1: %\documentclass[a4paper]{article}
2: \documentclass[prl,onecolumn,floats,showpacs,superscriptaddress]{revtex4}
3: \usepackage{epsfig}
4: \usepackage{amsmath,amssymb,amsthm}
5: \usepackage{graphicx}% Include figure files
6: \usepackage{psfrag}% use latex text in ps figures
7: \usepackage{bm}% bold math
8: %\usepackage{showkeys}
9: %\documentclass[prl,twocolumn,floats,showpacs,superscriptaddress]{revtex4}
10: %\usepackage{psfig}
11: %\usepackage{amssymb,amsthm}
12: %\usepackage{amsmath,amssymb,amsthm}
13: %\usepackage{graphicx}% Include figure files
14: %\usepackage{psfrag}% use latex text in ps figures
15: %\usepackage{bm}% bold math
16: %\usepackage{showkeys}
17: \newcommand{\lam}{\lambda}
18: \newcommand{\sig}{\sigma}
19: \newcommand{\al}{\alpha}
20: \newcommand{\LN}{L'}
21:
22: \begin{document}
23: \title{Optimal network topologies: \\
24: Expanders, Cages, Ramanujan graphs, Entangled networks and all that.}
25:
26: \author{Luca Donetti}
27: \affiliation{
28: Departamento de Electr\'onica and \\
29: Instituto de F{\'\i}sica
30: Te{\'o}rica y Computacional Carlos I, Facultad de Ciencias, Universidad de
31: Granada, 18071 Granada, Spain}
32:
33:
34: \author{Franco Neri}
35:
36: \affiliation{Dipartimento di Fisica, Universit\`a di Parma,
37: Parco Area delle Scienze 7/A, 43100 Parma, Italy}
38:
39: \author{Miguel A. Mu{\~n}oz}
40:
41: \affiliation{Departamento de Electromagnetismo y F{\'\i}sica de la Materia and \\
42: Instituto de F{\'\i}sica Te{\'o}rica y Computacional Carlos I,
43: Facultad de Ciencias, Universidad de Granada, 18071 Granada, Spain}
44:
45:
46: \date{\today}
47: \begin{abstract}
48: We report on some recent developments in the search for optimal
49: network topologies. First we review some basic concepts on spectral
50: graph theory, including adjacency and Laplacian matrices, and paying
51: special attention to the topological implications of having large
52: spectral gaps. We also introduce related concepts as ``expanders'',
53: Ramanujan, and Cage graphs. Afterwards, we discuss two different
54: dynamical features of networks: synchronizability and flow of random
55: walkers and so that they are optimized if the corresponding Laplacian
56: matrix have a large spectral gap. From this, we show, by developing a
57: numerical optimization algorithm that maximum
58: synchronizability and fast random walk spreading are obtained for a
59: particular type of extremely homogeneous regular networks, with long
60: loops and poor modular structure, that we call {\it entangled
61: networks}. These turn out to be related to Ramanujan and Cage
62: graphs. We argue also that these graphs are very good finite-size
63: approximations to Bethe lattices, and provide almost or almost optimal
64: solutions to many other problems as, for instance, searchability in
65: the presence of congestion or performance of neural networks. Finally,
66: we study how these results are modified when studying dynamical
67: processes controlled by a {\it normalized} (weighted and directed)
68: dynamics; much more heterogeneous graphs are optimal in this case.
69: %While this second case might be more
70: %relevant for flow properties in real-world networks the previous,
71: %entangled graphs have a huge number applications in computer science
72: %and technological networks.
73: Finally, a critical discussion of the limitations and possible
74: extensions of this work is presented.
75: \end{abstract}
76:
77: \pacs{89.75.Hc,05.45.Xt,87.18.Sn}
78:
79: \maketitle
80:
81:
82:
83: \section{I. Introduction}
84:
85: Imagine you are to design a network, be it a local computer network,
86: an infrastructure or communication network, an artificial neural
87: network, or whatever analogous example you can think of. Suppose also
88: that you have some restrictions to do so. First of all, the number of
89: nodes and the total amount of links are both fixed and, second, nodes
90: should be connected using the available links in such a way that the
91: resulting network topology is {\it optimal} in some sense. Of course,
92: the meaning of the word ``optimal'' depends on the task to be
93: performed by the network, or in other words, depends on the nature of
94: the dynamical process to be built on top of it.
95:
96: Along this paper we study optimization of dynamical processes in
97: networks, as exemplified by the following three cases.
98:
99: Suppose we are constructing an artificial neural network, so at each
100: node we locate a neuron (i.e. an oscillator) having some unspecified
101: dynamical properties. In order to enhance the neural net performance
102: we want the oscillators to be easily synchronizable
103: \cite{synchro}, i.e. to be able to reach as easily as possible a state
104: in which them all, or at least a large fraction of them, fire
105: (oscillate) at unison \cite{Torres,BJK,GG}, and that state to be as
106: robust as possible.
107:
108: As a second example, imagine having a communication or technological
109: network and wanting, for the sake of efficiency, any node to be
110: ``nearby'' any other one. The simplest strategy to do so, is by
111: constructing a star-like topology with a central node (or ``hub'' in
112: the network jargon) directly connected to all the rest. In this way,
113: any node is reachable from any other within two steps at most.
114: However, in cases of intense traffic flow, this might not be very
115: efficient, especially if the central hub gets overburdened or
116: congestioned. Also, the star-like topology is very fragile to sabotage
117: to the central node. So one could wonder what is the optimal topology
118: avoiding the use of a privileged central hub
119: \cite{Catalans,Congestion,Maritan}.
120:
121: For a third example, let us consider information packets traveling in
122: a network in such a way that they disperse jumping randomly between
123: contiguous nodes and let us require an optimum flow of
124: information. For that, we define an ensemble of random walkers
125: diffusing on the net, and impose the averaged first-passage time to be
126: minimized. Which is the optimal structure? Analogously, other
127: random-walk properties as the mixing rate \cite{Lovasz}, the mean
128: transit time or the mean return time \cite{dani} could be minimized.
129:
130: In all these situations the problem to be solved is very similar:
131: finding an optimal topology under some constraints. These and similar
132: problems have been addressed in the literature, especially in the
133: context of Computer Science and, more recently, in the emerging field
134: of Complex Networks
135: \cite{Strogatz,Laszlo,Newman,Porto,AleRomu,Boccaletti}.
136:
137: In this paper, we illustrate that the answer to these problems may
138: have a large degree of universality in the sense that, even if the
139: optimal topologies in each case are not ``exactly'' identical, they
140: may share some common features that we review here. To do so, we
141: translate these problems into the one of optimizing some invariant
142: property of a matrix encoding the network topology. In particular, we
143: can use adjacency, Laplacian, and normalized-Laplacian matrices
144: depending on the particular problem under scrutiny. Spectral analysis
145: techniques are employed to explore in a systematic way network
146: topologies and to look for optimal solutions in each specific problem.
147: In this context, we introduce and characterize what we call {\it
148: entangled networks}: a family of nets with very homogeneous properties
149: and a extremely intertwined or entangled structure which are optimal
150: or almost optimal from different points of view.
151:
152: Along this work, we focus mainly on un-weighted, un-directed,
153: networks, not embedded in a geography or physical coordinates, and
154: characterized by identical nodes. Changing any of these restrictions
155: may alter the nature of the emerging optimal topologies as will be
156: briefly discussed at the end of the paper.
157:
158: The paper is structured as follows. In {\bf section II} we introduce
159: topological matrices and other basic elements of spectral graph
160: theory, as for instance the ``spectral gap''. In {\bf section III} we
161: discuss the topological implications of having a large spectral gap
162: and illustrate the connections of this with topological optimization
163: problems: graphs with large spectral gaps are optimal (or almost
164: optimal) for some dynamical processes defined on networks. In {\bf
165: section IV} we present an explicit construction of optimal solutions,
166: called by graph theorists ``Ramanujan'' graphs; such explicit
167: constructions cannot be achieved for any number of nodes and
168: links. Owing to this, in {\bf section V} we discuss how to construct
169: optimal graphs in general (for any given number of nodes and edges) by
170: employing a recently introduced computational optimization method,
171: enabling us to construct the so called ``entangled networks''. We
172: compare entangled networks with Ramanujan graphs in the cases when
173: these last can be explicitly constructed, as well as with ``cage
174: graphs'', another useful concept in graph theory. In {\bf section VI}
175: we enumerate some network design problems where our results are
176: relevant. In {\bf section VII} we briefly elaborate on the connection
177: of Ramanujan and entangled networks in the limit of large sizes with
178: the Bethe lattice. In {\bf section VIII} we discuss how the previous
179: results are affected by analyzing dynamical processes controlled by
180: the normalized Laplacian rather than by the Laplacian. Finally, in
181: {\bf section IX} a critical discussion of our main results,
182: conclusions, and future perspectives is presented.
183:
184:
185: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
186: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
188:
189:
190: \section{II. Elements of spectral graph theory: spectral gaps}
191:
192: For the sake of self-consistency and to fix notation we start by
193: revising some basic concepts in graph theory.
194:
195: Given a network (or graph), it is possible to define the corresponding
196: {\it adjacency matrix}, $A$, whose elements $a_{ij}$ are equal to $1$
197: if a link between nodes $i$ and $j$ exists and $0$ otherwise. A
198: related matrix is the {\it Laplacian}, $L$, which takes values $-1$
199: for pairs of connected vertices and $k_i$ (the degree of the
200: corresponding node $i$) in diagonal sites. Obviously, $L = K - A$,
201: where $K$ is the diagonal connectivity (or degree) matrix. If the
202: graph is undirected both $A$ and $L$ are symmetric matrices. It is
203: easy to see that $\lambda_1=0$ is a trivial eigenvalue of $L$ with
204: eigenvector $(1,1,\ldots)$ and that the eigenvalues $\lambda_i$
205: satisfy
206: \begin{equation*}
207: 0 = \lam_1 \le \lam_2 \le \ldots \le \lam_N \le 2k_{\text{max}} ,
208: \end{equation*}
209: where $k_{\text{max}}$ is the largest degree in the graph. The proofs
210: of these and other elementary spectral graph properties can be found,
211: for instance, in \cite{Chung,Bollobas}. Finally, a {\it normalized
212: Laplacian matrix}, $\LN=K^{-1}L$, can be defined; each row is equal to
213: the analogous row in $L$ divided by the degree of the corresponding
214: node. This is useful to describe dynamical processes in which the
215: total effect of neighbors is equally normalized for all sites, while
216: in the absence of such a normalization factor sites with higher
217: connectivity are more strongly coupled to the others than loosely
218: connected ones. In this case we have $0\le\lambda'_i\le2$ for
219: $i=1\ldots N$, where the $\lambda'_i$ denote the eigenvalues of $\LN$.
220: Clearly, for {\it regular} graphs, that is, graphs where all nodes
221: have the same degree $k$, the two matrices $L$ and $\LN$ differ by a
222: multiple of the identity matrix and the eigenvalues are related by
223: $\lambda_i=k \lambda'_i$.
224:
225: Following the literature, here we consider mainly (but not only) the
226: Laplacian spectrum, but depending on the dynamical process under
227: consideration this choice will be changed.
228:
229: In order to give a first taste on the topological significance of
230: spectral properties, let us consider a network perfectly separated
231: into a number of independent subsets or {\it communities} having only
232: intra-subset links but not inter-subset connections. Obviously, its
233: Laplacian matrix (either the non normalized or the normalized one) is
234: block-diagonal. Each sub-graph has its own associated sub-matrix, and
235: therefore $0$ is an eigenvalue of any of them. In this way, the
236: degeneration of the lowest (trivial) eigenvalue of $L$ (or of $\LN$)
237: coincides with the number of disconnected subgraphs within a given
238: network.
239: For each of the separated components, the corresponding eigenvector
240: has constant components within each subgraph.
241:
242: On the other hand, if the subgraphs are not perfectly separated, but a
243: small number of inter-subset links exists, then the degeneracy will be
244: broken, and eigenvalues and eigenvectors will be slightly
245: perturbed. In particular relatively small eigenvalues will appear, and
246: their corresponding eigenvectors will take ``almost constant'' values
247: within each subgraph. For instance, if the number of subsets is $2$,
248: spectral methods have been profusely used for graph bipartitioning or
249: graph bisecting \cite{GN} as follows. One looks for the smallest
250: non-trivial eigenvalue; its corresponding eigenvector should have
251: almost constant components within each of the two subgroups, providing
252: us with a bipartitioning criterion. Therefore a graph with a ``small''
253: first non-trivial Laplacian eigenvalue, $\lambda_2$, customarily
254: called {\it spectral gap} (or also {\it algebraic connectivity}) has a
255: relatively clean bisection. In other words, the smaller the spectral
256: gap the smaller the relative number of edges required to be cut-away
257: to generate a bipartition. Conversely a large ``spectral gap''
258: characterizes non-structured networks, with poor modular structure, in
259: which a clearcut separation into subgraphs is not inherent. This is a
260: crucial topological implication of spectral gaps.
261:
262: A closely related concept is the {\it expansion property}. To
263: introduce it, let us consider a generic graph, $X$, and define the
264: {\it Cheeger or isoperimetric constant} as follows. First, one
265: considers all the possible subdivisions of the graph in two disjoint
266: subsets of vertices: $A$ and its corresponding complement $B$. The
267: Cheeger constant, $h(X)$, is defined as the minimum value over all
268: possible partitions of the number of edges connecting $A$ with $B$
269: divided by the number of sites in the smallest of the two subsets. The
270: Cheeger constant is larger than $0$ if and only if the graph is
271: connected, and is ``large'' if any possible subdivision has many links
272: between the two corresponding subsets. Hence, for graphs with poor
273: community structure, the Cheeger constant is large.
274:
275:
276: In the graph-theory literature the concept of {\it expander} is
277: introduced within this context \cite{Sarnak,Valette}: a family of
278: expanders is a family of {\it regular graphs} with degree $k$ and $N$
279: nodes, such that for $N \rightarrow \infty$ the Cheeger constant $h$
280: is always larger than a given positive number $\epsilon$. Note that
281: the word ``expansion'', refers to the fact that the topology of the
282: network-connections is such that any set of vertices connects in a
283: robust way (``expands through'') all nodes, even if the graph is
284: sparse. Obviously, for this to happen, $k$ should be larger or equal
285: than $3$, as for $k=2$ all graphs are linear chains and can be,
286: therefore, cut in two subgraphs by removing just one edge (arbitrarily
287: small Cheeger constant for large system sizes).
288:
289: The Cheeger constant and the expansion property can be formally
290: related to the spectral gap by the following inequalities \cite{AB}:
291: \begin{equation}
292: {\lam_2 \over 2} \leq h(X) \leq \sqrt{2k\lam_2}.
293: \label{inequality}
294: \end{equation}
295: Therefore, {\it a family of regular graphs is an expanding family if
296: and only if it has a lower bound for the spectral gap, and the
297: larger the bound the better the expansion}. Among the applications
298: of expander graphs are the design of efficient communication networks,
299: construction of error-correcting codes with very efficient encoding
300: and decoding algorithms, de-randomization of random algorithms, and
301: analysis of algorithms in computational group theory \cite{Sarnak}.
302: % Note that the optimal gap could not correspond to the optimal Cheeger constant.
303: % It's just a bound.
304:
305:
306: As a consequence of this, expansion properties are enhanced upon
307: increasing the spectral gap, but spectral gaps cannot be as large as
308: wanted, especially for large graph sizes. In the case of $k$-regular
309: graphs (the ones for which expanders have been defined) there is an
310: asymptotic upper bound for it derived by Alon and Boppana \cite{AB},
311: given by the spectral gap of the infinite regular tree of degree $k$
312: (called Bethe tree or Bethe lattice):
313: \begin{equation}
314: \lim_{N \rightarrow \infty} \lam_2 \leq k - 2 \sqrt{k-1}.
315: \label{inequality2}
316: \end{equation}
317:
318: Whenever a graph has $ \lam_2 \geq k - 2 \sqrt{k-1}$ (i.e. when the
319: spectral gap is larger than its asymptotic upper bound), it is called
320: a {\it Ramanujan graph}. Therefore, Ramanujan graphs are very good
321: expanders. In one of the following sections we will tackle the
322: problem of explicitly constructing Ramanujan graphs by following the
323: recently developed mathematical literature on this subject.
324:
325:
326: So far we have related the spectral graph to the ``compactness'' of a
327: given graph; a large gap characterizes networks with poor modular
328: structure, in which it is difficult to isolate sub-sets of sites
329: poorly connected with the rest of the graph or, in other words, large
330: gaps characterize expanders. In the following section we shift our
331: attention from topology to dynamics and discuss some other interesting
332: problems which turn out to be directly related to large spectral gaps.
333:
334:
335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
336: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
337: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
338:
339:
340: \section{III. Large spectral gaps and dynamical processes}
341:
342: In this section we review two different dynamical processes defined on
343: the top of networks and discuss how some of their properties depend on
344: the underlying graph topology \cite{Entangled}.
345:
346:
347: \subsection{Synchronization of dynamical processes}
348:
349: An aspect of complex networks that has generated a burst of activity
350: in the last few years, because of both its conceptual relevance and
351: its practical implications, is the study of {\it synchronizability} of
352: dynamical processes occurring at the nodes of a given network. How
353: does synchronizability depend upon network topology? Which type of
354: topology optimizes the stability of a globally synchronized
355: state\cite{synchro}?
356:
357: A first partial answer to this question was given in a seminal work by
358: Barahona and Pecora \cite{Pecora} who established the following
359: criterion to determine the stability of fully synchronized states on
360: networks. Consider a general dynamical process
361: \begin{equation}
362: \dot{x}_i = F(x_i) + \sig \sum_{j\in n.n. i} [ H(x_j) -H(x_i)]=
363: F(x_i) - \sig \sum_j L_{ij} H(x_j),
364: \label{pecora}
365: \end{equation}
366: where $x_i$ with $i \in {1,2,... ,N}$ are dynamical variables, $F$ and
367: $H$ are an evolution and a coupling function respectively, and
368: $\sigma$ is a constant. A standard linear stability analysis can be
369: performed by i) expanding around a fully synchronized state
370: $x_1=x_2=\ldots=x_N=x^s$ with $x^s$ solution of $\dot{x^s} = F(x^s)$,
371: ii) diagonalizing $L$ to find its $N$ eigenvalues, and iii) writing
372: equations for the normal modes $y_i$ of perturbations
373: \begin{equation}
374: \dot{y}_i =
375: \left[ F'(x^s) - \sig
376: \lam_i H'(x^s) \right] y_i,
377: \label{linear}
378: \end{equation}
379: all of them with the same form but different effective couplings
380: $\al=\sig \lam_i$. Barahona and Pecora noticed that the maximum
381: Lyapunov exponent for Eq. (\ref{linear}) is, in general, negative only
382: within a bounded interval $[\alpha_A,\alpha_B]$, and that it is a
383: decreasing (increasing) function below (above) (see fig. 1 in
384: \cite{Pecora}, and see also the related work by Wang and Chen
385: \cite{Wang} in which the case $[\alpha_A,\infty]$ is studied). Requiring all
386: effective couplings to lie within such an interval, $ \al_A < \sig
387: \lam_2 \le \ldots \le \sig \lam_N <
388: \al_B$, one concludes that a synchronized state is linearly stable if
389: and only if $ \frac{\lam_N}{\lam_2} < \frac{\alpha_B}{\alpha_A}$ for
390: the corresponding network. It is remarkable that the left hand side
391: depends only on the network topology while the right hand side depends
392: exclusively on the dynamics (through $F$ and $G$, and $x^s$).
393:
394: As a conclusion, the interval in which the synchronized state is
395: stable is larger for smaller eigenratios $\lam_N/\lam_2$, and
396: therefore one concludes that {\it a network has a more robust
397: synchronized state if the ratio $Q = \lam_N/\lam_2$ is as small as
398: possible} \cite{Wang}. Also, as the range of variability of $\lam_N$
399: is limited (it is related to the maximum connectivity \cite{Mohar})
400: minimizing $Q$ gives very similar results to maximizing the
401: denominator $\lam_2$ in most cases. Indeed, as argued in \cite{Wang}
402: in cases where the maximum Lyapunov exponent is negative in an
403: un-bounded from above interval, the best synchronizability is obtained
404: by maximizing the spectral gap.
405:
406: It is straightforward to verify that for normalized dynamics, i.e.,
407: problems where a quotient $k_i$ appears in the coupling function in
408: Eq.(\ref{pecora}), the Laplacian eigenvalues have to be replaced by
409: the normalized-Laplacian ones, so the gap refers to $\lambda'_2$ and
410: $Q$ becomes $Q_{norm}= \lam'_N/\lam'_2$.
411:
412: Summing up, {\it large spectral gaps favor stability of synchronized
413: states (synchronizability)}.
414:
415: \subsection{Random walk properties}
416:
417: One of the first and most studied models on graphs are random walks:
418: these are important both as a simple models of dispersion phenomena
419: and as a tool for exploring graph properties.
420:
421: Indeed random walk properties are strictly related to the underlying
422: network structure and, in particular, to the eigenvalue spectrum of
423: the previously introduced matrices. At each step the transition
424: probability from vertex $i$ to vertex $j$ is trivially given by
425: $P_{ij} = {A_{ij} \over k_i}$. This defines a transition matrix $P$
426: for random walk dynamics, that can be written as $P= K^{-1} A =
427: K^{-1} (K - L) = I -\LN$ (where $I$ the identity matrix); therefore
428: the eigenvectors of $P$ and $\LN$ coincide and the eigenvalues are
429: linearly related.
430:
431: In particular, the stationary probability distribution of the random
432: walk on a graph is given by the eigenvector corresponding to the
433: largest eigenvalue (1) of $P$, corresponding to $\lambda'=0$. Then it
434: is easy to see that the convergence rate of a given initial
435: probability distribution towards its stationary distribution (also
436: called ``mixing rate'') is controlled by the second largest eigenvalue
437: of $P$, and therefore by the spectral gap of the normalized Laplacian
438: matrix: the larger $\lambda'_2$ the faster the decay
439: \cite{Nielsen,Lovasz}.
440:
441:
442: Another relevant property is the first passage time, $\tau_{i,j}$
443: between two sites $i$ and $j$, defined as the average time it takes
444: for a random walker to arrive for the first time to $j$ starting from
445: $i$. It can be expressed in terms of the eigenvectors ($u_k$) and
446: eigenvalues of the normalized Laplacian matrix as
447: \begin{equation*}
448: \tau_{i,j} = 2 M \sum_{l=2...N} { \left(\frac{u_{l,i}}{\sqrt{k_i}}
449: -\frac{u_{l,j}}{\sqrt{k_j}}\right)^2 \frac1{\lambda'_l} },
450: \end{equation*}
451: where $M$ is the number of link of the network (see \cite{Lovasz}).
452: For $k-$regular graphs the expression of its graph average, $\tau$,
453: can be simplified in the following way:
454: \begin{equation}
455: \tau = 2 k \frac{N}{N-1} \sum_{l=2...N} \frac1\lambda_l
456: \label{fp}
457: \end{equation}
458: A large $\lambda_2$ (which imply that also the following eigenvalues
459: cannot be small) gives an important contribution for keeping $\tau$ small.
460: Therefore large spectral gaps are associated with short first-passage times.
461:
462: Finally, random walks move around quickly on graphs with large
463: spectral gap in the sense that they are very unlikely to stay long
464: within a given subset of vertices, $A$, unless its complementary
465: subgraph is very small \cite{Nielsen,Lovasz}. This idea has been
466: exploited by Pons and Latapy \cite{PL} to detect communities
467: structures in networks by associating them to regions where random
468: walks remain trapped for some time. In conclusion: random walks escape
469: quickly from any subset $A$ if the spectral gap is large.
470:
471: Summing up: {\it random walks move and disseminate fluently in large
472: spectral gap graphs}. Related results, more details, and proofs of
473: these theorems can be found, for example in \cite{Nielsen,Lovasz}.
474:
475:
476: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
479:
480: \section{IV. Explicit construction of Ramanujan Graphs}
481:
482: The summary of the preceeding section is that if we are aimed at
483: designing network topologies with good synchronizability or
484: random-walk flow properties, we need criteria to construct graphs with
485: large spectral gaps. On the other hand, as explained before,
486: Ramanujan graphs are optimal expanders, in the sense that a given
487: family of them, with growing $N$, will converge asymptotically from
488: above to the maximum possible value of the spectral gap. Even if this does
489: not imply that for finite arbitrary values of $N$ they are optimal,
490: i.e. that they have the {\it largest} possible spectral gap, they
491: provide a very useful approach for the optimization of the spectral
492: gap problem. Therefore, a good way to design optimal networks so is
493: to explicitly construct Ramanujan graphs by following the mathematical
494: literature on this respect \cite{Valette,LPS,Margulis,Chiu}.
495:
496: In this section we present a recipe for constructing explicitly
497: families of Ramanujan graphs while, in the following one, we will
498: construct large-spectral gap networks by employing a computational
499: optimization procedure. Readers not interested in the mathematical
500: constructions can safely skip this section.
501:
502:
503: \subsection{The recipe}
504:
505: In recent years some explicit methods for the construction of
506: Ramanujan graphs have appear in the mathematical literature
507: \cite{LPS,Margulis}. We describe here only one of them.
508: While the proof that these graphs, constructed by Lubotzky, Phillips
509: and Sarnak \cite{LPS} (and independently by Margulis \cite{Margulis})
510: are Ramanujan graphs is a highly non-trivial one, their construction,
511: following the recipe we describe in what follows, is relatively simple
512: and can be implemented without too much effort.
513:
514:
515: Let us consider a given group $G$ and let $S \in G$ be a subset of
516: group elements not including the identity. The {\it Cayley graph}
517: associated with $G$ and $S$ is defined as the directed graph having
518: one vertex for each group element and directed edges connecting them
519: whenever one goes from one group element (vertex) to the other by
520: applying a group transformation in $S$. The absence of the identity in
521: $S$ guarantees that self-loops are absent. The Cayley graph depends on
522: the choice of the generating set $S$ and it is connected if and only
523: if $S$ generates $G$. Indeed, this will be the only case we consider
524: here. Note that if for each element $s \in S$ its inverse $s^{-1}$
525: also belongs to $S$, then the associated Cayley graph becomes
526: undirected, which is the case in all what follows.
527:
528: In the construction we discuss here, the group $G$ is given by
529: either by $PGL(2,Z/qZ)$ or by $PSL(2,Z/qZ)$. Consider the group of $2
530: \times 2$ matrices with elements in $Z/qZ$ (i.e. integer numbers
531: modulo $q$, where $q$ is an odd prime number); the elements of
532: $PGL(2,Z/qZ)$ ($PSL(2,Z/qZ)$) are the equivalence classes of matrices
533: with non-vanishing determinant (determinant equal to one) with respect
534: to multiplications by multiples of the identity matrix (i.e. to
535: matrices differing in multiples of the identity are considered to be
536: equal).
537:
538: To specify the subset $S$ of Cayley graph generators, the integral
539: quaternions $H(Z) = \{ \alpha = a_0 + a_1 i + a_2 j + a_3 k: a_i \in Z
540: \} $ must be introduced. Quaternions can be casted in a matrix-like
541: form as \cite{Valette}:
542: \begin{equation}
543: \left( \begin{array}{cc}
544: a_0 + a_1 x +a_3 y & -a_1 y + a_2 + a_3 x \\
545: -a_1 y -a_2 + a_3 x & a_0 - a_1 x -a_3 y \end{array} \right)
546: \label{matrix}
547: \end{equation}
548: where $x,y$ are odd prime integers, satisfying $(x^2 + y^2 +1)\mod q =
549: 0$. Writing quaternions in this way, it turns out that their algebra
550: coincides with the standard algebra of matrices and the quaternion
551: norm $|\al|^2 = a_0^2+a_1^2+a_2^2+a_3^2$ is equal to the corresponding
552: matrix determinant. Now another odd prime number $p$ has to be
553: chosen. It can be shown that there are $8 (p+1)$ elements in $H(Z)$
554: satisfying $|\al|^2=p$. If $p \mod 4 = 1$ then by taking $a_0$ odd and
555: positive the number of solutions is reduced to $p+1$, while if $p \mod
556: 4 = 3$ then the number of solutions is reduced to $p+1$ by taking
557: $a_0$ even and requiring the first non-zero component to be positive.
558:
559: If $p$ is a perfect square modulo $q$ such solutions can be mapped to
560: a set $S$ of $p+1$ matrices belonging $PSL(2,Z/qZ)$ (to guarantee that
561: all such matrices are distinct, one can take $q>2\sqrt{p}$). A
562: Ramanujan graph is then built as the Cayley graph with $G=PSL(2,Z/qZ)$
563: using the set $S$ of generators we just constructed; the number of
564: nodes is given by $N=q(q^2-1)/2$ (the number of elements of
565: $PSL(2,Z/qZ)$) and the degree of each node is $k=p+1$ (the elements of
566: $S$). If, otherwise, $p$ is not a perfect square modulo $q$ we map the
567: solutions to a set $S$ of $PGL(2,Z/qZ)$ matrices and a Ramanujan graph
568: is build as the Cayley graph with $G=PGL(2,Z/qZ)$ and the set $S$ just
569: constructed. In this case $N=q(q^2-1)$ and $k=p+1$.
570:
571: Putting all this together, Ramanujan graphs can be built by the
572: following procedure \cite{Code,Valette}:
573:
574: \begin{itemize}
575: \item choose values of the odd prime numbers $q$ and $p$, with $q>2\sqrt{p}$;
576:
577: \item associate the matrices of $PGL(2,Z/qZ)$ or $PSL(2,Z/qZ)$
578: (depending on the relative values of $p$ and $q$) with graph nodes;
579:
580: \item find the solutions of $a_0^2+a_1^2+a_2^2+a_3^2=p$, a solution of
581: $(x^2 + y^2 +1)\mod q = 0$ and construct the matrices according to
582: (\ref{matrix});
583:
584: \item find the neighbors of each node by multiplying the corresponding
585: matrix by the just constructed $p+1$ matrices, and represent them by
586: edges.
587:
588: \end{itemize}
589:
590: In this way we construct regular graphs with a large spectral gap,
591: with some restrictions on the number of nodes $N$ and the degree $k$:
592: the degree can only be an odd prime number plus one, while the number
593: of nodes grows as the third power of $q$.
594:
595:
596:
597: \subsection{Topological properties of the resulting graphs}
598:
599: First of all, every Cayley graph is by construction a regular graph,
600: and so are the Ramanujan graphs built following the above procedure.
601: Moreover, Lubotzky, Phillips and Sarnak \cite{LPS} proved that
602: these graphs have a large girth (the {\it girth}, $g$, of a graph is
603: the length of the shortest loop or cycle, if any), providing a lower
604: bound for the $g$ which grows logarithmically with $q$ (its exact form
605: is different in the $PSL$ and $PGL$ cases). To gain some more
606: intuition about the properties of this family of graph, we plot in the
607: left of figure (\ref{Rama}) the smallest one, corresponding to $k=4$
608: ($p=3$) and $120$ nodes ($q=5$).
609: \begin{figure}
610: \centerline{\psfig{file=r3-5.eps,width=6.0cm}
611: %\hspace{0.5cm}
612: %\vspace{0.0cm}
613: \psfig{file=r2-3.eps,width=7.0cm}}
614: % \includegraphics[width=.55 \linewidth]{r3-5.eps}
615: %\hspace{2cm}
616: %\includegraphics[width=.55 \linewidth]{r2-3.eps}
617: %\vspace{-0.3cm}
618: \caption{
619: Ramanujan graphs with (left) degree $k=4$ ($p=3$) and $120$ nodes
620: ($q=5$), and (right) $k=3$ ($p=2$) and $24$ nodes ($q=3$).}
621: \label{Rama}
622: \end{figure}
623:
624: Some remarks on its topological properties are in order. The average
625: distance $3.714$ is relatively small (i.e. one reaches any of the
626: $120$ nodes starting from any arbitrary origin in less than four
627: steps, on average). The betweenness centrality \cite{betweenness} $
628: 161.5$ is also relatively small, and takes the same value for all
629: nodes. The clustering coefficient vanishes, reflecting the absence or
630: short loops (triangles), and the minimum loop size is large, equal to
631: $6$, and identical for all the nodes. In a nutshell: {\it the network
632: homogeneity is remarkable; all nodes look alike, forming rather
633: intricate, decentralized structure, with delta-peak distributed
634: topological properties.}
635:
636:
637: \subsection{Extension to $k=3$}
638:
639: P. Chiu, extended the previous method to consider also graphs with
640: degree $3$ (i.e. $p=2$; note that the previous construction was
641: restricted to odd prime values of $p$). To do this, one just needs to
642: consider the following $3$ generators \cite{Chiu}:
643: \begin{equation}
644: \left( \begin{array}{cc}
645: 1 & 0 \\ 0 & -1 \end{array} \right), \left( \begin{array}{cc}
646: 2+\sqrt{-2} & \sqrt{-26} \\ \sqrt{-26} & 2-\sqrt{-2} \end{array}
647: \right), \left( \begin{array}{cc} 2-\sqrt{-2} & -\sqrt{-26} \\ -
648: \sqrt{-26} & 2+\sqrt{-2} \end{array} \right),
649: \label{matrix7}
650: \end{equation}
651: acting on the elements of $PGL(2,Z/qZ)$ or $PGL(2,Z/qZ)$ as before.
652:
653: In the right part of figure (\ref{Rama}) we show the smallest
654: Ramanujan graph with degree $k=3$ ($p=2$) and $24$ nodes ($q=3$).
655: This is slightly less homogeneous than the one with $k=4$, but it is
656: as homogeneous as possible given the previous values of $N$ and
657: $k$. The average distance is also small in this case, $3.13$, the
658: betweenness is $24.50$ for all nodes, the clustering vanishes, and
659: the size of the minimum loop is $4$. While for edges, the edge
660: betweenness \cite{betweenness} is $27.31$ for $12 $ edges while it is
661: $22.34$, for the remaining $24$ edges; indicating that the net is
662: not fully homogeneous in this case, but not far from homogeneous
663: either.
664:
665: Summing up, the main topological features of these Ramanujan graphs is
666: that they are very homogeneous: properties as the average distance,
667: betweenness, minimum loop size, etc are very narrowly distributed
668: (they are delta peeks in many cases). Also, compared to generic random
669: regular graphs, their averaged distance is smaller and the
670: average minimum loop size is larger. A main limitation of Ramanujan
671: graphs constructed in this way is that, for a fixed connectivity, the
672: possible network sizes are restricted to specific, fast growing
673: values.
674:
675:
676: \section{V. Construction of networks by computational optimization}
677:
678: An alternative route to build up optimal networks with an arbitrary
679: number of nodes and an as-large-as-possible spectral gap or (almost
680: equivalently) an as-small-as-possible synchronizability ratio
681: $Q=\lambda_N/\lambda_2$ is by employing a computational optimization
682: process. Note, that enumerating all possible graphs with fixed $N$ and
683: $\langle k \rangle$, and looking explicitly for the minimum value of
684: $Q$ or largest gap, is a non-polynomial problem and, hence,
685: approximate optimization approaches are mandatory \cite{Entangled}.
686: In this section, we focus on optimizing the Laplacian spectral gap
687: while in section VIII we will tackle the normalized Laplacian case and
688: discuss the differences between the two of them.
689:
690: \subsection{The algorithm}
691:
692: The idea is to implement a modified {\it simulated annealing}
693: algorithm \cite{SA} which, starting from a random network with the
694: desired number of nodes $N$ and average connectivity-degree $ \langle
695: k \rangle$, and by performing successive rewirings, leads
696: progressively to networks with larger and larger spectral gaps or
697: smaller and smaller eigenratios $Q$. As the spectral gap will be very
698: large in the emerging networks, they will be typically Ramanujan
699: graphs (whenever they are regular).
700:
701: The computational algorithm is as follows \cite{Entangled}. At each
702: step a number of rewiring trials is randomly extracted from an
703: exponential distribution. Each of them consists in removing a randomly
704: selected link, and introducing a new one joining two random nodes
705: (self-loops are not allowed). Attempted rewirings are (i) rejected if
706: the updated network is disconnected, and otherwise (ii) accepted if
707: $\delta Q= Q_{final}-Q_{initial} <0$, or (iii) accepted with
708: probability
709: \cite{Penna} $p = \min\left( 1, [1-(1-q) \delta Q/T]^{1/(1-q)}\right)$
710: (where $T$ is a temperature-like parameter) if $\delta Q \geq 0$. Note
711: that, in the limit $q \to 1$ we recover the usual Metropolis
712: algorithm. Instead we choose $q=-3$ which we have verified to give the
713: fastest convergence, although the output does not essentially depend
714: on this choice \cite{Penna}. The first $N$ rewirings are performed at
715: $T=\infty$. They are used to calculate a new $T$ such that the largest
716: $\delta Q$ among the first $N$ ones would be accepted with some large
717: probability: $T=(1-q)\cdot(\delta Q)_{max}$. Then $T$ is kept fixed
718: for $100N$ rewiring trials or $10N$ accepted ones, whichever occurs
719: first. Afterwards, $T$ is decreased by $10\%$ and the process iterated
720: until there is no change during $5$ successive temperature steps,
721: assuming that a (relative) minimum of $Q$ has been found. It is
722: noteworthy that most of these details can be changed without affecting
723: significatively the final results, while two algorithm drawbacks are
724: that the calculation of eigenvalues is slow and that the dynamic can
725: get trapped into ``metastable states'' (corresponding to local but not
726: global extreme values) as we illustrate in \cite{Entangled}.
727:
728: Running the algorithm for different initial networks, we find that
729: whenever $N$ is small enough (say $N \lesssim 30$), the
730: output-topology is unique in most of the runs, while some dispersion
731: in the outputs is generated for larger $N$ ($N=2000$ is the larger
732: size we have optimized). This is an evidence that the $Q$ absolute
733: minimum is not always found for large graphs, and that the evolving
734: network can remain trapped in metastable states. In any event, the
735: final values of $Q$ are very similar from run to run, starting with
736: different initial conditions, as shown in fig.~\ref{Evolution}, which
737: makes us confident that a reasonably good and robust approximation to
738: the optimal topology is typically obtained, though, strictly speaking,
739: we cannot guarantee that the optimal solution has been actually found,
740: especially for large values of $N$.
741: \begin{figure}
742: \centerline{
743: \psfig{file=fig1.eps,width=8.5cm}}
744: \caption{Eigenvalue ratio, $Q$ as a function of the number of algorithmic
745: iterations, starting from different initial conditions (random
746: network, small-world, linear chain, and scale free network) with
747: $N=50$, and $\langle k \rangle=4$. The algorithm leads to topologies
748: as the one depicted in figure 4.}
749: \label{Evolution}
750: \end{figure}
751:
752:
753: \subsection{Topological properties of the emerging network}
754: In figures (\ref{Cages}) and (\ref{Entangled}) we illustrate the
755: appearance of the networks emerging out of the optimization procedure,
756: that we call {\it entangled networks}, for different values of $N$ and
757: $\langle k \rangle$.
758: \begin{figure}
759: \centerline{
760: \psfig{file=petersen.ps,width=3.5cm}
761: \hspace{2cm}
762: \psfig{file=mcgee.ps,width=3.5cm}}
763: \caption{Optimal, entangled, networks obtained as output of the optimization
764: procedure for $k =3$ with $N=10$ and $N=24$. The left one is a
765: Petersen cage graph (k=3, girth=5). The one to the right is a McGee
766: cage graph (k=3, girth=7).}
767: \label{Cages}
768: \end{figure}
769: Remarkably, for some small values of $N$ and $k$ (see figure
770: (\ref{Cages})), it is possible to identify the resulting optimized
771: networks with well-known ones in graph theory: {\it cage graphs}
772: \cite{Cages,Bollobas}.
773:
774: A $(k,g)$-cage graph is a $k$-regular graph with girth $g$ having the
775: minimum possible number of nodes. Cage graphs have a vast number of
776: applications in computer science and theoretical graph analysis
777: \cite{Cages,Bollobas}. For $k=3$ and $N=10, 14$, and $24$,
778: respectively, the optimal nets found by the algorithm are cage-graphs
779: with girth $5$, $6$, and $7$ respectively (called $Petersen$,
780: $Heawood$ and $McGee$ graphs in the mathematical literature; see
781: fig.(\ref{Cages}) and also the nice picture gallery and mathematical
782: details in \cite{Cages}). We also recover other cage graphs for small
783: values of $N$ for $k=4$ and $k=5$.
784: \begin{figure}
785: \centerline{
786: \psfig{file=ramanujan.eps,width=5.0cm}}
787: \caption{Optimal, entangled, network obtained as output of the optimization
788: procedure for $k=4$ with $N=50$.}
789: \label{Entangled}
790: \end{figure}
791: For some other small values of $N$, cage graphs do not exist. For
792: example, for $k=3$ and girths, $3, 4, 5, 6$, and $7$ the Cage graphs have
793: the sizes: $N=4, 6, 10, 14$, and $24$ respectively
794: \cite{Cages}. Therefore, for sizes as $N=12$ or $N=16$ there is no
795: cage with $k=3$. In these cases, our optimization procedure leads to
796: graphs similar to cages (see figure
797: \ref{NoCages}) in which the shortest loops are as large as possible,
798: and all of them have very similar lengths. In this sense, our family
799: of optimal graphs provides us with topologies ``interpolating''
800: between well known cage graphs.
801:
802: \begin{figure}
803: \centerline{
804: \psfig{file=g12-3.ps,width=3.5cm}
805: \hspace{2cm}
806: \psfig{file=g16-3.ps,width=3.5cm}}
807: \caption{Optimal, entangled, networks obtained as output of the optimization
808: procedure for $k =3$ with $N=12$ and $N=16$. They do not correspond to
809: cage-graphs.}
810: \label{NoCages}
811: \end{figure}
812:
813:
814: In the emerging entangled networks short loops are severely suppressed
815: as said before. This can be quantified by either the {\it girth} or
816: more accurately by the average length, $\langle \ell \rangle$, of the
817: shortest loop from each node. In particular, the clustering
818: coefficient (measuring the number of triangular loops in the net)
819: vanishes, as loops have typically more than three edges.
820:
821:
822: To have a more precise characterization of the emerging topologies,
823: especially for larger graphs (see figure (\ref{Entangled})), we
824: monitored different topological properties during the optimization
825: process, as shown in figure (\ref{evolution}).
826: \begin{figure}
827: \centerline{
828: \psfig{file=fig2.eps,width=10.0cm}}
829: \caption{Relation between the ratio $Q$ and (i) node-degree standard
830: deviation, (ii) betweenness standard deviation, (iii) average
831: node-distance, and (iv) average betweenness. The subscript ``norm''
832: stands for normalization with respect to the respective mean-values,
833: centering all the measured quantities around $1$. They all decrease on
834: average as the optimization procedure goes.}
835: \label{evolution}
836: \end{figure}
837: The node-degree standard deviation typically decreases as $Q$
838: decreases meaning that the optimal topology approaches a regular one.
839: The betweenness standard deviation also decreases with $Q$ implying
840: that optimal networks tend to be as homogeneous as possible, and that
841: all nodes play essentially the same role in flow properties. Also,
842: the average node distance and average betweenness are progressively
843: diminished on average: nodes are progressively closer to each other
844: and the network becomes less centralized as the optimization procedure
845: runs. It has also been numerically verified that the average first
846: passage times of random walk is reduced during the network
847: optimization.
848:
849: The fact that the betweenness distribution is very narrow, indicates
850: that (following an original idea by Girvan and Newman \cite{GN}) it is
851: very difficult to divide the network into communities. Indeed, in the
852: algorithm proposed in \cite{GN} to identify communities, links with
853: high betweenness centrality are progressively cut out; if all links
854: have the same centrality the method becomes useless and communities
855: are hardly distinguishable.
856:
857:
858: We have named the emerging structures {\it entangled networks}, to
859: account for their very intricate and interwoven topology, with
860: extremely high homogeneity in different topological properties, poor
861: modularity or community structure, and large loops. In particular, as
862: the spectral gap is very large and they tend to be regular, entangled
863: nets are usually Ramanujan graphs.
864:
865: We have also exploited our understanding of the entangled-topology to
866: generate optimal networks more efficiently. In particular, given the
867: convergence in all the explored cases to almost regular networks, we
868: have constructed an improved version of the algorithm in which we
869: start with random regular nets and rewire edges in such a way that the
870: original degree distribution is preserved (indeed, most of the
871: entangled networks depicted in this section are obtained using this
872: improved algorithm). For this, links are selected by pairs, an origin
873: node is selected for each one and the two end nodes are exchanged as
874: shown in figure \ref{rewiring}. We have observed numerically that the
875: convergence towards optimal solutions is faster when constraining the
876: optimization to the regular networks and that for large values of $N$
877: the final outputs have smaller values of $Q$ than the original ones
878: % generated
879: %by the non $k$-degree preserving version of the algorithm,
880: (owing to the fact that it is less likely to fall into metastable
881: states). The discussed topological traits remain unaffected.
882: \begin{figure}
883: \centerline{
884: \psfig{file=rewiring.xfig.eps,width=7.5cm}}
885: \caption{Rewiring attempt in the connectivity conserving
886: version of the algorithm.}
887: \label{rewiring}
888: \end{figure}
889:
890:
891:
892: \subsection{Comparing Cayley-graph Ramanujans with Entangled networks}
893:
894: \subsubsection{Small sizes: Cage graphs}
895:
896: If we are to design optimal networks with a small number of vertices
897: (say $N < 100$) then entangled networks (and hence, cage-graphs,
898: whenever they exist) are a better choice than Ramanujan nets based on
899: Cayley-graphs, as those constructed in the previous section. First,
900: these second do exist only for a few small values of $N$. Second,
901: because even when such Ramanujan graphs exist entangled network
902: outperform them, as they have a smaller value of $Q$ (larger spectral
903: gap) which has been optimized on-purpose.
904:
905:
906: \subsubsection{Large sizes}
907: For large values of $N$ (say $N >100$) entangled networks are
908: difficult to construct as the optimization algorithm becomes not
909: affordably time-consuming. Instead, Cayley-graph Ramanujan networks,
910: whenever they can be constructed, are a better choice. They are easy
911: and fast to construct by following the recipe described in a previous
912: section and they provide large spectral gaps, closer and closer to the
913: optimal value (asymptotic upper bound) as $N$ increases.
914:
915:
916: \section{VI. Related problems and similar topologies}
917:
918: In this section we illustrate how entangled networks and Ramanujan
919: graphs play a crucial role in different contexts, and emerge as
920: optimal (or close to optimal) solutions for a number of network
921: optimization problems.
922:
923:
924: \subsection{Local search with congestion}
925:
926: This is one of the examples illustrated in the introduction, and has
927: been recently tackled by Guimer\'a et al. \cite{Catalans}. By defining
928: an appropriate {\it search cost} function these authors explore which
929: is the ideal topology to optimize searchability and facilitate
930: communication processes. They arrive at the conclusion that, while in
931: the absence of traffic congestion a star-like (centralized) topology
932: is the optimal one (as briefly discussed in the introduction) when the
933: density of information-packets traveling through the net is above a
934: given threshold (i.e. when there is {\it flow congestion}) the optimal
935: topology is a highly homogeneous one. In this last, all nodes have
936: essentially the same degree, the same betweenness, and short loops are
937: absent (see figure 1 in \cite{Catalans}). These networks resemble
938: enormously the entangled and Ramanujan nets described above.
939:
940: For a similar comparison between centralized and de-centralized
941: transport in networks, see \cite{Congestion}. Also here,
942: decentralized, highly homogeneous, structures emerge as optimal ones
943: under certain circumstances.
944:
945:
946: \subsection{Network structures from selection principles}
947:
948: In a recent letter, Colizza et al. \cite{Maritan} have addressed the
949: search of optimal topologies by using different {\it selection
950: principles}. In particular they minimize a global cost function
951: defined as
952: \begin{equation}
953: H_{\alpha} = \sum_{i < j} d_{ij} (\alpha)
954: \end{equation}
955: with
956: \begin{equation}
957: d_{ij} (\alpha) = \min_P \sum_{p\in P: i \rightarrow j} k_p^{\alpha}.
958: \end{equation}
959: This is, the cost function is the sum over pairs of sites of their
960: relative $\alpha$-dependent distance. The distance between two nodes
961: is the minimum of the sum of $k_p^\alpha$ ($k_p$ is the degree of node
962: $p$) over all possible paths connecting the two nodes. In this way,
963: for $\alpha=0$ the standard distance is recovered. On the other hand,
964: for large values of $\alpha$, highly connected vertices (hubs) are
965: strongly penalized. In this way, and rephrasing the authors of
966: \cite{Maritan}, the generalized definition of the distance ``captures
967: the conflict between two different trends: the avoidance of long paths
968: and the desire to skip heavy traffic''. For small values of $\alpha$
969: (standard distance) the best topology include central hubs
970: (centralized communication), while for $\alpha >1$ the dominant
971: tendency is towards degree minimization, leading to open tree-like
972: structures. For intermediate cases, i. e. $\alpha=0.5$ topologies
973: extremely similar to entangled networks, with long loops and very high
974: homogeneity, have been reported to emerge (see figure 3c in
975: \cite{Maritan}).
976:
977:
978: \subsection{Performance of neural networks}
979:
980: Recently Kim concluded that neural networks with a clustering
981: coefficient as small as possible exhibit much better performance than
982: others \cite{BJK}. Entangled nets have a very low clustering
983: coefficient as only large loops exist and, therefore, they are natural
984: candidates to constitute an excellent topology to achieve good
985: performance and high capacity in artificial neural networks.
986:
987:
988: \subsection{Robustness and resilience}
989:
990: The problem of constructing networks whose robustness against random
991: and/or intentional removal is as high as possible has attracted a lot
992: of attention. In particular, in a recent paper \cite{2peak}, it has
993: been shown that for generalized random graphs in the limit
994: $N\to\infty$ the most robust topology (in the sense that the
995: corresponding percolation threshold is as large as possible) has a
996: degree distribution with no more than $3$ distinct node
997: connectivities; i.e. with a homogeneous degree-distribution as
998: homogeneous as possible.
999:
1000:
1001: To study the possible connection with our extremely homogeneous
1002: entangled networks, let us recall that the topology we have considered
1003: to initialize the improved version of the optimization algorithm
1004: (i.e. random $k$-regular graphs) is already the optimal solution for
1005: robustness-optimization against (combined) errors and attacks in
1006: random networks \cite{2peak}. A natural question to ask is whether
1007: further $Q$-optimization has some effect on the network
1008: robustness. This and related questions have been analyzed in
1009: \cite{Entangled}, where it was shown that indeed, $Q$ minimization
1010: implies a robustness improvement. This occurs owing to the generation
1011: of non-trivial correlations in entangled structures, which place them
1012: away from the range of applicability of the result in \cite{2peak}
1013: (they are not random networks). Hence, entangled networks are also
1014: extremely efficient from the robustness point of view. This conclusion
1015: also holds for {\it reliability} against link removal \cite{Myrvold}.
1016:
1017:
1018: \section{VII. The connection with the Bethe lattice}
1019:
1020: In this section we will show how entangled networks are related to the
1021: infinite regular tree called {\it Bethe lattice} (or Bethe tree) in
1022: the physics literature.
1023:
1024: Intuitively, whenever around a given node of a regular graph only
1025: large loops are present, the neighborhood of such node looks as the
1026: one in a Bethe tree, up to a distance equal to the half of the
1027: shortest loop length (see for instance the center of the graph
1028: depicted to the right of figure \ref{NoCages}). Therefore, if the
1029: girth diverges in a family of graphs, the neighborhood of each node
1030: tends to look locally like a Bethe, up to a growing distance.
1031:
1032: %% Just to provide with some further intuition, observe how the
1033: %% central part of the rightmost graph in figure \ref{NoCages}
1034: %% resembles a Cayley tree, and the full graph is a closure of it.
1035:
1036: We notice that the asymptotic spectral gap of regular graphs is
1037: bounded by the one of the Bethe tree, $\lambda_2 \leq k- 2
1038: \sqrt{k-1})$ (see equation (\ref{inequality2})) \cite{bethegap}.
1039: We can express this bound in terms of the spectral properties of the
1040: adjacency matrix (since the graphs are regular, the results can be
1041: easily translated to $L$ and $\LN$).
1042:
1043: The moments $m_n$ of the adjacency matrix eigenvalues $\mu_i$ can be
1044: trivially expressed through the trace of the $n$-th power of $A$ \cite{LPS}:
1045: \begin{equation}
1046: \label{eq:moms} m_n = \frac1N \sum_i \mu_i^n = \frac1N \sum_i
1047: (A^n)_{ii}.
1048: \end{equation}
1049: It is easy to see that the diagonal elements of $A^n$ represent the
1050: number of paths of length $n$ starting and ending at the corresponding
1051: node. Clearly, if $n$ is smaller than the girth of the graph the paths
1052: do not contain loops and their number is equal to the number of such
1053: paths for a node of the Bethe lattice and therefore equal for all the
1054: nodes. The eigenvalue moments for the infinite Bethe lattice can be
1055: obtained in a similar way, with the only difference that the
1056: eigenvalue distribution is now a continuous spectral density
1057: \cite{density}. The sum in equation~(\ref{eq:moms}) is now an integral
1058: and the adjacency matrix is a linear operator; however it is still
1059: possible to express the $n$-th moments as the number of length-$n$
1060: paths from a node to itself. As a result the moments $m_n$ of the
1061: adjacency matrix eigenvalues of a regular graph with girth $g$ are
1062: equal to the moment of the spectral density
1063: \begin{equation}
1064: \label{eq:specden} \rho(\mu) = \frac{k \sqrt{4(k-1)-\mu^2}}{2 \pi
1065: (k^2 -\mu^2)}\ , \quad \text{for } |\mu| \le 2\sqrt{k-1}
1066: \end{equation}
1067: of the Bethe tree \cite{bethedensity}, for every $n<g$.
1068:
1069: The previous argument can be generalized and it can be rigorously
1070: proven \cite{mckay} that for an infinite sequence $G_n$ of $k$-regular
1071: graphs, if the number of loops of length $l$ grows slower than the
1072: number of nodes for every $l$ then, the spectral density tends to the
1073: one in equation~(\ref{eq:specden}). Hence, we expect a fast
1074: convergence for graphs with large girth as the ones discussed before.
1075:
1076: To illustrate this, in figure \ref{spden} we plot the Bethe lattice
1077: integrated spectral density as a function of the eigenvalues for $k=4$
1078: and compare it with the one calculated for a Ramanujan graph with
1079: $6840$ nodes ($q=19$) and also $k=4$ ($p=3$). The agreement is very
1080: remarkable given the finite size of the Ramanujan graph.
1081:
1082: \begin{figure}
1083: \centerline{
1084: \psfig{file=spden.eps,width=7.5cm}}
1085: \caption{Integrated spectral density as a function of the adjacency matrix eigenvalues
1086: for the Bethe lattice (infinite size) and for a Ramanujan graph with
1087: $N=6840$, both with $k=4$.}
1088: \label{spden}
1089: \end{figure}
1090:
1091: Therefore, both from the (local) topological and the spectral point of
1092: view, a family of entangled networks (or Ramanujan graphs) with
1093: growing size and girth can be used as the best possible way to
1094: approach infinite Bethe lattices with a sequence of finite lattices.
1095: This would have applications in a vast number of problems in physics,
1096: for which exact solutions in the Bethe lattice exist, but comparisons
1097: with numerics in sufficiently large lattices are difficult to obtain.
1098: In particular, approximations of Bethe lattice obtained by truncating
1099: the number of generations (Cayley trees) include a large (extensive)
1100: amount of boundary effects (losing the original homogeneity) and
1101: are therefore not a convenient finite approximation. Good
1102: approximations should avoid strong boundary effects and preserve the
1103: Bethe lattice homogeneity, and entangled networks can play such a
1104: role.
1105:
1106: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1107:
1108: \section{VIII. A way out of homogeneity: weighted dynamics}
1109:
1110:
1111: While most of the literature on synchronization refers to Laplacian
1112: couplings as specified by equation (\ref{pecora}), different coupling
1113: functions can be relevant in some contexts. For example, another
1114: natural choice would be
1115: \begin{equation}
1116: \dot{x}_i =
1117: F(x_i) + {\sig\over k_i} \sum_{j\in n.n. i} [ H(x_j) -H(x_i)] =
1118: F(x_i) - {\sig\over k_i} \sum_j L_{ij} H(x_j),
1119: \label{norma}
1120: \end{equation}
1121: relevant in cases where the joint effect of the $k_i$ neighbors of
1122: node $i$ is normalized by the connectivity $k_i$. With this type of
1123: dynamics, the effect of neighbors has the same weight for all nodes,
1124: while in the absence of the normalization factor, $k_i$, sites with
1125: higher connectivity are more strongly coupled to their neighbors than
1126: loosely connected ones. This describes properly real-world situations,
1127: as for instance neural networks, where the influence of the
1128: neighboring environment on the node dynamics does not grow with the
1129: number of connections. Observe that, even if the underlying topology
1130: is unweighted and undirected, the normalization factor is such that
1131: {\it the dynamics is directed and weighted}, owing to the presence of
1132: $k_i$ in eq.(\ref{norma}).
1133:
1134: In a recent paper Motter and coauthors have undertaken a study of
1135: synchronizability by using a generalization of the previous normalized
1136: dynamics eq.(\ref{norma}), by employing a normalization factor
1137: $k_i^\beta$, with $\beta \geq 0$ \cite{Zhou}. These authors conclude
1138: that the most robust network synchronizability is obtained for
1139: $\beta=1$, which leads back to Eq.(\ref{norma}).
1140:
1141: {\it What is the influence of the normalization factor on the results
1142: reported on this paper?}
1143:
1144: First of all, the optimal synchronizability problem, as discussed in
1145: section III, can be straightforwardly translated for the new dynamics
1146: just by replacing Eq.(\ref{pecora}) by Eq.(\ref{norma}). The optimal
1147: topology for synchronizability in ``normalized'' dynamical processes
1148: is that minimizing the eigenratio $Q_{norm}=\lambda'_N/\lambda'_2$,
1149: ($\lambda'_i$ denote the normalized Laplacian eigenvalues). Hence, by
1150: employing the normalized dynamics, we have a new class of optimization
1151: problems, analogous but different to the ones studied before along
1152: this paper.
1153:
1154: We have implemented different versions of our modified simulated
1155: annealing algorithm (in its more general, not degree-conserving,
1156: version) to optimize either $Q_{norm}$ or the normalized spectral gap
1157: $\lambda'_2$.
1158: % Before proceeding let us emphasize that as the determinant of
1159: % $K^{-1} L - \nu I$ coincides with that of $K^{-1/2} L K^{1/2} - \nu
1160: % I$ the eigenvalues of the normalized Laplacian do not depend on
1161: % whether the Laplacian normalizing factor is $k_i$ or $\sqrt{k_i
1162: % k_j}$. This last one generates a symmetric matrix and simplifies the
1163: % computations.
1164: After going through the optimization algorithm with
1165: normalized-Laplacian eigenvalues, as described in section V, one
1166: observes that the emerging optimally synchronizable nets have a {\it
1167: non-homogeneous structure}. For some relatively small values of $N$
1168: and $\langle k \rangle=3$ the optimal topologies are shown in figure
1169: \ref{newgraphs}, while in figure \ref{newgraphs2} we depict an optimal
1170: net with $100$ nodes.
1171: \begin{figure}
1172: \centerline{
1173: \psfig{file=opt10.ps,width=4.0cm}
1174: \psfig{file=opt12.ps,width=4.0cm}
1175: \psfig{file=opt20.ps,width=4.0cm}}
1176: \caption{Optimal nets for normalized-Laplacian dynamics,
1177: with $\langle k \rangle=3$ and $N=10, 12$, and $20$ respectively.}
1178: \label{newgraphs}
1179: \end{figure}
1180:
1181: \begin{figure}
1182: \centerline{
1183: \psfig{file=rh100-2.1-0.eps,width=6.0cm}}
1184: \caption{Optimal net for normalized-Laplacian dynamics,
1185: with $\langle k \rangle=3$ and $N=100$.}
1186: \label{newgraphs2}
1187: \end{figure}
1188:
1189: Observe the remarkable difference with previously studied networks:
1190: here homogeneity is drastically reduced. To illustrate this, in figure
1191: \ref{degree} we plot the degree distribution obtained when optimizing
1192: for $\langle k \rangle=4$ with $N=200$ and $N=1000$ respectively. The
1193: resemblance between these two distributions seems to indicate that a
1194: limit degree-distribution exists. Roughly speaking it decays faster
1195: than exponentially, but still with a much higher degree of
1196: heterogeneity than before where we had delta-peaked distributions.
1197:
1198: The new emerging topologies exhibit a competition between the
1199: existence of central nodes and peripheral ones similarly to the
1200: networks studied in \cite{Congestion}. By maximizing the spectral gap
1201: $\lambda'_2$, rather than $\lambda'_N/\lambda'_2$ we obtain very
1202: similar results (not shown here). These new optimal topologies are
1203: very similar to those obtained by Colizza et al. \cite{Maritan} in
1204: cases where the minimization of node-degree dominates (see figure 3d
1205: in
1206: \cite{Maritan}).
1207:
1208:
1209: Let us remark that for the normalized-Laplacian, some eigenvalues
1210: bounds, analogous to those for $L$ exist; in particular \cite{Chung}
1211: \begin{equation}
1212: 0 < \lambda'_2 \leq {N \over N-1} \leq \lambda'_N \leq 2.
1213: \label{nl}
1214: \end{equation}
1215: However, concepts analogous to ``expanders'' or ``Ramanujan'' have not
1216: been defined for general, non-regular, graphs.
1217: \begin{figure}[h]
1218: \centerline{
1219: \psfig{file=degree2.ps,width=9.5cm,angle=-90}}
1220: \caption{Semi-logarithmic plot of the degree distribution of optimal weighted
1221: networks with $\langle k \rangle =4$ and (i) $N=200$ averaged over
1222: $17$ runs (black curve with down-triangles) and (ii) $N=1000$ for a
1223: single run (red curve with up-triangles).}
1224: \label{degree}
1225: \end{figure}
1226:
1227: Finally, it is interesting to observe that our results are in apparent
1228: contradiction with the conclusion by Motter et al. \cite{Zhou} that
1229: for their case $\beta=1$, corresponding to our normalized Laplacian,
1230: and large {\it sufficiently random networks} the eigenratio $Q$ does
1231: not depend on details of the network topology, but only on its mean
1232: degree \cite{Zhou}. If this was indeed the case for any network
1233: topology, our optimization procedure would be pointless. Instead,
1234: applying the minimization algorithm, starting from any arbitrary
1235: finite random network, we observe a progressive $Q$ optimization, and
1236: highly non-trivial non-random optimal structures are actually
1237: generated. This apparent contradiction is likely to be due to the
1238: building up of non-trivial correlations during the optimization
1239: process, which converts our network into highly non-random structures
1240: far away from the {\it sufficiently random} requirement in
1241: \cite{Zhou}.
1242:
1243: \section{IX. Discussion and Conclusions}
1244:
1245: In this paper we have reviewed recent developments in the design of
1246: optimal network topologies.
1247:
1248: First of all we have related the problem of finding optimal topologies
1249: for many dynamical and physical problems, such as optimal
1250: synchronizability and random walk flow on networks, to the search of
1251: networks with large spectral gaps. This connection allows us to
1252: relate optimal networks to expanders and Ramanujan graphs, well-known
1253: concepts in graph-theory which, by construction, have large spectral
1254: gaps. In one of the sections we have given, following the existing
1255: mathematical literature, a recipe to explicitly construct Ramanujan
1256: graphs.
1257:
1258: On the other hand, we have employed a simulated annealing algorithm
1259: which selects progressively networks with better synchronizability,
1260: i.e. networks with larger spectral gaps. When applied to problems with
1261: un-normalized Laplacian dynamics this algorithm leads to what we call
1262: ``entangled'' topologies. These are optimal expanders, and can be
1263: loosely described as being extremely homogeneous, having long loops,
1264: poor modularity, and short node-to-node distances. They are optimal or
1265: almost optimal for many different communication and flow processes
1266: defined on regular networks. In particular these topologies are
1267: relevant for the design of efficient communication networks,
1268: construction of error-correcting codes with very efficient encoding
1269: and decoding algorithms, de-randomization of random algorithms,
1270: traffic problems with congestion, and analysis of algorithms in
1271: computational group theory. Remarkably, they also provide a good
1272: finite-size approximation to Bethe lattices.
1273:
1274: Even though these topologies play an important role in human-designed
1275: networks, especially in Computer Science and Algorithmics, they do not
1276: seem to appear frequently in Nature. Indeed, most of the topologies
1277: described in recent years for biological, ecological, social, or
1278: technological networks exhibit a very heterogeneous scale-free degree
1279: distribution \cite{Strogatz,Laszlo,Newman,Porto,AleRomu,Boccaletti},
1280: at odd with the extremely homogeneous entangled topologies. To
1281: justify this, first of all we should emphasize that optimal entangled
1282: networks are obtained performing a {\it global} optimization
1283: procedure, which might be not very realistic. For example, by adding a
1284: new node to a given optimal network, in some cases the full entangled
1285: topology has to be altered to convey with the requirement of
1286: maximizing the spectral gap. This is certainly impractical for growing
1287: networks.
1288:
1289: On the other hand, we should also note that there are ingredients that
1290: could be (and actually are) relevant for the determination of optimal
1291: networks in real-world problems and that have not being considered
1292: here. A non exhaustive list of examples is:
1293:
1294: \begin{itemize}
1295: \item nodes might be non equivalent,
1296:
1297: \item links could support different variable weights,
1298:
1299: \item directed networks might be mandatory in some cases,
1300:
1301: \item in some real-world networks nodes are embedded into a geography
1302: and, therefore, the distances between them should be taken into
1303: account as a sort of quenched disorder (see for instance,
1304: \cite{Gastner,Barthelemy} for application to communication infrastructures).
1305:
1306: \end{itemize}
1307:
1308: However, our entangled networks are still relevant for any type of
1309: un-directed un-weighted networks in which none of these effects plays
1310: an important role, and where communication properties are to be
1311: maximized. Numerous examples have being cited above and along the
1312: paper.
1313:
1314: In the last section of the paper, we have ``enlarged our horizon'' by
1315: studying a different type of dynamical processes, that could be argued
1316: to be more realistic in some circumstances: instead of analyzing
1317: Laplacian coupling between the network nodes we shifted to the study
1318: of normalized-Laplacian coupling, which lead to weighted and directed
1319: dynamics. This is characterized by the presence of a ``normalized
1320: dynamics'' in which the relative influence of the neighbors on a given
1321: site does not depend on the site connectivity. In particular, this
1322: type of dynamics might be more adequate to describe random walks and
1323: general diffusion processes on general (not necessarily regular)
1324: networks. In this case, an optimization procedure analogous to the one
1325: described before, leads to much more heterogeneous optimal topologies,
1326: including hubs, and a much broader degree distribution. This makes the
1327: emerging optimal networks closer to real-world topologies than
1328: entangled networks, while still keeping the global-optimization
1329: perspective. Along this same line of reasoning, other alternative
1330: types of dynamics could be introduced, as for instance the
1331: ``load-weighted'' ones first proposed in
1332: \cite{Chavez,Zhou} as a generalization of the normalized-Laplacian
1333: dynamics. A very interesting, open question, is to define optimization
1334: processes along the lines of this paper, leading to scale free
1335: topologies.
1336:
1337: We hope this work will encourage further research on this fascinating
1338: question of optimal network topologies, their evolution, their
1339: application to human designed networks and their connection with
1340: real-world complex networks.
1341:
1342:
1343: \vspace{1cm}
1344:
1345: \begin{acknowledgments} We acknowledge useful discussions with
1346: D. Cassi with whom we are presently exploring the connection with the
1347: Bethe lattice. We are especially thankful to P. I. Hurtado, our
1348: coauthor in the paper where entangled networks were first introduced,
1349: for a very enjoyable collaboration and a critical reading of the
1350: manuscript. F. N. acknowledges the kind hospitality in Granada where
1351: this work was initiated. Finally, we acknowledge financial support
1352: from the Spanish MEyC-FEDER, project FIS2005-00791, and from Junta de
1353: Andaluc{\'\i}a as group FQM-165.
1354: \end{acknowledgments}
1355:
1356:
1357: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1358: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1359:
1360:
1361: \begin{thebibliography}{99}
1362:
1363:
1364: \bibitem{synchro}
1365: T. Nishikawa, A. E. Motter, Y.-C. Lai, and F. C. Hoppensteadt,
1366: % Heterogeneity in oscillator networks: are smaller worlds easier to
1367: % synchronize?
1368: Phys. Rev. Lett. {\bf 91}, 014101 (2003).
1369: H. Hong, B. J. Kim, M. Y. Choi, and H. Park
1370: % Factors that predict better synchronizability on complex networks
1371: Phys. Rev. E. {\bf 69}, 067105 (2004). H. Hong, M. Y. Choi, and
1372: B. J. Kim, Phys. Rev. E {\bf 65}, 026139 (2002); Phys. Rev. E
1373: {\bf 65}, 047104 (2002).
1374:
1375:
1376: \bibitem{Torres} J. J. Torres, M. A. Mu\~noz, J. Marro, and P. L. Garrido,
1377: %{\it Influence of Topology on a Neural Network performance}.
1378: Neurocomputing, {\bf 58-60}, 229 (2004); and references therein.
1379: % D. Stauffer, et al.
1380: % A. Aharony, L. da Fontoura Costa, and J. Adler,
1381: %Eur. Phys. J. B {\bf 32}, 395 (2003).
1382:
1383:
1384: \bibitem{BJK} B. J. Kim,
1385: % Performance of networks of artificial neurons: the role of clustering
1386: Phys. Rev. E {\bf 69}, 045101(R) (2004). See also, P. McGraw and
1387: M. Menzinger, Phys. Rev. E {\bf 72}, 015101(R) (2005).
1388:
1389:
1390: \bibitem{GG} G. Grinstein and R. Linsker,
1391: Proc. Natl. Acad. Sci. USA, {\bf 102}, 9948 (2005).
1392:
1393:
1394:
1395: \bibitem{Catalans}
1396: R. Guimera, A. Arenas, A. Diaz-Guilera, F. Vega-Redondo,
1397: A. Cabrales,
1398: % Optimal network topologies for local search with congestion
1399: Phys. Rev. Lett. {\bf 89}, 248701 (2002).
1400: %See also, I. Vragovi\'c, E. Louis, and A. Diaz-Guilera,
1401: %Cond-mat/0410174.
1402:
1403:
1404:
1405: \bibitem{Congestion} D. J. Ashton, T. C. Jarrett, and N. F. Johnson,
1406: Phys. Rev. Lett. {\bf 94}, 058701 (2005).
1407:
1408: \bibitem{Maritan}
1409: V. Colizza, J. R. Banavar, A. Maritan, and A. Rinaldo,
1410: Phys. Rev. Lett. {\bf 92}, 198701 (2004).
1411:
1412:
1413:
1414: \bibitem{Lovasz} L. Lov\'asz,
1415: %\emph{Random Walks on Graphs: a Survey},
1416: in \emph{Combinatorics, Paul Erd\"os is Eighty} (V2), Keszthely,
1417: Hungary, pp. 1-46, (1993).
1418:
1419:
1420:
1421: \bibitem{dani} E. M. Bollt and D. ben-Avraham,
1422: New Journal of Phys. {\bf 7}, 26 (2005).
1423: %cond-mat/0409465.
1424:
1425: \bibitem{Strogatz} S. H. Strogatz, Nature {\bf 410}, 268 (2001).
1426:
1427: \bibitem{Laszlo} A. L. Barab\'asi,
1428: %{\it Statistical mechanics of complex networks},
1429: Rev. Mod. Phys. {\bf 74}, 47 (2002).
1430:
1431: \bibitem{Newman} M. E. J. Newman,
1432: %{\it The structure and function on complex networks},
1433: SIAM Review {\bf 45}, 167 (2003).
1434:
1435: \bibitem{Porto} S. N. Dorogovtsev and J. F. F. Mendes, {\it Evolution of
1436: Networks: From Biological Nets to the Internet and WWW}, Oxford
1437: University Press, Oxford (2003).
1438:
1439: \bibitem{AleRomu} R. Pastor Satorras and A. Vespignani, {\it Evolution and
1440: Structure of the Internet: A Statistical Physics approach}, Cambridge
1441: University Press (2004).
1442:
1443: \bibitem{Boccaletti} S. Boccaletti, V. Latora, Y. Moreno,
1444: M. Chavez, and D.-U. Hwang, Phys. Rep. {\bf 424}, 175 (2006).
1445:
1446:
1447: \bibitem{Chung}
1448: F. Chung, {\it Spectral Graph Theory},
1449: Number 92 in CBMS Regional Conference Series in Mathematics. Am. Math.
1450: Soc., 1997. See also, F. Chung, L. Lu, and V. Vu,
1451: Internet Mathematics, {\bf I}, 257 (2004).
1452:
1453: \bibitem{Bollobas}
1454: B. Bollob\'as, {\it Extremal Graph Theory} Academic Press, New
1455: York. 1978. W. Tutte, {\it Graph Theory As I Have Known It}, Oxford
1456: U. Press, New York, (1998).
1457:
1458: \bibitem{GN} M. Girvan, M. E. J. Newman,
1459: Proc. Natl. Acad. Sci. USA {\bf 99}, 7821-7826 (2002).
1460: M. E. J. Newman, M. Girvan, Phys. Rev. E {\bf 69}, 026113 (2004).
1461: See also, L. Donetti and M. A. Mu\~noz,
1462: %{\it Detecting Network Communities: a
1463: %new systematic and powerful algorithm},
1464: J. Stat. Mech.: Theor. Exp. (2004) P10012;
1465: %L. Donetti and M. A. Mu\~noz, {\it
1466: %Improved spectral algorithm for the detection of network Communities},
1467: in ``{\it Modeling Cooperative behavior in the social sciences}'', AIP
1468: Conf. Proc. 779, 104 (2005); arXchic:Physics/0504059.
1469:
1470:
1471: \bibitem{Sarnak} P. Sarnak,
1472: % What is an expander?
1473: Notices Amer. Math. Soc. {\bf 51}, 762 (2004).
1474:
1475: \bibitem{Valette} G. Davidoff, P. Sarnak, and A. Valette, {\it Elementary Number Theory,
1476: Group Theory and Ramanujan Graphs}, London Math. Soc. Students Texts,
1477: 55, Cambridge (2003).
1478:
1479: \bibitem{AB} The original statement of the lower bound
1480: is due to N. Alon and R. Boppana, and appears in A. Nilli,
1481: % On the second eigenvalue of a graph
1482: Discrete Math., {\bf 91}, 207 (1991).
1483:
1484:
1485: \bibitem{Entangled} L. Donetti, P. I. Hurtado and M. A. Mu\~noz,
1486: Phys. Rev. Lett. {\bf 95}, 188701 (2005).
1487:
1488:
1489: \bibitem{Pecora}
1490: M. Barahona and L. M. Pecora,
1491: %Synchronization in small-world systems
1492: Phys. Rev. Lett. {\bf 89}, 054101 (2002). See also, L. M. Pecora and T. L.
1493: Carroll, Phys. Rev. Lett. {\bf 64}, 821 (1990); ibid, {\bf 80}, 2109
1494: (1998).
1495: % PRA 46 2374 1991
1496: L. M. Pecora, Phys. Rev. E {\bf 58} 347 (1998).
1497: L. M. Pecora and T. L. Carroll, Phys. Rev. Lett. {\bf 80}, 2109
1498: (1998).
1499:
1500:
1501:
1502:
1503: \bibitem{Wang} X. F. Wang and G. Chen,
1504: Int. J. Bifurcation Chaos Appl. Sci. Eng. {\bf 12}, 187
1505: (2002). X. F. Wang and G. Chen, IEEE Trans. Circuits and Systems I
1506: {\bf 49}, 54 (2002).
1507: %In this paper the case where the stability
1508: %interval is unbounded from above is studied. The larger stability
1509: %range (maximum synchronizability) is obtained my maximizing the
1510: %spectral gap.
1511:
1512:
1513: \bibitem{Mohar} B. Mohar,
1514: in {\it Graph Theory, Combinatorics, and Applications, Vol 2,}
1515: Ed. Y. Alavi, G. Chartrand, O. R. Oellermann, and A. J Schwenk,
1516: Wiley, New York, 1991. pp. 871.
1517:
1518:
1519:
1520: \bibitem{Nielsen}
1521: $http://www.math.ias.edu/~boaz/ExpanderCourse/$.
1522: See also, $http://www.qinfo.org/people/nielsen/blog/?p=222$
1523:
1524:
1525: \bibitem{PL} P. Pons and M. Latapy, Preprint. arXiv:physics/0512106.
1526:
1527:
1528: \bibitem{LPS} A. Lubotzky, R. Phillips, and P. Sarnak,
1529: Combinatorica, {\bf 8}, 261 (1988).
1530:
1531:
1532: \bibitem{Margulis} G. A. Margulis, Prob. of Info. Trans.
1533: {\bf 9}, 325 (1975). See also, O. Reingold, S. Vadhan, and
1534: A. Wigderson, Ann. of Math. {\bf 155}, 157 (2002).
1535:
1536:
1537:
1538:
1539:
1540: \bibitem{Chiu} P. Chui, Combinatorica, {\bf 12}, 275 (1992).
1541:
1542:
1543: \bibitem{Code} Our computer code to generate Ramanujan graphs is available
1544: upon request to any of the authors.
1545:
1546:
1547: %\bibitem{Legendre}
1548: %Suppose $m $ is an odd prime and $n$ is any integer. The Legendre
1549: %symbol $(n|m)$ is defined to be $+1$ if $n$ is a quadratic residue
1550: %(mod $m$), $-1$ if a is a quadratic non-residue (mod $m$) and, $0$ if
1551: %$m$ divides $n$. $ [[k < m]].$
1552:
1553:
1554: \bibitem{betweenness}
1555: The betweenness centrality is defined as the average number of
1556: shortest paths, connecting every possible couple of nodes, passing
1557: through each site. An edge-betweenness can be analogously defined for
1558: the number of shortest paths passing through each link.
1559:
1560: \bibitem{SA} S. Kirkpatrick, C. D. Gelatt, and M. P. Vecchi,
1561: Science {\bf 220} 671 (1983). N. Metropolis, A. W. Rosenbluth,
1562: M. N. Rosenbluth, A. H. Teller, and E. Teller, J. Chem. Phys. {\bf 21}
1563: 1087 (1953).
1564:
1565:
1566: \bibitem{Penna} T. J. P. Penna, Phys. Rev. E {\bf 51}, R1 (1995).
1567:
1568:
1569: \bibitem{Cages}
1570: Eric W. Weisstein. "Cage Graph." From MathWorld--A Wolfram Web
1571: Resource. $http://mathworld.wolfram.com/CageGraph.html$; and
1572: references therein. See also, R. C. Read and R. J. Wilson,{\it An
1573: Atlas of Graphs}, Oxford, England: Oxford University Press, pp. 263
1574: and 271-274, 1998.
1575:
1576:
1577: \bibitem{2peak} A.X.C.N. Valente, A. Sarkar, and H.A. Stone,
1578: % 2-Peak and 3-Peak Optimal Complex Networks
1579: Phys. Rev. Lett. {\bf 92}, 118702 (2004).
1580:
1581: \bibitem{Myrvold}
1582: W. Myrvold,
1583: % Reliable network syntesys: some recent results
1584: {\it Proceedings of the $8th$ Int. Conf. on Graph Theory,
1585: Combinatorics, Algorithms, and Applications}, Vol. II, pag. 650
1586: (1998).
1587:
1588: \bibitem{bethegap}
1589: See D. Cassi, Phys. Rev. B {\bf 45}, 454 (1992); and references
1590: therein.
1591:
1592: \bibitem{density} Note that as the tree is infinite, its spectrum is continuous, and
1593: therefore, it is more meaningful to refer to ``spectral densities''.
1594:
1595: \bibitem{bethedensity} H. Kesten, \textit{Trans. AMS} \textbf{92}, 336
1596: (1959).
1597:
1598:
1599:
1600: \bibitem{mckay} B. D. McKay, \textit{Linear Algebra and its
1601: applications} \textbf{40}, 203 (1981)
1602:
1603: \bibitem{Future} D. Cassi, F. Neri, L. Donetti, and M.A. Mu\~noz;
1604: preprint 2006.
1605:
1606: \bibitem{Zhou} A. E. Motter, C. Zhou, and J. Kurths,
1607: Phys. Rev E {\bf 71}, 016116 (2005); AIP Conference Proceedings
1608: {\bf 776}, 201 (2005); EuroPhys. Lett. {\bf 69}, 334 (2005).
1609:
1610: \bibitem{Gastner} M. T. Gastner and M. E. J. Newman,
1611: arXiv:cond-mat/0603278.
1612:
1613: \bibitem{Barthelemy} M. Barthelemy and A. Flammini,
1614: arXiv:physics/0601203.
1615:
1616: \bibitem{Chavez} M. Chavez, D.-U. Hwang, A. Amann,
1617: H. G. E. Hentschel, and S. Boccaletti, Phys. Rev. Lett. {\bf 94},
1618: 218701 (2005).
1619:
1620: \end{thebibliography}
1621:
1622:
1623: \end{document}
1624:
1625:
1626:
1627:
1628:
1629:
1630:
1631:
1632:
1633:
1634:
1635:
1636:
1637:
1638: