• Nenhum resultado encontrado

Arquitetura das Redes de Neurônios

N/A
N/A
Protected

Academic year: 2021

Share "Arquitetura das Redes de Neurônios"

Copied!
37
0
0

Texto

(1)

Arquitetura das

(2)

Tipos de arquitetura

• 

Ar/ficial

• 

Determinada por dados

(3)

Arquiteturas Ar/ficiais

• 

Grafo de Erdős–Rényi

• 

Mundo pequeno

• 

Livre de escala

• 

Solé & Valverde, 2004

(4)

Teoria de Grafos: algumas definições

Grafo: direcionado ou

não-direcionado

Arestas: binárias ou

ponderadas

www.brain-connec/vity-toolbox.net

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

8

Graphs can be displayed in matrix form or by embedding them in (usually

2D) space. Graph embedding and visualization is an extremely active

area of research in its own right@.

Network Architectures and Metrics

Graphs: Visualization

Visualizing a graph goes a long way towards understanding its structure.

pajek

Network Architectures and Metrics

Random Graphs

Erdös-Renyi (ER) random graphs: n nodes, k edges, all edges are

generated with

equal probability

p.

single network

1000 examples

p

Nodes in ER graphs have

normally distributed degrees

(characteristic scale)

n = 47, k = 505 (directed, binary)

Erdös & Renyi (1961) Publ Math Inst Hung Acad Sci 5, 17W61.

(5)

Redes complexas

WaZs & Strogatz (1998) Nature 393, 440

!

Regular

L grande, C grande

Mundo Pequeno

L pequeno, C grande

Erdős–Rényi

L pequeno, C pequeno

Aumento na aleatoriedade das conexões

(6)

Tipos de redes complexas

Sporns et al. (2004)

Trends Cogn Sci 8, 418

Livre de

Escala

Modular

Hierárquica

e modular

11

Network Architectures and Metrics

Small-World Networks

Clustering and path length alone are

insufficient

to fully characterize

network topology. There are

several types

of small-world architectures.

Regular graph with shortcuts C

no communities

Watts & Strogatz (1998) Nature 393, 440.

Kaiser et al. (2007) New J Phys 9, 110.

Sporns (2006) Biosystems 85, 55.

no communities

Graph with hierarchically arranged modules C

fractal (self-similar) adjacency matrix

Network Architectures and Metrics

Small-World Networks

Clustering and path length (actually all graph metrics) must be compared

to

appropriately constructed random (null) models.

What matters is the degree to which a given metric

deviates

from its value

i th

ll

d l

in the null model.

The absolute value of the metric provides little information and can vary

widely with network size and density.

Null models typically preserve the number of nodes and edges, as

well as the node degrees and degree sequence.

Rubinov and Sporns (2010) Neuroimage

The selection of a random model introduces important

assumptions and biases into the analysis (and thus also

the interpretation of the results)!

4

DEUTSCHE PHYSIKALISCHE GESELLSCHAFT

Figure 1.

Schematic view of a hierarchical cluster network with five clusters

containing five sub-clusters each.

(a) (b) (c)

Figure 2.

Spread of initially unlocalized activity in (a) random, (b)

small-world and (c) hierarchical cluster networks (i = 100, i

0

= 1000), based on 30

simulations for each network.

plotting figure

2

for various values of p, the simulation yielded sustained intermediate activity

(neither dying out nor spreading through the whole network) only if p was between 0.4 and

0.6. Consequently we chose p = 0.5 for all simulations. Note that for this p, the small-world

and hierarchical network had similar clustering coefficients and characteristic path lengths (see

table

1

) This means that the hierarchical cluster network also possessed small-world

characteristics [

17

]. Given the high probability of rewiring in the default small-world network,

one may wonder if this network still possessed strong local clustering comparable to the

hierarchical cluster network. To test this aspect, we measured the edge density within each

set of 10 consecutive network nodes in the small-world network, that is for 1000 sets in total.

New Journal of Physics9 (2007) 110 (http://www.njp.org/)

Kaiser et al. (2007) New J Phys 9, 110

dynamics of neurons and neuronal populations result in

patterns of statistical dependencies (functional

connec-tivity) and causal interactions (effective connecconnec-tivity),

defining three major modalities of complex brain networks

(

Box 2

). Human cognition is associated with rapidly

changing and widely distributed neural activation

pat-terns, which involve numerous cortical and sub-cortical

regions activated in different combinations and contexts

[12–15]

. Two major organizational principles of the

cerebral cortex are functional segregation and functional

integration

[16–18]

, enabling the rapid extraction of

information and the generation of coherent brain states.

Which structural and functional principles of complex

networks promote functional segregation and functional

integration, or, in general, support the broad range and

flexibility of cognitive processes?

In this review we examine recent insights gained about

patterns of brain connectivity from the application of novel

quantitative computational tools and theoretical models to

empirical datasets. Whereas many studies of single

neuron networks have revealed their complex morphology

and wiring

[19]

, our focus is on the large-scale and

intermediate-scale networks of the cerebral cortex,

allow-ing us to examine links between neural organization and

cognition arising at the ‘systems’ level. We divide this

review into three parts, devoted in turn to the

organiz-ation (structure), development (growth) and function

(dynamics) of brain networks.

Structural organization of cortical networks

Most structural analyses of brain networks have been

carried out on datasets describing the large-scale

connec-tion patterns of the cerebral cortex of rat

[20]

, cat

[21,22]

,

and monkey

[23]

– structural connection data for the

human brain is largely missing

[24]

. These analyses have

revealed several organizational principles expressed

within structural brain networks. All studies confirmed

that cerebral cortical areas in mammalian brains are

neither completely connected with each other nor

ran-domly linked; instead, their interconnections show a

specific and intricate organization. Methodologically,

investigations have used either graph theoretical

Box 1. Complex networks: small-world and scale-free architectures

For any number of n nodes and k connections, a RANDOM GRAPH (see

Glossary) can be constructed by assigning connections between pairs

of nodes with uniform probability. For many years, random graphs

(

Figure I

a) have served as a major class of models for describing the

topology of natural and technological networks. However, although

random graphs have yielded numerous and often surprising

math-ematical insights

[5]

, they are probably only poor approximations of

the connectivity structure of most complex systems.

Classical experiments

[71]

first revealed the existence of a small

world in large social networks. Small worlds are characterized by the

prevalence of surprisingly short PATHS linking pairs of nodes within

very large networks. In a seminal paper

[10]

, Watts and Strogatz

demonstrated the emergence of small world connectivity in networks

that combined ordered lattice-like connections with a small admixture

of random links (

Figure I

b). Combining elements of order and

randomness, such networks were characterized by high degrees of

local clustering as well as short path lengths, properties shared by

genetic, metabolic, ecological and information networks

[1–3]

.

The nodes in random graphs have approximately the same DEGREE

(number of connections). This homogeneous architecture generates a

normal (or Poisson) degree distribution. However, the degree

distributions of most natural and technological networks follow a

power law

[11]

, with very many nodes that have few connections and a

few nodes (hubs) that have very many connections (

Figure I

c). This

inhomogeneous architecture lacks an intrinsic scale and is thus called

SCALE-FREE. Scale-free networks are surprisingly robust with respect

to random deletion of nodes, but are vulnerable to targeted attack on

heavily connected hubs

[45]

, which often results in disintegration of

the network. A corollary of this finding is that the connection topology

of scale-free networks cannot be efficiently captured by random

sampling, as most nodes have few connections and hubs will tend to

be under-represented. Sampling is thus a crucial issue for determining

if brain networks have scale-free topology.

Systematic investigations of large-scale

[32–35]

and

intermediate-scale

[35,36]

structural cortical networks have revealed small-world

attributes, with path lengths that are close to those of equivalent

random networks but with significantly higher values for the

clustering coefficient. At the structural level, cortical networks

either do not appear to be scale-free

[35]

or exhibit scale-free

architectures with low maximum degrees

[44]

, owing to saturation

effects in the number of synaptic connections, which prevent the

emergence of highly connected hubs. Instead, functional brain

networks exhibit power law degree distributions as well as

small-world attributes

[52,62]

.

L=1.73 (0.06)

C=0.52 (0.05)

L=1.79 (0.04)

C=0.52 (0.04)

L=1.68 (0.01)

C=0.35 (0.03)

(a)

(b)

(c)

Figure I. Structure of random, small-world and scale-free networks. All networks have 24 nodes and 86 connections with nodes arranged on a circle. The characteristic

path length L and the clustering coefficient C are shown (mean and standard deviation for 100 examples in each case; only one example network is drawn). (a) Random

network. (b) Small-world network. Most connections are among neighboring nodes on the circle (dark blue), but some connections (light blue) go to distant nodes,

creating short-cuts across the network. (c) Scale-free network. Most of the 24 nodes have few connections to other nodes (red), but some nodes (black connections) are

linked to more than 12 other nodes. For comparison, an ideal lattice with 24 nodes and 86 connections has LZ1.96 and CZ0.64.

Review

TRENDS in Cognitive Sciences Vol.8 No.9 September 2004

419

www.sciencedirect.com

(7)

Redes complexas são comuns

15

Network Architectures and Metrics

Scale-Free Networks 9 Some Issues

Power laws have a long history (e.g. B

Pareto distributions

D, B

ZipfIs Law

D).

Power laws are found in extremely

disparate systems

(earthquakes,

popularity rankings, book sales, stock market fluctuations, population of

cities, word frequencies etc etc)

Statistical properties of power law probability distributions depend on

the value of the

exponent

(e.g. for 1 <

2, both mean and variance

are infinite, while for > 3, both are finite).

Power laws can be

difficult to determine

from finite data samples. Care

must be taken when binning data [For a rigorous treatment of this

Clauset et al. (2009) SIAM Review 51, 661

must be taken when binning data. [For a rigorous treatment of this

issue see Clauset et al. 2009]

True power laws exist only in

very large

(theoretically infinitely large)

networks 9 finite sizes can result in Bcut-offsD at the high end of the

degree distribution.

Network Architectures and Metrics

A Map of Network Architectures

Real-world networks occupy distinct locations in a continuous space of

possible network architectures.

Solé & Valverde (2004) Lect Notes Phys 650, 189.

Solé & Valverde (2004) Lect Notes Phys 650, 189

Internet

Tráfego aéreo

Rede

metabólica

Redes

sociais

(8)

Redes complexas no cérebro

as well as primate prefrontal cortex

[41]

. The algorithm

could be steered to identify clusters that no longer

contained any known absent connections, and thus

produced maximally interconnected sets of areas. The

identified clusters largely coincided with functional

cortical subdivisions, consisting predominantly of visual,

auditory, somatosensory-motor, or frontolimbic areas

[32]

.

On a finer scale, the clusters identified in the primate

visual system closely followed the previously proposed

dorsal and ventral visual streams, revealing their basis in

structural connectivity patterns.

In networks composed of multiple distributed clusters,

inter-cluster connections take on an important role. It can

be demonstrated that these connections occur most

frequently in all shortest paths linking areas with one

another

[42]

. Thus, inter-cluster connections can be of

particular importance for the structural stability and

efficient working of cortical networks. The degree of

CONNECTEDNESS

of neural structures can affect the

func-tional impact of local and remote network lesions

[43]

, and

this property might also be an important factor for

inferring the function of individual regions from

lesion-induced performance changes. Indeed, the cortical

net-works of cat and macaque are vulnerable to the damage of

the few highly connected nodes

[44]

in a similar way that

scale-free networks react to the elimination of hubs

[45]

.

Random lesions of areas, however, have a much smaller

impact on the characteristic path length.

Network growth and development

The physical structure of biological systems often reflects

their assembly and function. Brain networks are no

exception, containing structures that are shaped by

evol-ution, ontogenetic development, experience-dependent

refinement, and finally degradation as a result of brain

injury or disease.

Degree k

Counts (k)

Degree k

10

1

10

1

10

0

10

0

10

2

10

2

10

3

10

3

10

4

Counts (k)

0

500

700

800

0.3

0.4

0.5

0.6

1.7

1.8

1.9

Path length

Clustering coefficient

Lattice

Macaque visual cortex

Random

(a)

(b)

(c)

(d)

r

c

= 0.6

r

c

= 0.7

Figure 1. Small-world and scale-free structural and functional brain networks. (a) Characteristic path length and clustering coefficient for the large-scale connection matrix

(see Glossary) of the macaque visual cortex (red) (connection data from

[23]

, results modified from

[35]

). For comparison, 10 000 examples of equivalent random and lattice

networks are also shown (blue). Note that the cortical matrix has a path length similar to that for random networks, but a much greater clustering coefficient. (b) Cluster

structure of cat corticocortical connectivity, based on

[32]

and visualized using Pajek (

http://vlado.fmf.uni-lj.si/pub/networks/pajek/

). Bars indicate borders between nodes in

separate clusters. Cortical areas were arranged around a circle by evolutionary optimization, so that highly inter-linked areas were placed close to each other. The ordering

agrees with the functional and anatomical similarity of visual, auditory, somatosensory-motor and frontolimbic cortices. (c) A typical functional brain network extracted from

human fMRI data (from

[52]

). Nodes are colored according to degree (yellowZ1, greenZ2, redZ3, blueZ4, other coloursO4). (d) Degree distribution for two correlation

thresholds. The inset depicts the degree distribution for an equivalent random network (data from

[52]

).

Review

TRENDS in Cognitive Sciences Vol.8 No.9 September 2004

421

www.sciencedirect.com

Sporns et al. (2004) Trends Cogn Sci 8, 418

the system, summed over all subset sizes. Thus, complexity

provides a measure for the amount of information that is

integrated within a neural system (for a discussion of

complexity measures, see Box 2).

A schematic illustration of the notion of complexity,

based on the results of large-scale computer simulations

45,46

is given in Fig. 3 (for details see Fig. 3 legend). Complexity

is evaluated for the spontaneous activity of three simulated

examples of a cortical area that differ in the anatomical

pat-tern of their intra-areal (voltage-dependent) re-entrant

con-nections. Figure 3A shows the activity patterns that emerge

when neuronal groups are connected to each other

accord-ing to anatomical rules (i.e. specificity

47–49

, anisotropy

50

,

and fall-off with distance) derived from the actual

organiz-ation of primary visual cortex. The spontaneous dynamic

behavior of this system is complex: neighboring neurons of

similar orientation preference tend to fire synchronously

more often than neurons belonging to functionally

unre-lated groups, in agreement with the Gestalt laws of

similar-ity and continusimilar-ity. From one moment to the next, however,

the particular subsets of neuronal groups that are firing

to-gether changes, so that a large number of coherent patterns

is continuously generated. This results in a calculated

elec-troencephalogram (EEG) that shows waxing and waning

Review

T o n o n i e t a l . – C o m p l e x i t y a n d c o h e r e n c y

478

T r e n d s i n C o g n i t i v e S c i e n c e s – V o l . 2 , N o . 1 2 , D e c e m b e r 1 9 9 8

According to the Oxford English Dictionary, something is complex when it

constitutes ‘a whole… comprehending various parts united or connected

to-gether’, especially ‘parts or elements not simply coordinated, but involved in

various degrees of subordination’. While we think that we recognize

com-plexity when we see it, comcom-plexity is an attribute that is often employed

generically without any attempt at conceptual clarity or, even less,

quantifi-cation. Recently, scientific approaches to complexity have attempted to retain

the intuitive, common sense notion of complexity by emphasizing the idea

that complex systems are neither completely regular nor completely random.

For example, neither a random string nor a periodically repeating string of

let-ters is complex, while a string of English text certainly is. More generally, any

system of elements arranged at random (e.g. gas molecules) or in a completely

regular or homogeneous way (molecules in a crystal lattice) is not complex.

By contrast, the arrangement and interactions of neurons in a brain or of

molecules in a cell is obviously extremely complex (see Fig.).

A number of complexity measures have been proposed, but only a few satisfy

the requirement of attaining small values for both completely random and

completely regular systems. In neurobiology, for example, one often

encoun-ters the term ‘dimensional complexity’ or just ‘complexity’ referring to the

so-called correlation dimension of EEG signals

a

. Its value appears to increase, for

instance, from sleep to waking states, or with brain maturation

b,c

. The

corre-lation dimension is a measure developed in the context of nonlinear

dynam-ics, which should be proportional, roughly speaking, to the number of

inde-pendent neuronal populations giving rise to an EEG signal

d

. But because the

correlation dimension would be higher for complete independence than for

the mixture of functional segregation and integration that characterizes brain

dynamics, it violates the criterion for complexity mentioned above.

Complexity measures have been proposed in the context of algorithmic

information theory, which deals with the information necessary to generate

individual bit strings. For example, the well-known algorithmic (or

Kolmogorov) complexity is defined as the length of the shortest computer

program that generates a particular bit string

e

. While this measure is

appro-priately low for completely regular strings, it is highest for random strings,

and thus it too does not satisfy the above criterion for complexity.

Attempts at modifying the notion of algorithmic complexity in order

to capture ‘true complexity’ have recently been proposed

f,g

. The key idea is to

discount pure randomness or noise and measure complexity by the shortest

computer program capable of describing the remaining regularities. By

defi-nition, such a measure would be satisfactorily low both for random and

trivially regular strings, but would be high for systems incorporating a large

number of regularities that cannot be further reduced. Of course, insofar as

the notion remains algorithmic, it requires that the observer can distinguish

between what represents genuine organization and what is instead

random-ness or noise, and this may be difficult or even impossible. The length of the

description of the regularities is also highly dependent on the understanding

of the observer. Thus, a system that appears highly complex or random might

turn out to be considerably simpler once the organizing principles are

under-stood. Low-dimensional chaotic systems, for example, might appear random,

yet their behavior can be fully determined by as few as three equations.

The definitions of complexity considered in this review (see Fig. 2, main

article) are statistical measures that capture regularities based on the deviation

from independence (mutual information) among subsets of a system

h

. In this

way, noise can be distinguished from genuine regularities in a way that is

relatively independent of an observer’s understanding of the system’s

organ-ization. A degree of subjectivity remains, of course, in deciding which

variables to measure and in choosing the appropriate level of coarse-graining

for averaging. However, it is easy to show that these measures satisfy the

re-quirement of being low both for completely random and for trivially regular

(homogeneous) systems

h

.

References

a Babloyantz, A., Salazar, J.M. and Nicolis, C. (1985) Evidence of chaotic dynamics of

brain activity during the sleep cycle Phys. Lett. (A) 111, 152–156

b Anokhin, A.P. et al. (1996) Age increases brain complexity Electroencephalogr.

Clin. Neurophysiol. 99, 63–68

c Meyer-Lindenberg, A. (1996) The evolution of complexity in human brain

development: an EEG study Electroencephalogr. Clin. Neurophysiol. 99, 405–411

d Lutzenberger, W., Preissl, H. and Pulvermüller, F. (1995) Fractal dimension of

electroencephalographic time series and underlying brain processes Biol. Cybern.

73, 477–482

e Kolmogorov, A.N. (1965) Three approaches to the quantitative definition of

information Inf. Trans. 1, 3–11

f Crutchfield, J.P. and Young, K. (1989) Inferring statistical complexity Phys. Rev.

Lett. 63, 105–108

g Gell-Mann, M. and Lloyd, S. (1996) Information measures, effective complexity,

and total information Complexity 2, 44

h Tononi, G., Sporns, O. and Edelman, G.M. (1994) A measure for brain complexity:

relating functional segregation and integration in the nervous system Proc. Natl.

Acad. Sci. U. S. A. 91, 5033–5037

Box 2. Different kinds of complexity

Tononi et al. (1998) Trends Cogn Sci 2, 474

As organizações estruturais e

funcionais do cérebro têm

caracterís/cas de redes complexas –

como topologia de mundo pequeno,

hubs altamente conectados e

modularidade –, tanto na escala do

cérebro inteiro (revelada por técnicas

de neuroimagem em humanos) como

na escala celular (revelada por

estudos em animais).

Bullmore & Sporns (2009) Nat Rev Neurosci 10, 186

(9)

Modelos de redes complexas

para as redes cerebrais

(anatômicas e funcionais)

Redes de mundo pequeno são um modelo atraente para

a organização das redes anatômicas e funcionais do

cérebro porque a topologia de mundo pequeno permite

conciliar duas formas dis/ntas de processamento de

informação: segregada (especializada) e distribuída

(integrada).

Basset & Bullmore (2006) The Neuroscien/st 12, 512

(10)

Determinada por dados

• 

Incorpora informação quan/ta/va sobre a arquitetura

de redes cerebrais reais

The image cannot be displayed. Your computer may not have enough memory to open the image, or the image may have been corrupted. Restart your computer, and then open the file again. If the red x still appears, you may have to delete the image and then insert it again.

Mazza et al., 2004

Olfactory Epithelium

Olfactory Bulb

Olfactory cortex

Simões-de-Souza & Roque, 2004

Sirosh & Miikkulainen, 1994 2004

(11)

Escala espacial: macro ou microscópica

Global: connec/vity between cor/cal areas

Short-scale cortical connectivity: layers

5915756 – Introdução à Neurociência Computacional – Antonio Roque – Aula 1

15

• A estrutura organizacional das camadas (padrões de laminação e de conexões

excitatórias e inibitórias entre as células) parece ser basicamente a mesma em

todas as áreas corticais.

• A figura abaixo ilustra o padrão geral de conexões excitatórias entre camadas

corticais.

Local: connec/vity between neurons in cor/cal layers/

columns

FIG. 1. Schematic of the basic circuitry of the dentate gyrus and the changes to the net-work during sclerosis. A: relational representa-tion of the healthy dentate gyrus illustrating the network connections between the 8 major cell types: GC, granule cell; BC, basket cell; MC, mossy cell; AAC, axo-axonic cells; MOPP, molecular layer interneurons with axons in per-forant-path termination zone; HIPP, hilar inter-neurons with axons in perforant-path termina-tion zone; HICAP, hilar interneurons with ax-ons in the commissural/associational pathway termination zone; and IS, interneuron selective cells. Schematic shows the characteristic loca-tion of the various cell types within the 3 layers of the dentate gyrus. Note, however, that this diagram does not indicate the topography of axonal connectivity (present in both the struc-tural and functional dentate models) or the so-matodendritic location of the synapses (incor-porated in the functional network models). B1:

schematic of the excitatory connectivity of the healthy dentate gyrus is illustrated (only cell types in the hilus and granule cells are shown). Note that the granule cell axons (the mossy fibers) do not contact other granule cells in the healthy network. B2: schematic of the dentate

gyrus at 50% sclerosis shows the loss (indicated by the large✕symbols) of half the population of all hilar cell types and the 50% sprouting of mossy fibers that results in abnormal connec-tions between granule cells (note that, unlike in this simplified schematic, all granule cells formed sprouted contacts in the structural and functional models of sclerosis; thus progressive increase in sprouting was implemented by in-creasing the number of postsynaptic granule cells contacted by single sprouted mossy fibers; seeMETHODS). C: schematics of 3 basic network topologies: regular, small-world, and random. Nodes in a regular network are connected to their nearest neighbors, resulting in a high de-gree of local interconnectedness (high cluster-ing coefficient C), but also requircluster-ing a large number of steps to reach other nodes in the network from a given starting point (high aver-age path length L). Reconnection of even a few of the local connections in a regular network to distal nodes in a random manner results in the emergence of a small-world network, with a conserved high clustering coefficient (C) but a low average path length (L). In a random net-work, there is no spatial restriction on the con-nectivity of the individual nodes, resulting in a network with a low average path length L but also a low clustering coefficient C. 1568 DYHRFJELD-JOHNSEN ET AL.

J Neurophysiol•VOL 97 • FEBRUARY 2007 •www.jn.org

at CAPES - Usage on September 27, 2012

http://jn.physiology.org/

Downloaded from

equivalent random graph has the same numbers of nodes and links as the graph (representing a particular degree of sclerosis) to which it is compared, although the nodes have no representation of distinct cell types and possess uniform connection probabilities for all nodes. For example, the equivalent random graph for the control (0% sclerosis) structural model has about a million nodes and the same number of links as in the control structural model, but the nodes are uniform (i.e., there is no “granule cell node,” as in the structural model) and the links are randomly and uniformly dis-tributed between the nodes.

CONSTRUCTION OF THE FUNCTIONAL MODEL. The effects of struc-tural changes on network excitability were determined using a real-istic functional model of the dentate gyrus (note that “functional” refers to the fact that neurons in this model network can fire spikes, receive synaptic inputs, and the network can exhibit ensemble activ-ities, e.g., traveling waves; in contrast, the structural model has nodes that exhibit no activity). The functional model contained biophysically realistic, multicompartmental single-cell models of excitatory and inhibitory neurons connected by weighted synapses, as published previously (Santhakumar et al. 2005). Unlike the structural model, which contained eight cell types, the functional model had only four cell types, as a result of the insufficient electrophysiological data for simulating the other four cell types. The four cell types that were in the functional model were the two excitatory cells (i.e., the granule cells and the mossy cells) and two types of interneurons (the somatically projecting fast spiking basket cells

and the dendritically projecting HIPP cells; note that these repre-sent two major, numerically dominant, and functionally important classes of dentate interneurons, corresponding to parvalbumin- and somatostatin-positive interneurons; as indicated in Table 1, basket cells and HIPP cells together outnumber the other four interneuronal classes by about 2:1). Because the functional model had a smaller proportion of interneurons than the biological dentate gyrus, control simulations (involving the doubling of all inhibitory conductances in the network) were carried out to verify that the observed changes in network excitability during sclerosis did not arise from decreased inhibition in the network, i.e., that the conclusions were robust (see

RESULTSandAPPENDIXB3).

Although the functional model was large, because of computa-tional limitations, it still contained fewer neurons (a total of about 50,000 multicompartmental model cells) than the biological den-tate gyrus (about one million neurons) or the full-scale structural model (about one million nodes). Because of this 1:20 reduction in size, a number of measures had to be taken before examining the role of structural changes on network activity. First, we had to build a structural model of the functional model itself (i.e., a graph with roughly 50,000 nodes) and verify that the characteristic changes in network architecture observed in the full-scale struc-tural model of the dentate gyrus occur in the 1:20 scale strucstruc-tural model (graph) of the functional model as well. Second, certain synaptic connection strengths had to be adjusted from the experi-mentally observed values (see following text).

TABLE1. Connectivity matrix for the neuronal network of the control dentate gyrus

Granule Cells Mossy Cells Basket Cells Axo-axonicCells MOPPCells HIPP Cells HICAP Cells IS Cells

Granule cells X 9.5 15 3 X 110 40 20

(1,000,000) X 7–12 10–20 1–5 X 100–120 30–50 10–30

ref. [1–5] ref. [6] ref. [7] ref. [6–9] ref. [6,7,9] ref. [6] ref. [4,10,11] ref. [4,7,10,11] ref. [7]

Mossy cells 32,500 350 7.5 7.5 5 600 200 X

(30,000) 30,000–35,000 200–500 5–10 5–10 5 600 200 X

ref. [11] ref. [4,11–13] ref. [12,13] ref. [13] ref. [13] ref. [14] ref. [12,13] ref. [12,13] ref. [15]

Basket cells 1,250 75 35 X X 0.5 X X

(10,000) 1,000–1,500 50–100 20–50 X X 0–1 X X

ref. [16,17] ref. [4,16–19] ref. [11,16,17,19] ref. [16,17,20,21] ref. [18] ref. [18] ref. [18] ref. [18] ref. [10,20]

Axo-axonic cells 3,000 150 X X X X X X

(2,000) 2,000–4,000 100–200 X X X X X X

ref. [4,22] ref. [4,18,22] ref. [4,5,11,14,23] ref. [5,18] ref. [5,18] ref. [5,18] ref. [5,18] ref. [5,18] ref. [5,18,19]

MOPP cells 7,500 X 40 1.5 7.5 X 7.5 X

(4,000) 5,000–10,000 X 30–50 1–2 5–10 X 5–10 X

ref. [11,14] ref. [14] ref. [14,24] ref. [14,25] ref. [14,26] ref. [14,25] ref. [14,20,25] ref. [14,25] ref. [14,15]

HIPP cells 1,550 35 450 30 15 X 15 X

(12,000) 1,500–1,600 20–50 400–500 20–40 10–20 X 10–20 X

ref. [11] ref. [4,11,20] ref. [4,11,12,27,28] ref. [4,11,20] ref. [20,25] ref. [25] ref. [14,20,25] ref. [25] ref. [15,20]

HICAP cells 700 35 175 X 15 50 50 X

(3,000) 700 30–40 150–200 X 10–20 50 50 X

ref. [5,29,30] ref. [4,11,20] ref. [20] ref. [4,11,20] ref. [20] ref. [14,20] ref. [20] ref. [20]

IS cells X X 7.5 X X 7.5 7.5 450

(3,000) X X 5–10 X X 5–10 5–10 100–800

ref. [15,29,30] ref. [15] ref. [15] ref. [15,19] ref. [15] ref. [19] ref. [19] ref. [15] Cell numbers and connectivity values were estimated from published data for granule cells, Mossy cells, basket cells, axo-axonic cells, molecular layer interneurons with axons in perforant-path termination zone (MOPP), hilar interneurons with axons in perforant-path termination zone (HIPP), hilar interneurons with axons in the commissural/associational pathway termination zone (HICAP), and interneuron-selective cells (IS). Connectivity is given as the number of connections to a postsynaptic population (row 1) from a single presynaptic neuron (column 1). The average number of connections used in the graph theoretical calculations is given in bold. Note, however, that the small-world structure was preserved even if only the extreme low or the extreme high estimates were used for the calculation of L and C (for further details, seeAPPENDIXB1(3). References given correspond to:1Gaarskjaer (1978);2Boss et al. (1985);3West (1990); 4Patton and McNaughton (1995);5Freund and Buzsa´ki (1996);6Buckmaster and Dudek (1999);7Acsa´dy et al. (1998);8Geiger et al. (1997);9Blasco-Ibanez et

al. (2000);10Gulya´s et al. (1992);11Buckmaster and Jongen-Relo (1999);12Buckmaster et al. (1996);13Wenzel et al. (1997);14Han et al. (1993);15Gulya´s et

al. (1996);16Babb et al. (1988);17Woodson et al. (1989);18Halasy and Somogyi (1993);19Acsa´dy et al. (2000);20Sik et al. (1997);21Bartos et al. (2001);22Li

et al. (1992);23Ribak et al. (1985);24Frotscher et al. (1991);25Katona et al. (1999);26Soriano et al. (1990);27Claiborne et al. (1990);28Buckmaster et al. (2002a); 29Nomura et al. (1997a);30Nomura et al. (1997b).

1569 NETWORK REORGANIZATION IN EPILEPSY

J Neurophysiol•VOL 97 • FEBRUARY 2007 •www.jn.org

at CAPES - Usage on September 27, 2012

http://jn.physiology.org/

Downloaded from

(12)

Arquitetura microscópica (córtex)

• 

Subdivisão do neocórtex em 6 camadas

• 

As camadas diferem em termos de

densidades e /pos de células

(13)
(14)
(15)

Arquitetura microscópica (córtex)

conexões aferentes e eferentes

• 

Aferentes

– 

Externas (não-locais). Origem:

• 

Tálamo (aferentes tálamo-cor/cais)

principalmente para a camada IV

• 

Outras áreas cor/cais (aferentes cór/co-cor/cais) via matéria branca

principalmente para as camadas

superficiais

– 

Entradas vindas de neurônios cor/cais na

vizinhança local

• 

Eferentes

– 

Eferentes cór/co-cor/cais dos neurônios

piramidais das camadas II/III

– 

Eferentes cór/co-talâmicos dos

neurônios piramidais da camada VI

– 

Axônios de neurônios piramidais grandes

da camada V para o tronco encefálico e a

medula espinhal

Abeles, 1991

(16)

Arquitetura microscópica: conexões ver/cais

E

à

E

E

à

I

I

à

E

I

à

I

Binzegger et al., 2004

Cada figura mostra a proporção do total de sinapses no córtex visual primário do gato que

existe entre os /pos de neurônios indicados. Os números totais de sinapses da cada /po

também estão indicados.

13.6 x 10

10

2.1 x 10

10

2.4 x 10

10

0.4 x 10

10

(17)

Arquitetura microscópica (córtex):

Conexões horizontais

• 

Sinapses locais: feitas por colaterais dos axônios em um raio de

cerca de 0,5 mm (todos os /pos de neurônios).

• 

Conexões intrínsecas de longo alcance: feitas por neurônios

piramidais, alcançam distâncias de vários milímetros passando

pela matéria cinzenta.

• 

Conexões extrínsicas de longo alcance: feitas por neurônios

piramidais através da matéria branca.

Voges et al., 2010

(18)

Arquitetura microscópica (córtex):

Conexões locais

• 

A probabilidade de uma conexão sináp/ca entre dois

neurônios cor/cais adjacentes cai para zero a uma

distância horizontal de cerca de 0,5 mm

Hellwig, 2000

Boucsein et al., 2011, 2000

(19)

Arquitetura microscópica (córtex):

Conexões de longo alcance

• 

As conexões intrínsecas de longo alcance formam um

padrão de “retalhos”: os neuronios piramidais projetam seus

axônios para agrupamentos celulares

Lund et al., 2003

Voges et al., 2007

(20)

Arquitetura macroscópica

(21)

Arquitetura macroscópica (córtex)

Bressler & Menon, 2010

§

As áreas cerebrais são tratadas como nós da rede

§

As conexões entre os nós são reveladas por diferentes

(22)

Imagem por Tensor de Difusão

(diffusion tensor imaging – DTI)

• 

Cada nó corresponde a uma

área do cérebro

• 

Não permite a determinação

das conexões dentro de cada

área

• 

Conexões binárias (há/não

há, sem ponderação)

• 

Não permite determinar os

atrasos nas conexões

• 

Não permite dis/nguir entre

conexões ausentes ou

desconhecidas

Honey et al., 2007

(23)

Arquitetura macroscópica (córtex)

• 

Estrutura hierárquica e modular

• 

“Clube de ricos”

Van den Heuvel & Sporns, 2011

Meunier et al., 2010

(24)

Juntando tudo: modelos de

neurônios e sinapses em uma rede

com uma dada arquitetura

(alguns modelos selecionados)

(25)

Redes de Erdős–Rényi: modelo de Brunel

• 

Neurônios LIF: 80% excitatórios, 20% inibitórios

• 

Conec/vidade esparsa (# conexões k << # neurônios N)

N

E

= 10.000

N

I

= 2.500

p= 0,1

C

E

= conexões

recebidas pelas

células E = 1.000

C

I

= 250

(26)

Tipos de a/vidade em um

modelo de rede neuronal

(a)  Assíncrona regular: a a/vidade da

população é aproximadamente

constante e os neurônios

individuais disparam de forma

regular;

(b)  Síncrona regular: Tanto a

a/vidade populacional como a dos

neurônios individuais são

oscilatórias;

(c)  Síncrona irregular: a a/vidade

populacional oscila e os neurônios

individuais disparam de forma

irregular;

(d)  Assíncrona irregular: a a/vidade

populacional é aproximadamente

constante e os neurônios

individuais disparam de forma

irregular.

Vogels et al. (2005)

(27)

Estado balanceado

• 

O córtex opera em um estado

balanceado em que os valores

médios das correntes de entrada

excitatória e inibitória em um

neurônio se cancelam

mutuamente.

• 

Os disparos de um neurônio são

causados por flutuações em torno

da entrada média líquida.

• 

Isso explica os disparos irregulares

dos neurônios (parecendo ruído) no

estado assíncrono irregular (AI).

(28)

Redes de Erdős–Rényi: modelo de Vogels e AbboZ

(29)

Plas/cidade em um modelo de sistema sensorial:

Estudos sobre lesões

(30)

Modelo com arquitetura cor/cal microscópica

(31)
(32)

Modelo com arquitetura cor/cal microscópica

(modelos de neurônios individuais biofisicamente detalhados)

•  Projeto Blue Brain: modelo de uma coluna cor/cal do rato

jovem com 10.000 neurônios reconstruídos morfologicamente

interconectados por 3x10

7

sinapses

•  Roda no supercomputador paralelo Blue Gene: 8912

processadores. Uma simulação de um dado tempo biológico

leva 2x mais que esse tempo para ser simulada.

hZp://bluebrain.epfl.ch/

(33)

Modelo que combina informação macro e microscópica:

dados de DTI, estrutura microscópica e neurônios que

reproduzem diferentes padrões de disparo

Izhikevich & Edelman, 2008

Esquerda. Propagação de ondas no

modelo. Disparos dos neurônios

excitatórios (inibitórios) indicados

por pontos vermelhos (pretos).

Direita. Sensibilidade à adição de um

único disparo.

(34)

Topological Determinants of Epileptogenesis in Large-Scale Structural and

Functional Models of the Dentate Gyrus Derived From Experimental Data

Jonas Dyhrfjeld-Johnsen,1,* Vijayalakshmi Santhakumar,1,* Robert J. Morgan,1Ramon Huerta,2Lev Tsimring,2

and Ivan Soltesz1

1Department of Anatomy and Neurobiology, University of California, Irvine; and2Institute for Nonlinear Science, University of California,

San Diego, California

Submitted 6 September 2006; accepted in final form 5 November 2006

Dyhrfjeld-Johnsen J, Santhakumar V, Morgan RJ, Huerta R, Tsimring L, Soltesz I. Topological determinants of epileptogenesis in

large-scale structural and functional models of the dentate gyrus derived from experimental data. J Neurophysiol 97: 1566 –1587, 2007. First published November 8, 2006; doi:10.1152/jn.00950.2006. In temporal lobe epilepsy, changes in synaptic and intrinsic properties occur on a background of altered network architecture resulting from cell loss and axonal sprouting. Although modeling studies using idealized networks indicated the general importance of network to-pology in epilepsy, it is unknown whether structural changes that actually take place during epileptogenesis result in hyperexcitability. To answer this question, we built a 1:1 scale structural model of the rat dentate gyrus from published in vivo and in vitro cell type–specific connectivity data. This virtual dentate gyrus in control condition displayed globally and locally well connected (“small world”) archi-tecture. The average number of synapses between any two neurons in this network of over one million cells was less than three, similar to that measured for the orders of magnitude smaller C. elegans nervous system. To study how network architecture changes during epilepto-genesis, long-distance projecting hilar cells were gradually removed in the structural model, causing massive reductions in the number of total connections. However, as long as even a few hilar cells survived, global connectivity in the network was effectively maintained and, as a result of the spatially restricted sprouting of granule cell axons, local connectivity increased. Simulations of activity in a functional dentate network model, consisting of over 50,000 multicompartmental single-cell models of major glutamatergic and GABAergic single-cell types, re-vealed that the survival of even a small fraction of hilar cells was enough to sustain networkwide hyperexcitability. These data indicate new roles for fractionally surviving long-distance projecting hilar cells observed in specimens from epilepsy patients.

I N T R O D U C T I O N

The dentate gyrus, containing some of the most vulnerable cells in the entire mammalian brain, offers a unique opportu-nity to investigate the importance of structural alterations during epileptogenesis. Many hilar cells are lost in both hu-mans and animal models after repeated seizures, ischemia, and head trauma (Buckmaster and Jongen-Relo 1999; Ratzliff et al. 2002; Sutula et al. 2003), accompanied by mossy fiber (granule cell axon) sprouting. In temporal lobe epilepsy, loss of hilar neurons and mossy fiber sprouting are hallmarks of seizure-induced end-folium sclerosis (Margerison and Corsellis 1966; Mathern et al. 1996), indicating the emergence of a fundamen-tally transformed microcircuit. Because structural alterations in

experimental models of epilepsy occur concurrently with mul-tiple modifications of synaptic and intrinsic properties, it is difficult to unambiguously evaluate the functional conse-quences of purely structural changes using experimental tech-niques alone.

Computational modeling approaches may help to identify the importance of network architectural alterations. Indeed, prior modeling studies of idealized networks indicated the importance of altered network architecture in epileptogenesis (Buzsa´ki et al. 2004; Netoff et al. 2004; Percha et al. 2005). However, to test the role of structural changes actually taking place during epileptogenesis, the network models must be strongly data driven, i.e., incorporate key structural and func-tional properties of the biological network (Ascoli and Atkeson 2005; Bernard et al. 1997; Traub et al. 2005a,b). Such models should also be based on as realistic cell numbers as possible, to minimize uncertainties resulting from the scaling-up of exper-imentally measured synaptic inputs to compensate for fewer cells in reduced networks.

Within the last decade, large amounts of high-quality exper-imental data have become available on the connectivity of the rat dentate gyrus both in controls and after seizures. From such data, we assembled a cell type–specific connectivity matrix for the dentate gyrus that, combined with in vivo single cell axonal projection data, allowed us to build a 1:1 scale structural model of the dentate gyrus in the computer. We characterized the architectural properties of this virtual dentate gyrus network using graph theoretical tools, following recent topological studies of biochemical and social networks, the electric grid, the Internet (Albert et al. 1999; Baraba´si et al. 2000; Eubank et al. 2004; Jeong et al. 2000; Watts and Strogatz 1998), the

Caenorhabditis elegansnervous system (Watts and Strogatz 1998), and model neuronal circuits (Lin and Chen 2005; Masuda and Aihara 2004; Netoff et al. 2004; Roxin et al. 2004). To test the functional relevance of the alterations observed in our structural model, we enlarged, by two orders of magnitude, a recently published 500-cell network model of the dentate gyrus, incorporating multicompartmental models for granule cells, mossy cells, basket cells, and dendritically pro-jecting interneurons reproducing a variety of experimentally determined electrophysiological cell-specific properties (San-thakumar et al. 2005).

Taken together, the results obtained from these data-driven computational modeling approaches reveal the topological

* These authors contributed equally to this work.

Address for reprint requests and other correspondence: J. Dyhrfjeld-Johnsen, Department of Anatomy and Neurobiology, University of California, Irvine, CA 92697-1280 (E-mail: jdyhrfje@uci.edu).

The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

J Neurophysiol97: 1566 –1587, 2007.

First published November 8, 2006; doi:10.1152/jn.00950.2006.

1566 0022-3077/07 $8.00 Copyright © 2007 The American Physiological Society www.jn.org

at CAPES - Usage on September 27, 2012

http://jn.physiology.org/ Downloaded from

• 

Modelo estrutural em escala 1:1 do giro denteado (GD) do rato (~1 milhão de neurônios)

• 

Modelo funcional em escala 20:1 do GD com mais de 50.000 modelos de neurônios

compar/mentais reduzidos

• 

Usado para avaliar o efeito de diferentes níveis de esclerose neural sobre a excitabilidade do

GD

• 

GD normal tem estrutura de mundo pequeno (baixo L e alto C)

• 

A esclerose aumenta as caracterís/cas de mundo pequeno da rede (L diminui e C aumenta)

!

!

Modelos para estruturas não cor/cais:

giro denteado do hipocampo

(35)

Colunas: neurônio pré-sináp/cos

(36)

Latin American School on Computational Neuroscience

LASCON

(37)

NeuroMat

Research, Innova/on and Dissemina/on

Center for Neuromathema/cs

Referências

Documentos relacionados

The aim of the present study was to design a schedule to synchronize the cell cycle of A549 cell line (a non-small cell lung cancer) and to analyze the possible association between

Los islotes donde se fundaron México–Tenochtitlan y México–Tlatelolco, no era los mejores sitios para erigir una ciudad, sobre todo Tlatelolco, que era una zona

– Bloco de Natação ter contribuído para “publicitar” as respetivas piscinas, pelo facto destas serem um pouco resguardadas e não estarem à vista da população; também o facto

Para tanto foi realizado um comparativo do IPTU Verde de Caraguatatuba/SP e do ICMS-E com os incentivos fiscais semelhantes presentes nos municípios de Araraquara/SP e São Carlos/SP e

Biofilm production and antibiotic resistance of Staphylococcus epidermidis in catheter related bloodstream infections. Antibiotic resistance and biofilm production among the

J. é uma mulher de 28 anos, solteira, mora com a mãe e os irmãos na região metropolitana de Curitiba. Foi atendida num CAPS, o qual tem por.. objetivo atender pacientes com

O período 1968-1971 representa para o sector da Saúde uma época histórica de reforma alargada dos serviços públicos na tentativa de efectivar um sistema

9 – Porcentagem de frutos nas classes de diâmetro das cultivares Royal Gala, Fuji Suprema e Pink Lady®, sobre os porta-enxertos M-9 e combinação Maruba/M9. Vacaria, RS, safra