physics0701089/mem.tex
1: %\documentclass[aps,preprint,amsmath,amssymb]{revtex4}
2: \documentclass[aps,twocolumn,amsmath,amssymb]{revtex4}
3: %\documentclass[aps,pre,preprint,groupedaddress,floatfix]{revtex4}
4: 
5: %\documentclass[aps,twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
6: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
7: %\documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
8: \usepackage{graphicx}% Include figure files
9: \usepackage{dcolumn}% Align table columns on decimal point
10: \usepackage{bm}% bold math
11: \usepackage{amssymb}
12: \usepackage{amsmath}
13: %\nofiles
14: 
15: \bibliographystyle{apsrev}
16: 
17: \begin{document}
18: 
19: \title{On the Separability of Attractors in Grandmother Dynamic Systems
20: with Structured Connectivity}
21: 
22: 
23: \author{Luciano da Fontoura Costa}
24: \email{luciano@if.sc.usp.br}
25: \affiliation{Instituto de F\'{\i}sica de S\~{a}o Carlos,
26: Universidade de S\~{a}o Paulo, Av. Trabalhador S\~{a}o Carlense 400,
27: Caixa Postal 369, CEP 13560-970, S\~{a}o Carlos, S\~ao Paulo, Brazil}
28: 
29: \date{22nd Dec 2006}
30: 
31: 
32: \begin{abstract}
33: The combination of complex networks and dynamic systems research is
34: poised to yield some of the most interesting theoretic and applied
35: scientific results along the forthcoming decades.  The present work
36: addresses a particularly important related aspect, namely the
37: quantification of how well separated can the attractors be in dynamic
38: systems underlined by four types of complex networks
39: (Erd\H{o}s-R\'enyi, Barab\'asi-Albert, Watts-Strogatz and as well as a
40: geographic model).  Attention is focused on grandmother dynamic
41: systems, where each state variable (associated to each node) is used
42: to represent a specific prototype pattern (attractor).  By assuming
43: that the attractors spread their influence among its neighboring nodes
44: through a diffusive process, it is possible to overlook the specific
45: details of specific dynamics and focus attention on the separability
46: among such attractors.  This property is defined in terms of two
47: separation indices (one individual to each prototype and the other
48: considering also the immediate neighborhood) reflecting the balance
49: and proximity to attractors revealed by the activation of the network
50: after a diffusive process.  The separation index considering also the
51: neighborhood was found to be much more informative, while the best
52: separability was observed for the Watts-Strogatz and the geographic
53: models.  The effects of the involved parameters on the separability
54: were investigated by correlation and path analyses.  The obtained
55: results suggest the special importance of some measurements in
56: underlying the relationship between topology and dynamics.
57: \end{abstract}
58: 
59: 
60: \pacs{}
61: 
62: 
63: \maketitle
64: 
65: 
66: \vspace{0.5cm}
67: \emph{`Nothing in excess.' (Delphic proverb)}
68: 
69: 
70: \section{Introduction}
71: 
72: Most dynamic systems, from neuronal networks to pattern formation,
73: involves several interconnected variables, which can be properly
74: represented in terms of a graph (e.g.~\cite{West01,Bollobas90}) or
75: complex network (e.g.~\cite{Barabasi:surv, Dorogov:surv, Watts03,
76: Wasserman94, Newman03, Boccaletti05, Costa:survey}).  Such systems are
77: henceforth called \emph{complex dynamic systems}.  A state variable
78: $v(i)$ is normally associated to each node $i$, so that the complete
79: evolution of the system can be described by the $N
80: \times 1$ vector $\vec{v}^{(t)}$. In a neuronal network, for instance,
81: each neuron can be expressed as a vertex (or a node), while synapses
82: are represented by links (or edges) and the node activation by the
83: respective state variable.  Once the underlying connectivity of such a
84: system is represented in terms of its \emph{weight matrix}
85: $W$~\footnote{$W(j,i)=w$ whenever node $i$ is connected to node $j$ by
86: an edge with weight $w$.}, several important types of dynamics can be
87: subsumed as
88: 
89: \begin{equation}  \label{eq_dyn}
90: \vec{v}^{(t+1)} = f(W \vec{v}^{(t)}).  
91: \end{equation}
92: 
93: Any layer of the perceptron neuronal network
94: (e.g. ~\cite{Rosenblatt:1958,Haykin:1999}), for instance, is obtained
95: by substituting $f()$ by some abrupt function (e.g. hard limit or
96: sigmoid).  At the same time, the complete classic Hopfield model
97: (e.g.~\cite{Hopfield:1984,Haykin:1999}) can be represented by this
98: equation.  A simpler, linear model is obtained by making $f(x)=x$, so
99: that $\vec{v}^{(t+1)} = W \vec{v}^{(t)}$.  In case $W$ is also a
100: stochastic matrix (the transition matrix), this linear equation
101: subsumes all first order Markov chains, a particularly useful and
102: important dynamic model which is intrinsically associated to random
103: walks (e.g.~\cite{Rudnik:2004}), Markov chains
104: (e.g.~\cite{Bremaud:2001}) and diffusion
105: (e.g.~\cite{Crank:1980,Havlin:2000}).
106: 	
107: In many dynamic systems (e.g.~\cite{Haykin:1999,Boccara:2004}), the
108: codification of external stimuli as well as the results from the
109: network dynamics take place in the $N-$dimensional space defined by
110: the state variables.  For instance, one particularly pattern may be
111: represented as the state $\vec{v_1} = [1, 0, -1, 0.5, 1, 0]$, while
112: another pattern may be associated to $\vec{v_2} = [0, 1, -1, 0, 1,
113: 0]$.  This type of representation is called ~\emph{vector coding}
114: (e.g.~\cite{Georgopoulos:1986,Salinas:1994}). States which are near
115: any of such pattern-states are frequently understood to be associated
116: to that pattern.  For instance, in the case of the example above, a
117: third pattern similar to $\vec{v_1}$ will tend to be coded as a state
118: vector $\vec{v_3}$ such that $\delta = ||\vec{v_1}-\vec{v_3}||$ is
119: small.  Such a smooth, graded coding along the state space is
120: interesting because it allows for some robustness/redundancy and
121: flexibility/generality for the representation of the patterns, while
122: also favoring good dynamic properties such as where the patterns are
123: accessed through gradient-descent or similar methods
124: (e.g.~\cite{Haykin:1999,Rudnik:2004}).
125: 
126: However, it is also possible and interesting to consider the situation
127: where the patterns are associated to specific nodes, not to specific
128: points in the state space, such as in systems involving the so-called
129: \emph{grandmother cells} (e.g.~\cite{Barlow:1972, Gross:2002, 
130: Gazzaniga:1998}).  Extremely important systems including great part of
131: the mammals cortex are believed to be so organized.  In these cases, a
132: pattern is associated to a node $i$ such that high values of $v(i)$
133: signalizes the presence of that pattern.  This type of cells is
134: ubiquitous in several cortical areas (e.g.~\cite{McIlwain:1996,
135: Zeki:1990, Zeki:1999}).  For instance, cells which are highly specific
136: to hands and faces have been found in the inferior temporal
137: cortex~\cite{Gross:1969, Perrett:1982, Gross:2002}.  It is also known
138: (e.g.~\cite{Eichenbaum:1989,McIlwain:1996,Zeki:1999,Bosking:2002})
139: that much of the mammals cortical architecture is characterized by
140: spatial smoothness, in the sense that cells which are spatially close
141: tend to exhibit similar dynamics and response (i.e. state
142: correlations), except at eventual singularities involving
143: fractures~\cite{Zeki:1999,Bosking:2002}.  Such an organization reminds
144: of the smooth coding discussed above for dynamic systems where
145: patterns are represented by the overal state (i.e. vector coding).  It
146: is important to note that the smoothness property is not exclusive to
147: topographically organized systems~\footnote{By topographic system it
148: is henceforth meant that the network nodes have well-defined positions
149: along an $N-$dimensional space (usually the two-dimensional
150: space). The terms
151: \emph{geographic} or \emph{geometric networks} have also been used
152: in the graph theory and complex networks literatures.} such as the
153: mammals cortex.  In other words, even in networks where the nodes have
154: no specific position in an embedding space, neighboring grandmother
155: nodes may tend to have similar dynamics and response.  In such cases,
156: it is interesting to use the concept of progressive neighborhoods
157: around a node (e.g.~\cite{Faloutsos99,Newman_hier01,Cohen03,
158: Costa04,Costa06,Costa_Silva07})), defined by the immediate neighbors
159: of a node as well as its second neighbors, and so
160: on. Following~\cite{Costa04,Costa_Silva07,Costa_Silva07}, in this work
161: we organize the successive neighborhoods in terms of their
162: \emph{hierarchical level}.  Therefore, the immediate neighbors of a node $i$
163: are at hierarchy 1 of $i$, and so on.  In brief, the coding smoothness
164: property in grandmother systems is reflected by the fact that the
165: hierarchical neighbors of each node will tend to have responses
166: similar, though progressively diverging, to that of node $i$.
167: 
168: In dynamic networks with a finite number of nodes, the situation
169: considered herein, a particularly interesting problem arises as a
170: consequence of a tension between the grandmother coding and the
171: smoothness property, a phenomenon directly related to the limited
172: number of patterns which can be properly represented
173: (e.g.~\cite{Amit:1993}).  Before proceeding further, it is important
174: to provide a more objective characterization of such a problem.
175: Consider that $M$ prototypic, distinct, patterns are to be represented
176: in the network.  If each of such patterns is represented by a
177: respective prototypic neuron (node), in the sense that this neuron
178: will be the most highly activate when that pattern is invoked by the
179: network, we also need to allow for intermediate cells with graded
180: replies between pairs of nodes. Observe that such prototype nodes
181: typically act as attractors for the dynamics of the respective dynamic
182: system, being accessed, for instance, by using gradient descent and/or
183: random walks methods. If $M$ is too large, a point will be reached
184: where each network node will be required to represent each
185: prototypical pattern, running out of nodes for implementing the graded
186: responses.  Although this situation occurs at the very limit of the
187: network capacity, it is important to observe that it may still provide
188: a complete, invertible representation of the patterns (i.e. an
189: one-to-one mapping).  On the other hand, the smoothness of the coding
190: is undermined as the number of prototypical patterns increases.  In
191: addition to destroying the network potential for generality and
192: robustness, it will also become impossible to retrieve the patterns by
193: gradient-descent like mechanisms along the network.
194: Figure~\ref{fig:ex} illustrates two small topographic state spaces:
195: one (a) corresponding to the limit situation where each pattern is
196: associated to each node, the other (b) depicting a smooth map defined
197: by diffusion around five prototype nodes.  These two spaces were
198: assumed to have local connectivity, in the sense that each cell
199: communicates only with its most immediate neighbors in the orthogonal
200: lattice.  Observe that it is virtually impossible to devise an
201: effective retrieval or activation mechanism capable of getting to any
202: specific prototype node in Figure~\ref{fig:ex}(a). In addition, the
203: failure of any cell will imply in the irrecoverable lost of one
204: pattern (the origin of the \emph{gradmother cell} concept). On the
205: other hand, the state space in Figure~\ref{fig:ex}(b) allows a good
206: level of tolerance to failure (this also means some redundancy), at
207: the same time as simple dynamic mechanisms such as gradient descent
208: can be employed for retrieving and activating nodes.
209: 
210: 
211: \begin{figure}[htb]
212:   \begin{center} 
213:   \includegraphics[scale=0.5,angle=0]{im1_a.eps} \\
214:   (a) \\ \vspace{0.3cm}
215:   \includegraphics[scale=0.5,angle=0]{im1_b.eps} \\
216:   (b) \\
217:   \caption{Gray-level visualizations of a topographic state space
218:   with too many prototypical patterns (a) and of a smooth
219:   topographic state space (b). Both these spaces are underlined by
220:   4-neighborhood in the orthogonal lattice. The five prototype
221:   patterns in (b) have been marked in black for the sake of better
222:   visualization.~\label{fig:ex}}
223:   \end{center}
224: \end{figure}
225: 
226: For all that has been said above, it becomes very important to
227: quantify and study the pattern separability in dynamic systems such as
228: neuronal networks.  In the case of the dynamic systems which can be
229: fully represented by Equation~\ref{eq_dyn}, these two properties are
230: necessarily implied by the network connectivity, defined by the weight
231: matrix $W$, and the function $f()$.  The present work focuses
232: precisely on the investigation of the separability between patterns in
233: grandmother complex dynamic systems.  By considering a diffusive
234: process emanating from each prototype node, followed by a process
235: which assigns to each edge the difference between the activation
236: states at its head and tail nodes, a graded representation of the
237: prototype patterns can be obtained.  This allows full generality of
238: the reported investigations as an approximation to many types of
239: grandmother dynamic systems.  The accessibility to the original
240: prototypes is simulated through a preferential random walk (diffusion)
241: starting at every node of the network. After a given relaxation time,
242: the activation of the network is normalized so as to correspond to
243: occupancy probabilities and compared to the original prototypes
244: position.  This is done by considering two separation indices, one
245: individual, considering the resulting activity only at the prototype
246: nodes, and the other considering the activation not only at these
247: nodes, but also at their immediate neighbors.  These two indices, are
248: henceforth abbreviated as $s_{ind}$ and $s$, are obtained in terms of
249: the normalized geometric average of the activations at (or around)
250: each prototype node.  Therefore, the maximum value of these indices
251: (equal to one) is achieved only when the resulting activation is
252: totally and uniformly distributed among only the prototype nodes (for
253: $s_{ind}$) or among these nodes and their immediate neighbros (for
254: $s$).
255: 
256: The performed simulations and analyses first consider specific network
257: configurations, in order to gather insights about the influence of each
258: involved parameter, which are identified and discussed.  Subsequently,
259: a more systematic investigation of some parameter variations are
260: performed.  In order to infer the influence of the topologic
261: features on the dynamics properties (i.e. attractors separation), we
262: apply correlation analysis and path analysis.  A series of interesting
263: results are obtained regarding not only the separation indices, but
264: also the importance of several local and global topologic features
265: on the overall attractors separation.  Although the current work
266: concentrates on attractors in grandmother dynamic systems, several of
267: the results, measurements and methods can be immediately extended to
268: other relevant systems based on complex networks, including those
269: involving information transfer and retrieval.
270: 
271: This article starts by presenting a brief review of more closely
272: related works, and follows by describing the adopted notation and
273: methodologies.  Next, each of the four considered theoretic network
274: models are briefly reviewed and their generation described.  The
275: diffusion procedure and `differentiation' of the state values adopted
276: to implement the attraction basins are described next, as well as the
277: diffusion-base mechanism for activation/retrieval of prototypes.  The
278: two separation indices are motivated and mathematically defined.  The
279: simulation results and respective discussion are then presented,
280: followed by the correlation and path analyses of the effects of the
281: topologic features of the network on the respective attractors
282: separability.  Several interesting findings and effects are identified
283: and discussed.  The article concludes with a general synthesis of the
284: reported investigation and with identification of possible future
285: developments.
286: 
287: 
288: \section{A Brief Review of Previous Developments} \label{sec:review}
289: 
290: The subject of attractor characterization has been extensively
291: addressed in the literature, so we limit ourselves to reviewing a
292: small subset of those works which are more general or more directly
293: related to the investigations being considered in the present work.  
294: 
295: A comprehensive study of attractor networks has been presented by
296: Amit~\cite{Amit:1992}.  Torres et al. have considered the storage
297: capacity of attractor neural networks in which the synapses can undergo
298: depression~\cite{Torres:2002} and found that the memory capacity
299: decreases with the intensity of the depression. Amit and
300: Brunel~\cite{Amit:1993} investigated learning in attractor networks
301: considering the effects of stimuli in a network with wired-up
302: patterns.  Topographic neuronal networks involving grandmother cells
303: have been extensively studied by T. Kohonen and collaborators
304: (e.g.~\cite{Kohonen:1982, Kohonen:2003, Kohonen:2006}) in the form of
305: the self-organizing map (SOM) or Kohonen networks.  Here, the neurons
306: are spatially distributed and develop specificity to prototype pattern
307: stimuli by influencing the weights of neighboring nodes.  Therefore,
308: spatially adjacent smooth attraction basins are defined.
309: 
310: Several works have brought together the two important areas of complex
311: networks and dynamic systems, including neuronal networks.  At least
312: two excellent surveys have been written about this important
313: issue~\cite{Newman03, Boccaletti05}.  The problem of neuronal
314: structure and dynamics has also been a subject of growing interest,
315: including but by no means limited to~\cite{Costa_Salou:2005,
316: Peichl_Wassle:1983, Costa_BM:2003, SI_BM:2003, GA_hipp:2005,
317: Ooyen:1995, Ooyen:1999, Ooyen:2001, Ooyen:2004}.  Stauffer et
318: al.~\cite{Stauffer_etal:2003} investigated the performance of diluted
319: Hopfield networks underlined by the BA model.  The influence of the
320: network topology on the dynamics of neural networks has been
321: investigated also by Costa and Stauffer~\cite{Costa_Stauffer:2003},
322: who considered spatial neural networks and concluded that its
323: performance increased with the spatial uniformity of the cells
324: distribution.  Torres et al.~\cite{Torres:2004}, showed that the
325: pattern capacity of an attractor neural network with scale free
326: topology is higher than for a random-diluted network with the same
327: number of connections.  They also found that, at zero temperature, the
328: performance of scale free nets improves for larger values of the
329: power-law exponent. Models of non-randomly diluted neuronal networks
330: whose connectivity is determined as a function of the shape of its
331: individual neurons (as well as their relative spatial positions) has
332: been reported by Costa and
333: collaborators~\cite{Costa_BM:2003,Costa_EPJB:2004}.  It was found that
334: the shape of individual neurons can have a great influence on the
335: respective memory capacity, suggesting that shapes more similar to
336: real neurons tend to have better attractor properties.  The influence
337: of the network topology on the recovery of patterns in recurrent
338: neuronal networks was addressed by Castillo et
339: al.~\cite{Castillo_etal:2004}. Their work showed that the retrieval
340: properties can be enhanced by considering connectivity more structured
341: than in random networks.  Morelli et al.~\cite{Morelli:2004}
342: investigated the memory capacity in associative networks and found
343: that the best performance is obtained at an intermediate level of
344: disorder. Zhou and Lipowski~\cite{Zhou_Lipowski:2005} investigated,
345: through analytic and simulation means, a general class of dynamic
346: systems on scale free networks with binary states.  They reported
347: important variations of performance with respect to the scale free
348: exponential coefficient. The effect of structured connections on the
349: interactive statistical mechanics algorithm for minimization of the
350: Bethe free energy (associated to Ising models) has been studied by
351: Ohkubo et al.~\cite{Ohkubo_etal:2005}.  The adaptation of the Sznajd
352: dynamics to take place over the network connectivity instead of its
353: states has been reported by Costa~\cite{L_sznaj:2005}, yielding
354: network realizations which are a consequence of an inherent dynamical
355: process.  Lu et al.~\cite{Lu_etal:2006} considered the effect of
356: regular, random, small-world and scale free topologies on Hopfield
357: networks.  They reported, among other findings, that the performance
358: improved with the local order of the connections, which seems to be in
359: agreement with~\cite{Costa_Stauffer:2003}. The periodicity of activity
360: in networks with small-world and scale-free topologies were
361: investigated by Paula et al.~\cite{Paula_etal:2006} who concluded,
362: among other findings, that periodic activity appears only for
363: relatively small networks. Perotti et al.~\cite{Perotti_etal:2006}
364: studied the interesting problem of associative memory on a growing
365: diluted Hopfield model which converges to a small-world, scale free
366: topology and showed that the performance of such a network is higher
367: than that of a randomly diluted network with the same
368: connectivity. More recently, Davey et al.~\cite{Davey:2006}
369: investigated sparse small world associative memory considering
370: Perceptron training under small world connectivity ranging from local
371: to global and found that non-symmetric connectivity networks exhibited
372: superior performance.  By considering random walks as a reference
373: model for implementing dynamics in complex networks, Costa et
374: al. investigated the correlations between the topology (node degree)
375: and activation (frequency of visits to nodes at equilibrium).  They
376: found that while full correlation is guaranteed for undirected
377: networks, it can vary substantially in directed networks such as
378: biological neuronal networks and the Internet, implying that topologic
379: hubs are not necessarily hubs of activity.  That work also identified
380: a relationship between scale free networks and the Zipf's
381: law~\cite{Newman_Zipf:2005}. A study of self-organizing models
382: underlined by complex networks related to mental processes has been
383: reported by Wedemann et al.~\cite{Wedemann:2006}. The investigation of
384: the relationship between topology and dynamics at higher spatial
385: scales (e.g. cortical areas) has also been addressed in an increasing
386: number of works (e.g.~\cite{Sporns:2002, Hilgetag:2002,
387: Hilgetag_etal:2000, Sporns_etal:2000, Sporns_etal:2004,
388: Costa_Sporns:2006, Costa_Sporns:2007, Costa_etal:2006}).
389: 
390: Increasing attention has also been drawn on the synchronization of
391: oscillations as the means for pattern encoding and retrieval
392: (e.g.~\cite{Arecchi:2004}).  Among the works related to the separation
393: of attractors, Arecchi~\cite{Arecchi:2005} has considered a metric
394: structure for the percept space, while taking into account the
395: separation between states.  Wang et al. investigated the influence of
396: the node degree distribution in the synchronization of two-layer
397: neural networks.  The criticality of coupling parameters on the
398: synchronization of an ensemble of identical neural networks with
399: small-world topology has been addressed by Wang et
400: al.~\cite{Wang_etal:2005}.
401: 
402: All in all, as far as the influence of connectivity on the performance
403: of dynamic systems for storing patterns is concerned, several of the
404: above reviewed works seem to indicate that better results tend to be
405: obtained by considering non-random connectivity, but at a level of
406: order that ranges from low to intermediate.
407: 
408: 
409: \section{Notation, Basic Concepts and Methodology}
410: 
411: This section covers the concepts and methods used in the current
412: investigation.  Its subsections present the basic concepts and
413: measurements in complex networks, the four considered theoretic
414: network models, the procedure suggested to extend the prototype
415: influence through their successive neighborhoods in order to establish
416: the attraction basins, the diffusive way to activate the attractors,
417: the separation measurements, and the correlation and path analyses
418: considered in this work.
419: 
420: 
421: \subsection{Complex Networks Concepts and Topologic Measurements} 
422: \label{sec:conc}
423: 
424: A \emph{graph} $\Gamma$ involves a set $V$ of $N$ nodes interconnected
425: by a set $U$ of $E$ edges, i.e. $\Gamma = (V,U)$.  Each directed edge
426: linking a node $i$ to a node $j$ is represented as $(i,j)$.  Such a
427: graph can be conveniently expressed in terms of its \emph{adjacency
428: matrix} $K$, so that $K(j,i)=1$ whenever an edge $(i,j)$ exists
429: (otherwise, $K(j,i)=0$).  A graph such that $K(i,j)=K(j,i)=1$ is said
430: to be non-oriented.  Although the present work focuses on this type of
431: graph, all reported developments can be extended to directed graphs.
432: The graphs considered henceforth are also devoid of self-connections
433: (i.e. $K(i,i)=0$ for every node $i$).  A \emph{complex network} is
434: henceforth understood as a graph exhibiting a particularly intricated
435: structure, in the sense of differing from a random graph of the
436: Erd\H{o}s-R\'enyi type~\cite{ErdosRenyi:1959,ErdosRenyi:1960}.
437: However, because of statistic fluctuations, a random graph can also
438: exhibit intricated structure.
439: 
440: Given a generic node $i$, some measurements can be associated to it
441: (e.g.~\cite{Barabasi:surv,Newman03,Costa:survey}).  Its \emph{degree}
442: $k(i)$ is defined as the number of edges attached to it, so that
443: 
444: \begin{equation}
445:   k(i) = \sum_{p=1}^{N} K(p,i). 
446: \end{equation}
447: 
448: The immediate neighbors of a node $i$ correspond to those nodes which
449: are directly attached to it. The \emph{clustering coefficient} of a
450: node $i$ expresses the degree of connectivity among its immediate
451: neighbors and can be calculated as
452: 
453: \begin{equation}
454:   cc(i) = \frac{e(i)}{e_{max}},
455: \end{equation}
456: 
457: where $e(i)$ is the number of undirected edges between the immediate
458: neighbors of $i$ and $e_{max}$ is the maximum possible number of such
459: connections, given as $e_{max} = e(i)(e(i)-1)/2$.
460: 
461: The \emph{shortest path} between any two nodes $i$ and $j$ corresponds
462: to the minimal set $SP$ of adjacent edges connecting those two nodes.
463: A network in which all nodes can be reached, through paths, from any
464: other node is henceforth said to be \emph{connected}.  The respective
465: \emph{shortest path length} $sp$ is defined as the number of edges in $SP$.
466: Therefore, the immediate neighbors of a node $i$ can be alternatively
467: defined as the nodes which are at shortest path length of 1 from node
468: $i$.  The second neighborhood (or neighbors of second hierarchy) are
469: those nodes which are at shortest path length of 2 from $i$, and so
470: on.  The maximum shortest path length between any pair of nodes is
471: defined as the network \emph{diameter}, henceforth abbreviated as
472: $diam$.
473: 
474: Given a set $R$ of $M$ reference (or seed) nodes, it is possible to
475: obtain the respective \emph{Voronoi} tessellation~\cite{Stoyan:1996}
476: of the network~\cite{Costa_Silva07} with respect to the seeds, which
477: partitions the $N$ nodes into $M$ connected subgraphs associated to
478: each of the reference nodes.  This can be achieved by, for each node
479: $i$, identifying which of the reference nodes is closest (in the sense
480: of shortest path) to $i$ and assigning it to that region.  Given two
481: nodes $i$ and $j$, whose respective immediate neighbors are
482: represented by the sets $n(i)$ and $n(j)$, we define their
483: \emph{common neighbors} as the set $c(i) = n(i) \cap n(j)$.  
484: Given the Voronoi partition of a network, it is interesting to devise
485: a measurement capable of expressing the uniformity of the areas $A(i)$
486: (i.e. number of nodes) of each of the $M$ partitioned regions $i$.
487: This can be conveniently achieved by considering the \emph{geometric
488: average} $\left[ A \right]$ of the areas, i.e.
489: 
490: \begin{equation}
491:   \left[ A \right] = \left( \prod_{i=1}^{M} A(i) \right)^{1/M},
492: \end{equation}
493: 
494: Observe that the geometric average implies a high penalty on higher
495: variations of the Voronoi areas, therefore providing a more strict
496: quantification of the homogeneity of those areas.
497: Figure~\ref{fig:avgs} illustrates the comparison between the
498: arithmetic and geometric average considering two measurements $p$
499: (with $0 \leq p \leq 1$) and $q = 1-p$.  The ability of the geometric
500: average in quantifying the similarity between $p$ and $q$ is evident
501: from this figure.  While dispersion-related measurements
502: (e.g. standard deviation and entropy of the state activations) could
503: be used, they could not be interepreted as quantifications of the
504: activations.
505: 
506: 
507: \begin{figure}[htb]
508:   \begin{center} 
509:   \includegraphics[scale=0.55,angle=0]{avgs.eps} \hspace{0.5cm}
510:   \caption{The arithmetic ($*$) and geometric ($+$)
511:   averages between the values $p$ and $q = 1-p$. The
512:   geometric average allows the quantification of the 
513:   uniformity between the values of $p$ and $q$, with a
514:   peak at $p = q = 0.5$.~\label{fig:avgs}}
515:   \end{center}
516: \end{figure}
517: 
518: As the geometric average $u$ for the $M$ Voronoi areas varies as $0
519: \leq u \leq 1/M$, it is convenient to redefined $\left[ A \right]$ as
520: 
521: \begin{equation}
522:  \left[ A \right] = M \left( \prod_{i=1}^{P} A(i) \right)^{1/M}.
523: \end{equation}
524: 
525: The \emph{matching index}~\cite{Sporns:2002,Hilgetag:2002} of a pair
526: of nodes $i$ and $j$, expressing the relative degree of overlap
527: between the immediate neighborhoods of those nodes, can be calculated
528: as
529: 
530: \begin{equation}
531:   mi(i,j) = \frac{n(i) \cap n(j)}{n(i) \cup n(j)}.
532: \end{equation}
533: 
534: This measurement gives the fraction of immediate neighbors of $i$ and
535: $j$ which are common neighbors to them both.  For instance, for the
536: situation depicted in Figure~\ref{fig:common}, we have that the
537: matching index of the pair of nodes $P1$ and $P2$ is given as
538: $m(P1,P2) = 2/8 = 0.25$.
539: 
540: 
541: \subsection{Complex Networks Theoretic Models}
542: 
543: Four theoretic models of complex networks are considered in the
544: present work: Erd\H{o}s-R\'enyi (ER), Barab\'asi-Albert (BA),
545: Watts-Strogatz (WS) and a geographic model (GG).  Reflecting their
546: different natures and organizing principles, these models correspond
547: to a significant portion of the complex networks found in nature,
548: providing therefore a representative choice of models for the current
549: study of attractor separability. In order to account for a more
550: coherent comparison between the separability of attractors in these
551: four models, each comparison always consider $N$ and $\left< k
552: \right>$ for each model to be as similar as possible~\footnote{Observe
553: that it is not possible to ensure identical values of such parameters
554: in all models. For instance, the average degree $\left< k \right>$
555: will naturally vary within a limited interval in all models.}.  In
556: this work, the average degree of the BA model (defined by the
557: parameter $m$) is always taken as a reference for defining the average
558: degree of the other models.  The methodology for constructing such
559: models are described in the following subsections.
560: 
561: 
562: \subsubsection{Erd\H{o}s-R\'enyi}
563: 
564: \emph{Random networks}, also called Erd\H{o}s-R\'enyi --
565: ER, were among the first models of stochastic networks to be
566: extensively
567: studied(e.g.~\cite{Rapoport:1957,ErdosRenyi:1959,ErdosRenyi:1960}).
568: These networks are characterized by constant probability of having a
569: connection between any of the possible pairs of nodes, and are
570: therefore related to Poisson processes.  The connectivity of ER
571: networks can generally be well approximated in terms of its average
572: degree, implying that such networks are similar to regular networks,
573: characterized by having the same degree at any of its nodes. As a
574: consequence of its indiscriminate connectivity and largely regular
575: organization, the ER type of networks does not typically provide a
576: good model of natural structures and phenomena, where the connections
577: tend to follow more purposive and specific rules.
578: 
579: In an ER graph, each possible connection between each possible pair of
580: nodes, has constant probability $\gamma$ of existence.  Such networks
581: can be easily created through Monte Carlo simulation by making
582: $K(i,j)=K(j,i)=1$, with $i \neq j$, with probability $\gamma$.
583: Observe that the construction of ER networks consider only two
584: parameters: $N$ and $\gamma$.  In order to obtain values of $\left< k
585: \right>$ similar to the BA reference, we enforce $\gamma = 2m/(N-1)$,
586: where $m$ is a parameter of the BA model (see next subsection).
587: 
588: 
589: \subsubsection{Barab\'asi-Albert}
590: 
591: The so-called \emph{Barab\'asi-Albert
592: model}~\cite{Albert:1999,Barabasi:Linked,Barabasi:surv} -- BA, belongs
593: to the important class of \emph{scale free} networks.  This type of
594: network is characterized by the fact that the loglog plot of their
595: node degrees tends to a straight line, implying the absence of any
596: characteristic scale.  In other words, the node distribution in such a
597: network follows a \emph{power law}.  One of the most important
598: properties of scale free networks
599: (e.g.~\cite{Barabasi:surv,Barabasi:Linked}), when compared to models
600: such as the ER, is the higher probability of existence of \emph{hubs},
601: i.e. nodes with particularly high degree.  Such special nodes are
602: particularly important in defining the connectivity and topologic
603: features of the network, such as the average shortest path length.
604: For instance, the fact that a hub connects to many nodes immediately
605: implies that the shortest path between any of these nodes will be at
606: most equal to 2 edges.  Several important natural and human-made
607: structures -- including the Internet, WWW, protein interaction and
608: even scientific collaborations -- have been found to exhibit the scale
609: free property (e.g.~\cite{Barabasi:surv,Barabasi:Linked,Newman03}).
610: The BA model incorporates the so-called
611: \emph{rich-get-richer} paradigm because of its attachment of links being
612: preferential to the degree of existing nodes.
613: 
614: In the current work, BA networks are generated starting with $m0$
615: randomly connected nodes.  At each subsequent step, a new node with
616: $m$ edges is added to the network, with each of the $m$ edges being
617: attached to previous network nodes preferentially to their degree.
618: Therefore, each new connection is more likely to be established with
619: previous nodes with high degree, implementing the `rich get richer'
620: paradigm.  As with the ER model, the construction of BA networks also
621: involves only two parameters: $N$ and $m$.
622: 
623: 
624: \subsubsection{Watts-Strogatz}
625: 
626: Historically, the \emph{small world networks} of Watts and
627: Strogatz~\cite{Watts_Strogatz:1998,Watts03} followed the random
628: networks of Erd\H{o}s, R\'enyi and collaborators.  Small world
629: networks are exactly as implied by their name, i.e. the average
630: shortest path length between their nodes tends to be small.  At the
631: same time, they also tend to be characterized by relatively high
632: clustering coefficient, implying that they local connectivity is
633: relatively high.  The small world property, which has been found to be
634: present in many interesting networks including ER and BA, has
635: important implications for the separation of attractors because it
636: implies that many prototype nodes will be near one another and,
637: consequently, possibly less separated as far as the dynamics is
638: concerned.  The WS model considered in this work, however, presents
639: some specific topologic organization which has potential
640: implications for the distribution of the prototypes.  More
641: specifically, this model is characterized by a relatively high
642: regularity and uniformity of local connectivity.
643: 
644: The Watts-Strogatz networks used in the current work have been
645: constructed by starting with a ring of nodes where each node is
646: connected to its $m$ clockwise and anti-clockwise nodes.  After such
647: an initial network is obtained, $\alpha$\% of the existing connections
648: are rewired at random.  This network model involves three parameters:
649: $N$, $m$ and $\alpha$.  All configurations in this work assume
650: $\alpha = 10$\%.
651: 
652: 
653: \subsubsection{A Geographic Model}
654: 
655: Geographic networks (e.g.\cite{Costa_Stauffer:2003, Costa:2003,
656: Spatial_Satorras:2004, Costa_Diambra:2005, Spatial_Barrat:2005,
657: Spatial_Newman:2006})) -- also called spatial, geometric or topologic
658: -- are characterized by the fact that their nodes have well-defined
659: spatial positions within an embedding space.  Frequently, the
660: connectivity in such networks is considered to be highly influenced by
661: the spatial adjacencies and/or spatial proximity between its nodes, in
662: the sense that two nodes which are adjacent or near one another will
663: have higher chances of being connected.  Therefore, geographic models
664: in small dimensional spaces (e.g. 2D or 3D) tend not to be small
665: world.  Such a property is particularly important as far as the
666: separation of the prototype nodes is concerned because, in principle,
667: this type of network allows more space distribution of nodes which are
668: relatively further apart.  Here, we consider one of the simplest
669: possible approaches to obtaining a geographic model, which involves
670: the distribution of the $N$ nodes uniformly along a 2D space followed
671: by the interconnection of all nodes which are closer than a given
672: distance $d$.
673: 
674: In order to build a GG network, we start with an empty $L \times L$
675: discrete space $S$, such that each of its positions is expressed as
676: $S(x,y)$, where $i$ and $j$ are integer values so that $1 \leq x, y
677: \leq N$.  This space is henceforth understood as a Poisson 
678: field with density $\rho$, in the sense that any region with area $a$
679: (i.e. number of discrete elements $(x,y)$) will have, in the average,
680: a total of $a \rho$ points marked as $S(x,y)=1$~\cite{Stoyan:1996}.
681: The network nodes are selected by considering each position $(x,y)$ in
682: the space $S$ with probability $\rho = N/L^2$, implying an average
683: total number of nodes $N$.  Then, each of such nodes, marked as
684: $S(x,y)=1$, is connected to all other nodes to be found up to a
685: maximum Euclidean distance $d$.  Therefore, the average degree of the
686: network can be defined by controlling $d$.  More specifically, we make
687: $d = L\sqrt{2m/(N \pi)}$, so that every disk of radius $r$ in $S$
688: centered at each node will contain, in the average, $\left< k \right>
689: = 2m$, where $m$ is the BA parameter taken as the reference.  The
690: growth parameters of such a network model therefore are again limited
691: to only $N$ and $m$.  In order to ensure a relatively small variation
692: of $N$, every generated network with a total number of nodes smaller
693: than 90\% of the desired value of $N$ were discarded.
694: Figure~\ref{fig:exs} illustrates a GG complex network obtained by
695: using the methodology described above for $N=50$ and $m=3$.
696: 
697: 
698: \subsection{Generating the Attraction Basins}
699: 
700: As indicated in the Introduction, in order to avoid the intricacies
701: and specificities of how each distinct dynamic system represents
702: attractors, here we resort to a simple methodology involving diffusion
703: of activity from the prototype nodes, followed by the transformation
704: of the so obtained activity into a derivative
705: network~\cite{Costa_simple:2004,Costa_Travieso:2005}.
706: Although not reproducing in detail the attraction basins which would
707: be otherwise produced by diverse specific dynamics, this approach does
708: ensure the smoothness of coding, i.e. the property that nodes which
709: are topologic close will tend to have similar state dynamics.  The
710: details of such a methodology are presented as follows.
711: 
712: Let the $M$ prototype patterns be associated with respective nodes
713: chosen with uniform probability among the $N$ network nodes.  Let
714: $\vec{P}$ be the vector such that $P(i)=1$ if and only $i$ is one of
715: the prototype nodes, with $P(i)=0$ otherwise.  In order to allow a
716: probabilistic interpretation, we normalize this vector as $\vec{p} =
717: \vec{P} / \sum_{i=1}^{N} p(i)$.  This vector will act as the fixed
718: source of probabilities during the diffusion.  The adjacency matrix
719: describing the network is also normalized into its respective transfer
720: matrix $W$, i.e.:
721: 
722: \begin{equation}
723:   W(i,j) = K(i,j) / k(i), 
724: \end{equation}
725: 
726: where $k(i)$ is the degree of node $i$. Note that all the sums of $W$
727: along each of its columns will now be equal to 1, i.e. $W$ is a
728: \emph{stochastic matrix}.  In order to ensure that the matrix $K$ is
729: connected (irreducible)~\cite{Bremaud:2001}, all the nodes which do
730: not belong to the main connected component are excluded from the
731: network at the end of its respective construction.  Because the
732: considered average node degrees are relatively high, and well above
733: the percolation critic density, very few nodes are removed through
734: such a procedure.
735: 
736: Now, the final distribution of occupancy $\vec{\Omega^{T}}$ of each
737: node $i$ after $T$ interactions can be calculated by applying
738: recursively ($T$ times) the following set of equations:
739: 
740: \begin{eqnarray} 
741:   \vec{a} = \left( \vec{\Omega}^{t} + \vec{p}  \right),    \nonumber \\
742:   \vec{b} = \left( \vec{a} / \sum_{i=1}^{N} a(i) \right),   \nonumber \\
743:   \vec{\Omega}^{t+1} = W \vec{b}.
744: \end{eqnarray}
745: 
746: The number $T_t$ of total interactions is henceforth defined as
747: corresponding to $3$ times the diameter of the respective network,
748: i.e. $T_t = 3 diam$.  Figure~\ref{fig:exs}(b) illustrates the
749: occupancy states obtained for a GG network with $N=100$, $\left< k
750: \right> = 6$ and $M=2$ after $T_t=27$ interactions.
751: 
752: After the occupancy state of each node is obtained for the network,
753: its respective \emph{derivative network} $\Delta$
754: (e.g.~\cite{Costa_simple:2004, Costa_Travieso:2005}) is obtained by
755: applying the equation below for each edge $(i,j)$ existing in the
756: original network.
757: 
758: \begin{eqnarray}
759:   aux = \Omega(j) - \Omega(i)  \nonumber \\
760:   \left\{ \begin{array}{ll}
761:     aux>0  \Rightarrow &  \Delta(j,i) = aux  \nonumber \\
762:     aux<0  \Rightarrow &  \Delta(j,i) = aux/10  \nonumber \\
763:     aux=0  \Rightarrow &  \Delta(j,i) = -1  \nonumber \\
764:     \end{array}
765:   \right.
766: \end{eqnarray}
767: 
768: The following substitution is applied afterwards:
769: 
770: \begin{equation}  \label{eq:subst}
771:     \Delta(j,i) = -1  \Rightarrow  \Delta(j,i)=max(\Delta)/\beta.
772: \end{equation}
773: 
774: All cases in this work assumes $\beta=1000$, but this parameters has
775: been found not to be critical.  
776: 
777: Note that the values $\Delta(j,i)$ correspond to the weights of the
778: respective edges $(i,j)$.  Those edges which connect a node with small
779: induced activity to a node with higher activity will have larger
780: weights.  More specifically, in case $aux>0$ the weights are directly
781: proportional to the difference of activations.  Edges leading from
782: higher to smaller activations have one tenth of the reciprocal edge.
783: Because several adjacent nodes may result with equal activations, the
784: eventual previous connectivity between then will be preserved though
785: at an incremental value proportional to the maximum activation
786: (Equation~\ref{eq:subst}).  Observe that the matrix $\Delta$ is not
787: symmetric.
788: 
789: 
790: \subsection{Activating the Network}
791: 
792: Once the derivative network defined by the weight matrix $\Delta$ has
793: been calculated, its activation can be easily achieved through a
794: random walk preferential to the weights of the edges (see
795: also~\cite{Toro:2004}, where information is transmitted considering
796: the gradient of a node).  In order to do so, we obtain the stochastic
797: version of $\Delta$ by applying the following equation to each of its
798: edge $(i,j)$
799: 
800: \begin{equation}
801:   \delta(j,i) = \Delta(j,i) / \sum_{j=1}^{N} \Delta(j,i).
802: \end{equation}
803: 
804: The activation after $T$ steps is now given as
805: 
806: \begin{equation}
807:   \vec{\alpha} = 1/N \delta^T \vec{1},
808: \end{equation}
809: 
810: where $\vec{1}$ is the $N \times 1$ vector of ones.  As before, we
811: assume that the total number of interactions $T_r$ is equal to 3 times
812: the network diameter.  Therefore, the final activation corresponds to
813: the near equilibrium occupancy of each node after starting from any
814: node.  Figure~\ref{fig:exs}(c) shows the activations obtained by the
815: method described above with respect to the network in
816: Figure~\ref{fig:exs}(b).  It is interesting to note that the peaks of
817: the obtained activations do not necessarily correspond to the original
818: prototype nodes.  This is an interesting consequence of the fact that
819: the activation of the derivative network is affected not only by the
820: attraction basins, but also by the
821: \emph{in-strength}~\footnote{In a directed weighted network such as the 
822: derivative network, the \emph{in-strength} of a node $i$ is defined as
823: being equal to the sum of the weights of each inbound edge.} and the
824: local connectivity of each node.  More
825: specifically~\cite{Costa_etal:2006}, in case the in-strength is equal
826: to the out-strength for all nodes, the equilibrium activation will be
827: directly proportional to the respective in-strength.  However, this
828: result is not guaranteed for asymmetric connectivity such as in the
829: derivative network.
830: 
831: \begin{figure*}[htb]
832:   \begin{center} \includegraphics[scale=0.75,angle=0]{ex_GG.eps}
833:   \hspace{0.5cm} \includegraphics[scale=0.75,angle=0]{ex_diff.eps} \\
834:   (a) \hspace{8cm} (b) \vspace{0.5cm} \\
835:   \includegraphics[scale=0.75,angle=0]{ex_rw.eps} \hspace{0.5cm}
836:   \includegraphics[scale=0.75,angle=0]{ex_vor.eps} \\ (c) \hspace{8cm}
837:   (d) \\ \caption{Example of GG network obtained for $N=50$ and $m=3$
838:   (with $\left< k \right> = 5.5$) (a).  The
839:   respective occupancy states, defining the basis of the prototype
840:   nodes (striped), after $27$ interactions (b). Note the smooth
841:   distribution of states through both the topologic neighborhoods. 
842:   The activation induced by the random walks
843:   preferential to the weights of the respective derivative network
844:   (c).  The Voronoi tessellation defined by the two prototype nodes
845:   (represented as striped) (d).~\label{fig:exs}} \end{center}
846: \end{figure*}
847: 
848: 
849: \subsection{Separation Indices}
850: 
851: Given a specific grandmother dynamic system, it is important to
852: quantify in an objective manner the separability of its attractors.
853: Recall that the system incorporates $M$ prototype patterns, associated
854: to respective prototype nodes $p = 1, 2,
855: \ldots, M$.  In this work we propose two separation indices,  
856: $s_{ind}$ and $s_{ngh}$, defined by taking into account diffusion
857: dynamics along the complex network underlying the dynamic system.  For
858: simplicity's sake, the latter is henceforth called simply as
859: \emph{separation index} and abbreviated as $s$.  
860: 
861: Once the activations have been obtained in each of the dynamic states
862: associated to the nodes, we define the \emph{individual separation
863: index} $s_{ind}$ as being proportional to the geometric average of the
864: activations $v(p)$ at each prototype node $p = 1, 2,\ldots, M$, i.e.
865: 
866: \begin{equation}
867:   s_{ind} = M \left( \prod_{i=1}^{M} v(i) \right)^{1/M},
868: \end{equation}
869: 
870: where the proportionality factor $M$ is introduced in order to ensure
871: that $0 \leq w \leq 1$ instead of $0 \leq w \leq 1/M$.  As with the
872: quantification of the uniformity of the Voronoi areas
873: (Section~\ref{sec:conc}), the geometric average will reach its maximum
874: when all activations $v(i)$ have the same value.
875: 
876: However, as preliminary simulations (see Section~\ref{sec:fixed})
877: showed that this index tends to be too small, an alternative separation
878: index has been considered which takes into account also the immediate
879: neighborhoods of each prototype node.  Consider the situation
880: illustrated in Figure~\ref{fig:common}.  This figure shows two
881: prototype nodes $P1$ and $P2$ as well as their common $C$ immediate
882: neighboring nodes.  Because these nodes are at the same shortest path
883: distance from $P1$ and $P2$, they do not contribute to the
884: discrimination between those prototype nodes and are therefore not
885: considered in the calculation of the neighborhood separation index,
886: which is more formally defined as follows.
887: 
888: Let $Z(i)$ be the set including the respective prototype node $i$ and
889: its immediated neighbors which are not common to the immediated
890: neighborhoods of any of the other prototype nodes.  The probability
891: $p(i)$ at this set of nodes corresponds to the sum of the normalized
892: activations of the states in $Z(i)$.  Ideally, all prototype nodes
893: should result with probability equal to $1/M$, implying that all
894: prototypes are equally accessible and therefore maximally separable.
895: In order to quantify how the obtained network state approaches such a
896: reference separation, we define the neighborhood separation index
897: $s_{ngh}$ as corresponding to the geometric average of the total
898: probabilities at each set $Z(i)$, for all $M$ prototype nodes $i$,
899: i.e.
900: 
901: \begin{equation}
902:   s = s_{ngh} = M \left( \prod_{i=1}^{M} p(i) \right)^{1/M},
903: \end{equation}
904: 
905: where the multiplying constant $M$ is, as before, included in order to
906: imply that $0 \leq s \leq 1$.  The maximum separation between
907: attractors is therefore obtained whenever $s=1$.  
908: 
909: 
910: \begin{figure}[htb]
911:   \begin{center} \includegraphics[scale=0.5,angle=0]{common.eps}
912:   \caption{A portion of a hypothetical network showing two prototype
913:   nodes ($P1$ and $P2$).  The nodes belonging to both immediate
914:   neighborhoods are identified by the region $C$, while the nodes
915:   belonging to the respective immediate neighborhoods minus the common
916:   nodes are enclosed by the regions $N1$ and $N2$,
917:   respectively.~\label{fig:common}} \end{center}
918: \end{figure}
919: 
920: 
921: \subsection{Correlation Analysis}
922: 
923: Given several measurements of the topology of the network under
924: analysis, as well as the separation indices, a first insight about
925: their possible relationship and redundancy can be obtained by
926: considering the \emph{Pearson correlation coefficient}.  Let us
927: express each measurement as a random variable $X$ of which we have
928: $N_X$ observations.  First, we
929: \emph{standardize}~\cite{Johnson_Wichern:2002} the variable $X$ as
930: 
931: \begin{equation}
932:   \tilde{X} = \frac{X - \mu_X}{\sigma(X)},
933: \end{equation}
934: 
935: where $\mu_X$ and $\sigma(X)$ are the estimate of the average and
936: standard deviation of $X$.  The new, normalized variable has average
937: zero and unit variance. The Pearson correlation coefficient $r(X,Y)$
938: between any pair of normalized measurements $X$ and $Y$ can now be
939: estimated as
940: 
941: \begin{equation}
942:   r(X,Y) = \frac{1}{N_X-1} \sum_{i=1}^{N} \tilde{X} \tilde{Y}.
943: \end{equation}
944: 
945: Observe that $-1 \leq r(X,Y) \leq 1$, while $r(X,Y)=0$ means lack of
946: correlation between the two measurements.  It is important to stress
947: that uncorrelation does not imply statistic independence.  At the same
948: time, correlation can by no means be understood as a certain
949: indication of \emph{causality}.  In case two measurements result with
950: high absolute value of the Pearson correlation coefficient, they are
951: said to be \emph{correlated}, indicating redundancy of measurements.
952: However, even correlated measurements can contribute to the
953: characterization and discrimination between the
954: networks~\cite{Costa:survey}.
955: 
956: 
957: \subsection{Path Analysis}
958: 
959: While the Pearson correlation coefficient quantifies, in a normalized
960: fashion, the joint variation of two measurements, such a pairwise
961: measurement does not consider information about additional
962: measurements.  A series of sound statistic methods, ranging from
963: \emph{multivariate regression} (e.g.~\cite{Kleinbaum:1998,
964: Montgomery:2001}) to the more sophisticated
965: \emph{Structural Equation Modeling} (SEM) (e.g.~\cite{Raykov:2000, 
966: Kline:2005}), can be considered in order to obtain a more
967: representative characterization of the relationship between multiple
968: random variables (including latent variables) or measurements.  In
969: this work we considered the \emph{Path Analysis} methodology
970: (e.g.~\cite{Wright:1920, Raykov:2000, Kline:2005}), understood here as
971: a particular case of the SEM framework, in order to obtain indication
972: about the influence of the several measurements of the topology of the
973: networks on the respective attractor separation indices.
974: 
975: Path analysis was largely developed by S. Wright
976: (e.g.~\cite{Wright:1920}) in order to model explanatory relationships
977: between observed variables.  This methodology is similar to the
978: solution of a system of equations implied by substituting the model
979: generated covariance matrix into the sample
980: covariance~\cite{Raykov:2000}.  One of the interesting features of
981: this approach is that it considers the influence of the covariances
982: between all variables, sometimes being closely related to multivariate
983: linear regression.  The path analysis performed in this work considers
984: the structural relationship between the topologic and dynamic
985: properties of the investigated networks as shown in
986: Figure~\ref{fig:diagram}.  A relatively simple relationship between
987: the measurements has been considered, where the topologic features are
988: understood to cause the dynamic properties, namely the individual and
989: neighborhood attractor separation indices ($s_{ind}$ and $s$,
990: respectively).  Each of the measurements are associated to a reference
991: number (upper righthand side of each box).  The parameters
992: $\gamma_ij$, which are in principle not known, express the importance
993: of the topologic measurements with respect to the separation indices.
994: 
995: \begin{figure}[htb]
996:   \begin{center} 
997:   \includegraphics[scale=0.7,angle=0]{diagram_path.eps} 
998:   \caption{The structural relationship between the topologic
999:   and dynamic measurements considered for the path analysis
1000:   reported in this work.  The covariances between variables
1001:   3 to 9 are not shown for simplicity's sake.~\label{fig:diagram}}
1002:   \end{center}
1003: \end{figure}
1004: 
1005: Equations~\ref{eq:sind} and~\ref{eq:s} express the relationship
1006: between the considered variables, reflecting the fact that the
1007: regression coefficients $\gamma_ij$ establish the weights of the
1008: influences of each topologic variable onto the two separation
1009: indices. The environment
1010: LISREL~\footnote{http://www.ssicentral.com/lisrel/index.html} was used
1011: in this work in order to perform path analysis.
1012: 
1013: \begin{eqnarray} 
1014:   s_{ind} \propto \gamma_{13}\left<cc\right> + \gamma_{14}\left<sp\right>
1015:   +\gamma_{15}\sigma{sp} + \gamma_{16}Vi \nonumber \\
1016:   +\gamma_{17}\left<mi\right> + \gamma_{18}\sigma(mi) +\gamma_{19}diam  
1017:   \label{eq:sind} \\ 
1018:   \nonumber \\ s \propto \gamma_{13}\left<cc\right> +
1019:   \gamma_{14}\left<sp\right> +\gamma_{15}\sigma{sp} +
1020:   \gamma_{16}Vi \nonumber \\ +\gamma_{17}\left<mi\right> +
1021:   \gamma_{18}\sigma(mi) +\gamma_{19}diam   \label{eq:s}
1022: \end{eqnarray}
1023: 
1024: 
1025: \section{Results and Discussion}
1026: 
1027: This section starts by identifying each involved parameter which can
1028: affect the simulations and follows by presenting the results obtained
1029: for specific configurations of $N$ and $m$ and an analysis of the
1030: variation of the parameters.  The statistic analysis of the
1031: relationship between the topologic and dynamic measurements by using
1032: correlation and path analyses is also reported.
1033: 
1034: 
1035: \subsection{Involved Parameters and Their Expected Effects}
1036: 
1037: An important first issue to be considered while investigating the
1038: attractors separation in complex dynamic networks concerns the
1039: identification of all involved parameters which can influence the
1040: results.  The parameters involved in our simulations, as well as their
1041: expected effects, are described in the following:
1042: 
1043: \emph{The network model and its intrinsic parameters:} Each theoretic
1044: model of complex network considered in this work is likely to yield
1045: different attractors separations.  Those networks characterized by the
1046: small world property (i.e. ER, BA and WS) are, in principle, likely to
1047: produce less separated attraction basis because of the relatively
1048: small shortest path distance, expected in the average, between the
1049: prototype nodes (however, see Section~\ref{sec:res_path})).  While the
1050: ER, BA and GG models can be generated by considering just two
1051: parameters (i.e. $N$ and $m$), the WS network also involves a third
1052: parameters corresponding to the percentage of rewirings, which is
1053: fixed as $\alpha = 10$\%.
1054: 
1055: \emph{The network size $N$:} This is an important global parameter of
1056: every complex network, corresponding to its total number of nodes.
1057: One of the most important aspects related to this parameter are the
1058: so-called \emph{finite size effects}, namely the fact that the network
1059: properties change considerably when moving from large (possibly
1060: infinite) to small values of $N$.  As an extreme example, the
1061: connectivity of any network will decrease when $N$ approaches just one
1062: or two nodes.  Although in this work we are more interested in finite
1063: size networks, it is often useful to try to extrapolate from
1064: properties measured for smaller values of $N$ to the infinite limit,
1065: as done in Section~\ref{sec:effs}.  Because of the computational
1066: demand required to simulate hundreds of realizations for each
1067: configuration, the current work is limited to relatively small values
1068: of $N$.  As far as the attractors separation is concerned, it is
1069: expected that, for fixed $m$ and $M$, the separability will tend to
1070: increase with $N$ because of the additional space thus allowed for the
1071: representation and distribution of the prototypes.
1072: 
1073: \emph{The network average node degree $\left< k \right>$:}  
1074: This parameter, which in this work is defined with respect to the
1075: reference $m$, is particularly important in defining the overall
1076: degree of connectivity in the network, especially in those structures
1077: which are not scale free and therefore have specific degree scales.
1078: Generally, smaller values of $\left< k \right>$ tend to imply longer
1079: shortest paths, possibly improving the attractors separation (however,
1080: see Section~\ref{sec:res_path}).  This parameter is not systematically
1081: investigated in this work, which mostly considers fixed $m = 3$
1082: (however, a situation with $m=10$ is considered in
1083: Section~\ref{sec:fixed}).
1084: 
1085: \emph{The number $M$ of prototype patterns to be represented:} 
1086: This parameter is directly involved in the separability of the
1087: prototype patterns, in the sense that the larger the number of
1088: prototypes, the less separated they tend to be.  
1089: 
1090: \emph{The number $T_t$ of interactions used for diffusion around each
1091: prototype node:} The diffusion of activity emanating from the
1092: prototype nodes has been verified to converge quickly for all
1093: considered networks, so that the assumption of the total number of
1094: interactions to be given as $T_t = 3 diam$ practically ensures the
1095: resulting activity to correspond to its equilibrium state.
1096: 
1097: \emph{The number $T_r$ of interactions considered in the retrieval
1098: random walk:} Again, the assumed total number of interactions used in
1099: the activation of the attraction basins is large enough to imply near
1100: equilibrium states.
1101: 
1102: 
1103: \subsection{Fixed Configurations} \label{sec:fixed}
1104: 
1105: In order to get a better understanding about the separability of the
1106: attractors in the considered four theoretic models and to obtain
1107: preliminary indications about the effects of the involved parameters,
1108: we considered a set of preliminary simulations as described in the
1109: following.  All results presented in this section considered
1110: $T_t=T_r=3 diam$ and 500 realizations of each configuration.
1111: 
1112: So as to have the first glimpses of the separability in each of the
1113: four complex networks models, we considered $N=100$, $\left< k \right>
1114: = 3$, and a relatively small number of prototypes $M=3$.  The
1115: individual separation indices $s_{ind}$ and $s$ obtained for each of
1116: the four models are shown in terms of their respective population
1117: histograms in Figures~\ref{fig:caso_sind} and~\ref{fig:caso1}.  The
1118: respective averages and standard deviations are also shown inside each
1119: graph.  The first important result is that the individual separation
1120: index $s_{ind}$ allows considerably less resolution than the
1121: neighborhood index $s$, with most of its values being close to zero
1122: (Fig.~\ref{fig:caso_sind}).  As this effect has been confirmed for
1123: other configurations, we henceforth focus attention on the
1124: neighborhood separation index $s$.  Interestingly, similar
1125: distributions have been obtained for the pairs of models ER/BA and
1126: WS/GG in Figure~\ref{fig:caso1}, the former being characterized by
1127: smaller averages of $s$.  Observe also the presence of cases where
1128: $s=0$ in the WS and GG cases.  The largest average, implying the
1129: better average attractor separability, was obtained for the WS model,
1130: followed by the GG networks, which also implied a significant number
1131: of null separability indices.  This phenomenon is caused whenever all
1132: immediate neighbors of the prototype nodes are common.  These results
1133: also show that a reasonable separation between the three prototype
1134: attractors have been obtained for the WS and GG cases.
1135: 
1136: \begin{figure*}[h]
1137:   \begin{center} 
1138:   \includegraphics[scale=0.55,angle=0]{caso_sind.eps}
1139:   \caption{The population histograms of the separation index $s_{ind}$
1140:   for the four considered theoretic models of complex networks with
1141:   $N=100$, $M=3$ and $\left< k \right> = 6$ (i.e. $m=3$).  It is clear
1142:   from these histograms that the separation index $s_{ind}$ does not
1143:   provide good resolution.~\label{fig:caso_sind}}
1144: \end{center}
1145: \end{figure*}
1146: 
1147: \begin{figure*}[htb]
1148:   \begin{center} 
1149:   \includegraphics[scale=0.55,angle=0]{caso1.eps} 
1150:   \caption{The population histograms of the separation index $s$
1151:   for the four considered theoretic models of complex networks
1152:   with $N=100$, $M=3$ and $\left< k \right> = 6$.
1153:   Similar results were obtained for the ER/BA and WS/GG models.
1154:   The former case involves smaller average and standard deviation of
1155:   the $s$ index.~\label{fig:caso1}}
1156:   \end{center}
1157: \end{figure*}
1158: 
1159: Next, we verify the possible influence of the network size $N$ by
1160: considering the same configuration as before, but with $N=200$.  The
1161: respective results are shown in Figure~\ref{fig:caso2}.  It is clear
1162: from these results that the larger size of the network had relatively
1163: little influence on the separation indices, suggesting the finite size
1164: effect to be small for this number of parameters.  
1165: 
1166: \begin{figure*}[htb]
1167:   \begin{center} 
1168:   \includegraphics[scale=0.55,angle=0]{caso2.eps} 
1169:   \caption{The population histograms of the separation index $s$
1170:   for the four considered theoretic models of complex networks
1171:   with $N=200$, $M=3$ and $\left< k \right> = 6$.
1172:   Similar results were obtained for the ER/BA and WS/GG models.
1173:   The results are similar to those obtained for $N=100$ (see
1174:   Figure~\ref{fig:caso1}).~\label{fig:caso2}}
1175:   \end{center}
1176: \end{figure*}
1177: 
1178: Now we turn our attention to the influence of the number of prototype
1179: patterns to be represented.  The same configuration as in
1180: Figure~\ref{fig:caso1} is considered, except that $M=10$.  The
1181: obtained results are shown in Figure~\ref{fig:caso3}.  It is clear
1182: that increase by more than threefold of the number of prototype
1183: patterns implied not only a substantially more cases such that $s=0$,
1184: but also the distributions to be strongly left-shifted in comparison
1185: with the respective cases in Figure~\ref{fig:caso1}.  This effect
1186: holds for all the four considered network models. The main reason
1187: behind such a substantial decrease of performance is that simply there
1188: was not much space left, in the average, between the prototype
1189: attractors.
1190: 
1191: \begin{figure*}[htb]
1192:   \begin{center} 
1193:   \includegraphics[scale=0.55,angle=0]{caso3.eps} 
1194:   \caption{The population histograms of the separation index $s$
1195:   for the four considered theoretic models of complex networks
1196:   with $N=100$, $M=3$ and $\left< k \right> = 6$.
1197:   A substantial decrease of the attractors separation has
1198:   been implied in all cases.~\label{fig:caso3}}
1199:   \end{center}
1200: \end{figure*}
1201: 
1202: Because the last configuration exhibited an accentuated loss of
1203: separability because of lack of space in the network, it is
1204: interesting to reconsider the effect of increasing $N$ for this
1205: situation.  The results obtained for the same previous configuration,
1206: but now with $N=200$, are shown in Figure~\ref{fig:caso4}.  The
1207: separation index increased substantially for all the four network
1208: models.  It could be expected that such improvements tend to decrease
1209: for still larger $N$, until reaching a regime where little improvement
1210: is observed.  In such a state, the prototype nodes would be
1211: sufficiently far away one another so that their separation no longer
1212: depends on $N$.  Additional investigation about the change of the
1213: average and standard deviation of $s$ are to be found in
1214: Section~\ref{sec:effs}.
1215: 
1216: \begin{figure*}[htb]
1217:   \begin{center} 
1218:   \includegraphics[scale=0.55,angle=0]{caso4.eps} 
1219:   \caption{The population histograms of the separation index $s$
1220:   for the four considered theoretic models of complex networks
1221:   with $N=100$, $M=10$ and $\left< k \right> = 6$.
1222:   A substantial decrease of the attractors separation has
1223:   been implied in all cases.~\label{fig:caso4}}
1224:   \end{center}
1225: \end{figure*}
1226: 
1227: Finally, we check for the possible effect of the average node degree
1228: on the separation index.  Again, the same configuration as the network
1229: in Figure~\ref{fig:caso1} is adopted, but now with $\left< k
1230: \right> = 20$ (i.e. $m=10$) instead of $\left< k \right> = 6$.  
1231: Figure~\ref{fig:caso5} depicts the respectively obtained results,
1232: which indicate a clear reduction of the attractors separability.  Such
1233: an effect is possibly a consequence of the fact that once the network
1234: become too intensely connected, the shortest path between the
1235: prototype nodes will be reduced and the separability undermined.
1236: 
1237: \begin{figure*}[htb]
1238:   \begin{center} 
1239:   \includegraphics[scale=0.55,angle=0]{caso5.eps} 
1240:   \caption{The population histograms of the separation index $s$
1241:   for the four considered theoretic models of complex networks
1242:   with $N=100$, $M=10$ and $\left< k \right> = 20$.  The effect
1243:   of increasing the average degree was to substantially 
1244:   decrease the attractors separation in all cases.~\label{fig:caso5}}
1245:   \end{center}
1246: \end{figure*}
1247: 
1248: 
1249: \subsection{Finite Size Effects}  \label{sec:effs}
1250: 
1251: The simulations discussed in the previous section seem to have
1252: indicated that increases of the value of $N$ would tend to improve the
1253: attractors separation until reaching a regime where so much space is
1254: available that the patterns no longer feels the finite size of the
1255: network.  In order to obtain further insights about this effect, the
1256: configuration involving $m=3$ and $M=3$ was simulated for $N = 50,
1257: 100, 150, 200, 250$ and the results are given in
1258: Figure~\ref{fig:finite}.  This figure shows the average (a) and
1259: standard deviation (b) of the neighborhood separation index in terms
1260: of $N$.  It is clear from these results that the increase of $N$ does
1261: enhance the attractors separation until reaching a plateau, where a
1262: possible 'unsaturation' occurs, indicating that the finite size
1263: effects are over for the specific parameters $m$ and $M$.  At the same
1264: time, $\sigma(s)$ decreases, also reaching a relatively low plateau.
1265: Similar results could be expected for other configurations.  In the
1266: case of a larger value of $M$, it is expected that $\left< k
1267: \right>$ will increase more steeply along the smaller values of $N$,
1268: reaching a similar plateau at larger values.  Therefore, as the
1269: 'unsaturation' effect was further corroborated by the additional
1270: analysis, such a behavior can be considered as a guideline for
1271: choosing a proper value of $N$ given $M$ and $m$, for instance by
1272: choosing $N$ where $\left< k
1273: \right>$ reaches a fixed proportion of its plateau value.
1274: 
1275: \begin{figure*}[htb]
1276:   \begin{center} 
1277:   \includegraphics[scale=0.55,angle=0]{finite_size.eps} \\
1278:   (a) \hspace{4cm} (b) \\
1279:   \caption{The average and standard deviation of the neighborhood
1280:   separation index ($\left< k \right>$ and $\sigma(s)$, respectively)
1281:   as a function of $N$ for the four models of networks considering
1282:   $m=3$ and $M=3$.  While the averages increase and then tend to 
1283:   a plateau, the standard deviations decrease and also stabilize.
1284:   ~\label{fig:finite}}
1285:   \end{center}
1286: \end{figure*}
1287: 
1288: 
1289: \subsection{Correlation Analysis}
1290: 
1291: In addition to studying the effects of the involved topologic
1292: measurements on the attractors separation, it is also particularly
1293: useful to try to identify in which ways the separation index is
1294: related to them.  This may allow the prediction of the separability
1295: without performing dynamic simulation, i.e. by considering only the
1296: topologic measurements.  Table~\ref{tab:pears} show the Pearson
1297: correlation coefficients obtained considering all pair os topologic
1298: measurement (i.e. $\left<cc\right>$, $\left<sp\right>$, $\sigma(sp)$,
1299: $Vi$, $\left<mi\right>$, $\sigma(mi)$ and $diam$) and the two
1300: separation indices ($s_{ind}$ and $s$).  
1301: 
1302: As previously verified in~\cite{Costa:survey}, the pattern of
1303: correlations resulted not similar for each model of network. In
1304: addition, the obtained results indicate several high absolute
1305: correlation values.  In the case of $ER$ networks, the most intense
1306: negative correlation (-0.79) was obtained between $\left<sp\right>$
1307: and $\left<mi\right>$, which was indeed expected as longer shortest
1308: path lengths tend to reduce the matching index.  While similar high
1309: correlation (-0.78) was observed also for $BA$, this specific
1310: correlation was relatively smaller for the $WS$ and $GG$ cases (-0.50
1311: and -0.42, respectively).  This fact was reflected in the intense
1312: negative correlation (-0.77) between $s$ and $\left<mi\right>$ and
1313: positive correlation (0.63) with $\left<sp\right>$ for the $ER$ model.
1314: A similar effect can be observed for the $BA$, $WS$ and $GG$ networks.
1315: Observe that $\left<mi\right>$ is highly correlated (0.72 for ER) with
1316: $\sigma(mi)$, suggesting a dependence between the standard deviation
1317: and average of this measurement. The index quantifying the uniformity
1318: of the Voronoi tessellation was found to be negatively correlated with
1319: $\left<mi\right>$ and $\sigma(mi)$, as expected, because higher $mi$
1320: values are favored by more irregular Voronoi areas.  Interestingly,
1321: low correlation values were observed between $s$ and the network
1322: diameter and clustering coefficients in all cases.  It is also
1323: interesting to note that rather different patterns of correlations
1324: were obtained between $s_{ind}$ and $s$ and the topologic measurements
1325: for all network models.
1326: 
1327: \begin{table*}[htb]
1328: \centering
1329: \vspace{1cm}
1330: {\bf ER} \hspace{0.5cm}
1331: \begin{tabular}{||c||c|c|c|c|c|c|c|c|c||}
1332:  \hline  \hline
1333:  &  $s_{ind}$  &  $s$ & $\left<cc\right>$ & $\left<sp\right>$ & 
1334:    $\sigma(sp)$ & $Vi$ & $\left<mi\right>$ & 
1335:    $\sigma(mi)$ & $ diam $   \\   \hline
1336:  $s_{ind}$         & 1   &     &     &     &     &     &     & &  \\  \hline
1337:  $s$               &-0.11& 1   &     &     &     &     &     & &  \\  \hline
1338:  $\left<cc\right>$ &-0.06&-0.11& 1   &     &     &     &     & &  \\  \hline
1339:  $\left<sp\right>$ &-0.51& 0.63&-0.11& 1   &     &     &     & &  \\  \hline
1340:  $\sigma(sp)$      & 0.31&-0.38& 0.07&-0.11& 1   &     &     & &  \\  \hline
1341:  $Vi$              &-0.17& 0.40&-0.03& 0.19&-0.19& 1   &     & &  \\  \hline
1342:  $\left<mi\right>$  & 0.51&-0.77& 0.08&-0.79& 0.49&-0.44& 1 & &  \\ \hline
1343:  $\sigma(m)i$       & 0.37&-0.52& 0.09&-0.51& 0.44&-0.52& 0.72& 1 &  \\  \hline
1344:  $diam$            & 0.02& 0.00&-0.06& 0.03&-0.01& 0.10&-0.01&-0.05& 1 \\ \hline  \hline
1345: \end{tabular} \hspace{0.5cm} (a) \vspace{0.3cm} \\
1346: {\bf BA} \hspace{0.5cm}
1347: \begin{tabular}{||c||c|c|c|c|c|c|c|c|c||}  % BA
1348:  \hline  \hline
1349:  &  $s_{ind}$  &  $s$ & $\left<cc\right>$ & $\left<sp\right>$ & 
1350:    $\sigma(sp)$ & $Vi$ & $\left<mi\right>$ & 
1351:    $\sigma(mi)$ & $ diam $   \\   \hline
1352:  $s_{ind}$         & 1   &     &     &     &     &     &     & &  \\  \hline
1353:  $s$               &-0.02& 1   &     &     &     &     &     & &  \\  \hline
1354:  $\left<cc\right>$ &-0.08&-0.26& 1   &     &     &     &     & &  \\  \hline
1355:  $\left<sp\right>$ &-0.53& 0.62&-0.13& 1   &     &     &     & &  \\  \hline
1356:  $\sigma(sp)$      & 0.25&-0.10&-0.16& 0.01& 1   &     &     & &  \\  \hline
1357:  $Vi$              &-0.10& 0.81&-0.26& 0.54&-0.12& 1   &     & &  \\  \hline
1358:  $\left<mi\right>$  & 0.44&-0.69& 0.13&-0.78& 0.40&-0.65& 1 & &  \\ \hline
1359:  $\sigma(mi)$       & 0.47&-0.35& 0.06&-0.58& 0.28&-0.46& 0.61& 1 &  \\  \hline
1360:  $diam$            & 0.03& 0.09&-0.24& 0.08&-0.08& 0.07&-0.06&-0.03& 1 \\ \hline  \hline
1361: \end{tabular} \hspace{0.5cm} (b) \vspace{0.3cm} \\
1362: {\bf WS} \hspace{0.5cm}
1363: \begin{tabular}{||c||c|c|c|c|c|c|c|c|c||}   %WS
1364:  \hline  \hline
1365:  &  $s_{ind}$  &  $s$ & $\left<cc\right>$ & $\left<sp\right>$ & 
1366:    $\sigma(sp)$ & $Vi$ & $\left<mi\right>$ & 
1367:    $\sigma(mi)$ & $ diam $   \\   \hline
1368:  $s_{ind}$         & 1   &     &     &     &     &     &     & &  \\  \hline
1369:  $s$               &-0.58& 1   &     &     &     &     &     & &  \\  \hline
1370:  $\left<cc\right>$ &-0.04& 0.09& 1   &     &     &     &     & &  \\  \hline
1371:  $\left<sp\right>$ &-0.46& 0.54& 0.10& 1   &     &     &     & &  \\  \hline
1372:  $\sigma(sp)$      & 0.43&-0.16& 0.06& 0.31& 1   &     &     & &  \\  \hline
1373:  $Vi$              &-0.55& 0.63& 0.10& 0.53&-0.09& 1   &     & &  \\  \hline
1374:  $\left<mi\right>$  & 0.89&-0.80&-0.04&-0.50& 0.36&-0.69& 1 & &  \\ \hline
1375:  $\sigma(mi)$       & 0.42&-0.54&-0.05&-0.37& 0.13&-0.37& 0.55& 1 &  \\  \hline
1376:  $diam$            &-0.01& 0.02& 0.29& 0.18& 0.16& 0.07&-0.02&-0.02& 1 \\ \hline  \hline
1377: \end{tabular} \hspace{0.5cm} (c) \vspace{0.3cm} \\
1378: {\bf GG} \hspace{0.5cm}
1379: \begin{tabular}{||c||c|c|c|c|c|c|c|c|c||}  %GG
1380:  \hline  \hline
1381:  &  $s_{ind}$  &  $s$ & $\left<cc\right>$ & $\left<sp\right>$ & 
1382:    $\sigma(sp)$ & $Vi$ & $\left<mi\right>$ & 
1383:    $\sigma(mi)$ & $ diam $   \\   \hline
1384:  $s_{ind}$         & 1   &     &     &     &     &     &     & &  \\  \hline
1385:  $s$               &-0.16& 1   &     &     &     &     &     & &  \\  \hline
1386:  $\left<cc\right>$ &-0.06& 0.04& 1   &     &     &     &     & &  \\  \hline
1387:  $\left<sp\right>$ &-0.04& 0.42& 0.13& 1   &     &     &     & &  \\  \hline
1388:  $\sigma(sp)$      & 0.16&-0.10& 0.11& 0.57& 1   &     &     & &  \\  \hline
1389:  $Vi$              &-0.38& 0.71& 0.08& 0.40&-0.09& 1   &     & &  \\  \hline
1390:  $\left<mi\right>$  & 0.40&-0.87&-0.06&-0.42& 0.16&-0.71& 1 & &  \\ \hline
1391:  $\sigma(mi)$       & 0.33&-0.72&-0.06&-0.31& 0.20&-0.58& 0.82& 1 &  \\  \hline
1392:  $diam$            &-0.07& 0.04& 0.19& 0.31& 0.30& 0.12&-0.10&-0.05& 1 \\ \hline  \hline
1393: \end{tabular} \hspace{0.5cm} (d) \vspace{0.3cm}
1394: \caption{Pearson correlation coefficients obtained for the
1395: ER (a), BA (b), WS (c) and GG (d) networks with $N=100$, $m=3$ and
1396: $M=3$ considering the individual ($s_{ind}$) and neighborhood
1397: ($s_{ngh}$) separability indices and seven measurements of the
1398: topology of the network (i.e. average clustering coefficient
1399: ($\left<cc\right>$), average shortest path between prototype nodes
1400: ($\left<sh_path\right>$), standard deviation of the shortest paths
1401: between the prototype nodes ($\sigma(sh_path)$)), average Voronoi
1402: index ($Vi$), average matching index between prototype
1403: nodes ($\left<m\right>$), standard deviation of the matching index
1404: between prototype nodes ($\sigma(mi)$) and diameter of the networks
1405: (diam)}.\label{tab:pears}
1406: \end{table*}
1407: 
1408: 
1409: 
1410: \subsection{Structural Equation Modeling (SEM) / Path Analysis}  
1411: \label{sec:res_path}
1412: 
1413: Although the correlation analysis described in the previous section
1414: can provide interesting information about the pairwise relationship
1415: between the separation indices and the topologic measurements, such
1416: results are limited because they do not reflect the general
1417: relationship between all the topologic measurements and the separation
1418: indices.  In addition, some of the high observed absolute
1419: correlation values can be a consequence of
1420: spurious~\cite{Raykov:2000,Kline:2005} effects between the variables.
1421: In order to gather additional insights about the how the dynamic
1422: parameters (i.e. the separation indices) are influenced (and even to a
1423: large extent defined) by the topology of the network, while
1424: considering all measurement co-variations, we performed path analysis
1425: considering the structural dependence between measurements as
1426: expressed in Figure~\ref{fig:diagram}.  It is expected that, by
1427: considering all dispersions, the path analysis can provide a more
1428: objective and filtered indication of the influences of the topologic
1429: measurements on the attractors separation.
1430: 
1431: The considered data were respective to ER, BA, WS and GG networks with
1432: $N=100$, $M=3$ and $m=3$. After estimation of the covariance of the
1433: measurements, coding and execution in the LISREL environment, the
1434: regression coefficients and residues ($E1$ and $E2$), as well as the
1435: relationship between these residuals ($C_{12}$) were obtained.  The
1436: results are given in Table~\ref{tab:pathan}.  
1437: 
1438: A series of interesting insights have been derived from these results.
1439: As with the correlations, the dependencies between the topology and
1440: separation are strictly specific to each network model, a dependency
1441: which may also change for other configurations with different values
1442: of the parameters $M$ and $m$.  In the case of the ER networks, the
1443: standard deviation of the matching index ($\sigma(mi)$) resulted
1444: particularly influent (negative influence = -0.96) on the $s$ index.
1445: At the same time, the average matching index ($\left<k\right>$) was
1446: found to have a strong influence on $s$.  Except for relatively
1447: smaller negative influence of the $\left< cc
1448: \right>$ on $s$, no particularly strong influences are observed with
1449: respect to the other topologic measurements.  Such a strong influence
1450: of the matching index can be understood because non-zero matching
1451: index are obtained only in extreme cases, where the prototype nodes
1452: are too close.  Therefore, nonzero matching indices are a strong and
1453: secure indication of poorly separated attractors.  For this reason,
1454: the matching index dominated the path analysis and implied smaller
1455: influences for most other measurements, including the shortest path.
1456: This effect can also be observed for the BA and WS cases.  However, it
1457: is interesting to note the weak influence of $\sigma(mi)$ on $s$ in
1458: the case of GG networks.  For the BA case, the strongest influence on
1459: $s$ was identified for the Voronoi index $Vi$, which is compatible
1460: with the fact that higher uniformity of the Voronoi tessellation by
1461: the prototype nodes tends to promote better separation between those
1462: nodes.  Interestingly, a particularly strong influence of $Vi$ on the
1463: $s$ has been observed only for the BA model.  This effect can be
1464: related to the fact that the BA provides the poorest general
1465: separation between attractors as a possible consequence of the
1466: generalized connectivity implemented by the hubs.  In such cases,
1467: where a larger number of non-zero matching indices are therefore
1468: obtained, the Voronoi separation may become more relevant as a
1469: predictor of the separability.  In the case of WS models, strong
1470: influence (positive = 0.65) was obtained for the $\left<cc\right>$.
1471: This effect is particularly interesting because it could be expected
1472: that, by promoting smaller shortest paths, this measurement would be
1473: inversely related to the separability, which is indeed the case for
1474: the respective ER and BA path analysis results.  Indeed, a very weak
1475: dependency with this variable had been revealed by the Pearson
1476: correlations.  However, the positive influence of $\left<cc\right>$ on
1477: $s$ can be a consequence of the fact that GG networks with higher
1478: average clustering coefficients will tend to have more intense local
1479: connectivity which could allow the activity to diffuse more uniformly
1480: and to concentrate effectively around the prototype nodes.  Such an
1481: effect would be more definite in the WS case because of the higher
1482: uniformity of local connections implied by its ring structure.  The
1483: influences of the topologic features on the $s$ obtained for the GG
1484: cases are dominated by $\left<mi\right>$ as discussed above.
1485: 
1486: 
1487: \begin{table*}[htb]
1488: \centering
1489: \vspace{1cm}
1490: \begin{tabular}{|c|c||c|c|c|c|c|c|c|c|c|}
1491:   \hline  
1492:     & &  $\left<cc\right>$ & $\left<sp\right>$ & 
1493:     $\sigma(sp)$ & $ Vi$ & $\left<mi\right>$ & 
1494:     $\sigma(mi)$ & $ diam $ &  $Res.$ &  $C$ \\   \hline  \hline
1495:     ER  & 
1496:     $s_{ind}$   & $\gamma_{13}=-0.01$ & $\gamma_{14}=-0.02$ & 
1497:                 $\gamma_{15}=0.01$    & $\gamma_{16}=-0.01$ 
1498:                 & $\gamma_{17}=0.01$ & $\gamma_{18}=0.01$ & 
1499:                 $\gamma_{19}=0.00$ & $E1=0.00$ & $C_{1,2}=0.01 $ \\ 
1500:     & $s$       & $\gamma_{23}=-0.43$ & $\gamma_{24}=0.04$ & 
1501:                 $\gamma_{25}=-0.03$ & $\gamma_{26}=0.25$ & 
1502:                 $\gamma_{27}=-0.96$ & $\gamma_{28}=1.13$ & 
1503:                 $\gamma_{29}=-0.01$ & $E2=0.00$ & \\  \hline \hline
1504:     BA  & 
1505:     $s_{ind}$   & $\gamma_{13}=-0.02$ & $\gamma_{14}=-0.02$ & 
1506:                 $\gamma_{15}=0.01$    & $\gamma_{16}=0.03$ 
1507:                 & $\gamma_{17}=-0.02$ & $\gamma_{18}=0.06$ & 
1508:                 $\gamma_{19}=0.00$ & $E1=0.00$ & $C_{12}=0.00$ \\ 
1509:     & $s$       & $\gamma_{23}=-0.21$ & $\gamma_{24}=0.06$ & 
1510:                 $\gamma_{25}=0.01$ & $\gamma_{26}=0.75$ & 
1511:                 $\gamma_{27}=-0.39$ & $\gamma_{28}=0.73$ & 
1512:                 $\gamma_{29}=0.00$ & $E2=0.00$ & \\  \hline \hline
1513:     WS  & 
1514:     $s_{ind}$   & $\gamma_{13}=-0.05$ & $\gamma_{14}=0.00$ & 
1515:                 $\gamma_{15}=0.01$    & $\gamma_{16}=0.03$ 
1516:                 & $\gamma_{17}=0.17$ & $\gamma_{18}=-0.69$ & 
1517:                 $\gamma_{19}=0.00$ & $E1=0.00$ & $C_{12}=0.00$\\ 
1518:     & $s$       & $\gamma_{23}=0.65$ & $\gamma_{24}=0.01$ & 
1519:                 $\gamma_{25}=0.01$ & $\gamma_{26}=0.08$ & 
1520:                 $\gamma_{27}=-0.55$ & $\gamma_{28}=-3.54$ & 
1521:                 $\gamma_{29}=-0.16$ & $E2=0.00$ & \\  \hline  \hline
1522:     GG  & 
1523:     $s_{ind}$   & $\gamma_{13}=-0.03$ & $\gamma_{14}=0.00$ & 
1524:                 $\gamma_{15}=0.00$    & $\gamma_{16}=-0.02$ 
1525:                 & $\gamma_{17}=0.05$ & $\gamma_{18}=-0.01$ & 
1526:                 $\gamma_{19}=0.00$ & $E1=0.00$ & $C_{12}=0.01$ \\ 
1527:     & $s$       & $\gamma_{23}=-0.10$ & $\gamma_{24}=0.00$ & 
1528:                 $\gamma_{25}=0.00$ & $\gamma_{26}=0.17$ & 
1529:                 $\gamma_{27}=-0.87$ & $\gamma_{28}=-0.08$ & 
1530:                 $\gamma_{29}=-0.01$ & $E2=0.00$ & \\  \hline 
1531: \end{tabular}
1532: \caption{The influences of each considered topologic measurement
1533: on the attractors separation indices as revealed by path
1534: analysis considering $N=100$,$M=3$ and $m=3$.}\label{tab:pathan}
1535: \end{table*}
1536: 
1537: 
1538: While the previous path analysis has revealed a series of insights
1539: about the influence of the network topology on the dynamic separation
1540: of its attraction basins, it was strongly biased by the critical
1541: influence of the matching index.  In order to try to get further
1542: insights on the structure/dynamics relationship for the four
1543: considered models, we repeated the path analysis while not including
1544: $\left<mi\right>$ and $\sigma(mi)$.  The results are summarized in
1545: Table~\ref{tab:pathan2}.  Interestingly, except for relatively small
1546: increases with respect to $Vi$ and $\left<cc\right>$, the obtained
1547: influences remained similar, with small effects observed for the
1548: shortest path measurements.  This can be understood as a confirmation
1549: that the separation between the attractors is critically defined by
1550: the overlap between the hierarchical neighbors of the prototype
1551: nodes, especially their immediate neighbors (implying higher matching
1552: indices).  In the cases where the matching index is zero, the lack of
1553: congruence between the original attractors and the distribution of
1554: diffusive activity should be largely explained by the higher
1555: influences observed in Table~\ref{tab:pathan2}, which involve the
1556: average clustering coefficient of the network and the Voronoi index
1557: for the prototype nodes.  Indeed, in the cases where the prototype
1558: nodes are not topologically too close one another, the lack of
1559: agreement between the activity distribution and the original
1560: attractors (as in Figure~\ref{fig:ex}b and c) will depend particularly
1561: on the properties of the local connectivity as expressed by the
1562: clustering coefficient and the node degree.
1563: 
1564: 
1565: \begin{table*}[htb]
1566: \centering
1567: \vspace{1cm}
1568: \begin{tabular}{|c|c||c|c|c|c|c|c|c|}
1569:   \hline  
1570:     & &  $\left<cc\right>$ & $\left<sp\right>$ & 
1571:     $\sigma(sp)$ & $Vi$ & $ diam $ &  
1572:     $Res.$ &  $C$ \\   \hline  \hline
1573:     ER  & 
1574:     $s_{ind}$   & $\gamma_{13}=-0.01$ & $\gamma_{14}=-0.02$ & 
1575:                 $\gamma_{15}=0.01$    & $\gamma_{16}=-0.01$ 
1576:                 & $\gamma_{17}=0.00$ & $E1=0.00$ & $C_{12}=0.00$  \\ 
1577:     & $s$       & $\gamma_{23}=-0.26$ & $\gamma_{24}=0.14$ & 
1578:                 $\gamma_{25}=-0.09$ & $\gamma_{26}=0.43$ & 
1579:                 $\gamma_{27}=-0.01$ & $E2=0.00$ & \\  \hline \hline
1580:     BA  & 
1581:     $s_{ind}$   & $\gamma_{13}=-0.02$ & $\gamma_{14}=-0.02$ & 
1582:                 $\gamma_{15}=0.01$    & $\gamma_{16}=0.03$ 
1583:                 & $\gamma_{17}=0.00$ & $E1=0.00$ & $C_{12}=0.00$ \\ 
1584:     & $s$       & $\gamma_{23}=-0.22$ & $\gamma_{24}=0.08$ & 
1585:                 $\gamma_{25}=-0.01$ & $\gamma_{26}=0.79$ & 
1586:                 $\gamma_{27}=-0.00$ & $E2=0.00$ & \\  \hline \hline
1587:     WS  & 
1588:     $s_{ind}$   & $\gamma_{13}=0.03$ & $\gamma_{14}=-0.01$ & 
1589:                 $\gamma_{15}=0.01$    & $\gamma_{16}=-0.05$ 
1590:                 & $\gamma_{17}=0.00$ & $E1=0.00$ & $C_{12}=0.00$\\ 
1591:     & $s$       & $\gamma_{23}=0.43$ & $\gamma_{24}=0.03$ & 
1592:                 $\gamma_{25}=-0.03$ & $\gamma_{26}=0.41$ & 
1593:                 $\gamma_{27}=-0.01$ & $E2=0.00$ & \\  \hline  \hline
1594:     GG  & 
1595:     $s_{ind}$   & $\gamma_{13}=-0.01$ & $\gamma_{14}=0.00$ & 
1596:                 $\gamma_{15}=0.0$    & $\gamma_{16}=-0.04$ 
1597:                 & $\gamma_{17}=0.00$ & $E1=0.00$ & $C_{12}=0.01$ \\ 
1598:     & $s$       & $\gamma_{23}=-0.21$ & $\gamma_{24}=0.04$ & 
1599:                 $\gamma_{25}=-0.03$ & $\gamma_{26}=0.51$ & 
1600:                 $\gamma_{27}=-0.01$ & $E2=0.00$ & \\  \hline 
1601: \end{tabular}
1602: \caption{The influences of each considered topologic measurement
1603: on the attractors separation indices as revealed by path
1604: analysis considering $N=100$,$M=3$ and $m=3$.}\label{tab:pathan2}
1605: \end{table*}
1606: 
1607: 
1608: \section{Conclusions}
1609: 
1610: The investigation about the relationship between the structure and
1611: function of networks (e.g.~\cite{Newman03, Boccaletti05,
1612: Costa_Salou:2005}) represents one of the most interesting perspectives
1613: for obtaining insights about complex dynamic systems.  As briefly
1614: reviewed in this article, several works have addressed the problem of
1615: how the structure of the connectivity may affect and largely define
1616: the properties of dynamic systems.  One particularly important aspect
1617: which has received relatively lesser attention concerns the
1618: separability between different grandmother attractors, each
1619: representing a prototype pattern or state.  This issue is critically
1620: relevant because it has great impact on the capacity of the network
1621: for proper representation of patterns, the level of
1622: redundancy/robustness of such representations, the degree of
1623: generalization for recognition of not previously trained prototypes,
1624: as well as the effectiveness during retrieval and activation of such
1625: prototypes.  While other types of coding can be used in dynamic
1626: system, grandmother representation stands out as particularly
1627: important because it seems to be the way a great part of the primates
1628: cortex is organized.  
1629: 
1630: The current work has reported an approach to the characterization of
1631: the separability between prototype patterns which incorporates a
1632: number of special features.  First, we have considered four
1633: representative theoretic models of complex networks -- namely the
1634: random networks of Erd\H{o}-R\'enyi, the scale free model of
1635: Barab\'asi-Albert, the small-world networks of Watts-Strogatz, as well
1636: as a simple, non-small-world topographic model.  Second, by using a
1637: generic diffusive process, followed by the calculation of the
1638: respective derivative network in order to obtain the attraction
1639: basins, we obtained a methodology which completely avoids the
1640: intricacies and specifities implied by each type of dynamic system.
1641: While it remains to be verified how good accurate and general such an
1642: approximation is, it does allow the definition of smooth attraction
1643: basins around each prototype node in a way which is remindful of many
1644: important dynamical systems such as Kohonen's self-organizing maps and
1645: the primates cortex.  Once the attraction basins are so defined, a
1646: simple diffusive scheme emanating from each network node is employed
1647: in order to obtain the general activation of the network.  Although
1648: ideally such a procedure should activate equally only the original
1649: prototype nodes, the structured connectivity of the networks will act
1650: so as to produce non-uniform activation, where just a fraction of the
1651: overall activation coincides with the prototype nodes.  The
1652: disagreement between the original prototypes and the induced activity
1653: has been quantified in terms of two separation indices, namely the
1654: individual and neighborhood indices.  Because the former consider only
1655: the agreement and balance between the induced activation and the
1656: original prototypes, it resulted to be too small and therefore with
1657: low resolution for quantifying the attractors separation.  By
1658: considering also the immediate neighborhood of the prototype nodes,
1659: the second index allowed a more informative indication of the
1660: attractors separability.
1661: 
1662: The specific topologic features of each of the four considered
1663: theoretic network models can have different effects for the separation
1664: of the attraction basins.  Therefore, we performed an investigation of
1665: the effects of the most important parameters of the simulations over
1666: the respective performance.  We verified that the separation tends to
1667: decrease with the number of prototypes and increase with the size of
1668: the network.  At the same time, more intense general connectivity, as
1669: expressed by the average node degree, also tended to undermine the
1670: attractors separation.  Special attention was given to the finite
1671: sizes implied by the parameter $N$.  The obtained results seem to
1672: suggest that the separability tends to increase with $N$ up to a
1673: regime where the finite size of the network is no longer felt
1674: (`unsaturation').  Therefore, by identifying the region where such a
1675: plateau of separability is reached, it is possible to obtain
1676: near-optimum values of $N$ for given $M$ and $m$.  Next, we applied
1677: correlation analysis in orde not only to identify the
1678: inter-relationships between each topologic measurement, but also
1679: between these and the two separation indices.  The presence of such
1680: correlations can not only help to understand the origin of the
1681: attractors separation, but also allow the prediction of such a
1682: property from measurements of the network topology, without the need
1683: of simulations.  Several tendencies were identified through such an
1684: analysis, including the tendency of the index $s$ to be strongly
1685: proportional to the uniformity of the Voronoi areas, as expressed by
1686: the Voronoi index $Vi$.  A positive, though weaker, correlation was
1687: also identified between $s$ and the average shortest path length
1688: $\left<sp\right>$ of the networks.  A strong negative correlation was
1689: identified between $s$ and the matching index $mi$.  Although all such
1690: behaviors are compatible with what could be expected, meaningfully
1691: different correlations differences were observed for each network
1692: model.  Interestingly, the network model allowing the best overall
1693: separation of attractors was found to be the Watts-Strogatz, followed
1694: by the geographic, random and scale-free models.  It is conjectured
1695: here that the superior properties of the WS model stem from its
1696: enhanced uniformity and low randomness of local connectivity as well
1697: as by the fact that, although being a small world model, there are
1698: relatively very few long range connections between any pair of nodes.
1699: Similar properties, though at the expense of a weaker local order and
1700: uniformity of local connections, are characteristic of the geographic
1701: model, which came in second in performance.  As a matter of fact, the
1702: WS and GG models showed similar behavior as far as the histograms of
1703: separation index were concerned.  The same was observed for the ER and
1704: BA cases.  Therefore, from the perspective of attractors separation,
1705: the WS/GG and ER/BA models seem to represent two different classes of
1706: systems, with the latter providing rather poorer separability.
1707: Interestingly, the small world property can not explain such a
1708: partition, as the GG is not a small world model as the other three
1709: cases.  Consequently, it seems that the common denominators in the
1710: pairs WS/GG and ER/BA seems to be more strongly related also to the
1711: uniformity of the local connectivity.  Such an explanation would be
1712: largely in agreement with several of the results of works
1713: investigating the effect of connectivity on memory, as reviewed in
1714: Section~\ref{sec:review}.
1715: 
1716: In order to try to learn more about the effects of the topological
1717: features of the network on the dynamical property of attractors
1718: separation, path analysis was also applied assuming a simple
1719: structural relationship.  To our best knowledge, this is the first
1720: time such an insightful analysis has been applied for the study of
1721: complex networks.  One of the main advantages of such an approach over
1722: the Pearson correlation coefficients is that here the dispersions of
1723: \emph{all} the considered measurements are taken into account instead
1724: of the pairwise relationships underlying the correlations.
1725: Interestingly, the path analysis yielded influences of the topologic
1726: measurements on the separation indices which were often substantially
1727: different from those suggested by the Pearson correlation
1728: coefficients.  Of particular interest was the near null influence
1729: assigned to the average and standard deviation of the shortest path
1730: length between the prototype nodes.  This is all the most surprising
1731: as it seems intuitively reasonable to discuss much of the effect of
1732: the topology of the network over the attractors separation by
1733: considering such measurements.  More specifically, networks where the
1734: prototype nodes are topologically close one another would tend to have
1735: poorer separation.  However, the performed path analysis substantially
1736: emphasized, for all the four considered models, the importance of the
1737: average matching index on the attractors separation.  On second
1738: thoughts, this is indeed reasonable because the presence of non-zero
1739: matching index is a certain indication of overlap of the attraction
1740: basins.  Even when repeating the path analysis while leaving out the
1741: matching index measurements, the other parameters (especially the
1742: shortest path length) did not result with higher influences.  Other
1743: interesting insights, including influences which were specific to
1744: network models, were also allowed by the path analysis, corroborating
1745: therefore the potential value of such a statistical approach in order
1746: to get insights about the relationship between structure and dynamics
1747: of complex dynamic systems.  However, it is important to keep in mind
1748: that every statistical methodology should be understood as a source of
1749: insights to be further investigated and corroborated rather than
1750: spelling definitive facts.
1751: 
1752: Despite the relative comprehensiveness of the presently reported
1753: investigation, and perhaps as its consequence, a series of future
1754: developments can be suggested.  First, it would be interesting to
1755: extend the reported investigation to larger network sizes and to
1756: consider additional measurements such as the standard deviation of the
1757: clustering coefficient and the spectral structure of the adjacency and
1758: weight matrix.  Second, the sources of the deviations of the induced
1759: activation from the prototype nodes could be further investigated by
1760: considering simulations involving a single attractor.  In such cases,
1761: all the loss of `separability' would necessarily be a consequence of
1762: the probability leakage from the prototype node into its neighbors.
1763: It would be particularly interesting to verify, through path analysis,
1764: which of the topological measurements of the network would be more
1765: influence on the attractor activation.  It would also be worth
1766: investigating the potential of using the \emph{percolation
1767: transform}~\cite{Costa_percol:2004, Costa_bioinfo:2005} over the
1768: network as the means to explain and predict the attractors
1769: separability.  Other promising possibilities include the extension of
1770: the matching index to take into account hierarchical levels larger
1771: than one, i.e. including also the second and higher neighborhoods.
1772: 
1773: Several real problems are closely related to the issue of attractor
1774: separation and prototype activation.  One real problem which could be
1775: particularly interesting to be addressed is the phenomenon of
1776: \emph{facilitation} of neurons (represented by nodes).  By
1777: facilitation of a neuron it is meant that that neuron will become more
1778: likely to engage into activity.  For instance, the definition of
1779: temporary priorities in the primates brain could be related to the
1780: reinforcement of one or more attractors, so that they become more
1781: likely to be revised along time (e.g. through a random walk).
1782: Interestingly, as suggested by the current work, the facilitation of a
1783: given node would be highly dependent on its local connectivity.  Such
1784: studies could be eventually extended to higher level mental dynamics,
1785: such as those underlying attention and even pathologies
1786: (e.g.~\cite{Wedemann:2006}). It is also reasonable to expect that the
1787: definition of attractors is a process which co-evolves with the
1788: network topology.  In this sense, it would be interesting to try to
1789: identify growing schemes where the connectivity is affected by the
1790: success of the establishment of prototype attractores, and vice-versa.
1791: Another related investigation would be the identification of topologic
1792: organizations of networks allowing optimal or near-optimal attractors
1793: separation.
1794: 
1795: 
1796: \vspace{1cm}
1797: {\bf List of Symbols}
1798: 
1799: \begin{trivlist}
1800: 
1801: \item  $\Gamma =$ a graph or complex network;
1802: 
1803: \item  $N =$ number of nodes in a network;
1804: 
1805: \item  $K =$ the adjacency matrix of a complex network;
1806: 
1807: \item  $k(i) =$ degree of a network node $i$;
1808: 
1809: \item  $m =$ the parameter in the BA model defining its average
1810:                node degree.  This parameter is considered as
1811:                reference for all network models in this work;
1812: 
1813: \item  $diam =$ the diameter of the network;
1814: 
1815: \item  $cc(i) =$ clustering coefficient of a network node $i$;
1816: 
1817: \item  $sp(i,j)$ shortest path between nodes $i$ and $j$ in a complex network;
1818: 
1819: \item  $mi(i,j) = $ the matching index of nodes $i$ and $j$;
1820: 
1821: \item  $Vi =$ the Voronoi index of separation between the areas of
1822:               influence of the prototype nodes;
1823: 
1824: \item $\left< a \right> =$ the arithmetic average of the property $a$;
1825: 
1826: \item $\left[ a \right] =$ the geometric average of the property $a$;
1827: 
1828: \item $\sigma(a) =$ the standard deviation of the random variable $a$;
1829: 
1830: \item $M =$ the number of prototype patterns to be represented in a network;
1831: 
1832: \item $\vec{\alpha} =$ the activations in a network after a long 
1833:       random walk;
1834: 
1835: \item $s_{ind} =$ the separability index between prototype nodes in a 
1836:       network at the level of individual nodes;
1837: 
1838: \item $s =$ the separability index between prototype nodes in a network
1839:             considering also the immediate neighborhood of such nodes;
1840: 
1841: \end{trivlist}
1842: 
1843: 
1844: \begin{acknowledgments}
1845: Luciano da F. Costa thanks CNPq (308231/03-1) and FAPESP (05/00587-5)
1846: for sponsorship.
1847: \end{acknowledgments}
1848: 
1849: %\bibliographystyle{plain}
1850: %\bibliographystyle{unsrt}
1851: \bibliography{mem}
1852: 
1853: \end{document}
1854: