Andrew M. Oster- Mathematical Models of Cortical Development

  • Upload
    neerfam

  • View
    221

  • Download
    0

Embed Size (px)

Citation preview

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    1/136

    MATHEMATICAL MODELS OF

    CORTICAL DEVELOPMENT

    by

    Andrew M. Oster

    A dissertation submitted to the faculty ofThe University of Utah

    in partial fulfillment of the requirements for the degree of

    Doctor of Philosophy

    Department of Mathematics

    The University of Utah

    December 2006

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    2/136

    Copyright c Andrew M. Oster 2006

    All Rights Reserved

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    3/136

    THE UNIVERSITY OF UTAH GRADUATE SCHOOL

    SUPERVISORY COMMITTEE APPROVAL

    of a dissertation submitted by

    Andrew M. Oster

    This dissertation has been read by each member of the following supervisory committeeand by majority vote has been found to be satisfactory.

    Chair: Paul C. Bressloff

    Alessandra Angelucci

    Aaron L. Fogelson

    James P. Keener

    Richard M. Normann

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    4/136

    THE UNIVERSITY OF UTAH GRADUATE SCHOOL

    FINAL READING APPROVAL

    To the Graduate Council of the University of Utah:

    I have read the dissertation of Andrew M. Oster in its final formand have found that (1) its format, citations, and bibliographic style are consistent andacceptable; (2) its illustrative materials including figures, tables, and charts are in place;

    and (3) the final manuscript is satisfactory to the Supervisory Committee and is readyfor submission to The Graduate School.

    Date Paul C. BressloffChair, Supervisory Committee

    Approved for the Major Department

    Aaron J. BertramChair/Dean

    Approved for the Graduate Council

    David S. ChapmanDean of The Graduate School

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    5/136

    ABSTRACT

    We extend classical activity-based developmental models of ocular dominance column

    (ODC) formation in primary visual cortex (V1) to include cortical growth and cortexs

    laminar structure. We show that with cortical growth the OD pattern exhibits a sequence

    of pattern forming instabilities as the size of the cortex increases. Each instability results

    in the insertion of an additional OD column such that over the course of development, the

    mean width of an ODC is approximately preserved, consistent with recent experimental

    observations of postnatal growth in cat. The other biologically motivated extension we

    make is to consider a multilayer representation of V1 with thalamic and vertical connec-

    tions taken to be modifiable by activity. By including the layer-specific thalamic input to

    V1 along with the interlaminar projections to a correlationbased Hebbian learning rule,

    our model allows for the joint development of OD columns and cytochrome oxidase (CO)

    blobs in primate V1. The developed OD map in layer 4C is inherited by layer 2/3 via the

    vertical projections. Competition between these projections and the direct thalamic input

    to layer 2/3 then results in the formation of CO blobs superimposed upon the OD map.

    The spacing of the OD columns is determined by the spatial profile of the intralaminar

    connections within layer 4, while the spacing of CO blobs depends both on the width of

    the ODCs inherited and the spatial distribution of intralaminar connections within the

    superficial layer. These mathematical models of cortical development demonstrate that

    simple models with key aspects of the biological system give us insight into the underlying

    mechanisms for observed patterns in V1.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    6/136

    To Dennis Stanton, my high school math teacher,

    and to Thomas ONeil, my undergraduate advisor.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    7/136

    CONTENTS

    ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

    NOTATION AND SYMBOLS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

    ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

    CHAPTERS

    1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    2. A SURVEY OF THE STRUCTURE AND DEVELOPMENT OFVISUAL CORTEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

    2.1 Structure of the primary visual cortex . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52.1.1 Retinotopic map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52.1.2 Receptive fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82.1.3 Feature maps. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92.1.4 Cortical circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

    2.2 Development of ocular dominance columns . . . . . . . . . . . . . . . . . . . . . . . . . 162.2.1 Refinement of columns during the critical period . . . . . . . . . . . . . . . . 172.2.2 Early development of OD columns . . . . . . . . . . . . . . . . . . . . . . . . . . . 182.2.3 Development of CO blobs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

    2.3 The Hebbian synapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.3.1 Excitatory synapses and NMDA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.3.2 Neurotrophins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

    3. MODELING CORTICALDEVELOPMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

    3.1 LGN afferents terminating at a single neuron . . . . . . . . . . . . . . . . . . . . . . . 273.1.1 Swindale model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283.1.2 Subtractive normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293.1.3 Competition for neurotrophins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

    3.2 ODC development on a cortical sheet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323.2.1 Subtractive normalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    3.2.2 Inclusion of component . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373.2.3 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

    4. A DEVELOPMENTAL MODELOF OCULAR DOMINANCECOLUMN FORMATIONON A GROWINGCORTEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    8/136

    4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424.2 Developmental model on a growing domain . . . . . . . . . . . . . . . . . . . . . . . . 454.3 Linear stability analysis on a fixed domain . . . . . . . . . . . . . . . . . . . . . . . . . 51

    4.3.1 Stationary front . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

    4.3.2 Single stationary bump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 544.3.3 Periodic pattern . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 584.4 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 604.5 Correlationbased Hebbian model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 654.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

    5. LAMINAR NETWORK MODEL FOR THEJOINT DEVELOPMENT OF OCULARDOMINANCE COLUMNS ANDCYTOCHROME OXIDASEBLOBS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

    5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

    5.2 Developmental model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 725.2.1 Laminar architecture of V1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 725.2.2 Reduced twolayer cortical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 755.2.3 Mathematical formulation of model . . . . . . . . . . . . . . . . . . . . . . . . . . 77

    5.3 Development of OD columnsand CO blobs in layer 2/3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    5.3.1 O(1) analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 805.3.2 O() analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

    5.4 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 865.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

    6. A THEORY FOR THE ALIGNMENT OF CORTICAL FEATURE

    MAPS DURING DEVELOPMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 936.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 936.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

    6.2.1 Developmental model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 946.2.2 Linear stability analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

    6.2.2.1 Homogeneous case ( = 0) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 986.2.2.2 Inhomogeneous case ( > 0) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

    6.2.3 Commensurability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1026.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

    6.3.1 Pinning in a onedimensional network . . . . . . . . . . . . . . . . . . . . . . . . 1046.3.2 Pinning in a twodimensional network . . . . . . . . . . . . . . . . . . . . . . . . 109

    6.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

    REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

    vii

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    9/136

    NOTATION AND SYMBOLS

    Uniformly Used

    r, (x) cortical position (in 1d)w feedforward weights to layer 4W upperbound of synaptic weights into layer 4V postsynaptic activityI feedforward inputJ recurrent weight function strength of recurrent weight function

    conversion factor for membrane potential to output firing rateL operator corresponding to the convolution with 1

    J (r)G intracortical interaction function the domain (i.e., cortex)C correlation matrixS spatial correlations of inputH(r) effective intracortical interaction function, H(r) = G(r)S(r)p,q wavenumbers used as an eigenvalue and sometimes as the growth factor for spatial modes general subtractive normalization terma vector of 1s

    f Fourier transform of f space constant of the weight functionA amplitude of weight function coefficient of inhibition noise term measure of L/R control over all V1c perturbation of solution used in linearization

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    10/136

    Cortical Growth cortical growth termHt interaction function that varies with timeH effective interaction function on original domain, depends upon growth growth rate, taken to be small ratio of initial to adult length of cortexu flow of cortexd 2 standard width of an ODC 2 growth functionL 1d length of cortex function used for stability test an auxiliary function effective space constant of the weight function

    after mapping to original domain

    Laminar model for V1 development

    k koniocellular weightsm 1 weight of interlaminar projectionK upperbound of synaptic weights for koniocellular inputM upperbound of weight for vertical interlaminar projectionB correlation matrix between K and M inputC correlation matrix for koniocellular input subtr. normal term in layer 2/3 dynamicsL,R degree of L/R eye control at a cortical point with [0, 1]L,R convolution over layer 4 of the weight function of the L/R eye density

    Ki mexican hat, O(), component of theith interaction function (Ki(r) = Ji(r)S(r))

    CO blobs intrinsically defined via molecular markers

    strength of binocular stabilizing term in modified SwindaleWb L/R density during initial binocular stated 2 distance between blobsv patchy distribution of the CO marker strength of modulation due to the patchy distribution

    W baseline level for the maximal density of feed forward afferentsQ reciprocal lattice vector 2 disorder of CO marker lattice degree of pinning patch size of molecular marker

    1,2 sometimes used as a summation index, used differently in other chapters

    ix

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    11/136

    ACKNOWLEDGEMENTS

    I would like to thank Paul Bressloff for his excellent direction that helped bring this

    work to completion. Additionally, the many conversations with Alessandra Angelucci and

    Jenny Lund were immensely helpful. I would also like to thank James Keener (Jim) for

    the exciting classes he taught during my earlier years at the University of Utah. Lastly,

    I would like to thank the University of Utah and the National Science Foundation (RTG

    0354259) for their generous support during my studies. Cheers.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    12/136

    CHAPTER 1

    INTRODUCTION

    When studying the large-scale functional and anatomical structure of cortex, two

    distinct questions naturally arise: (I) how did the structure develop? and (II) what forms

    of spontaneous and stimulus-driven neural dynamics are generated by such a cortical

    structure? It turns out that in both cases the Turing mechanism for spontaneous pattern

    formation plays an important role. Turing originally considered the problem of how

    animal coat patterns develop, suggesting that chemical markers in the skin comprise a

    system of diffusion-coupled chemical reactions among substances called morphogens [164].

    He showed that in a two-component reaction-diffusion system, a state of uniform chemical

    concentration can undergo a diffusion-driven instability leading to the formation of a

    spatially inhomogeneous state. Ever since the pioneering work of Turing on morphogenesis

    [164], there has been a great deal of interest in spontaneous pattern formation in physical

    and biological systems [46, 130]. In the neural context, Wilson and Cowan [179] proposed

    a nonlocal version of Turings diffusiondriven mechanism, based on competition between

    short-range excitation and longer-range inhibition. Here interactions are mediated, not

    by molecular diffusion, but by long-range axonal connections. Since then this neural

    version of the Turing instability has been applied to a number of problems concerning the

    dynamics and development of cortex. Examples in visual neuroscience include the Marr

    Poggio model of stereopsis [120], developmental models of retinotopic, ocular dominance

    and isoorientation maps [177, 155, 156, 158, 124], and cortical models of geometric visual

    hallucinations [53, 24].

    In this thesis we focus on the role of pattern formation in models of the activitydriven

    development of primary visual cortex. We begin by briefly outlining how this fits in with

    other stages of neural development. During embryogenesis, a thin sheet from the ectoderm

    on the dorsal surface becomes specified as neural tissue. From this initial substrate, called

    the neural plate, the nervous system develops. Through a combination of cell movement,

    changes in cell shape, and differential cell adhesion, the lateral edges of the neural plate

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    13/136

    2

    elevate and later fuse to form a hollow tube, called the neural tube. During the neural

    tubes formation, a portion from the edges of the neural plate becomes pinched off to form

    the neural crest. Attractive and repulsive cues guide these cells away from the forming

    neural tube and act as the seeds to the peripheral nervous system. Other moleculesdiffusing from the middle layer of the embryo create molecular gradients that migrating

    cells can use to navigate and modify gene expression. The caudal portion of the tube

    is to become the spinal cord, whereas the rostral portion goes on to divide to form two

    telencephalic vesicles and a diencephalon vesicle, upon which two evaginations develop

    that form the optic vesicles that later become cups along the inner walls of which the

    retinas form.

    Retinal cells then form axons that travel through the optic tract guided by attractive

    and repulsive molecular concentration gradients. In general, molecular gradients play akey role in the navigation of the axonal projections to their destination cortical areas.

    The projections from the retinas travel to a portion of the thalamus and are then

    forwarded on to the primary visual cortex (V1). The mapping from the retina to the

    cortex, called the retinotopic map, preserves the the topography of the visual field. In

    1947, Roger Sperry [153] demonstrated in the frog that molecular cues in the form of

    chemical gradients play a role in the development of the retinotopic map, i.e., axon-target

    recognition relied on chemical matching. However, the initial termination of the axonal

    projections to V1 is wide and disperse yet has some general order owing to arrangementvia molecular gradients within V1. The activity within the system is involved in a process

    that selectively prunes and refines connections so as to create effective connections. The

    mapping of the visual field to the retinas onto the visual cortex is a prototypical example

    of the ubiquitous cortical maps in cortex. For example, in the somatosensory system in

    rodent, the twodimensional array of whiskers on the snout projects through midbrain

    nuclei to form a somatotopic map in the somatosensory cortex [140], called the barrel

    field because of its honeycomb-like structure that resembles an array of barrels. Similar

    to the retinotopic map, the somatotopic map preserves the spatial arrangement of the

    whisker array and encodes for the frequency and force of stimulation.

    Not only are neurons in mature V1 selective to stimuli from specific regions in the

    visual fields, they also respond preferentially to a variety of features associated with

    visual stimuli, including orientation, ocular dominance, spatial frequency, and direction

    selectivity. As one progresses tangentially across V1, the response properties vary in a

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    14/136

    3

    nearly continuous fashion. Each feature has a corresponding selectivity map, that is,

    across the cortex the selectivity or preference for a particular feature varies. Since there

    are multiple feature maps across V1, these feature maps overlay one another in an intricate

    manner, discussed further in Chapter 2. Furthermore, neurons through the cortical layershave similar response properties, so that visual cortex is thought to be organized in

    a columnar fashion. However, this conjectured homogeneity is an oversimplification,

    as highlighted in Chapter 5. Hubel and Wiesel [82, 84] conjectured that the feature

    preferences of cortical neurons are generated by the convergence of thalamic afferents on to

    input layer 4, and then passed on to other layers through vertical interlaminar projections.

    Along these lines, the formation of feature preference maps could be understood in terms

    of the development of feedforward connections from thalamus to layer 4 as many models

    for cortical development assume (see the reviews of [158, 165]).The most studied cortical map is that of ocular dominance (OD), characterized by

    a significant influence of, say, left-eye over right-eye activation determining a neurons

    response properties, that is, the feedforward connections originating from left-eye thala-

    mic regions are stronger than the connections from the right-eye thalamic regions. The

    segregation of ocular streams may play a crucial role in stereopsis, i.e., depth perception.

    OD maps appear not to be predetermined and depend critically upon the driving activity

    during early development. As such, the OD maps for different species take on a variety

    of patterns from a stripe-like pattern in macaque monkey and humans where the leftand right eye drives are believed to be approximately balanced to a patchy pattern in

    cats where the contralateral eye more powerfully drives V1 during early development.

    OD maps from animals that have been monocularly deprived at young age, either via

    strabismus, suturing, or enucleation, are qualitatively and quantitatively different from

    OD maps from normally raised animals. This dependence on activity in the system

    motivates a class of models for the development of cortical maps, i.e., activitydriven

    developmental models (see the reviews of [158, 165]). In this work, we extend the

    existing activitydriven developmental models of Swindale and Miller [155, 124] to include

    biological substrates that could affect development, specifically, cortical growth, the

    laminar structure of cortex, and molecular markers.

    We begin this work by giving a review of the visual system, principally for macaque

    monkey and cat, and its plasticity in Chapter 2. In the following Chapter, we review

    prevalent models for OD formation and highlight a mathematical sleightofhand used

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    15/136

    4

    in all of these models that has important consequences for the underlying pattern forming

    process. Although considerable cortical growth occurs after the initial formation of the

    OD pattern, previous models have neglected the possible effects this growth has on the

    OD map. Experimentalists [138] have shown that the OD pattern in cat V1 does notmerely expand with growth, since the width of an OD stripe in adult cortex is similar in

    size to that of a kitten. Motivated by work by Crampin et al. [44] on the reaction-diffusion

    of pigment cells, postulated to determine stripe formation during the growth of marine

    angelfish, in Chapter 4 we consider the effects of cortical growth on OD map formation.

    The resulting OD map undergoes a series of pattern forming instabilities as the cortex

    grows, leading to the insertion of additional OD stripes.

    In Chapter 5, we address the ubiquitous assumption of homogeneity within cortical

    columns. For instance, staining for cytochrome oxidase (CO) in the superficial layers2/3 of macaque monkey V1 results in a periodic distribution of blob-like formations.

    These stains have been given a variety of names: CO blobs, CO patches, and CO puffs.

    Fascinatingly, there is a strong association (particularly in macaque) with the array of CO

    blobs and the centers of the OD stripes, suggesting a strong relationship between these

    structures. We consider a multilayer, activitydependent model for the joint development

    of ocular dominance columns and cytochrome oxidase blobs in primate V1. We include

    both thalamic and interlaminar vertical projections and take both to be modifiable by

    activity. By including this more detailed biologically motivated framework, we are able toobtain CO blob distributions, defined by direct thalamic input to the superficial layers,

    consistent with experimental data that are aligned with the underlying OD pattern.

    Finally, in Chapter 6, we take a single hybrid layer representation of V1 and consider

    an alternate approach that assumes the existence of a periodic distribution (lattice) of

    molecular markers. The molecular markers will go on to intrinsically define the CO blob

    positions and weakly affects the developing OD pattern so that it aligns to the same

    underlying lattice. The virtue of the laminar approach in Chapter 5, instead of the

    aligning of the CO blobs and OD pattern to a periodic distribution of molecular marker

    as in Chapter 6, is that the former more closely mirrors detectable biological substrates.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    16/136

    CHAPTER 2

    A SURVEY OF THE STRUCTURE AND

    DEVELOPMENT OF VISUAL CORTEX

    2.1 Structure of the primary visual cortex

    The primary visual cortex (V1) is the first cortical area to receive visual information

    from the retina (see Figure 2.1(a)). The output from the retina is conveyed by ganglion

    cells whose axons form the optic nerve. The optic nerve conducts the output spike trains

    of the retinal ganglion cells to the lateral geniculate nucleus (LGN) of the thalamus, which

    acts as a relay station between retina and primary visual cortex. Prior to arriving at the

    LGN, some ganglion cell axons cross the midline at the optic chiasm. This allows the

    left and right sides of the visual fields from both eyes to be represented on the right and

    left sides of the brain, respectively. Note that signals from the left and right eyes are

    segregated in the LGN and in input layers of V1, as seen in Figure 2.1(b). This means

    that the corresponding LGN and cortical neurons are monocular, in the sense that they

    only respond to stimuli presented to one of the eyes but not the other (ocular dominance).

    2.1.1 Retinotopic map

    One of the striking features of the visual system is that the visual world is mapped

    onto the cortical surface in a topographic manner. This means that neighboring points

    in a visual image evoke activity in neighboring regions of visual cortex. Moreover, one

    finds that the central region of the visual field has a larger representation in V1 than

    the p eriphery, partly due to a nonuniform distribution of retinal ganglion cells. The

    retinotopic map is defined as the coordinate transformation from points in the visualworld to locations on the cortical surface. In order to describe this map, we first need

    to specify visual and cortical coordinate systems. Since objects located a fixed distance

    from one eye lie on a sphere, we can introduce spherical coordinates with the north

    pole of the sphere located at the fixation point, the image point that focuses onto the

    fovea or center of the retina. In this system of coordinates, the latitude angle is called

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    17/136

    6

    R

    R

    R

    L

    LL lateral

    geniculatenucleusM

    P {{

    RL

    KK

    KK

    KK

    V1

    (b)

    V1

    retina

    opticchiasm

    LGN

    retina

    LGN

    (a)

    Figure 2.1: Schematic of the wiring of the visual system. (a) Visual pathwaysfrom the retina through the lateral geniculate nucleus (LGN) of the thalamusto the primary visual cortex (V1). (b) A cartoon of the layers of LGN (inmacaque) and the LGN afferents to V1 terminating in a segregated fashion.We have labeled the magnocellular and parvocellular pathways and picturethe intercalated koniocellular pathway by magenta colored strips, discussedin section 5.2.1. The M, P, and K pathways in the macaque correspond toanalogous X, Y and W pathways in the cat visual system.

    the eccentricity and the longitudinal angle measured from the horizontal meridian is

    called the azimuth . In most experiments the image is on a flat screen such that, if

    we ignore the curvature of the sphere, the pair (, ) approximately coincides with polar

    coordinates on the screen. One can also represent p oints on the screen using Cartesiancoordinates (X, Y). In primary visual cortex the visual world is split in half with the

    region 90o 90o represented on the left side of the brain, and the reflection of thisregion represented on the right side brain. Note that the eccentricity and Cartesian

    coordinates (X, Y) are all based on measuring distance on the screen. However, it is

    customary to divide these distances by the distance from the eye to the screen so that

    they are specified in terms of angles. The structure of the retinotopic map in monkey is

    shown in Figure 2.2, which was produced by imaging a radioactive tracer that was taken

    up by active neurons while the monkey viewed a visual image consisting of concentriccircles and radial lines. The fovea is represented by the point F on the left hand side

    of the cortex, and eccentricity increases to the right. Note that concentric circles are

    approximately mapped to vertical lines and radial lines to horizontal lines.

    Motivated by Figure 2.2, we assume that eccentricity is mapped onto the horizontal

    coordinate x of the cortical sheet, and is mapped onto its y coordinate An approximate

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    18/136

    7

    S

    F H

    3

    2

    1

    1 cm

    H

    2

    S

    1

    F

    1

    3

    0 00

    Figure 2.2: A deoxyglucose autoradiograph from the left side primary visualcortex of a macaque monkey brain. The radioactive trace displays the activitypattern evoked by the image shown to the left. Adapted from [160]

    equation for the retinotopic map can then be obtained through specification of a quantity

    known as the cortical magnification factor M(). This determines the distance across a

    flattened sheet of cortex separating the activity evoked by two nearby image points. First

    suppose that the two image points in question have eccentricities and + but thesame azimuthal coordinate . The corresponding distance on cortex is x = M() so

    that

    dx

    d= M() (2.1)

    Using experimental data such as shown in Figure 2.2 suggests that

    M() =

    0 + (2.2)

    with 12 mm and 0 1o in macaque monkey. It follows that

    x = ln(1 + /0) (2.3)

    assuming x = 0 when = 0. Similarly, for two image points with the same eccentricity

    but different azimuthal coordinates we find that

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    19/136

    8

    dy

    d=

    180oM() (2.4)

    and hence

    y = a(0 + )1800

    (2.5)

    The minus sign appears because the visual field is inverted in cortex. For eccentricities

    greater than 1o,

    x ln(/0), y 180o

    (2.6)

    and the retinotopic map can be approximated by a complex logarithm [146]. That

    is, introducing the complex representations Z = (/0)ei/180o

    and z = x + iy thenz = log Z.

    2.1.2 Receptive fields

    Neurons in the retina, LGN and primary visual cortex respond to light stimuli in

    restricted regions of the visual field called their classical receptive fields (RFs). Patterns

    of illumination outside the RF of a given neuron cannot generate a response directly,

    although they can significantly modulate responses to stimuli within the RF via long

    range cortical interactions. The RF is divided into distinct ON and OFF regions. In an

    ON (OFF) region illumination that is higher (lower) than the background light intensity

    enhances firing. The spatial arrangement of these regions determines the selectivity of

    the neuron to different stimuli.

    The receptive fields of LGN cells are circular and are described as ON-center and

    OFF-center. ON-center means that a cell responds to a light stimulus centered in the

    receptive field with a darker region surrounding the contrasting disc of light in the center.

    The higher the contrast between the center of the receptive field and its boundary, the

    greater the response. An OFF-center receptive field has the roles of the light and dark

    regions reversed; see Figure 2.3(a). Because of the circular shape of LGN receptive fields,

    LGN cells are thus invariant to orientation. For example, consider an LGN cell with an

    OFF-center receptive field. If a dark bar were placed across the center of the RF, the

    neurons firing rate would increase. Keeping the bar centered in the RF yet changing its

    orientation has no effect on the firing rate.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    20/136

    9

    V1 neurons, on the other hand, are generally orientation selective, i.e., a neuron will

    only respond strongly to a bar in its receptive field if it is at a certain preferred orientation.

    Orientation preference found in primary visual cortex is due to the fact that, unlike the

    circular receptive fields for LGN cells, the receptive fields of V1 neurons are elongatedwith their principal axis at some orientation; see Figure 2.3(b). Thus, a neuron in V1 will

    respond strongly to a bar at an orientation that matches the orientation of its receptive

    field, whereas a bar at an oblique angle to the preferred orientation will elicit a reduced

    response. Hubel and Wiesel [82, 84] made the observation that typical V1 receptive fields

    could be constructed from the convergence of multiple feed forward inputs from LGN

    cells with overlapping circular receptive fields onto a single V1 cell; see Figure 2.3(c). A

    sinusoidal grating is a commonly used stimulus where both the orientation and spatial

    frequency (spacing of the grating) can be varied. V1 receptive fields that are also selectivefor spatial frequency can be constructed from multiple LGN circular receptive fields.

    2.1.3 Feature maps

    In recent years much information has accumulated about the spatial distribution of

    orientation selective cells in V1 [61]. Figure 2.4 shows a typical arrangement of such

    cells, obtained via microelectrodes implanted in cat V1. The first panel shows how

    orientation preferences rotate smoothly over the surface of V1, so that approximately

    every 300m the same preference reappears, i.e., the distribution is periodic in the

    orientation preference angle. The second panel shows the receptive fields of the cells,

    -

    --

    --

    ---

    -

    --

    --

    + +

    ++

    Presynaptic cellsfrom the LGN

    V1 simple cell

    (a) (b) (c)

    On-center Off-center

    V1 RF

    Figure 2.3: Typical receptive fields for LGN and V1 cells, (a) and (b)respectively. Note that LGN RFs are circularly symmetric, whereas V1 RFsare oriented and ovular. In (C), a sample construction of an elongated V1receptive field from LGN inputs [82].

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    21/136

    10

    and how they change with V1 location. The third panel shows more clearly the rotation

    of such fields with translation across V1. One also finds that cells with similar feature

    preferences tend to arrange themselves in vertical columns so that to a first approximation

    the layered structure of cortex can be ignored. For example, electrode track 1 in Figure2.4 is a vertical penetration of cortex that passes through a single column of cells with the

    same orientation preference and ocular dominance. Thesituation regarding orientation

    columns in macaque V1 is more complicated [115, 114]. For example, input layer 4 has

    an additional sublaminar structure that reflects amongst other things the division of the

    LGN afferents into parvocellular (P) and magnocellular (M) pathways (see Figures 2.1(b)

    and 2.5). One finds that many cells in layer 4C are not orientation selective: orientation

    preference emerges in a graded fashion as one moves to mid and upper layer 4 C. The M

    1 2 3

    y

    x

    2

    y

    x

    1 3

    Figure 2.4: Orientation tuned cells in layers of cat V1 which is shown incross-section. Note the constancy of orientation preference at each corticallocation [electrode tracks 1 and 3], and the rotation of orientation preferenceas cortical location changes [electrode track 2]. Redrawn from [61].

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    22/136

    11

    PM

    M2

    P

    M1

    Figure 2.5: The parvocellular (P) and magnocellular (M) pathways ofmacaque V1. The P pathway innervates layer 4C, consisting of cells withsmaller receptive fields and slower responses. These send axons to layer 4A,which feeds into the CO blob regions of superficial layers 2/3. The M pathwayinnervates layer 4C, consisting of cells that have larger receptive fields andfaster responses. Upper 4C cells connect to layer 4B which itself connects tothe blob regions of layers 2/3. Midlayer 4C has a mixture of M and P neurons

    and connects to the interblob regions of layers 2/3.

    pathway is thought to contribute primarily to motion perception, whereas the P pathway

    contributes primarily to form and color perception. However, there is some mixing of

    the two pathways. Inhomogeneities in the laminar structure of cortex will be considered

    further in Chapter 5.

    A more complete picture of the twodimensional distribution of both orientation

    preference and ocular dominance in layers 2/3 has been obtained using optical imaging

    techniques [14, 18, 12]. The basic experimental procedure involves shining light directly

    onto the surface of the cortex. The degree of light absorption within each patch of cortex

    depends on the local level of activity. Thus, when an oriented image is presented across

    a large part of the visual field, the regions of cortex that are particularly sensitive to

    that stimulus will be differentiated. In the case of macaque V1, the topography revealed

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    23/136

    12

    by these methods has a number of characteristic features [133], which are illustrated

    in Figure 2.6: (i) Orientation preference changes continuously as a function of cortical

    location, except at singularities or pinwheels. (ii) There exist linear zones, approximately

    750 750 m2

    in area, bounded by pinwheels, within which isoorientation regions formparallel slabs. (iii) Linear zones tend to cross the borders of ocular dominance stripes at

    right angles; pinwheels tend to align with the centers of ocular dominance stripes.

    Another important example of nonuniformity through the layers is the occurrence

    of cytochrome oxidase (CO) blobs in superficial layers of primate and cat V1 [79, 76,

    128]. These are regions of higher metabolic activity that receive a distinct class of direct

    thalamic inputs [36, 71], with the density of CO staining being highly correlated with

    the density of the thalamic afferents [107]. The spatial distribution of CO blobs within

    cortex is also correlated with a number of stimulus feature preferences. For example, inold world monkeys such as macaques the blobs are found at evenly spaced intervals along

    the center of OD columns [76], and neurons within the blobs tend to be less binocular

    and less orientation selective [108]. The latter is probably due to their association with

    orientation pinwheels; see Figure 2.6. The blobs are also linked with low spatial frequency

    domains [159]. The arrangement of CO blobs is reflected anatomically by the distribution

    of intrinsic horizontal connections (see section 2.3), which tend to link blobs-toblobs

    and interblobstointerblobs [109, 183, 182], and by extrinsic corticocortical connections

    Figure 2.6: Map of iso-orientation contours (yellow lines), ocular dominanceboundaries (dark gray lines) and CO blob regions (shaded areas) of macaqueV1. Adapted from [12].

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    24/136

    13

    linking blobs to specific compartments in V2 and other extrastriate areas [109, 150].

    Taken together these observations suggest that the CO blobs are sites of functionally and

    anatomically distinct channels of visual processing. Some of the functional properties

    of CO blobs, such as association with low spatial frequencies, persist in species otherthan old world monkeys. However, the spatial relationship b etween CO blobs and OD

    columns is often less clear. For example, OD columns are less regular in cats compared

    to macaques, and the periodicity of the CO blobs (around 1mm) appears to be too large

    to provide sufficient coverage of CO blobs across all OD columns. Nevertheless, Murphy

    et al. [128] found that CO blobs are more numerous near the centers of OD columns,

    with nearby blobs tending to merge across OD borders; see Figure 2.7(a). The spatial

    relationship between CO blobs and OD columns is weak or lacking in new world primates

    such as the squirrel monkey [77] (pictured in Figure 2.7(b)), which also exhibit less strongOD segregation compared with their old world counterparts.

    (a) (b)

    1 mm

    Figure 2.7: OD patterns and CO distributions in other mammalian systems.(a) Cat OD map with the centers of the CO blobs represented by gray circlessuperimposed over the OD pattern. The scale bar is 1mm. Adapted from[128]. In (b), OD segregation in squirrel monkey with superposition of theCO patches outlined. Adapted from [77].

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    25/136

    14

    2.1.4 Cortical circuits

    It turns out that the majority of synapses onto a cortical neuron arise from other

    cortical cells rather than from feedforward LGN inputs. These include intralaminar con-

    nections, interlaminar connections and feedback connections from higher cortical areas.There are at least two distinct classes of intralaminar connection. First, there is a local

    circuit operating at submillimeter dimensions in which cells make connections with most

    of their neighbors in a roughly isotropic fashion. It has been suggested that such circuitry

    provides a substrate for the recurrent amplification and sharpening of the tuned response

    of cells to local visual stimuli [151, 10]. The other circuit connects cells separated by

    several millimeters of cortical tissue. The axons of these connections make terminal arbors

    only every 0.7 mm or so along their tracks [139, 63], such that local populations of cells

    are reciprocally connected in a patchy fashion to other cell populations. Optical imagingcombined with labeling techniques has generated considerable information concerning

    the pattern of these connections in superficial layers of V1 [118, 183, 19], see Figure 2.8.

    In particular, one finds that the patchy horizontal connections tend to link cells with

    similar feature preferences. Stimulation of a neuron via lateral connections modulates

    rather than initiates spiking activity [75, 161], suggesting that the long-range interactions

    provide local cortical processes with contextual information about the global nature of

    stimuli. As a consequence the horizontal connections have been invoked to explain a wide

    variety of context-dependent visual processing phenomena [62, 58].One of the most conspicuous features of cortical circuits is their laminar organization.

    Neurons within a layer send projections to only a subset of the other cortical layers with

    a high degree of accuracy. Additionally, it is believed that each cortical layer provides its

    primary output to just one other layer [32]. This assumption allows us to simplify the

    circuitry into subunits (see Chapter 5). Additionally, the response properties of neurons

    may vary through the layers due to a multitude of possible inputs to a particular cortical

    circuit, which could consist of inputs due to interlaminar projections, projections from

    other cortical areas, or from intralaminar connections. In Chapter 5, we provide a detailed

    review of the properties of the layer specific thalamic drive to V1 as well as outline the

    interlaminar circuits.

    Finally, there are extensive feedforward and feedback connections linking V1 to ex-

    trastriate areas such as V2, V3 and MT [143, 26]. As one proceeds to higher cortical areas

    the receptive field size of neurons increases. Hence, feedback from these areas provides

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    26/136

    15

    (a)

    (d)(c)

    (b)

    Figure 2.8: Camera-lucida drawing of biocytin injection into layer 3 at blob(a) and interblob (b) regions adapted from [183]. The biocytin patches areoutlined in the drawings (a) and (b). Corresponding representations of (a)and (b) are given in (c) and (d), respectively. In (c,d), the mocha areas arethe injection sites at blob/interblob regions, respectively, and the gray areasare the biocytin patches. The CO blobs are outlined in dashed lines. Wecan see that interblob regions connect to other interblob regions (12 out of17) and in general that blobs connect to blobs (8 out of 12). The scale barsrepresent 200 m and the arrows point to blood vessels, which are used as

    fiduciary landmarks.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    27/136

    16

    information to V1 neurons from a much larger region of visual space than expected

    from the classical feedforward receptive fields. It is also important to note that the

    feedforward and feedback connections are layer specific and reciprocal. Ongoing studies

    of feedback connections from points in extrastriate areas back to area V1 [5, 6], show thatthe feedback connectional fields are also distributed in highly regular geometric patterns,

    having a topographic spread of up to 13mm that is significantly larger than the spread of

    intrinsic lateral connections. It is likely that the patchiness again signifies that feedback

    correlates cells with similar feature preferences [148].

    2.2 Development of ocular dominance columns

    The existence of a set of overlapping cortical feature maps raises a number of inter-

    esting questions. What are the anatomical substrates for cortical feature maps and howdo they develop, to what extent are these maps genetically predetermined, and what role

    does neural activity play? Since our work is mainly concerned with the development

    of ocular dominance columns (and the joint development of CO blobs), we focus our

    discussion on this particular feature.

    Recall that ocular dominance is characterized by one eye predominantly driving a

    section of cortex. This, in turn, is intimately related to the segregation of the left and right

    LGN afferents terminating in layer 4 of the primary visual cortex. Thus, understanding

    the development of LGN afferent termination sites in V1 is key to the understanding of

    how OD columns develop. Hubel and Wiesel [175], who first detected ocular dominance

    columns using electro-physiological recordings and transneuronal tracers, theorized that

    molecular cues or genetic markers predetermine the initial wiring of LGN afferents and,

    consequently, the arrangement of ocular dominance columns in early development [83].

    They further postulated that neural activity subsequently refines the ocular dominance

    pattern during a critical period later in development. More recently, experimental ev-

    idence has come to light that seems to suggest that activity-dependent plasticity may

    still play a role in the initial development of OD columns, e.g., the existence of prenatal

    retinal waves [117] and recurrent thalamocortical activity [47] could both be underlying

    mechanisms for activitydependent development (reviewed in [28]).

    The idea that neural activity fashions the OD map draws upon a seminal conjecture

    of Donald Hebb regarding learning and synaptic plasticity [70]: When an axon of cell A

    is near enough to excite cell B or repeatedly or persistently takes part in firing it, some

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    28/136

    17

    growth process or metabolic change takes place in one or both cells such that As efficiency,

    as one of the cells firing B, is increased. A more modern interpretation of this postulate

    is that synaptic modification is driven by correlations in the firing activity of presynaptic

    and postsynaptic neurons. As stated, Hebbs rule is unstable and would result in allthe synaptic weights saturating: thus some sort of normalization constraint is required in

    order to maintain stability, i.e., a modified Hebbian rule is one in which the strengthening

    of one synapse comes at a cost to others. A Hebbianlike competition b etween LGN

    afferents is the backbone of the conjecture that activity-dependent plasticity plays the

    main role in the development of V1, a view contrary to a genetically predetermined

    structure of V1 as Hubel and Wiesel proposed. In this section, we present a brief history

    of the discovery of ocular dominance columns and arguments that support and challenge

    the above hypotheses regarding genetic versus activitybased mechanisms.

    2.2.1 Refinement of columns during the critical period

    In the late 1970s, Hubel and Wiesel performed a series of experiments examining

    the cellular mechanism by which patterned visual stimulation affects the development of

    visual perception [86, 87, 103, 104, 81] during a critical period in later development. They

    examined both kittens and monkeys by making electro-physiological recordings of neural

    responses in V1 elicited from stimuli presented to one or both eyes. They found that

    neurons in V1 responded almost exclusively to input from a single eye, in other words,

    most neurons are effectively monocularly driven. Nearby neurons in cortex were typically

    driven by the same eye, yielding patterns of left and right eye driven regions of cortex,

    ocular dominance patterns. In a key experiment [86], Hubel and Wiesel sutured shut

    an eye of an infant monkey, and hence eliminated much of the retinal activity, referred

    to as monocular deprivation (MD). After 6 months, the stitches were removed, and the

    previously sutured eye had been rendered blind. In a normal animal the ocular dominance

    stripes are of equal width, but in the monkey with one eye sutured the width of the stripes

    had changed: the stripes driven by the sutured eye were dramatically thinner than the

    remaining eye (see Figure 2.9). In binocular deprivation, on the other hand, the resulting

    ocular dominance pattern closely resembled the normal case. This suggested some sort

    of activitydriven competition between the left- and right-eye afferents for termination

    sites in V1 during the critical period.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    29/136

    18

    (a) (b)

    Figure 2.9: Activity or retinal drive during the critical period greatly altersthe final architecture of V1. In (a), the OD pattern in a normally rearedmonkey with OD stripes of approximately equal width. In (b), the OD patternfrom a monocularly deprived monkey that displays a stripe-like pattern, butthe stripes corresponding to the remaining open eye (white) are significantlywider, [86].

    2.2.2 Early development of OD columns

    LeVay et al. [103] used transneuronal tracers in multiple stages of development in

    order to ascertain how OD patterns develop in cat. The technique of tracking transneu-

    ronal transport of tritiated amino acids begins with the injection of a tracer into the

    retina. The tracer then proceeds through the retinal ganglia to the LGN and continues

    through the LGN afferents to terminate in layer 4C of V1. Experimenters use this

    technique to visualize the ocular dominance columns. LeVay et al. found that beforethe onset of the critical period, the injected tracers stained continuous rather than

    alternating bands of cortex, implying binocular input to V1 cells (see Figure 2.10).

    This calls into question Hubel and Wiesels conjecture of an innate architecture early in

    development. Sequential experiments in time unveiled alternating band patterns emerging

    at the onset of the critical period and b ecoming more distinct as the animal aged. This

    suggests that binocularity is the initial state of V1 and that, in time, activity-dependent,

    competition-based plasticity rules develop the OD pattern, i.e., afferent arbors initially

    overlap and are selectively pruned or strengthened during the critical period. However, itmay still be the case that there is an initial bias in the L/R thalamocortical connections

    as depicted in Figure 2.11. In subsequent experiments, LeVay et al. [104] discovered that

    leakage or spillover of the tracer occurs between the LGN layers of young animals (more

    severely in younger animals), suggesting that perhaps spillover in the tracing process was

    the cause of the continuous band, rather than the absence of a genetically predetermined

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    30/136

    19

    2 weeks

    3 weeks

    5.5 weeks

    13 weeks

    1 mm

    Figure 2.10: Audiographs of four stages in the development of the visualcortex in cat show the postnatal development of ocular dominance columns.Initially V1 appears binocularly driven and through the critical period theOD bands become distinct [103].

    segregation of LGN afferents. So there may in fact exist an initial segregation of LGN

    afferents, but due to complications in labeling it is not detectable. Nevertheless, LeVay

    et al. concluded that in the cat the two sets of afferents were intermixed initially since

    spillover would not have been sufficient to mask a columnar pattern had it been present

    [104].

    The time period for the initial formation of ocular dominance columns varies from

    species to species. In macaque monkeys, connections from LGN to the primary visual

    cortex begin to segregate into stripes prenatally. The finding of OD columns in prenatalmacaque monkeys supports the argument for genetically-driven development. However,

    local correlations in the firing of retinal ganglion cells have been found to exist in dark-

    reared mammals [121]. Furthermore, other studies found prenatal, spontaneously gener-

    ated, correlated patterns of activity from the retina, or retinal waves, independently

    generated from each eye [117]. These waves could, in theory, drive an activity-dependent,

    competition-based developmental process for prenatal pruning of LGN afferents.

    In cats, the formation of OD columns was originally thought to occur at the beginning

    of their critical period, postnatal day 21 (P21) [103]. Stryker and Harris [154] performedpivotal experiments on cats where they made binocular injections of tetrodotoxin (TTX)

    to block all forms of retinal activity from P14 to P45, in order to test the hypothesis

    that the segregation of overlapping afferents requires activity to induce competition. At

    P45, there was no evidence of OD columns [154]. The label in layer 4 of an animal

    at P45 was continuous, prompting Stryker and Harris to conclude that blocking retinal

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    31/136

    20

    Birth

    2 wk

    3 wk

    6 wk

    Figure 2.11: Assuming that molecular cues bias the afferents into overlapping,alternating bands of left and right control, throughout the critical periodactivity drives the further segregation of inputs to a final state of alternatingmonocularly driven bands of cortex. Timeline given as speculated in cat.Adapted from [91].

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    32/136

    21

    activity prevented activity-driven competition that should have segregated the afferents.

    Though the natural conclusion at the time, subsequent experiments by Crair et al. [42],

    using more advanced optical and single-unit recording techniques, provided evidence

    that OD columns are present at postnatal day 14, a week before the onset of the catscritical period. Thus an alternate conclusion to Stryker and Harris experiments is that

    geniculocortical afferents were already segregated in P14 and that TTX desegregated the

    existing OD columns.

    More support for the claim that OD columns are present before P21 is derived from

    comparisons to the ferrets developmental process. Ferret and cat development takes place

    on a similar time course [89], with ferrets having a 21-day lag behind cat with respect to

    birth dates. This makes ferrets good candidates for developmental studies because more

    of the developmental process can be observed. Using transneuronal tracing, OD columnswere found in ferrets at P37 (2 days after critical period onset) [57, 141] and at P14 in

    cat (a week before critical period onset) [47]. However using methods of direct injection,

    ferret thalamus is found to have segregated LGN columns at P16, implying cats have

    them 5 days before birth [47].

    2.2.3 Development of CO blobs

    In macaque both CO blobs and OD columns emerge prenatally, so that at birth the

    pattern of OD columns and their spatial relationship with blobs is adult-like. CO blobs

    in cat are normally first visible around 2 weeks of age [126], which is approximately

    coincident with the earliest observation of OD columns in cat [43]. Thus it is possible

    that CO blobs and OD columns develop at about the same time and thus interact with

    each other. In contrast to OD columns, however, CO blobs are not significantly altered

    by visual experience during the critical period. Modifying visual experience by either

    dark-rearing, monocularly or binocularly depriving a kitten has little or no effect on the

    cytochrome oxidase blob lattice [126]. In primates, further experiments by Murphy et

    al. [127] found no significant difference in the spacing of CO blobs between normal and

    strabismic macaque monkeys, monkeys that have one eye that cannot focus on an object

    because of an imbalance of the eye muscles. The independence of the CO blob lattice to

    visual experience could be interpreted in at least two ways:

    1. The CO blob lattice is defined intrinsically via molecular markers and acts as a

    roadmap to the developing visual cortex

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    33/136

    22

    2. The CO blob lattice is not predefined but there is a mechanism which constricts

    the blobs to the center of the OD regions. MD may change the width of the OD

    regions but the centers of the OD regions remain unchanged.

    We consider the first interpretation in Chapter 6 and the second in Chapter 5.

    2.3 The Hebbian synapse

    As we have already mentioned, the activity-driven development of OD columns in-

    volves competition between left and right eye LGN afferents. A possible cellular substrate

    for such competition is a set of synapses modified according to a form of Hebbs rule. Here

    we briefly review some of the biophysical evidence for a Hebbian synapse and its role in

    cortical development.

    2.3.1 Excitatory synapses and NMDA

    The basic stages of synaptic processing induced by the arrival of an action potential

    at an axon terminal are shown in Figure 2.12. (See [29] for a more detailed description).

    An action potential arriving at the terminal of a presynaptic axon causes voltage-gated

    Ca2+ channels within an active zone to open. The influx of Ca2+ produces a high con-

    centration of Ca2+ near the active zone [60, 11], which in turn causes vesicles containing

    neurotransmitter to fuse with the presynaptic cell membrane and release their contents

    into the synaptic cleft (a process known as exocytosis). The released neurotransmittermolecules then diffuse across the synaptic cleft and bind to specific receptors on the

    postsynaptic membrane. These receptors cause ion channels to open, thereby changing

    the membrane conductance and membrane potential of the postsynaptic cell.

    The predominant fast, excitatory neurotransmitter of the vertebrate central nervous

    system is the amino acid glutamate, whereas in the peripheral nervous system it is acetyl-

    choline. Glutamate-sensitive receptors in the postsynaptic membrane can be subdivided

    into two major types, namely, NMDA and AMPA [29]. At an AMPA receptor the

    postsynaptic channels open very rapidly. The resulting increase in conductance peakswithin a few hundred microseconds, with an exponential decay of around 1 msec. In

    contrast to an AMPA receptor, the NMDA receptor operates about 10 times slower

    and the amplitude of the conductance change depends on the postsynaptic membrane

    potential. If the postsynaptic potential is at rest and glutamate is bound to the NMDA

    receptor then the channel opens but it is physically obstructed by Mg2+ ions (see Figure

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    34/136

    23

    Action potential innerve terminalopens Ca channels

    Ca entry causesvesicle fusion andtransmitter release

    Transmitter

    Receptor-channels open,Na enters the postsynapticcell and vesicles recycle

    Presynaptic

    nerveterminal

    Receptor-channel

    2+

    2+

    Ca2+

    PostsynapticcellNa

    +

    +

    Na+

    Na+

    Figure 2.12: Schematic of synaptic transmission. Adapted from [91].

    2.13). As the membrane is depolarized, the Mg2+ ions move out and the channel becomes

    permeable to Na+ and Ca2+ ions. The NMDA receptor is thus well placed to act as a

    detector of correlations in the firing activity of presynaptic and postsynaptic neurons,

    which forms the basic principle of a Hebbian synapse. Indeed, the rapid influx of calcium

    ions due to the opening NMDA channels is thought to be the critical trigger for the onset

    of long term potentiation (LTP) and long term depression (LTD), two major components

    of bidirectional synaptic plasticity [16, 48, 15, 106, 119]. LTP is a p ersistent increase in

    synaptic efficacy produced by high-frequency stimulation of presynaptic afferents or by the

    pairing of low frequency presynaptic stimulation with robust postsynaptic depolarization.

    LTD is a long-lasting decrease in synaptic strength induced by low-frequency stimulation

    of presynaptic afferents.

    The existence of LTP and LTD has been well documented in the visual cortex (e.g.,

    see the review of Malenka and Bear [119]) and that the NMDA receptors play a key

    role. One of the most striking demonstrations of the role of the NMDA receptors in the

    development of the visual system occurs in the frog. We compare cases where the NMDA

    receptors are blocked or overly activated to the control case. In normal development

    of the frog retinotectal map, the L/R eye afferents pass through the optic chiasm and

    terminate in the R/L hemispheres of the tectum, which is the analog of V1. However,

    with the transplant of an additional right eye, the two right eye afferents traverse the

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    35/136

    24

    Resting membranepotential

    NMDAR AMPAR

    Mg2+

    Na+

    Na+

    Ca2+

    Glu

    Depolarized membranepotential

    NMDAR AMPAR

    Mg2+

    Na+

    Na+

    Ca2+

    Glu

    Figure 2.13: Schematic of NMDA receptor.

    optic track to the left hemisphere tectum. The afferents then compete for termination

    sites in the tectum resulting in a stripe-like ocular dominance pattern, as pictured in

    Figure 2.14(a). Constantine-Paton et al. (reviewed in [39]) studied the role of the NMDA-

    type glutamate receptors by coadministration of the NMDA channel blocker MK801, i.e.,

    effectively turning off the NMDA receptors. With the NMDA receptors blocked, the two

    right-eye pathways fail to segregate, as seen in Figure 2.14(c). On the other hand, an

    application of NMDA sharpens the segregation [39] as shown in Figure 2.14(d) compared

    to the untreated tectum at the same magnification shown in Figure 2.14(b).

    2.3.2 Neurotrophins

    Neurotrophins are proteins that initiate essential biological functions required for

    the maintenance and/or strengthening of neural connections. Included in the class of

    neurotrophins are nerve growth factor (NGF), brain-derived neurotrophic factor (BDNF),

    neurotrophin-3 (NT-3), and NT-4/5. For example, a sufficient amount of neurotrophic

    factor is required for sustaining axonal arbors and determines the degree of the arboriza-

    tions (e.g., [40]). Interestingly, neurotrophins also affect synaptic weights. Thus, it

    has been postulated that the regulation of the neurotrophins and their receptors could

    be a mechanism involved in plasticity. There is an array of examples from multiple

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    36/136

    25

    (a) (b)

    (c) (d)

    Figure 2.14: Demonstration of the effects of NMDA. In (a) we show anuntreated tectum and the resulting segregated pattern, whereas in (c) thetectum is treated with MK801, an NMDA antagonist. We note that theretinotectal projections are disperse and without order. In (b) is a moremagnified view of the untreated tectum. In (d) we show the effects of anNMDA agonist and note the sharpness of the pattern when contrasted with(b). Scale bars in (a,c) are 200 m and 100 m in (b,d). Adapted from [39].

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    37/136

    26

    cortical areas that demonstrate the substantial role of neurotrophins on plasticity. In

    the visual cortex, BDNF has been shown to facilitate LTP [3, 88] and to attenuate LTD

    in layer 2/3 synapses of young adult rats [2, 88, 95, 100], suggesting that one role of

    BDNF during development is to modulate the properties of synaptic plasticity, enhancingsynaptic strengthening and reducing synaptic weakening processes which contribute to

    the formation of specific synaptic connections [88].

    Interestingly, neuroelectrical activity regulates the expression of neurotrophins and

    their receptors, i.e., the expression of neurotrophins is associated with active synapses.

    For instance, the regulation by BDNF of dendritic arborization requires both neuronal

    activity and the Ca2+ influx due to Mg2+ NMDA receptors [122]. The interdependence

    of the expression of neurotrophic factor and postsynaptic activity makes it a leading

    candidate as a principal mechanism for activity-dependent synaptic plasticity [122].It has been suggested that neurotrophins could be thought of as a resource for which

    postsynaptic connections compete (e.g., see [116] and the review [165] on modeling).

    Specifically, neurotrophins play a role in activity-dependent synaptic modification in the

    development of ocular dominance columns [116, 30, 80]. For example, the continuous

    infusion of the neurotrophins NT 4/5 or BDNF to V1 prevents the formation of OD

    columns in cat [30], presumably because the LGN axon branches fail to retract, yet the

    application of a NT antagonist prevents OD column segregation by eliminating inputs to

    both eyes [31]. In monocular deprivation experiments on rats, Maffei et al. showed thatwith additional NGF injected, the OD distribution fails to produce shifts towards open-eye

    dominance[116], i.e., the OD distribution remains approximately the same, suggesting

    that with an abundance of neurotrophic factor, the L/R neural afferents need not compete

    for resource. These results support the notion that neurotrophins are a substrate for which

    postsynaptic connections compete, yet may also, because of their regulation by acivity,

    be involved in a type of Hebbian plasticity.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    38/136

    CHAPTER 3

    MODELING CORTICAL

    DEVELOPMENT

    3.1 LGN afferents terminating at a single neuron

    Consider a single cortical neuron that receives two inputs, IL and IR, with correspond-

    ing synaptic weights wL and wR. These could represent, for example, LGN afferents from

    the left and right eyes. Suppose that the membrane potential of the cell evolves according

    to the linear equation

    vdV

    dt= V + w I = V + wLIL + wRIR, (3.1)

    where w = (wL, wR)T, I = (IL, IR)

    T, and v is a membrane time constant. Assuming

    that IL and IR are in units of1V , then wL and wR have the same units as the membrane

    potential V. Since development takes place on a much slower time scale than the dynamics

    of the feedforward input, we take V to be at steady-state. Thus,

    V = wLIL + wRIR. (3.2)

    Recall Donald Hebbs conjecture that if input from neuron A often contributes to the

    firing of neuron B, then the synapse connecting neuron A to B should be strengthened

    [70], i.e., if neuron j drives neuron B to fire then wj , the synaptic weight between j and

    B should increase. These dynamics are described by

    wd

    dt wLwR

    = V

    ILIR

    . (3.3)

    where w determines the timescale of the weight dynamics with w V. Equation(3.3) is the mathematical representation of the basic Hebb rule. Note that activity of

    the postsynaptic cell is specified by the membrane potential V, which is consistent with

    the basic operation of the NMDA receptor (see section 2.3). Because development takes

    place on a slower time scale then the feedforward dynamics, it is the long term statistics

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    39/136

    28

    of the activity pattern that matter. Therefore, we consider an averaged version of the

    Hebb rule.dw

    dt=< V I >, (3.4)

    where < . > denotes the ensemble average over the distribution of inputs.

    From equation (3.2), we can replace V with w I and obtaindw

    dt= Cw =

    < IL IL > < IR IL >< IL IR > < IR IR >

    w, (3.5)

    where C is the input correlation matrix with matrix elements given by Ci,j =< Ij Ii >.

    Within the context of ODC formation, C represents the correlation of the activities within

    an eye and between the eyes. Ocular dominance occurs when the synaptic connections

    originating from one eye are driven to zero while the connections from the alternate

    eye grow or strengthen. As it stands, the simple Hebb rule does not generate such

    competition. Moreover, it is inherently unstable. We now describe some of the most

    common modifications of Hebbs rule that remove these limitations.

    3.1.1 Swindale model

    Swindale rectified the problem of unbounded synaptic weights by introducing a logistic

    term to the right hand side of equation (3.5) [155], i.e.,

    dw

    dt= Cw

    F(w) (3.6)

    where F(w) = w(W w) ensures that the left and right synaptic densities remain bothnonnegative and bounded above by some maximum density W.

    One expects sameeye correlations CRR and CLL to be positive and stronger than the

    oppositeeye correlations CRL and CLR. Swindale took the eyes to be anticorrelated, i.e.,

    CLR, CRL < 0 in order to induce competition b etween the L/R eyes. The reversal in sign

    of opposite eye interactions is supposed to reflect negative statistical correlations between

    left and right eye inputs. However, the existence of negative correlations is difficult to

    justify from a neurobiological perspective. The problem of negative correlations can

    be avoided by using a linear Hebbian model with subtractive normalization (see below)

    instead of the Swindale model [124]. It turns out that both models exhibit very similar

    behavior and can be analyzed in almost an identical fashion.

    As a further simplification, suppose that the total synaptic weight at a point in cortex

    is constant with wL + wR = W. This condition implies dwR/dt = dwL/dt, which is

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    40/136

    29

    guaranteed if the eyes are symmetrically anticorrelated, CRR = CLR and CLL = CRL.With this correlation structure, the correlation matrix C has a single positive eigenvalue

    with corresponding eigenvector, e = (1, 1)T, and growth in this direction corresponds

    to the segregation of the L and R pathways. When intracortical interactions are included,this model yields an OD column pattern consistent with experimental observations, see

    section 3.2 and [155].

    3.1.2 Subtractive normalization

    Another approach is to assume that the total synaptic weight that connects to each

    point in cortex remains conserved. We can accomplish this by subtracting an equal

    amount from each synaptic weight according to the subtractive normalization scheme.

    dwdt = Cw (w)a (3.7)

    where a = (1, 1)T and (w) enforces the conservation constraint. Equation (3.7) is

    supplemented with additional constraints that ensure that the weights remain positive

    and bounded

    0 wL,R W. (3.8)

    For illustrative purposes, suppose that the left and right eye inputs are symmetric.

    That is, we take CLL = CRR = CS and CLR = CRL = CD so that C is a positive

    symmetric matrix. Note that unlike the Swindale model, no assumption of anticorrelatedeyes is needed. We then set

    (w) = [wL + wR] . (3.9)

    Exploiting the fact that the input correlation matrix C has eigenvalues = CS CDwith corresponding eigenvectors e = (1, 1), it is straightforward to show that thevector equation (3.7) decomposes into the pair of decoupled equations

    wdw+(r, t)

    dt= (CS+ CD 2)w+(t) (3.10)

    wdw

    (r, t)

    dt = (CS CD)w(t) (3.11)

    with w = wL wR. Conservation of total synaptic density is thus achieved by setting = +/2 so that dw+(t)/dt = 0 for all t. For a more general choice of correlation matrix

    C, the subtractive normalization term takes the form

    (w) =1

    a a (a Cw) . (3.12)

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    41/136

    30

    Note that the subtractive normalization must be modified when there are more than

    two populations vying for termination sites. We discuss this further in Chapter 5. This

    phenomenological model could be interpreted as a simple model for the left and right eye

    afferents competing for a common pool of resources, neurotrophins, which are requiredfor the maintenance and strengthening of synapses.

    3.1.3 Competition for neurotrophins

    Instead of introducing a mathematical abstraction to maintain total synaptic strength

    (as in subtractive normalization), Ermentrout and Osan [55] consider a single pool of

    resources that synaptic populations require in order to maintain or strengthen synaptic

    connections (akin to a previous more detailed approach in [65]), which could, for example,

    correspond to a p ool of neurotrophic factor at the postsynaptic location, see section 2.3.2.

    One assumes mass action kinetics between the resource and the amount of synaptic

    weight,

    fK+

    Kw (3.13)

    where f is the available pool of substance needed for producing lasting synaptic connec-

    tions and w is the strength of a synapse. Suppose that the total amount of neurotrophic

    factor is conserved (both bound and unbound). So that

    f + w = W (3.14)

    where in this context W denotes a constant corresponding to the total amount of resource,

    an upperbound on synaptic weight. Without any loss of generality, we can let W = 1.

    Now consider a pair of synaptic connections each having its own pool of resources.

    Then the weight dynamics are given by

    dwidt

    = K+(1 wi) Kwi, i = L,R. (3.15)

    The positive influence of trophic factor on LTP has been suggested by experiments

    in the rat hippocampal system in knockout mice [98]. Additionally, in slices of rat visual

    cortex, Korte et al. showed that a type of neurotrophin, BDNF, plays a modulatory

    role in synaptic plasticity [99]. In simple terms, these experiments suggest that the more

    trophic factor available, the faster its uptake and subsequent increases in connection

    strength. Motivated by the experiments of Korte et al., we take the forward binding rate

    to be a nonnegative monotonically increasing sigmoidal function that depends upon a

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    42/136

    31

    Hebbian term, that is high pre- and post-synaptic activity leads to a strengthening of

    cortical connections. The steepness of the binding rate could be related to the uptake of

    neurotrophic factor by the presynaptic afferent (see [65] for a detailed description). Yet

    since there is only a finite amount of resource, we assume an active form of decay wherethe backward binding rate, i.e., decay of synaptic weights, is a function of the averaged

    postsynaptic activity and similarly represent it as a nonnegative monotonically increasing

    sigmoidal function. Then the dynamics can be written as

    dwidt

    = K+

    j=L,R

    Ci,jwj

    (1 wi) K j=L,R

    wj

    wi, i = L,R, (3.16)so that the forward binding rate has an LTP-like term and the backward binding rate

    has an LTD-like behavior. Note that we are considering L/R inputs that are statistically

    equivalent, so that when one averages over the developmental period, the L/R eyes have

    the same amount of presynaptic drive. Thus, the total postsynaptic activity due to

    L/R drive is proportional to the feedforward synaptic weights. If the synaptic weights

    competed for the same pool, then 1 wL,R would be replaced by 1 wL wR, thoughconsidering the first choice does not lead to a substantial difference in the models

    behavior.

    A binocular steady state, i.e., wL = wR = w, exists provided the function g(w),

    g(w) = K+(+w)(1

    w)

    K(2w)w, (3.17)

    where + = CS+ CD, has a root in the interval (0,1). Since K are both nonnegative

    functions, g(0) = K+(0) > 0 and g(1) = K(2) < 0. Thus by the intermediate valuetheorem, there exists w (0, 1) such that g(w) = 0. In order to study the stability of thisfixed point, we linearize equation (3.16) about the binocular fixed point, w, to obtain

    dcLdt

    = AccL + BccRdcRdt

    = BccL + AccR (3.18)

    where

    Ac = CSK+

    (+w)(1 w) K+

    (+w) K

    (2w) K

    (2w)wBc = CDK

    +(+w)(1 w) K(2w)w, (3.19)

    and K

    = dKdw . The system yields two eigenvectors, e = (1, 1)T, corresponding tosymmetric and antisymmetric growth (note antisymmetric growth is synonymous with

    competition) with the symmetric eigenvalue

    s = Ac+ Bc = (CS+ CD)K+(+w)(1 w) K+(+w) K(2w) K(2w)w (3.20)

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    43/136

    32

    and the antisymmetric eigenvalue

    a = Ac Bc = (CS CD)K+(+w)(1 w) K+(+w) K(2w). (3.21)

    In order for the L and R pathways to segregate, we require s < 0 and a > 0. SincedK

    dw > 0, there are two ways that this could occur: either CD < 0, which corresponds

    to negative correlations between the eyes and is not a realistic assumption, or K(2w)

    becomes very large. The physical meaning of the latter case is tied to a requirement for

    strong activity-dependent synaptic depression. That is, when the postsynaptic activity

    is high (equivalent to the feedforward weights being strong), the decay term increases

    sharply resulting in a decrease in the total synaptic weight.

    We consider as an example, the case CS > 0, CD = 1 CS > 0. Take the reaction

    rates to be of the form

    K+(u) =1

    1 + e(u0.5)and K(u) =

    1

    1 + e(u/20.5). (3.22)

    In this example, the binocular fixed point is located at w = 1/2. The steepness of

    the forward and backward reaction rates is determined by , which we will treat as a

    bifurcation parameter. Large corresponds to strong LTP with strong active decay of the

    synaptic weights. From a near 0 value, as increases the binocular state loses stability at

    a subcritical pitchfork bifurcation (plotted in Figure 3.1(a)), meaning that for large the

    binocular state quickly destabilizes and goes to either a left or right eye dominated state

    dependent upon the initial conditions. Since the bifurcation is a subcritical pitchfork,

    there exists a bistable region where both the binocular and L/R eye dominated states are

    stable. In order to study the size of the basins of attraction for the fixed points, we plot

    nullclines and some sample trajectories in phase space (pictured in Figure 3.1(b)). We

    find that only a narrow set of initial conditions around the origin result in a binocular

    state. So while the binocular state in this parameter regime is stable, any significant

    perturbation should result in a L or R eye dominated state.

    3.2 ODC development on a cortical sheet

    Extending the analysis of a single postsynaptic neuron with multiple inputs to a system

    of postsynaptic neurons modeled continuously is relatively straightforward. Suppose we

    have a two-dimensional cortical sheet. To model the interactions between neighboring

    cells, we introduce a recurrent weight function J that depends upon cortical distance; we

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    44/136

    33

    wL

    1.0

    0.75

    0.25

    0.5

    05 10 15 20 25 30

    (a) (b)

    0 1.20.80.4

    1.2

    0.8

    0.4

    0

    wR

    wL

    w =0L.

    w =0R.

    Figure 3.1: Stability properties of synaptic weights in a system where thereis competition for a pool of neurotrophic factor. (a) Bifurcation diagramwith CS = 0.8 and CD = 0.2 that demonstrates that increasing results in

    a subcritical pitchfork bifurcation at 13.33. On the unstable branches,it undergoes fold bifurcations at values of 10.665. Hence, there exists anarrow parameter regime that exhibits bistability. Note that solid lines rep-resent stable solutions, whereas the dashed line represents unstable solutions.(b) Phase diagram for solutions in wL, wR space in the bistable regime with = 12.4. Stable fixed points are circled, whereas the unstable fixed points aredenoted by triangles. Additionally, we plot a few sample tra jectories, shownin blue, about the origin. Both (a) and (b) were obtained using XPP.

    denote cortical position by r. The recurrent weight function is meant to represent the

    lateral cortico-cortical interactions. As outlined in section 2.1.4, the lateral circuitry has

    at least two spatial components: the local, submillimeter connections and the nonlocal

    patchy connections that link cortical points several millimeters apart. In this work, we

    consider only the local connectivity and assume it is an innate feature of V1, so that J

    represents solely the effects due to the submillimeter interactions, which are postulated

    to amplify and sharpen tuned responses [151] and does not vary with time or cortical

    location. Further, assume that cortico-cortical interactions are weak. Thus, we introduce

    a small parameter to scale recurrent connection strengths.

    Analogous to equation (3.1), the activity of postsynaptic cells changes with time as

    vdV

    dt(r, t) = V(r, t) + I(r) w(r) +

    J(r r)V (r, t)dr, (3.23)

    where V, I, and w now vary with space and is a factor that converts membrane potential

    to output firing rate, assuming these neurons operate in a linear regime.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    45/136

    34

    Since development takes place on a much slower timescale than the dynamics of

    cortical activity, we can take V to be given by its steadystate value. However, calculating

    the steadystate explicitly requires inverting the nonlocal linear operator

    LV(r) = V(r)

    J(rr )V (r)dr. In the case of weak intracortical interactions, this inversion can becarried out by performing a perturbation expansion in (see appendix). The firstorder

    approximation is thus

    V(r) =

    G(r r) wL(r)IL(r) + wR(r)IR(r) dr (3.24)

    with G(r r) (r r) +

    J (r r) and is the Dirac delta function. Note that weexplicitly include the -function component of the inverted operator L1. Usually thisterm is ignored and G is treated as a smooth weight function [124, 158]. However, in

    order to preserve the invertibility of L, it is necessary to include the -component and torestrict the strength of intracortical interactions as determined by (see appendix). If is

    too large there does not exist a stable steadystate solution and the linear approximation

    of correlationbased learning breaks down. As we show below, the -component can have

    a significant effect on the weight dynamics. In particular, it leads to a sensitivity to initial

    conditions, namely, the initial balance of left and right eye afferents.

    Substitute equation (3.24) into equation (3.4) and assume that the input correlations

    are of the form

    IL(r)IL(r) IL(r)IR(r)IR(r)IL(r) IR(r)IR(r) = S(r r)C, C = CS CDCD CS (3.25)to obtain the subsequent averaged Hebb rule

    wdw

    dt=

    G(r r)S(r r)Cwdr, (3.26)

    where S represents the spatial correlations of inputs and is taken to be a Gaussian and

    w = (wL, wR)T. Thus, we can incorporate S into G and express the averaged Hebb rule

    as

    wdw

    dt = H(r r)Cwdr, (3.27)where H(r) = G(r)S(r).

    The averaged Hebbs rule in a spatially extended system has the same problems as

    its single cell representation, i.e., a lack of competition and being unbounded. As such,

    we provide the derivation for subtractive normalization on a cortical sheet, but note that

    the underlying principles hold for the other normalization schemes.

  • 8/3/2019 Andrew M. Oster- Mathematical Models of Cortical Development

    46/136

    35

    3.2.1 Subtractive normalization

    Introducing a subtractive normalization term to (3.27) leads to t