letters, 86:3662–3665, 2001.

A. Kuhn, S. Rotter, and A. Aersten. Correlated input spike trains and their effects on the

response of a leaky integrate-and-fire neuron.Neucomputing, 44-46:121–126, 2002.

R. Moreno, J. de la Rocha, A. Renart, and N. Parga. Response of spiking neurons to

correlated inputs.Physical Review Letters, 89(28):288101, 2002.

W. Rall and I. Segev. Functional possibilities for synapses on dendrites and dendritic spines.

In G. Edelman and J.D. Cowan, editors,Synaptic Function, pages 605–636, New York,

1987. John Wiley.

I Segev and W Rall. Excitable dendrites and spines: earlier theoretical insights elucidate

recent direct observations.Trends Neurosci, 21(11):453–60, 1998.

E De Schutter. Using realistic models to study synaptic integration in cerebellar purkinje

cells. Rev Neurosci, 10(3-4):233–45, 1999.

JC Magee. Dendritic integration of excitatory synaptic input.Nat Rev Neurosci, 1(3):

181–90, 2000.

Thomas Natschlager. Efficient computation in Networks oif Spiking neurons- Simulation

and theory. PhD thesis, Technishe Universitat Graz, 1999.

Page 224:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

210 Bibliography

M. Abeles. Corticonics: Neural circuits of the cerebral cortex. Cambridge University

Press, Cambridge, 1991.

AU Larkman, JJ Jack, and KJ Stratford. Quantal analysis of excitatory synapses in rat

hippocampal ca1 in vitro during low-frequency depression.J Physiol, 505:457–71, 1997.

C Stricker, AC Field, and SJ Redman. Statistical analysis of amplitude fluctuations in

EPSCs evoked in rat CA1 pyramidal neurones in vitro.J Physiol (Lond), 490(2):419–

441, 1996. URLhttp://www.jphysiol.org/cgi/content/abstract/490/

2/419 .

B. Katz. Neural transmitter release: from quantal secretion to exocytosis and beyond. the

fenn lecture.J Neurocytol, 25(12):677–86, 1996.

Santiago Ramon y Cajal.Histology of the nervious system of man and vertebrates. Oxford

University Press, 1995.

H. von Gersdor, R. Schneggenburger, S. Weis, and E. Neher. Presynaptic depression at a

calyx synapse: the small contribution of metabotropic glutamate receptors.J. Neurosci.,

17:8137–8146, 1997.

Laurence O. Trussell. Modulation of transmitter release at giant synapses of

the auditory system. Current Opinion in Neurobiology, 12(4):400–404, Au-

gust 2002. URL http://www.sciencedirect.com/science/article/

B6VS3-468D8Y1-1/1/cb2dbf04e3530071ec8a3adf70f2cc02 .

H. Korn, D. S. Faber, Y. Burnod, and A. Triller. Quantal analysis and synaptic efficacy in

the cns.Trends Neurosci, 4:125–130, 1984.

L. M. Ricciardi.Diffusion processes and Related topics in biology. Springer-Verlag, Berlin,

1977.

D. J. Amit and N. Brunel. Model of global spontaneous activity and local structured delay

activity during delay periods in the cerebral cortex.Cerebral Cortex, 7:237–252, 1997.

X-J. Wang. Synaptic reverberation underlying mnemonic persistent activity.Trends in

Neuroscience, 24:455–463, 2001.

A. Rauch, G. La Camera, H.R. Luscher, W. Senn, and S. Fusi. Neorcortical pyramidal cells

respond as integrate-and-fire neurons to in vivo-like input currents. In Press, 2002.

Page 225:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

N. Berretta and R. S. Jones. A comparison of spontaneous EPSCs in layer II and layer IV-V

neurons of the rat entorhinal cortex in vitro.J Neurophysiol, 76(2):1089–1100, 1996. URL

http://jn.physiology.org/cgi/content/abstract/76/2/1089 .

M. V. Tsodyks and T. J. Sejnowski. Rapid state switching in balanced cortical network

models.Network, 6:111, 1995.

C. vanVreeswijk and H. Sompolinsky. Chaos in neural networks with balanced excitatory

and inhibitory activity.Science, 274:1724–1726, 1996.

Alfonso Renart.Models of multi-areal cortical processing. PhD thesis, Universidad Au-

tonoma de Madrid, 2000.

G. Silberberg, M. Bethge, M. Tsodyks H. Markram, and K. Pawelzik.Submitted, 10:815–

819, 2002.

I.S. Gradshtein, I.M. Ryzhik, and A. Jeffrey.Table of integrals, series, and products. Acam-

demic Press, Inc., London, 1980.

Page 226:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

212 Bibliography

Page 227:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

List of Figures

1.1 Electron micrograph of a synapse from the stratum radiatum in CA1 in the

hippocampus of an adult mouse.. . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Picture of the morphology of a synapse. . . . . . . . . . . . . . . . . . . 3

2.1 Release probability as a function of the number of vesicles in the RRP. . . 16

2.2 Schematic picture of the synaptic dynamics seen as a system composed of

two pools of vesicles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.3 Examples of input spike trains with exponential autocorrelations.. . . . . . 21

2.4 Population ofM neurons making single contacts onto a target cell. . . . . 28

2.5 Model of the population distribution of synaptic parameters,D(U,N0). . . 31

2.6 Synaptic transfer function: the synaptic response rateνsr vs the input rateν.

Squared coefficient of variation of the IRI’sCV 2iri vs τv . . . . . . . . . . . 33

2.7 Why do correlated trains saturate for higher input rates than Poisson trains?36

2.8 Connected conditional rateCcr(∆) of the synaptic responses. . . . . . . . 38

2.9 Distribution of synaptic responsesρiri(∆|ν) for several values of the number

of docking sitesN0 and different recovery rates1/τv. . . . . . . . . . . . . 39

2.10 Synaptic response rateνsr vs the input rateν for different values ofN0 . . . 41

3.1 Relative reconstruction errorε versus input rate. . . . . . . . . . . . . . . 60

3.2 InformationI(∆; ν) and information rateR(∆; ν) νr I(∆; ν) versus input

rateν. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

3.3 Input Gamma distributionsf(ν) . . . . . . . . . . . . . . . . . . . . . . . 64

3.4 InformationI(∆; ν) versus the mean input rateν , for three different input

distributionsf(ν). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

3.5 Optimization of the recovery time constantτv regarding the Fisher informa-

tion per responseJsr . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

3.6 Optimization of the recovery time constantτv , and comparison of the Fisher

information per responseJsr, per unit timeJ and the mutual information

I(∆; ν) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

213

Page 228:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

214 List of Figures

3.7 Analysis of the mutual informationI(∆; ν) and entropiesH(∆) and〈H(∆|ν)〉νas a function ofτv (smallτv scale). . . . . . . . . . . . . . . . . . . . . . . 72

3.8 Analysis of the mutual informationI(∆; ν) and entropiesH(∆) and〈H(∆|ν)〉νas a function ofτv (largeτv scale) . . . . . . . . . . . . . . . . . . . . . . 73

3.9 Optimization ofτv regarding the Fisher information per response, for differ-

ent values of the readily releaseable pool sizeN0 . . . . . . . . . . . . . . 75

3.10 Optimization ofT regarding the Fisher information per response, for dif-

ferent values of the readily releaseable pool sizeN0 (whereτv in each case

varies such thatτv = N0 T .) . . . . . . . . . . . . . . . . . . . . . . . . . 76

3.11 Optimization ofτv andT regarding the Mutual informationI(∆; ν) , for

different values of the readily releaseable pool sizeN0 . . . . . . . . . . . 77

3.12 Optimal recovery timeτopt as a function of the input rateν. . . . . . . . . . 78

3.13 Optimal recovery timeτopt of the Fisher information per response, as a func-

tion of the inputCVisi. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

3.14 Optimal recovery timeτopt of the Fisher information per response, as a func-

tion of the time scaleτc of the input correlations. . . . . . . . . . . . . . . 80

3.15 Optimal recovery timeτopt of the Fisher information per response, as a func-

tion of the number of docking sitesN0. . . . . . . . . . . . . . . . . . . . . 81

3.16 Optimal recovery timeτopt of the mutual informationI(∆; ν) , as a function

of the number of docking sitesN0. . . . . . . . . . . . . . . . . . . . . . . 82

3.17 Optimization of the recovery time constantτv regarding the mutual informa-

tion per unit energyI(∆; ν) . . . . . . . . . . . . . . . . . . . . . . . . . 83

3.18 Optimization of the recovery time constantτv regarding the Fisher informa-

tion per unit energyI(∆; ν) . . . . . . . . . . . . . . . . . . . . . . . . . 84

3.19 Analysis of the transfer functionνr(ν) for two values of the release proba-

bility U . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

3.20 Optimization of the release probability ,U , regarding the Fisher information

per second, when the output code is the number of responses.. . . . . . . . 88

3.21 3-D plot of the optimal release probability ,Uopt , as a function ofν andτv,

when the input is Poisson.. . . . . . . . . . . . . . . . . . . . . . . . . . 89

3.22 Optimal release probability ,Uopt , as a function ofν, when the input is

correlated.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

3.23 Optimization of the release probability regarding the mutual information

I(∆; ν) for different values of the input correlations magnitudeCV and the

recovery time constantτv . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

3.24 Optimization of the population distributionD(U,N0, τv) . . . . . . . . . . 94

Page 229:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

List of Figures 215

4.1 Model of a single synaptic contact.. . . . . . . . . . . . . . . . . . . . . . 110

4.2 Diagram of the temporal evolution of a system composed of a single synaptic

contact. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .111

4.3 Illustration of the different ways two neurons may be connected. . . . . . 113

4.4 Schematic picture of the calyx of Held synapse. . . . . . . . . . . . . . . 114

4.5 Schematic description of a model connection with two synaptic contacts

(M = 2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .116

4.6 Schematic description of two synaptic contacts belonging to different pre-

synaptic neurons whose activity is correlated.. . . . . . . . . . . . . . . . 118

5.1 Two examples showing the simulation of the membrane potential evolution

of a LIF neuron in a sub-threshold balanced regime (top) and in a supra-

threshold situation (bottom).. . . . . . . . . . . . . . . . . . . . . . . . . 140

5.2 Current varianceσ2w (per contact) as a function of the number of contactsM

and the input rateν whenCM is kept fixed. . . . . . . . . . . . . . . . . . 142

5.3 Correlation magnitudeα as a function of the number of contactsM and the

input rateν. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .143

5.4 Current parameters as a function ofν for several values of(C, M) where

CM is invariant. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

5.5 Positionνmax of the variance maximum and the ratioσ2w(νmax)σ2

limas a function

of the number of contactsM and the input rateν, when the productCM is

held constant.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .146

5.6 Current parameters as a function ofν for several values of the correlationρ

where all theC = 3750 neurons make a mono-synaptic connectionM = 1. 148

5.7 Current varianceσ2w (per contact) as a function of the number of contactsM

and the input rateν whenMJ = J ′ is kept fixed. . . . . . . . . . . . . . . 149

5.8 Current parameters as a function of the input rateν for several values of

(M, J) whereMJ = J ′ is constant. . . . . . . . . . . . . . . . . . . . . . 151

5.9 Numerical results and theoretical prediction of thenon-monotonicresponse

of a LIF neuron for input current examples with differentMs, whileCM is

held constant.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .153

5.10 Evolution of the afferent spikes, synaptic releases, total afferent current and

membrane potential (I).. . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

5.11 Evolution of the afferent spikes, synaptic releases, total afferent current and

membrane potential (II). . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

5.12 Evolution of the afferent spikes, synaptic releases, total afferent current and

membrane potential (III).. . . . . . . . . . . . . . . . . . . . . . . . . . . 159

Page 230:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

216 List of Figures

5.13 Evolution of the afferent spikes, synaptic releases, total afferent current and

membrane potential (IV).. . . . . . . . . . . . . . . . . . . . . . . . . . . 160

5.14 Numerical results and theoretical prediction of thenon-monotonicresponse

of a LIF neuron for examples of input current with differentMs, whileMJ

is held constant.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .162

B.1 Diagram showing the nomenclature and logic of the terms in whichρiri(t) is

expanded . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .174

E.1 Diagram of the temporal evolution of a system composed of two synaptic

contacts belonging to the same pre-synaptic neuron.. . . . . . . . . . . . . 190

Page 231:  · Contents Table of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv Acknowledgments

List of Tables

2.1 Parameters and functions of the model of a pre-synaptic terminal. . . . . . 42

2.2 Parameters and functions of the population of synapses distribution. . . . . 43

2.3 Parameters and functions used to model the input spike statistics.. . . . . . 44

2.4 Parameters and functions used to model the synaptic response statistics. . 45

217