1: %\documentclass[aps,pre,preprint,groupedaddress,showpacs,floatfix]{revtex4}
2: \documentclass[aps,pre,twocolumn,groupedaddress,showpacs,floatfix]{revtex4}
3: %\documentclass[aps,pre,twocolumn,groupedaddress]{revtex4}
4: %\usepackage{graphicx}
5: \usepackage{amsmath}% Include figure files
6: \usepackage{dcolumn}% Align table columns on decimal point
7: \usepackage{bm}% bold math
8: %\def\BibTeX{\rm B{\sc ib}\TeX}
9: %\def\btt#1{{\tt$\backslash$#1}}
10: \input epsf
11: \epsfclipon
12: \begin{document}
13: % \draft command makes pacs numbers print
14: \title{Fluctuation-dissipation relations for complex networks}
15: \author{Agata Fronczak, Piotr Fronczak and Janusz A. Ho\l yst}
16: \affiliation{Faculty of Physics and Center of Excellence for
17: Complex Systems Research, Warsaw University of Technology,
18: Koszykowa 75, PL-00-662 Warsaw, Poland}
19: \date{\today}
20:
21: \begin{abstract}
22: In the paper, we study fluctuations over several ensembles of
23: maximum-entropy random networks. We derive several
24: fluctuation-dissipation relations characterizing susceptibilities
25: of different networks to changes in external fields. In the case
26: of networks with a given degree sequence, we argue that the
27: scale-free topologies of real-world networks may arise as a result
28: of self-organization of real systems into sparse structures with
29: low susceptibility to random external perturbations. We also show
30: that the ensembles of networks with noninteracting links (both
31: uncorrelated and with two-point correlations) are equivalent to
32: random networks with hidden variables.
33: \end{abstract}
34:
35: \pacs{89.75.-k, 64.60.Ak}
36:
37: \maketitle
38:
39: %______________________________________________________________________________ S: Introduction
40: \section{Introduction}
41:
42: \par Recently, the statistical properties of real networks
43: (including biological, social and technological systems) have
44: attracted a large amount of attention among physicists (see e.g.
45: \cite{Handbook, Dorogbook, BArev}). It has been realized that
46: despite functional diversity, most of real web-like systems share
47: important structural features e.g. small average path length, high
48: clustering and scale-free degree distribution. A number of network
49: models have been proposed to embody the fundamental
50: characteristics. The models can be roughly divided into two
51: classes: static (homogeneous, equilibrium) and evolving (causal,
52: nonequilibrium). The second class of causal networks encompasses,
53: in particular, the famous BA model \cite{BAScience}, whereas
54: configuration model \cite{Molloy1995,NewPRE2001,Fron2005a} and the
55: large group of networks with hidden variables
56: \cite{SodPRE2002,BogPRE2003, Fron2005b} belong to the first class
57: of static networks. Although very intuitive, the mentioned
58: representatives of static random networks are {\it not properly}
59: defined from the point of view of the equilibrium statistical
60: mechanics. Below we briefly outline what the mentioned lack of
61: {\it seemliness} means with reference to random networks.
62:
63: \par To start with, let us concentrate on the phrase {\it random
64: network}. What does it mean that a network is random? One possible
65: answer is that there is a large amount of randomness in the
66: process of network construction. It treats to all the examples of
67: homogeneous and causal networks quoted in the previous paragraph.
68: The answer however suffers a few disadvantages among which the
69: most striking is the issue of quantification of the randomness.
70: Another answer to the asked question could be that random network
71: is a member of a statistical {\it ensemble} of networks and the
72: probability of the occurrence of a given network in random
73: sampling is proportional to its statistical weight. Without a
74: doubt, the last treatment directly follows principles of the
75: equilibrium statistical mechanics.
76:
77: \par At the moment, a simple example could be the configuration model.
78: In this model, the total number of nodes is fixed to $N$ and
79: degrees of all nodes $i=1,2,\dots,N$ create a specific degree
80: sequence $\{k_i\}$. Until now, nothing has been said about
81: connections between nodes. As a rule, random graphs with a given
82: degree sequence are constructed in the following way: i. first,
83: attach to each node $i$ a number $k_i$ of {\it stubs} (ends of
84: edges); ii. next, choose pairs of these stubs uniformly at random
85: and join them together to make complete edges. For sure, such a
86: procedure represents a large randomness justifying the phrase {\it
87: random networks}. On the other hand, however, the second among the
88: mentioned possible meanings of the phrase, treating the resulting
89: networks as members of the ensemble of graphs with the desired
90: degree sequence, seems to be more familiar to physicists.
91:
92: \par As a truth, the concepts of statistical mechanics (including
93: statistical ensembles, partition functions, averages over ensemble
94: and so forth) has already been applied to analysis of complex
95: networks. Although, the majority among the recently submitted
96: articles still define network models through construction
97: procedures, there has also been published several interesting
98: papers on the {\it genuine} statistical mechanics of random
99: networks (cf. \cite{BurdaPRE2001,
100: BergPRL2002,Bauer2002,BurdaPRE2003,DorogNuc2003,PallaPRE2004,
101: Farkas2004,ParkPRE2004,Bogacz2005,Bialas2005}). The general idea
102: is similar to all the above mentioned papers. Statistical ensemble
103: of networks is defined by specifying a set of networks
104: $\mathcal{G}$ which one wants to study (e.g. simple graphs,
105: digraphs, weighted graphs) and a rule that associates probability
106: distribution $P(G)$ with these networks $G\in\mathcal{G}$. The
107: differences between the quoted approaches consist in different
108: weight assignment strategies.
109:
110: \par In this contribution we extend the information-theoretic approach
111: to random networks that was very recently proposed by Park and
112: Newman \cite{ParkPRE2004} (see also \cite{Bauer2002}). Information
113: theory \cite{Jayens1,bookCover} provides a criterion for setting
114: up probability distributions over a given ensemble on the basis of
115: partial knowledge and leads to a type of statistical inference. It
116: is the least biased estimate possible on the given information
117: i.e. it is maximally non-committal with regard to missing
118: information. Since the procedure consists in entropy maximization
119: under constraints imposed by the physical conditions of the
120: ensemble, it is also known as maximum-entropy estimate
121: \cite{bookKapur}.
122:
123: \par In this paper, following Park and Newman \cite{ParkPRE2004}, we
124: use Shannon entropy in order to establish probability distribution
125: over analyzed networks \cite{footnote1}. Park and Newman have
126: presented a few exact solutions (in the sense of weighted averages
127: over ensembles) of specific network models including undirected
128: networks with a given degree sequence and networks incorporating
129: arbitrary but independent edge probabilities. Here, we analyze
130: these exactly solved models from the point of view fluctuations
131: over ensembles. We discuss several fluctuation-dissipation
132: relations for the mentioned ensembles. We also show that the
133: quoted maximum-entropy models are equivalent to random networks
134: with hidden variables \cite{BogPRE2003}.
135:
136: %______________________________________________________________________________ S: General definitions
137: \section{General definitions}
138:
139: \par In this section we review the fundamentals of maximum-entropy random
140: networks due to Park and Newman \cite{ParkPRE2004}.
141:
142: \par In order, to correctly define statistical ensemble of networks one
143: has to start with specifying a set of graphs $\mathcal{G}$ which
144: one wants to study. In the following, we restrict ourselves to
145: labelled simple graphs \cite{footnote2} with a fixed number of
146: nodes $N$. Let us remind that a simple graph has at most one link
147: between any pair of vertices and it does not contain self-loops
148: connecting vertices to themselves. Note also that there exists
149: one-to-one correspondence (isomorphism) between simple graphs and
150: symmetric matrices of size $N$ with elements $\sigma_{ij}$ equal
151: either $0$ or $1$.
152:
153: \par Once the set $\mathcal{G}$ of possible networks has been
154: established, in the next step one has to decide what kind of
155: constraints should be imposed on the ensemble. The choice may be,
156: for example, encouraged by properties of real networks like high
157: clustering, significant modularity or scale-free degree
158: distribution. In fact, due to the mentioned isomorphism between
159: graphs and matrices only such ensembles can be exactly solved
160: which constraints are simply expressed in terms of adjacency
161: matrix.
162:
163: \par Now, suppose that one would like to establish probability
164: distribution over $\mathcal{G}$ in such a way that the expected
165: values (i.e. averages over the ensemble) of several observables
166: $\{x_i(G)\}$, $i=1,2,\dots,r$ were respectively equal to
167: $\{\langle x_i\rangle\}$. Due to maximum entropy principle the
168: best choice for probability distribution $P(G)$ is the one that
169: maximizes the Shannon entropy
170: \begin{equation}\label{entropy}
171: S=-\sum_{G}P(G)\ln P(G),
172: \end{equation}
173: subject to the constraints
174: \begin{equation}\label{war1}
175: \langle x_i\rangle=\sum_{G}x_{i}(G)P(G)
176: \end{equation}
177: for $i=1,2,\dots,r$, plus the normalization condition
178: \begin{equation}\label{norm}
179: \sum_{G}P(G)=1.
180: \end{equation}
181: The Langrangian for the above problem is given by the below
182: expression
183: \begin{eqnarray}\label{Lang}
184: \mathcal{L}&=&-\sum_{G}P(G)\ln P(G)+\alpha(1-\sum_{G}P(G))\\&&+
185: \sum_{i=1}^{r}\theta_i(\langle x_i\rangle-\sum_{G}x_{i}(G)P(G)),
186: \end{eqnarray}
187: where the multipliers $\alpha$ and $\theta_i$ are to be determined
188: by (\ref{war1}) and (\ref{norm}).
189:
190: \par Differentiating $\mathcal{L}$ with respect to $P(G)$ and then
191: equating the result to zero one obtains the desired probability
192: distribution over the ensemble of graphs with given properties
193: (\ref{war1})
194: \begin{equation}\label{PG0}
195: P(G)=\frac{e^{-H(G)}}{Z},
196: \end{equation}
197: where $H(G)$ is the network Hamiltonian
198: \begin{equation}\label{H0}
199: H(G)=\sum_{i=1}^{r}\theta_{i}x_{i}(G),
200: \end{equation}
201: and $Z$ represents the partition function (normalization constant)
202: \begin{equation}\label{Z0}
203: Z=\sum_{G}e^{-H(G)}=e^{\alpha+1}.
204: \end{equation}
205:
206: Finally, in order to complete the section devoted to general
207: considerations it is useful to define the free energy of the
208: ensemble
209: \begin{equation}\label{F0}
210: F=-\ln Z.
211: \end{equation}
212: The last quantity is of wide use in the rest of the paper.
213:
214: \par Now, let us examine the introduced formalism with a few examples.
215: In the next section, we will analyze fluctuations over the below
216: presented ensembles.
217:
218: %______________________________________________________________________________ SubS: Microcanonical
219: \subsection{Microcanonical ensemble of random networks}
220:
221: \par At the beginning, let us study the equivalent of the
222: microcanonical ensemble for maximum-entropy random networks.
223: Maximizing Shannon entropy (\ref{entropy}) subject to only
224: normalization condition (\ref{norm}), i.e. omitting other
225: constraints (\ref{war1}), one obtains the uniform distribution
226: over all simple graphs of size $N$
227: \begin{equation}\label{micPG}
228: P(G)=\frac{1}{\Omega},
229: \end{equation}
230: where $\Omega=2^{N\choose 2}$ represents the total number of the
231: considered networks i.e. the total number of $0-1$ symmetric
232: matrices of size $N$. The uniform distribution (\ref{micPG}) means
233: that each graph in the ensemble have the same weight regardless of
234: its properties.
235:
236: \par Since all graphs in the ensemble are equiprobable one can simply
237: argue that the probability of a graph having $m$ links is given by
238: \begin{equation}\label{micPm}
239: P(m)=\frac{{{N\choose 2}\choose m}}{2^{N\choose 2}},
240: \end{equation}
241: and respectively
242: \begin{equation}\label{micsm}
243: \langle m\rangle=\sum_{m=0}^{N\choose 2}mP(m)=\frac{{N\choose
244: 2}}{2}
245: \end{equation}
246: Similarly, the probability of an arbitrary node to have $k$
247: nearest neighbors equals $P(k)={N-1\choose k}/2^{N-1}$, and the
248: average connectivity is $\langle k\rangle=(N-1)/2$.
249:
250: \par In fact, the considered microcanonical ensemble of random networks
251: is equivalent to the ensemble of classical random graphs with the
252: connection probability $p=1/2$. Ensembles of classical random
253: graphs with an arbitrary linkage probability will be considered in
254: the next subsection.
255:
256: %______________________________________________________________________________ SubS: Classical RG
257: \subsection{Classical random graphs}
258:
259: \par Now, let us consider an ensemble of networks with an expected
260: number of links $\langle m\rangle$ (as stressed before the
261: ensemble is equivalent to random graphs introduced by
262: Erd\"os-R\'enyi). The Hamiltonian (\ref{H0}) for this ensemble is
263: given by
264: \begin{equation}\label{Her}
265: H(G)=\theta m(G),
266: \end{equation}
267: where $\theta$ represents a field or an inverse temperature whose
268: value is fixed and depends only on $\langle m\rangle$. Park and
269: Newman \cite{ParkPRE2004} have shown that the partition function
270: (\ref{Z0}) for the ensemble is equal to
271: \begin{equation}\label{Zer}
272: Z=(1+e^{-\theta})^{N\choose 2},
273: \end{equation}
274: and respectively the free energy (\ref{F0}) can be written as
275: \begin{equation}\label{Fer}
276: F=-\ln Z=-{N\choose 2}\ln (1+e^{-\theta}).
277: \end{equation}
278:
279: \par Having probability distribution (\ref{PG0}) over the ensemble one
280: can, for example, find the relation between the average number of
281: links $\langle m\rangle$ and $\theta$
282: \begin{equation}\label{mer}
283: \langle m\rangle=\frac{\partial F}{\partial\theta}
284: =\frac{{N\choose 2}}{e^\theta+1}.
285: \end{equation}
286: Now, since $\theta$ is fixed one can reexpress the last formula in
287: terms of the linkage probability $p$ that is known from the theory
288: of classical random graphs
289: \begin{equation}\label{mer1}
290: \langle m\rangle={N\choose 2}p,
291: \end{equation}
292: where
293: \begin{equation}\label{per}
294: p=\frac{1}{e^\theta+1}.
295: \end{equation}
296:
297: \par Finally, let us point out that in the limit of very small fields
298: $\theta\rightarrow 0$ (high temperatures) the ensemble of random
299: networks with an expected number of links is equivalent to the
300: microcanonical ensemble of random networks (\ref{micPG})
301: introduced in the previous subsection.
302:
303: %______________________________________________________________________________ SubS: P(k)
304: \subsection{Networks with a given degree sequence}
305:
306: \par At the moment, suppose that one would like to deal with random
307: networks with an expected degree sequence
308: \begin{equation}\label{Pk_seq}
309: \{\langle k_i\rangle\}\;\;\;\;\;\mbox{for}\;\;\;\;\;
310: i=1,2,\dots,N.
311: \end{equation}
312: In this case, the network Hamiltonian is given by
313: \begin{equation}\label{PkH}
314: H(G)=\sum_{i=1}^{N}\theta_{i}k_{i}(G),
315: \end{equation}
316: where the multipliers $\theta_i$ represent a kind of potential
317: assigned to each node and they only depend on the expected degrees
318: $\langle k_i\rangle$ (see Eqs. (\ref{Pkki}) and (\ref{Pkki1})).
319: The partition function for the considered ensemble can be written
320: as \cite{ParkPRE2004}
321: \begin{equation}\label{PkZ}
322: Z=\prod_{i<j}(1+e^{-(\theta_i+\theta_j)}),
323: \end{equation}
324: and the free energy is
325: \begin{equation}\label{PkF}
326: F=-\sum_{i<j}\ln(1+e^{-(\theta_i+\theta_j)}),
327: \end{equation}
328:
329: \par Performing weighted averages over the ensemble one can easily
330: prove the below identities: the first one describing the
331: connection probability between two nodes $i$ and $j$
332: \begin{equation}\label{Pkpij}
333: p_{ij}=\frac{1}{e^{(\theta_i+\theta_j)}+1},
334: \end{equation}
335: and the second identity for the average connectivity of a node
336: characterized by the local field $\theta_i$
337: \begin{equation}\label{Pkki}
338: \langle k_i\rangle(\theta_i)=\frac{\partial F}{\partial\theta_i}
339: =\sum_{j=1}^{N}\frac{1}{e^{(\theta_i+\theta_j)}+1}=\sum_{j=1}^{N}p_{ij}.
340: \end{equation}
341: The both expressions show the reverse relation between the two
342: parameters (i.e. $\langle k_i\rangle$ and $\theta_i$)
343: characterizing each node. The relation consist in the statement:
344: small degrees correspond to large multipliers and vice versa,
345: nodes with a large number of connections possess small
346: multipliers.
347:
348: \par Park and Newman \cite{ParkPRE2004} have pointed out that
349: instead of studying networks with an expected degree sequence
350: (\ref{Pk_seq}) one can deal with networks characterized by an
351: expected degree distribution $P(\langle k_i\rangle)$
352: \cite{footnote3}. The authors have argued that one can produce any
353: degree distribution by a judicious choice of the distribution of
354: multipliers $\rho(\theta_i)$. In fact, due to (\ref{Pkki}),
355: $\rho(\theta_i)$ resulting in the desired $P(\langle k_i\rangle)$
356: can be determined from the below expression
357: \begin{equation}\label{PkRt}
358: \rho(\theta_i)=P(\langle k_i\rangle)\left|\frac{d\langle
359: k_i\rangle}{d\theta_i}\right|,
360: \end{equation}
361: where $\langle k_i\rangle(\theta_i)$ is given by (\ref{Pkki}).
362: There are, however, a few subtleties related to the transition
363: between the sequence $\{\langle k_1\rangle,\langle
364: k_2\rangle\,\dots,\langle k_N\rangle\}$ and $P(\langle
365: k_i\rangle)$. First, the Eq. (\ref{PkRt}) defines $\rho(\theta_i)$
366: as an implicit function which, except for a very few cases, can
367: not be explicitly calculated. Second, performing such a transition
368: one has to keep in mind that our phase space consists of labelled
369: graphs in which every node $i=1,2,\dots,N$ has assigned its own
370: multiplier $\theta_i$ (i.e. is distinguishable). Using
371: $\rho(\theta_i)$ makes an impression that the nodes lose their
372: identities. In such a case, there exists a threat of widening the
373: original phase space.
374:
375: \par To proceed further, let us consider {\it sparse} networks. In
376: the case, connection probabilities (\ref{Pkpij}) factorize
377: \begin{equation}\label{Pkpij1}
378: p_{ij}\simeq e^{-(\theta_i+\theta_j)}=\frac{\langle k_i
379: \rangle\langle k_j\rangle}{\langle k\rangle N},
380: \end{equation}
381: where
382: \begin{equation}\label{Pkki1}
383: \langle k_i\rangle(\theta_i) \simeq e^{-\theta_i}\sqrt{\langle
384: k\rangle N}.
385: \end{equation}
386: As shown in \cite{BogPRE2003,Fron2005b} such ensembles are
387: equivalent to uncorrelated networks. The relation (\ref{Pkki1})
388: between expected degrees and its multipliers makes the ensembles
389: very simple for both Monte Carlo simulations and analytical
390: treatment. In particular, the distribution of multipliers
391: (\ref{PkRt}) corresponding to $P(\langle k_i\rangle)$ is simply
392: \begin{equation}\label{PkRt1}
393: \rho(\theta_i)=\langle k_i\rangle P(\langle k_i\rangle),
394: \end{equation}
395: where $\langle k_i\rangle$ is given by (\ref{Pkki1}).
396:
397: \par There exist, however, some side effects of the approximation.
398: First, since the connection probability $p_{ij}\leq 1$,
399: (\ref{Pkpij1}), thus the assumption of sparse networks is only
400: valid for networks with non-negative Lagrange multipliers (i.e.
401: $\theta_i,\theta_j\geq 0$). The restriction causes failure of the
402: approach in the case of scale-free networks $P(k)\sim k^{-\gamma}$
403: with $2<\gamma<3$. To existence of hubs $k_{max}\sim
404: N^{1/(\gamma-1)}$ \cite{footnote4}, i.e. nodes with negative
405: multipliers (see comment after Eq. (\ref{Pkki})), spontaneously
406: develop degree correlations \cite{MaslovPhysA2004,ParkPRE2003}. It
407: was argued \cite{BurdaPRE2003,BogEPJ2004,CatanPRE2005} that one
408: can omit the correlations by applying the so-called structural
409: cutoff i.e. forcing the largest degree to scale as
410: $k_{max}\sim\sqrt{N}$. At the moment, let us stress that the
411: structural cutoff in uncorrelated networks naturally emerges form
412: the Eq. (\ref{Pkki1}) when $\theta_i\rightarrow 0$.
413:
414: %______________________________________________________________________________ SubS: Correlations
415: \subsection{Networks with two-point correlations}
416:
417: \par In order to study random networks with two-point correlations one
418: may consider a class of Hamiltonians (\ref{H0}) constructed on the
419: basis of an expected linkage probability
420: \begin{equation}\label{corH}
421: H(G)=\sum_{i<j} \Theta_{ij}\sigma_{ij}(G).
422: \end{equation}
423: In the last expression $\sigma_{ij}(G)$ is an element of the
424: adjacency matrix representing the graph $G$ and $\Theta_{ij}$
425: characterizes field coupled to the hypothetical link $\{i,j\}$.
426: The partition function and the free energy for the ensemble are
427: given by
428: \begin{equation}\label{corZF}
429: Z=\prod_{i<j}(1+e^{-\Theta_{ij}}),\;\;\;\;\;
430: F=-\sum_{i<j}\ln(1+e^{-\Theta_{ij}}).
431: \end{equation}
432: Comparing (\ref{PkZ}) and (\ref{corZF}) one can see that the
433: previous ensemble of networks with an expected degree sequence is
434: a special case (for $\Theta_{ij}=\theta_i+\theta_j$) of networks
435: with arbitrary two-point correlations. Similarly to (\ref{Pkpij})
436: one can also find that
437: \begin{equation}\label{corpij}
438: p_{ij}=\langle\sigma_{ij}\rangle=\frac{\partial
439: F}{\partial\Theta_{ij}}= \frac{1}{e^{\Theta_{ij}}+1}.
440: \end{equation}
441:
442: %_____________________________________________________________________ S: Fluctuations,Correlations,Responses
443: \section{Fluctuations and responses}
444:
445: \par In classical thermodynamics, fields interacting with a system
446: have conjugate variables which represent the response of the
447: system to perturbation of the corresponding field. For example,
448: the response of a gas to a change in volume is a change in
449: pressure. The pressure $p$ is the conjugate variable to the volume
450: $V$. Similarly, the magnetization $M$ of a magnet changes in
451: response to the applied field $B$. The mentioned relations are
452: produced by terms in the Hamiltonian of the form $YX$, where $Y$
453: is a field and $X$ is the conjugate variable to which it couples.
454: Note, that the above also holds for maximum-entropy random
455: networks being the subject of the paper, where (see Eq.
456: (\ref{H0}))
457: \begin{equation}\label{H00}
458: H(G)=\sum_{i=1}^{r}\theta_ix_i(G).
459: \end{equation}
460:
461: \par Taking advantage of (\ref{PG0})-(\ref{F0}), expectation
462: values $\langle x_{i}\rangle$ of observables $x_{i}$ can be
463: calculated as first derivatives of the free energy with the
464: appropriate field $\theta_i$ (cf.
465: (\ref{mer}),(\ref{Pkki}),(\ref{corpij}))
466: \begin{equation}\label{sxi0}
467: \langle x_i\rangle=\sum_{G}x_i(G)P(G)=\frac{\partial F}{\partial
468: \theta_i}.
469: \end{equation}
470: Similarly, second derivatives of the free energy $F$ give the mean
471: square fluctuations of the variables
472: \begin{equation}\label{varxi0}
473: \langle x_{i}^2\rangle-\langle x_i\rangle^2=-\frac{\partial^2
474: F}{\partial\theta_i^2}.
475: \end{equation}
476: Now, inserting (\ref{sxi0}) into (\ref{varxi0}) one obtains a very
477: important result
478: \begin{equation}\label{fdt0}
479: \langle x_{i}^2\rangle-\langle x_i\rangle^2=-\frac{\partial
480: \langle x_i\rangle}{\partial\theta_i}=\chi_i^{(\theta)},
481: \end{equation}
482: that is known as the {\it fluctuation-dissipation theorem} (FDT).
483: The theorem states that fluctuations in an observable $x_{i}$ are
484: proportional to the susceptibility $\chi_i^{(\theta)}$ of the
485: observable to its conjugate field $\theta_i$. Let us remind that
486: the susceptibility $\chi_i^{(\theta)}$ measures the strength of
487: the response of $x_{i}$ to changes in $\theta_i$. In reality, due
488: to practical purposes, it is often simpler to analyze the
489: susceptibility $\chi_i^{(\phi)}$ to other field $\phi_i$ that
490: directly depends on $\theta_i$ (\ref{fdt0})
491: \begin{equation}\label{fdt0a}
492: -\frac{\partial \langle
493: x_i\rangle}{\partial\theta_i}=-\frac{\partial \langle
494: x_i\rangle}{\partial\phi_i}\frac{\partial\phi_i}
495: {\partial\theta_i}=\frac{\partial\phi_i}{\partial\theta_i}\chi_i^{(\phi)},
496: \end{equation}
497: where $\partial\phi_i/\partial\theta_i$ is the transitional
498: derivative.
499:
500: Probably the best known example of the theorem (\ref{fdt0}) is the
501: one arising from fluctuations of energy in the canonical ensemble
502: \begin{equation}\label{fdtE}
503: \langle E^2\rangle-\langle E\rangle^2=-\frac{\partial
504: E}{\partial\beta}=kT^2C_{V},
505: \end{equation}
506: where $C_{V}=\partial\langle E\rangle/\partial T$ is the heat
507: capacity (or thermal susceptibility), whereas
508: $kT^2=\partial\beta/\partial T$ is the respective transitional
509: derivative. Another example relates fluctuations in the
510: magnetization to the magnetic susceptibility
511: \begin{equation}\label{fdtM}
512: \langle M^2\rangle-\langle M\rangle^2=
513: \frac{1}{\beta}\frac{\partial M}{\partial B}=
514: \frac{1}{\beta}\chi^{(B)}.
515: \end{equation}
516:
517: The fluctuation-dissipation theorems (\ref{fdt0})-(\ref{fdtM}) are
518: interesting for a number of reasons. First, they join both
519: microscopic description (left-hand side) and macroscopic,
520: properties (right-hand side) of the considered systems. Second,
521: they relate the actual state (fluctuations) of the systems to
522: their future behavior (response). Third, due to FDT phase
523: transitions certified by singularities in susceptibilities can
524: also be reported by large scale fluctuations.
525:
526: Extending the idea of the susceptibility one can consider what
527: happens with a variable $x_i$ when one changes the value of a
528: field $\theta_j$. To study the problem one can define a
529: generalized susceptibility $\chi_{ij}^{(\theta)}$ which is a
530: measure of the response of $\langle x_i\rangle$ to the variation
531: of the field $\theta_j$
532: \begin{equation}
533: \chi_{ij}^{(\theta)}=-\frac{\partial\langle
534: x_{i}\rangle}{\partial\theta_j}
535: \end{equation}
536: Again, the susceptibility $\chi_{ij}^{(\theta)}$ is a second
537: derivative of the free energy
538: \begin{equation}
539: \chi_{ij}^{(\theta)}=-\frac{\partial^2
540: F}{\partial\theta_i\partial\theta_j} =\langle
541: x_ix_j\rangle-\langle x_i\rangle\langle x_j\rangle.
542: \end{equation}
543: The issue of generalized susceptibilities is of special interest
544: in lattice systems (\ref{H00}), where the variables $x_{i}$ for
545: $i=1,2,\dots,r$ may represent to the same observable but measured
546: in different spatial points. Then, susceptibility
547: $\chi_{ij}^{(\theta)}$ is just the two-point correlation function
548: between sites $i$ and $j$.
549:
550: In the following, we will concentrate on fluctuations over a few
551: selected ensembles of random networks.
552:
553: %______________________________________________________________________________ SubS: Classical RG
554: \subsection{Classical random graphs}
555:
556: At the beginning, let us consider the ensemble of classical random
557: graphs. By definition, the average number of links $\langle
558: m\rangle$ is fixed in the ensemble. As stressed at the beginning
559: of the section, there exist, however, fluctuations around the
560: average. In fact, the probability of a graph $G$ with $m$ links is
561: given by
562: \begin{equation}\label{er1PG}
563: P(G)=\frac{e^{-\theta m(G)}}{Z}=p^m(1-p)^{{N\choose 2}-m}.
564: \end{equation}
565: The variance of the above distribution calculated from
566: (\ref{fdt0}) is very similar to (\ref{fdtE})
567: \begin{eqnarray}\label{erfdtm}
568: \langle m^2\rangle-\langle m\rangle^2&=&-\frac{\partial\langle
569: m\rangle}{\partial\theta}\\&=&-\frac{\partial\langle
570: m\rangle}{\partial p}\frac{\partial
571: p}{\partial\theta}=p(1-p)C_{m}\nonumber,
572: \end{eqnarray}
573: where $C_{m}=\partial\langle m\rangle/\partial p={N\choose 2}$ is
574: the link capacity in classical random graphs. Note that for a
575: given network size $N$ the link capacity does not depend on the
576: linkage probability $C_m(p)=const$ (classical ideal gases reveal
577: the analogous behavior $C_{V}(T)=const$).
578:
579: %______________________________________________________________________________ SubS: P(k)
580: \subsection{Networks with a given degree sequence}
581:
582: Now, let us continue with random networks characterized by an
583: expected degree sequence (\ref{Pk_seq}).
584:
585: Taken advantage of (\ref{Pkpij}) and (\ref{Pkki}),
586: fluctuation-dissipation relations (\ref{fdt0}) for the ensemble
587: may be written in the following form
588: \begin{eqnarray}\label{Pk_fdt1}
589: \chi_i^{(\theta)}=-\frac{\partial\langle
590: k_i\rangle}{\partial\theta_i}&=&\langle k_i^2\rangle-\langle
591: k_i\rangle^2\\&=&\sum_{j}p_{ij}(1-p_{ij})=\langle
592: k_i\rangle-\sum_{j}p_{ij}^2\nonumber.
593: \end{eqnarray}
594: At the moment, before delving into the discussion of the last
595: expression, let us note that the susceptibility
596: $\chi_{i}^{(\theta)}$ is also given by the below formula
597: \begin{eqnarray}
598: \chi_{i}^{(\theta)}=-\sum_{j}\frac{\partial\langle k_i\rangle}
599: {\partial p_{ij}}\frac{\partial
600: p_{ij}}{\partial\theta_i}=-\sum_{j}\frac{\partial
601: p_{ij}}{\partial\theta_i}C_{ij},
602: \end{eqnarray}
603: where $C_{ij}=\partial\langle k_i\rangle/\partial p_{ij}$
604: represents the link capacity. Here, since $C_{ij}=1$ the
605: fluctuations in nodes degrees result only form the transitional
606: derivative $\partial p_{ij}/\partial\theta_i$.
607:
608: The importance of the two above identities lies in the fact that
609: from the fluctuations over degrees of nodes characterized by the
610: same local field $\theta_i$, one can deduce on the future behavior
611: of the nodes in the face of possible changes in $\theta_i$. Large
612: (small) fluctuations correspond to high (low) local susceptibility
613: $\chi_i^{(\theta)}$.
614:
615: Now, let us note, that in the case of small degrees (see also the
616: assumption of sparse networks (\ref{Pkki1})), the
617: fluctuation-dissipation relation (\ref{Pk_fdt1}) gets a simplified
618: form
619: \begin{equation}\label{Pk_fdt2} \langle k_i^2\rangle-\langle
620: k_i\rangle^2\simeq\langle k_i\rangle,
621: \end{equation}
622: indicating the Poissonian fluctuations. Since however, in the case
623: of sparse homogeneous networks one can omit the last term in
624: (\ref{Pk_fdt1}), in sparse scale-free networks with the
625: characteristic exponent $2<\gamma<3$, the mentioned term is
626: dominated by hubs and the nodes susceptibilities
627: $\chi_i^{(\theta)}$ are much smaller than their expected degrees
628: $\langle k_i\rangle$. The total network susceptibility decreases
629: making the system resistant against random changes in the
630: landscape of multipliers and simultaneously susceptible to
631: behavior of supernodes \cite{HavPRL2000,HavPRL2001}.
632:
633: In order to establish the better understanding of the statement
634: included in the last paragraph, let us consider a trivial network
635: consisting of $N$ nodes with expected degrees $\langle k\rangle=1$
636: and one {\it supernode} with the tunable desired degree $\langle
637: k^{*}\rangle=1,2,\dots,N$ (in the sequel, the parameters denoted
638: by the star apply to the supenode). Adjusting the degree of the
639: supernode makes possible to smoothly pass between regular graphs
640: (for $\langle k^{*}\rangle=1$) and star networks (for $\langle
641: k^{*}\rangle=N$) (see Fig. \ref{fig1}). The transition enables the
642: understanding of what the reduced network susceptibility consists
643: in.
644:
645: \begin{figure} \epsfxsize=7.5cm \epsfbox{fig1.eps}
646: \caption{Schematic representation of networks possessing $N=19$
647: nodes with one nearest neighbor $\langle k\rangle=1$ and the
648: supernode (the gray one) with the tunable connectivity $\langle
649: k^*\rangle=1,2, \dots,N$.}\label{fig1}
650: \end{figure}
651:
652: First, let us find the Lagrange multipliers $\theta$ and
653: $\theta^{*}$ corresponding to the nodes of the considered
654: ensemble. Taking advantage of (\ref{Pkki}) one can show, that the
655: parameters fulfill the set of equations
656: \begin{displaymath}
657: \left\{
658: \begin{array}{ccl}1&=&
659: \dfrac{N-1}{e^{2\theta}+1}+\dfrac{1}{e^{\theta+\theta^{*}}+1}\\\langle
660: k^{*}\rangle&=&\dfrac{N}{e^{\theta+\theta^{*}}+1}.
661: \end{array}
662: \right.
663: \end{displaymath}
664: Solving the above equations for $\theta$ and $\theta^*$ one gets
665: (see Fig. \ref{fig2}a)
666: \begin{displaymath}
667: \left\{
668: \begin{array}{ccl}
669: \theta&=&\dfrac{1}{2}\ln\left[\dfrac{N^2-2N+\langle k^{*}\rangle}
670: {N-\langle k^{*}\rangle}\right]\\\theta^{*}&=&\ln\left[
671: \dfrac{N-\langle k^{*}\rangle}{\langle k^{*}\rangle}\right]-
672: \theta\nonumber.
673: \end{array}
674: \right.
675: \end{displaymath}
676: Next, inserting the multipliers into (\ref{Pkpij}) and then taking
677: advantage of (\ref{Pk_fdt1}) one obtains the susceptibilities of
678: expected degrees due to changes in the nodes intensive parameters.
679: The relative susceptibilities are respectively given by
680: \begin{eqnarray}\label{Pk_fdtex1}
681: \chi=-\frac{1}{\langle k\rangle}\frac{\partial\langle k
682: \rangle}{\partial\theta}=1-\frac{(N-\langle
683: k^*\rangle)^2}{N^2(N-1)}-\frac{\langle k^*\rangle^2}{N^2}
684: \end{eqnarray}
685: for the bulk of nodes, and
686: \begin{eqnarray}\label{Pk_fdtex2}
687: \chi^{*}=-\frac{1}{\langle k^*\rangle}\frac{\partial\langle k^*
688: \rangle}{\partial\theta^*}=1-\frac{\langle k^*\rangle}{N}
689: \end{eqnarray}
690: for the supernode.
691:
692: \begin{figure} \epsfxsize=7.8cm \epsfbox{fig2.eps}
693: \caption{(a) Lagrange multipliers $\theta$ and $\theta^*$ as a
694: function of $\langle k^*\rangle$, (b) relative susceptibilities
695: $\chi$ and $\chi^{*}$ as a function of $\langle k^*\rangle$. Here,
696: we have assumed $N=100$.} \label{fig2}
697: \end{figure}
698:
699: The behaviour of susceptibilities (\ref{Pk_fdtex1}) and
700: (\ref{Pk_fdtex2}) is depicted at the Fig. \ref{fig2}b. One can see
701: that the susceptibilities decrease to $0$ when the expected degree
702: of the hub $\langle k^*\rangle$ approaches $N$. In the region of
703: the vanishing susceptibilities, the small changes in the fields
704: $\theta$ and $\theta^*$ poorly affect the topological features
705: (i.e. $\langle k^*\rangle$) of the considered networks (c.f. Fig.
706: \ref{fig2}a). The last statement supports our previous claim of
707: the resilience of such networks against random external
708: perturbations. The large $\langle k^*\rangle$ makes that the
709: supernode accumulates most of the links present in the system and
710: causes that there exist relatively small number of network
711: realizations which both: i. fulfill physical constraints of the
712: ensemble and ii. possess significant statistical weights. For
713: example, only one such a realization exists in the limiting case
714: of the star network with $k^*=N$.
715:
716:
717: %_____________________________________________________________________ S: Hidden
718: \section{Equivalence of maximum entropy networks and networks with hidden variables}
719:
720: In the section, we continue the thread of Poissonian fluctuations
721: (\ref{Pk_fdt2}).
722:
723: Random networks with hidden variables are simply defined through
724: the construction procedure that consists of only two steps:
725: \begin{itemize}
726: \item[i.] first, prepare $N$ nodes and assign them hidden
727: variables independently drawn from the probability distribution
728: $R(h)$,
729: \item[ii.] next, each pair of nodes $\{i,j\}$ link with a
730: probability $r_{ij}$.
731: \end{itemize}
732: One can show, that the uncorrelated networks with hidden variables
733: arise from the factorized connection probability
734: \begin{equation}\label{rij}
735: r_{ij}=\frac{h_{i}h_{j}}{\langle h\rangle N},
736: \end{equation}
737: whereas networks with two-point correlations require more
738: sophisticated expressions for $r_{ij}$.
739:
740: Even comparing the above short review to our previous results on
741: sparse networks with an expected degree sequence (cf. Eqs.
742: (\ref{Pkpij1}) and (\ref{rij})), allows one to deduce on the
743: equivalence of the two approaches. In the course of the section,
744: we will argue that the claimed equivalence also holds for networks
745: with two-point correlations. We will prove it by recovering the
746: so-called Poissonian propagators characterizing both correlated
747: and uncorrelated sparse networks with hidden variables
748: \cite{BogPRE2003}.
749:
750: \subsection{Networks with a given degree sequence}
751:
752: At the moment, it is clear that although the expected degree of
753: the node $i$ is $\langle k_i\rangle$, but due to ensemble
754: fluctuations its actual degree $k_i$ changes from network to
755: network. Our aim is to find the so-called propagator
756: $P(k_i/\theta_i)$ i.e. the degree distribution of the node given
757: that it is characterized by the multiplier $\theta_i$.
758:
759: At the beginning, let us reformulate the probability of a graph
760: $G$ in the ensemble
761: \begin{equation}
762: P(G)=\frac{e^{-H(G)}}{Z},
763: \end{equation}
764: where $H(G)$ and $Z$ are respectively the graph Hamiltonian
765: (\ref{PkH}) and the partition function (\ref{PkZ}). Taking
766: advantage of the connection probability $p_{ij}$ (\ref{Pkpij}),
767: $P(G)$ can be written in a similar form as in the case of
768: classical random graphs (\ref{er1PG})
769: \begin{equation}\label{Pk1PG}
770: P(G)=\prod_{i<j}\Phi(i,j),
771: \end{equation}
772: where
773: \begin{equation}
774: \Phi(i,j)=p_{ij}^{\;\;\sigma_{ij}}(1-p_{ij})^{(1-{\sigma_{ij}})}
775: \end{equation}
776: whereas $\sigma_{ij}$ are elements of the adjacency matrix
777: describing $G$ and they are equal to either $1$ or $0$ depending
778: on whether $i$ and $j$ are connected or not.
779:
780: In the following, without the loss of generality, we will
781: concentrate on the node $i=1$. Having $P(G)$ one can calculate the
782: probability $P(\{\sigma_{1j}\})$ of the node to have a given
783: linkage profile $\{\sigma_{1j}\}$ (e.g. \{0,0,0,1,0,1,\dots,0\})
784: \begin{eqnarray}
785: P(\{\sigma_{1j}\})&=&\sum_{G^*}P(G)\nonumber\\&=&\prod_{j}\Phi(1,j)
786: \prod_{2\leq i<j}\;\sum_{\sigma_{ij}=0}^{1}\Phi(i,j)\nonumber
787: \\&=&\prod_{j}\Phi(1,j)\label{Pk2PG},
788: \end{eqnarray}
789: where the first sum runs over all networks $G^{*}$ with the fixed
790: sequence $\{\sigma_{1j}\}$ (i.e. the fixed neighborhood of the
791: node $i=1$). Now, in order to obtain $P(k_1/\theta_1)$ one has to
792: sum the probabilities (\ref{Pk2PG}) over different sequences
793: $\{\sigma_{1j}\}^*$ representing the same degree
794: $k_1=\sum_{j}\sigma_{1j}$
795: \begin{eqnarray}\label{prop0}
796: P(k_1/\theta_1)=\sum_{\{\sigma_{1j}\}^*}P(\{\sigma_{1j}\}).
797: \end{eqnarray}
798:
799: Up to this point, the derivation of $P(k_1/\theta_1)$ has been
800: exact. Now, before proceeding with approximations let us test the
801: formula (\ref{prop0}) against the simplest ensemble i.e. networks
802: with an expected homogenous degree sequence. In the ensemble, all
803: nodes have the same desired degree and also
804: $\forall_{i=1}^{N}\;\theta_i=\theta$. It is easy to check that the
805: degree distribution of an arbitrary node (\ref{prop0}) is given by
806: \begin{equation}\label{proper}
807: P(k/\theta)={N-1\choose k}p^{k}(1-p)^{N-1-k},
808: \end{equation}
809: where the binomial factor in the front of the expression arises
810: from the fact that there exist ${N-1\choose k}$ different
811: connection profiles corresponding to degree $k$ and
812: $p=(e^{2\theta}+1)^{-1}$ (\ref{Pkpij}) (please do not confuse it
813: with (\ref{per}), where $\theta$ has a different meaning !). One
814: should not be surprised with the last result. If it is not
815: obvious, let us stress that the ensemble of networks with an
816: expected homogeneous degree sequence is in fact equivalent to the
817: ensemble of classical random graphs. To become familiar with the
818: statement compare the formulas (\ref{Zer}) and (\ref{PkZ}).
819:
820: Now, in order to recover the claim of equivalence between the
821: considered maximum-entropy models and random networks with hidden
822: variables one has to apply the mean field approximation to the
823: expression (\ref{prop0})
824: \begin{eqnarray}\label{prop1}
825: P(k_1/\theta_1)\simeq{N-1\choose k_1}\langle p_{1j}\rangle^{k_1}
826: (1-\langle p_{1j}\rangle)^{(N-1-k_1)},
827: \end{eqnarray}
828: where $\langle p_{1j}\rangle=\sum_jp_{1j}/(N-1)=\langle
829: k_{1}\rangle/(N-1)$ (\ref{Pkki}). The assumption of sparse
830: networks enables further simplification of the distribution
831: \begin{eqnarray}\label{prop2}
832: P(k_1/\theta_1)\simeq\frac{e^{-\langle k_1\rangle}\langle
833: k_1\rangle^{k_1}}{k_1!},
834: \end{eqnarray}
835: recovering the result previously derived by S\"{o}derberg
836: \cite{SodPRE2002} for uncorrelated networks with hidden variables.
837: The Poissonian propagator (\ref{prop2}) indicates the mentioned
838: equivalence of the considered maximum-entropy networks and the
839: well-known class of uncorrelated random networks with hidden
840: variables (Lagrange multipliers characterizing nodes correspond to
841: hidden attributes).
842:
843: \par As pointed out in \cite{BogPRE2003}, the result is indeed very
844: strong since it also holds for random networks with two-point
845: degree correlations (see Eq. (\ref{prop3}) and also Eq. (23) in
846: \cite{BogPRE2003}). The key point to notice with reference to the
847: last expression is that in the region of small degrees, due to the
848: Poissonian fluctuations (c.f. (\ref{Pk_fdt2}) and (\ref{prop2})),
849: the {\it real} degree distributions observed in single
850: realizations of the considered networks strongly differ from the
851: desired degree distribution $P(\langle k_i\rangle)$. On the other
852: hand, in the limit of large degrees the real distribution and the
853: expected one converge. The interplay between the two distributions
854: has been carefully studied in our previous paper \cite{Fron2005b}.
855:
856: %______________________________________________________________________________ SubS: Correlations
857: \subsection{Networks with two-point correlations}
858:
859: One can show that probability of a graph $G$ in the ensemble
860: (\ref{corH}) can be transformed into the same form (\ref{Pk1PG})
861: as the one for random networks with an expected degree sequence,
862: where the linkage probability is given by (\ref{corpij}).
863: Performing the same calculations as in the case of ensembles
864: analyzed in the previous subsection, one can prove that in the
865: limit of sparse networks the degree distribution of a specific
866: node is Poissonian (\ref{prop2})
867: \begin{equation}\label{prop3}
868: P(k_1/\{\Theta_{1,i}\})\simeq\frac{e^{-\langle k_1\rangle}\langle
869: k_1\rangle^{k_1}}{k_1!},
870: \end{equation}
871: where $\langle k_1\rangle=\sum_{i}p_{1i}$ and $p_{1i}$ represents
872: the connection probability given by (\ref{corpij}). Again, the
873: last formula supports the claimed equivalence between the analyzed
874: maximum-entropy networks with two-point correlations and the class
875: of correlated networks with hidden variables \cite{BogPRE2003}.
876:
877: %______________________________________________________________________________ S: Conclusions
878: \section{Conclusions}
879:
880: In this paper we have extended the information-theoretic approach
881: to random networks that was very recently proposed by Park and
882: Newman \cite{ParkPRE2004}. We have concentrated on fluctuations
883: over ensembles of undirected networks with non-interacting edges,
884: including random networks with a given degree sequence and
885: networks characterized by two-point correlations. We have studied
886: a few fluctuation-dissipation relations characterizing
887: susceptibilities of different networks to changes in the external
888: fields. In the case of networks with a given degree sequence, we
889: have argued that the scale-free topologies of real networks may
890: arise as a result of self-organization of real systems into sparse
891: structures with the low susceptibility to random external
892: perturbations. Finally, we have also shown that maximum-entropy
893: networks are equivalent to random networks with hidden variables.
894:
895: \section{Acknowledgments}
896:
897: A.F. acknowledges financial support from the Foundation for Polish
898: Science (FNP 2005). A.F. and J.A.H. were partially supported by
899: the State Committee for Scientific Research in Poland (Grant No.
900: 1P03B04727). The work has also been supported by the European
901: Commission Project CREEN FP6-2003-NEST-Path-012864.
902:
903: \begin{thebibliography}{7}
904: \bibitem{Handbook} S.Bornholdt and H.G.Schuster, {\it Handbook of Graphs and Networks}, Wiley-Vch (2002).
905: \bibitem{Dorogbook} S.N. Dorogovtsev and J.F.F.Mendes, {\it Evolution of Networks}, Oxford Univ.Press (2003).
906: \bibitem{BArev} R.Albert and A.L.Barab\'asi, Rev. Mod. Phys. {\bf 74} 47 (2002).
907: \bibitem{BAScience} A.L. Barab\'asi and R. Albert, Science {\bf 286}, 509 (1999).
908: \bibitem{Molloy1995} M. Molloy and B. Reed, Ran. Struct. and Algor. {\bf 6}, 161 (1995).
909: \bibitem{NewPRE2001} M.E.J. Newman, S.H. Strogatz and D.J. Watts, Phys. Rev. E {\bf 64}, 026118 (2001).
910: \bibitem{Fron2005a} A. Fronczak, P. Fronczak and J.A. Ho\l yst, cond-mat/0502663.
911: \bibitem{SodPRE2002} B. S\"{o}derberg, Phys. Rev E {\bf 66}, 066121 (2002).
912: \bibitem{BogPRE2003} M. Bogu\~{n}\'{a} and R. Pastor-Satorras, Phys. Rev. E {\bf 68}, 036112 (2003).
913: \bibitem{Fron2005b} A. Fronczak and P. Fronczak, cond-mat/0503069.
914: \bibitem{BurdaPRE2001} Z. Burda, J.D. Correia and A. Krzywicki, Phys. Rev. E {\bf 64}, 046118 (2001).
915: \bibitem{BergPRL2002} J. Berg and M. L\"{a}ssig, Phys. Rev. Lett. {\bf 89}, 228701 (2002).
916: \bibitem{Bauer2002} M. Bauer and D. Bernand, cond-mat/0206150.
917: \bibitem{BurdaPRE2003} Z. Burda and A. Krzywicki, Phys. Rev. E {\bf 67}, 046118 (2003).
918: \bibitem{DorogNuc2003} S.N. Dorogovtsev, J.F.F. Mendes and A.N. Samukhin, Nucl. Phys. B {\bf 666}, 396 (2003).
919: \bibitem{PallaPRE2004} G. Palla, I. Der\'enyi, I. Farkas and T. Vicsek, Phys. Rev. E {\bf 69}, 046117 (2004).
920: \bibitem{Farkas2004} I. Farkas, I. Der\'enyi, G. Palla and T. Vicsek, Springer Lect. Notes Phys. {\bf 650}, 163 (2004).
921: \bibitem{ParkPRE2004} J. Park and M.E.J. Newman, Phys. Rev. E {\bf 70}, 066117 (2004).
922: \bibitem{Bogacz2005} L. Bogacz, Z. Burda and B. Wac\l aw, cond-mat/0502124.
923: \bibitem{Bialas2005} P. Bia\l as, Z. Burda, B. Wac\l aw, cond-mat/0503548.
924: \bibitem{Jayens1} E.T. Jaynes, Phys. Rev. {\bf 106}, 620 (1957).
925: \bibitem{bookCover} T.M. Cover and J.A. Thomas, {\it Elements of Information Theory}, John Willey $\&$ Sons (1991).
926: \bibitem{bookKapur} J.N. Kapur, {\it Maximum-Entropy Models in Science and Engineering}, John Willey $\&$ Sons (1989).
927: \bibitem{footnote1} Of course, other information measures including those nonextensive can also be applied (see \cite{Wilk2003}).
928: \bibitem{Wilk2003} G. Wilk and Z. W\l odarczyk, Acta Phys. Polon. B {\bf 35}, 871 (2004).
929: \bibitem{footnote2} One can also decide to study digraphs, weighted graphs, regular graphs and so forth.
930: \bibitem{footnote3} The issue of an expected degree distribution in maximum-entropy networks has been worked out in \cite{Bauer2002}.
931: \bibitem{footnote4} In a finite network, the cut-off of degree distribution $P(k)\sim k^{-\gamma}$ may be estimated from $\int_{k_{max}}^{\infty}P(k)dk=1/N$ yielding $k_{max}=N^{1/(1-\gamma)}.$
932: \bibitem{MaslovPhysA2004} S. Maslov, K. Sneppen and A. Zaliznyak, Physica A {\bf 333}, 529 (2004).
933: \bibitem{ParkPRE2003} J. Park and M.E.J. Newman, Phys. Rev. E {\bf 68}, 026112 (2003).
934: \bibitem{BogEPJ2004} M. Bogu\~{n}\'{a}, R. Pastor-Satorras and A. Vespignani, Euro. Phys. J. B {\bf 38}, 205 (2004).
935: \bibitem{CatanPRE2005} M. Catanzaro, M. Bogu\~{n}\'{a} and R. Pastor-Satorras, Phys. Rev. E {\bf 71}, 027103 (2005).
936: \bibitem{HavPRL2000} R. Cohen, K. Erez, D. ben-Avraham and S. Havlin, Phys. Rev. Lett. {\bf 85}, 4626 (2000).
937: \bibitem{HavPRL2001} R. Cohen, K. Erez, D. ben-Avraham and S. Havlin, Phys. Rev. Lett. {\bf 86}, 3682 (2001).
938:
939: \end{thebibliography}
940:
941: \end{document}
942: