cond-mat0509238/AEH.tex
1: \documentclass[12pt]{article}
2: 
3: \usepackage{fullpage}
4: \usepackage{graphicx}
5: \usepackage{graphics}
6: \usepackage{epsfig}
7: \usepackage{latexsym}
8: \usepackage{amsfonts}
9: \usepackage{amssymb}
10: \usepackage{amsmath}
11: 
12: \newcommand{\ul}{\underline}
13: 
14: \title{Condensation transitions in a model for a directed network with
15: weighted links}
16: 
17: \author{A. G. Angel, T. Hanney, M. R. Evans \\  {\small
18:       School of Physics, University of Edinburgh, Mayfield Road,
19:       Edinburgh, EH9 3JZ, UK} } 
20: 
21: \begin{document}
22: 
23: \maketitle
24: 
25: \begin{abstract}
26: An exactly solvable model for the rewiring dynamics of weighted,
27: directed networks is introduced. Simulations indicate that the model
28: exhibits two types of condensation: (i) a phase in which, for each
29: node, a finite fraction of its total out-strength condenses onto a
30: single link; (ii) a phase in which a finite fraction of the total
31: weight in the system is directed into a single node.  A virtue of the
32: model is that its dynamics can be mapped onto those of a zero-range
33: process with many species of interacting particles --- an exactly
34: solvable model of particles hopping between the sites of a
35: lattice. This mapping, which is described in detail, guides the
36: analysis of the steady state of the network model and leads to
37: theoretical predictions for the conditions under which the different
38: types of condensation may be observed. A further advantage of the
39: mapping is that, by exploiting what is known about exactly solvable
40: generalisations of the zero-range process, one can infer a number of
41: generalisations of the network model and dynamics which remain exactly
42: solvable.
43: \end{abstract}
44: 
45: \section{Introduction} \label{introduction}
46: Networks, and in particular weighted networks, have a long history in
47: the social, natural and engineering sciences. The many and varied
48: examples of networks range from systems such as transportation networks to
49: ecological, biochemical and social networks \cite{AB02,DM03}.
50: Basically, a network is a graph containing nodes connected by links
51: (or edges) --- the network is said to be `weighted' when weights are
52: assigned to the links. The interest in networks within the physics
53: community was initially with regard to the statistical properties of
54: the connectivity structure of unweighted networks (i.e., networks
55: where links are restricted to have weight one or zero only, indicating
56: a connection or no connection respectively). Studies focussed, for
57: example, on the degree distribution, i.e., the distribution of the
58: number of links connected to a node, and the clustering coefficient
59: which probes how trios of nodes are correlated.  Subsequently simple
60: models for growing networks exhibiting interesting statistical
61: properties were proposed \cite{BA99}.
62: 
63: Recently, weighted networks have also become of interest to this
64: community. These networks may be defined through their adjacency
65: matrix $w_{ij}$, which gives the weight of the link from node $i$ to
66: node $j$. The links may be directed or undirected corresponding to an
67: asymmetric or symmetric adjacency matrix.  Data are available for many
68: real weighted networks, of either directed or undirected kind, leading
69: to studies of the statistical properties of Scientific Collaboration
70: Networks (SCN) \cite{Newman01a,Newman04} and the World-wide Airport
71: Network (WAN) \cite{GA04,BBPV05}. Typical quantities used to describe
72: the topology of general weighted networks are: the distribution of
73: in-strength and out-strength, which are defined as the total weight
74: going into or out of a node due to all the connected links; weighted
75: clustering coefficients; generalisations of the idea of minimal
76: spanning trees to weighted networks \cite{NR02,MAB04}.  An interesting
77: aspect of weighted networks is how the weights of the links are
78: correlated to the topology of the network: the weights may be
79: determined solely by the global topology, they may be independent, or
80: they may be correlated in some more complicated way.  For example, in
81: Transportation Networks, where the weight represents the flow through
82: a link, the weights are determined solely by the topology; in the SCN
83: it appears that the degree of a node and weights of its links are
84: uncorrelated; in the WAN, correlations between the node degree and the
85: weights of its links lead the strength to grow faster than the degree.
86: 
87: Attention has recently focussed on constructing simple models of
88: `growing' weighted networks, with the aim being to investigate the
89: resulting degree and weight distributions and the relation between the
90: two. Initial work had the weight of links determined by the degree of
91: the node they were attached to \cite{MAB04,YJBT01,AK05,AKR05}.  To
92: loosen this coupling between weight and degree, a model with dynamical
93: evolution of weights during growth was introduced in
94: \cite{BBV04,BBV04a, BBV04b}.  Moreover, models of growing networks in
95: which the dynamics depend on the weights of the links rather than the
96: degree of the nodes were proposed in \cite{DM04}.  For a review of
97: recent work on weighted networks see \cite{BBPV05}.
98: 
99: An alternative class of evolving network models, which is the focus of
100: this work, is that of `rewiring' or `equilibrium' networks
101: \cite{DM03,BCK01,DMS03}. In these models, usually considered for
102: unweighted networks, the number of nodes is fixed and the dynamics
103: involves the rewiring of links between nodes, although the effect of
104: creating and destroying links has been examined recently
105: \cite{AELM05}. For rewiring dynamics, one is interested in the
106: steady-state properties of an ensemble of networks.  Under certain
107: conditions on the rewiring rules \cite{EH05} one can obtain a
108: condensed phase where one node (the condensate) attracts a finite
109: fraction of the available links. Thus there is a single extensively
110: connected node while the other nodes exhibit a power-law background in
111: the degree distribution. Hence the network is scale free if one
112: discounts the condensate. The transition is reminiscent of the
113: `gelation' transition in growing networks from the scale free phase to
114: a phase with a dominant `hub' \cite{KRL00,KR01,BB01}.
115: 
116: These rewiring networks may be related to interacting particle systems
117: such as the zero-range process \cite{Evans00} for which condensation
118: transitions have been widely studied --- for a recent review see
119: \cite{EH05}. In the zero-range process, particles hop between the
120: sites of a lattice with a rate that depends on the number of particles
121: at the site of departure. The idea behind the mapping is to identify
122: the nodes of the network with the sites of the lattice model and the
123: links of the network become the particles --- the rewiring of links
124: corresponds to particles hopping from site to site. At the simplest
125: level, this mapping is approximate in that the number of particles at
126: a site represents only the degree of the corresponding node, but does
127: not determine the other nodes to which a node is attached. However,
128: the steady states of the two systems have the same form.
129: 
130: The aim of the present work is to introduce a model for the rewiring
131: dynamics of a directed, weighted network. By virtue of an exact
132: mapping between the network model and a set of coupled interacting
133: particle systems, we are able to analyse exactly conditions on the
134: rewiring rates under which the network will exist in some kind of
135: condensed phase. Specifically, we define the network model in Section
136: \ref{network model}, the key features of which are:
137: \begin{enumerate}
138: \item The network is directed.
139: 
140: \item The out-strength of each node is conserved under the dynamics
141: but the in-strength evolves.
142: 
143: \item The dynamics are governed by the weights of links and
144: the in-strength of the nodes.
145: 
146: \end{enumerate}
147: In addition, the weights are integer variables, although this
148: constraint may be relaxed (as we discuss in Section
149: \ref{generalisations}). Thus we define a weighted network model with a
150: fixed number of nodes where strong nodes and links with large weight
151: may tend to attract more weight.
152: 
153: In Section \ref{mapping}, we show how this model enjoys a mapping to
154: an interacting particle system known as a multi-species zero-range
155: process \cite{EH03,HE04}. The idea is that, since the out-strength of
156: each node is conserved, the dynamics at each node is basically the
157: exchange of weight between the links coming out of that node. Thus
158: each node may be related to an interacting particle system in which
159: the exchange of particles between sites represents the exchange of
160: weight between links. However, since the dynamics at each node also
161: depends on the in-strength, these interacting particle systems are
162: coupled.  This mapping of the weighted network to a set of coupled
163: interacting particle systems contrasts with the previous mappings of
164: an unweighted network to a single interacting particle system
165: discussed in \cite{DM03,BCK01,DMS03,EH05}.
166: 
167: The results of simulations, presented in Section \ref{simulations},
168: indicate two distinct types of condensation within this model: the
169: first is when, for any given node, a finite fraction of the
170: out-strength condenses onto one weight out of that node; the second is
171: when a finite fraction of the total weight in the system is directed
172: onto one particular node.  The first scenario gives a realisation of a
173: transition in the `disparity', the idea of `disparity', discussed in
174: the context of internet traffic \cite{BGG03} and simple growing
175: weighted network models \cite{B05}, being that one weight from a node
176: is dominant over the others from that node. The second scenario is
177: related to the usual condensation transition occurring in unweighted
178: rewiring networks, wherein a single `hub' acquires a finite fraction
179: of the total number of links in the system.
180: 
181: We exploit the mapping to the interacting particle system in Section
182: \ref{theory}, in order to solve and analyse steady state properties of
183: the network model exactly. In particular, we derive conditions on the
184: rewiring rates which lead to the two types of condensation observed in
185: simulations. The mapping also enables one to identify a number of
186: possible generalisations of the network model and its rewiring
187: dynamics, which we discuss in Section \ref{generalisations}. Finally,
188: we conclude in Section \ref{conclusion}.
189: 
190:  
191: \section{Model}
192: \label{network model}
193: 
194: We study a model  of a dynamically evolving, directed network of $L$
195: nodes with weighted links. Since the weights are integers we will
196: denote the weight associated with the link from node $k$ into node $l$
197: by an integer $n_{kl}\geq 0$ (instead of the usual $w_{kl}$). When
198: $n_{kl}=0$, this represents no link from $k$ to $l$. We depict such a
199: network with $L=4$ nodes in Figure \ref{fig:dwnet}.
200: 
201: The dynamics of the network are such that one unit of the weight
202: associated with the link pointing from node $k$ into node $l$ is
203: rewired to point into another randomly selected node $l'$ with a rate
204: $u_k(\underline{n}_l)$, where $\underline{n}_l \equiv n_{1l}, \ldots,
205: n_{Ll}$. In general, this rate is a function of the weights associated
206: with all of the links pointing into node $l$, and it depends on the
207: source node $k$: we will specialise to a rate of the form
208: \begin{equation} \label{rate}
209: u_k(\underline{n}_l) = u^s (n_{kl}) u^c(\sum_{k=1}^L n_{kl})\;,
210: \end{equation}
211: which depends on the weight of the link being rewired, through the
212: function $u^s(n_{kl})$, and the total in-strength of node $l$, through
213: the function $u^c(\sum_{k=1}^L n_{kl})$.  So for the example in Figure
214: \ref{fig:dwnet}, the rate at which the one unit of weight for the link
215: connecting node $C$ to node $A$ is rewired to another node is
216: $u_C(n_{AA}, n_{BA}, n_{CA}, n_{DA})=u^s(3) u^c(10)$. After the
217: rewiring, $n_{CA}=2$ and any of $n_{CB}$, $n_{CC}$ or $n_{CD}$ have
218: increased by one. These dynamics conserve the {\it out-strength}
219: $M_k=\sum_{l=1}^L n_{kl}$ which is the total weight of all links
220: pointing out of node $k$; the total weight of all links pointing into
221: node $l$, i.e., the {\it in-strength} $X_l=\sum_{k=1}^L n_{kl}$, is not
222: conserved.  For simplicity we will take $M_k = M$ for all $k$. Later,
223: in Section \ref{generalisations}, we will consider several
224: generalisations of the model and dynamics. In particular we consider
225: rewiring dynamics which depend on the node to which a link is being
226: rewired. This allows one to make a connection with a form of preferential
227: attachment in which links are preferentially rewired to nodes with
228: large in-strengths \cite{BBV04}. We proceed with the dynamics defined
229: above in order to make a clear connection to well-studied interacting
230: particle systems.  We stress that either alternative for the dynamics
231: yields the same behaviour and transitions.
232: \begin{figure}
233: \begin{center}
234: \begin{equation}
235: \setlength{\arraycolsep}{4pt}
236: \raise-11.1ex\hbox{\includegraphics[width=3.5cm]{dwnet2.eps}}\quad
237:   \raise-1.8ex\hbox{$\equiv$}\quad
238:     \begin{matrix}
239:       \hbox to 4.5em{\vphantom{\Big|}\footnotesize A\hfil B\hfil C\hfil D}
240:       \\
241:       \begin{pmatrix}
242:         0 & 0 & 0 & 5 \\
243:         3 & 2 & 0 & 0 \\
244:         3 & 0 & 0 & 2 \\
245:         4 & 0 & 1 & 0
246:       \end{pmatrix}
247:       &
248:       \begin{matrix}
249:         \mbox{\footnotesize A} \\
250:         \mbox{\footnotesize B} \\
251:         \mbox{\footnotesize C} \\
252:         \mbox{\footnotesize D}
253:       \end{matrix}
254:     \end{matrix} \nonumber
255: \end{equation}
256: \caption{A picture of a weighted directed network with $L=4$ nodes,
257:   labelled $A$, $B$, $C$ and $D$. The direction and weight of each
258:   link is shown. The corresponding adjacency matrix is also shown.} 
259: \label{fig:dwnet}
260: \end{center}
261: \end{figure}
262: 
263: 
264: \subsection{Mapping to a system of interacting particles} \label{mapping}
265: To make the connection with interacting particle systems, we use the
266: adjacency matrix, $A$, to encode the network: the value of element
267: $(k,l)$ of the matrix is $n_{kl}$, the weight of the link pointing out
268: of node $k$ into node $l$, as shown for example in Figure
269: \ref{fig:dwnet}. Note that here, because the network is directed,
270: $A\neq A^T$. The idea is to represent the adjacency matrix by a
271: lattice, where site $(k,l)$ of the lattice represents element $(k,l)$
272: of the adjacency matrix. The value of the element $(k,l)$, $n_{kl}$,
273: is then understood as the number of particles at site $(k,l)$ of the
274: lattice. The lattice is therefore two-dimensional, since it represents
275: the elements of a matrix, even though the network model may typically
276: be defined in a fully connected geometry (i.e.\ a link pointing into
277: one node is rewired to point into any other randomly selected node).
278: 
279: The network dynamics then causes the elements of the adjacency matrix
280: to change. Thus the corresponding particle configurations of the
281: lattice model also evolve according to dynamical rules, rules which
282: have their origin in the dynamics of the network model. 
283: For the network dynamics
284: described above the corresponding interacting particle system
285: is a generalisation of the much studied zero-range process \cite{EH05},
286: for which the steady state is exactly solvable.
287: 
288: 
289: \subsection{Model as a ZRP with $L$ species of particles}
290: The interacting particle system is defined on a square lattice (since
291: the adjacency matrix is a square matrix) with $L\times L$ sites. On
292: this lattice $n_{kl}$ labels the number of particles at the site
293: located at row $k=1,\ldots,L$, column $l=1,\ldots,L$.  There are $M$
294: particles in each row, a total of $LM$ in the system. Under the
295: network dynamics described in Section \ref{network model}, both $M$
296: and $L$ are conserved. Rewiring single units of the weight of a link
297: connecting node $k$ to node $l$ corresponds, in the interacting
298: particle system, to a particle in row $k$ and column $l$ hopping to
299: another randomly selected site \emph{within the same row} with a rate
300: $u_k(\underline{n}_l)$, given by (\ref{rate}).
301: 
302: We remark that these are the dynamics of a ZRP with $L$ species of particles,
303: labelled $k=1,\ldots,L$. Each site $l=1,\ldots,L$ of a one-dimensional
304: lattice contains $n_{kl}$ particles of species $k$. A particle of
305: species $k$ then hops with a rate $u_k(\underline{n}_l)$ to any other site on
306: the lattice, i.e.\ a rate which depends upon the number of particles of
307: each species at the departure site $l$. Thus we identify rows of the
308: two-dimensional lattice with different particle species (so the total
309: number of particles of each species is conserved) and the columns of
310: the two-dimensional lattice are identified with sites of a
311: one-dimensional lattice, as illustrated in Figure \ref{fig:lats}.
312: \begin{figure}
313: \begin{center}
314: \includegraphics[scale=0.5]{latts.eps}
315: \caption{A picture of the interacting particle systems which represent
316: the network depicted in Figure \ref{fig:dwnet}. On the left-hand side
317: is the two-dimensional lattice model with the occupation numbers
318: $n_{kl}$ entered in row $k$ and column $l$ of the lattice. On the
319: right-hand side is the corresponding multi-species ZRP (here there are
320: four species) --- note that there is no ordering amongst the particles
321: at a particular site.}
322: \label{fig:lats}
323: \end{center}
324: \end{figure}
325: This mapping is useful because much is known about the exact steady states
326: of zero-range processes and many of its generalisations. The model
327: defined above can be solved exactly in the steady state: the analysis
328: which follows exploits this knowledge and one expects that many
329: variations of the network model will also be exactly solvable, those
330: which are can be inferred from the corresponding zero-range process.
331: 
332: Although the model we consider can be thought of as a ZRP with $L$
333: species of particles, we will proceed with the terminology of the rows
334: and columns of a two-dimensional lattice, since this retains closer
335: contact with the adjacency matrix of the original network model.
336: 
337:  
338: 
339: \section{Simulations} \label{simulations}
340: 
341: In this section we present results from simulations of the model that
342: show the two types of condensation: {\em site condensation}, where a
343: finite fraction of the particles in each row condenses onto a single
344: randomly  located
345: site of the row;  and {\em column condensation}, 
346: where a finite
347: fraction of all the particles in the system condenses into a single
348: column.  In the network context this corresponds to a finite
349: fraction of the weight from each node being contained in a single link
350: from that node; and a finite fraction of the weight pointing from all
351: nodes to a single node, respectively.  We also present simulations
352: which exhibit characteristics of both
353: types of condensation which are 
354: interesting from a networks perspective.
355: 
356: The model is simulated using a simple Monte Carlo algorithm.  At each
357: time step the following update procedure is followed:
358: \begin{enumerate}
359: \item A site of the lattice $(k,l)$ is chosen at random.
360: \item If the site is occupied then a particle will be removed from
361: this site with probability $u(\underline{n}_l) \delta t$, where
362: $u$ is the hop rate (\ref{rate}) and $\delta t$ is the time interval,
363: chosen such that the probability of a hop occurring is always less
364: than or equal to one.
365: \item If a particle was removed, then a second site $(k,m)$ (within 
366: the same row) is chosen at random and a particle is added to this site.
367: \end{enumerate}
368: 
369: All simulations were run on lattices of size $100 \times 100$, i.e.\
370: on networks with 100 nodes, and for $\mathcal{O}(10^{7})$
371: Monte Carlo Sweeps.  The first half of each run was used to relax to
372: the steady state and the second half to measure probability
373: distributions.  Distributions were measured for
374: \begin{itemize}
375: \item the number of particles at a site (link weight)
376: \item  the number of particles in a column
377: (node in-strength)
378: \item  the number of occupied sites in a column and
379: row (in- and out-degree respectively).
380: \end{itemize}
381: Typical configurations of the lattice were also output.
382: 
383: In order to compare with theory (see Section \ref{theory}), the
384: following hop rates were chosen to show the two condensation types.
385: Recall that we consider a hop rate for particles from site $(k,l)$ of
386: the form (\ref{rate}). For site condensation we make the choice 
387: \begin{equation} \label{site rate}
388: u^s(n) = (1+b^s/n)  \qquad \textrm{and} \qquad u^c(n) = 1\,,
389: \end{equation}
390: with $b^s = 4$ and for $n>0$. For column condensation 
391: we make the choice 
392: \begin{eqnarray} \label{col rate}
393: u^c(n) = \left\{ \begin{array}{ll}
394: 1 + b^c & \textrm{for $n \leq L$}\,, \\ 
395: 1 + b^c L/n & \textrm{for $n > L$}\,,
396: \end{array} \right. \qquad \textrm{and} \qquad u^s(n) = 1\,,
397: \end{eqnarray}
398: with $b^c = 1.05$.
399: 
400: \subsection{Site Condensation} \label{site condensation}
401: The simulations for the site condensation case were run with $175$
402: particles in each row starting from a random initial configuration. The
403: site condensation can be seen clearly from the site distribution
404: (black circles, Figure~\ref{simfigsc} (a)).  The peak at around
405: $n=150$ represents the site condensates, it has an area of order $1/L$
406: and there are $L^2$ sites indicating that $L$ sites have occupations
407: of around $150$ particles.  As the number of particles in each row is
408: fixed at $175$, there can be at most one such site in a row and an
409: occupation of this size represents an appreciable fraction of the
410: total number of particles available to a site, hence implying a
411: condensation in the thermodynamic limit.  Thus we have a single
412: condensed site in each row.  Note also that the site distribution has
413: an apparently power-law background before the condensate peak, this is
414: reminiscent of the behaviour of a single-species ZRP \cite{EH05},
415: indeed as the column attraction has effectively been switched off, the
416: system is a collection of single-species ZRPs.  The column
417: distribution (grey crosses, Figure~\ref{simfigsc} (a)) shows that, in
418: the absence of any coupling between the rows, the condensed sites are
419: randomly distributed among the columns --- as one would expect.  The
420: first peak in the distribution at around $n=50$ shows that some
421: columns do not contain any condensed sites, just those from the
422: power-law background.  The second, third, $\ldots$ peaks correspond to
423: columns with one, two, $\ldots$ condensed sites in them, plus a
424: background from the rest of the sites.  This is exactly what one would
425: expect if the condensed sites were randomly distributed in the columns
426: and a random sampling of the site distribution to create a column
427: distribution agrees closely with the simulation.  The typical
428: configuration (Figure~\ref{simfigscconf}) bears this out, condensates
429: can be seen on individual sites, but no column ordering can be
430: discerned.  Although not shown here, systems with weak column
431: attraction showed similar behaviour, i.e. as though no column
432: attraction was present.
433: 
434: The degree distributions for the site condensation case
435: (Figure~\ref{simfigsc} (b)) appear to be binomial in nature: a
436: binomial distribution constructed from the measured probability of a
437: site having zero occupation matches very closely the data shown.  Thus
438: in the network context, for the site condensation we have connectivity
439: similar to that of a random graph (which also has a binomial 
440: connectivity distribution). In addition to this we have
441: a condensed link weight from each site with a power-law background,
442: and an in-strength which is a random sum of link weights.
443: 
444: \begin{figure}%[htb]
445: \begin{center}
446: \includegraphics[width=12cm]{sitecond_sc_idod.eps}
447: \caption{Distributions from simulation of the model with 
448: the hop-rate chosen to display site condensation.  
449: (a) Site and column occupation distributions, corresponding 
450: to the distributions of weight among links and in-strength 
451: among nodes in the network context respectively. 
452: (b) In- and out-degree distributions.}
453: \label{simfigsc}
454: \end{center}
455: \end{figure}
456: 
457: \begin{figure}%[htb]
458: \begin{center}
459: \includegraphics[width=12cm]{configsitecond.eps}
460: \caption{Typical configuration of the system, output from simulation 
461: of the model with the hop-rate chosen to display site condensation.}
462: \label{simfigscconf}
463: \end{center}
464: \end{figure}
465: 
466: \subsection{Column Condensation}  \label{column condensation}
467: The simulations for the column condensation case were run with $1000$ 
468: particles in each row. The initial condition was taken with all
469: particles on the same site.  
470: The column condensation can be seen clearly from the column
471: distribution (grey crosses, Figure~\ref{simfigcc} (a)).  The peak at
472: around $n=85000$ corresponds to the condensate, it has an area of
473: order $1/L$ and there are $L$ columns, indicating that a single column
474: has an occupation of around $85000$, an appreciable fraction of the
475: total number of particles in the system $L\times M = 100 \times 1000$.
476: The other, lower peak at around $n=100$ represents the columns which do
477: not contain a condensate.  The site distribution (black circles,
478: Figure~\ref{simfigcc} (b)) also shows condensation onto sites in the
479: peak at around $n=850$.  If the column interaction were to be switched
480: off leaving $L$ uncoupled single-species ZRPs, it is known that the
481: chosen $u^s(n)$ ($= \mathrm{constant}$) would not give condensation in 
482: this fully connected homogeneous system, see for example \cite{EH05}.  
483: Thus in this case it is
484: the column interaction that has induced a site condensation.  
485: Note that both the column and site distributions have apparent 
486: power-law pieces, with equal or close exponents.  The theory 
487: presented in Section~\ref{theory} allows one to construct the 
488: critical column distribution at the transition point --- shown as 
489: the dotted line in Figure~\ref{simfigcc}~(a).  This fits the 
490: non-condensate background part of the measured column distribution 
491: very well. The typical configuration of the system, 
492: Figure~\ref{simfigccconf},
493: shows the column condensate comprises site condensates all in the
494: same column.
495: 
496: The out-degree distribution for the column condensation case (grey
497: triangles, Figure~\ref{simfigcc}~(b)) is similar to the site
498: condensation case, being distributed according to a binomial
499: distribution.  However, the in-degree distribution
500: (Figure~\ref{simfigcc}~(b), black squares) is significantly different,
501: in particular it has a broader tail and also a single site which is
502: connected to all others, as shown by the lone data point at degree $L$
503: with probability $1/L$ ($L=100$ in this case). Thus in the network
504: context the out-connectivity resembles that of a
505: random graph, but
506: the in-degree distribution displays a broader tail and there is a
507: single node to which all others point --- the node corresponding to
508: the condensed column. Furthermore, all links pointing to this node
509: hold a finite fraction of the weight available and the node holds a
510: finite fraction of the total in-strength available to it.
511: Broadly-tailed degree distributions are often observed in real
512: networks, weighted and unweighted alike.  While the tail observed 
513: in these simulations is not as broad as many observed, it is at least 
514: broader than that of the equivalent random graph.  Also apart from the
515: condensed pieces the in-strength and weight distributions display
516: power-law tails, again something that has been observed in many real
517: networks.  Thus, as with equilibrium networks \cite{DM03,DMS03},
518: certain distributions may take a power-law form at a critical point.
519: Power-law strength and weight distributions and non-power-law degree
520: distributions have been observed for a traffic network in
521: \cite{MBCV05}.
522: 
523: As mentioned at the beginning of this subsection, we employed a fully
524: condensed initial condition. The reason is that, for a random initial
525: configuration, the dynamics slows down as the system reaches a state
526: in which several columns contain a large number of particles; the
527: single column condensate is attained only on prohibitively long
528: timescales.
529: 
530: \begin{figure}%[htb]
531: \begin{center}
532: \includegraphics[width=12cm]{colcond_sc_idod.eps}
533: \caption{Distributions from simulation of the model with the
534: hop-rate chosen to display column condensation.
535: (a) Site and column occupation distributions, corresponding to 
536: the distribution of weight among links and in-strength among
537: nodes in the network context respectively.
538: (b) In- and out-degree distributions.}
539: \label{simfigcc}
540: \end{center}
541: \end{figure}
542: 
543: \begin{figure}%[htb]
544: \begin{center}
545: \includegraphics[width=12cm]{configcolcond.eps}
546: \caption{Typical configuration of the system output from simulation of the 
547: model with the hop-rate chosen to display column condensation.}
548: \label{simfigccconf}
549: \end{center}
550: \end{figure}
551: 
552: 
553: \subsection{Further behaviour}
554: In the preceding sections the somewhat extreme cases of the absence of
555: site or column attraction in the presence of the other ($u^s$ or $u^c$
556: set equal to one) were considered.  This was to allow for a direct
557: comparison with theory.  Cases where some forms of non-trivial
558: attraction are present in both $u^s$ and $u^c$ are difficult to
559: compare directly with theory.  However, such systems do display
560: interesting behaviour and so in this section we present some numerical
561: data from simulations of such a system.
562: 
563: Simulations were run on a system with $175$ particles in each row and
564: the hop-rate given by: $u^c(n) = 1 + b^c/L$ for $n\leq L$, $u^c(n) = 1
565: + b^c/n$ for $n>L$ and $u^s(n) = 1 + b^s/n$, with $b^c = 16$ and $b^s
566: = 2.5$.  If $u^c$ were set equal to one, then no condensation would
567: occur at this particle number \cite{EH05}.  If $u^s$ were set equal to
568: one, then no condensation would take place at
569: any finite particle number (see Section~\ref{theory}). Results from
570: simulations show that this system, with both site and column
571: attraction present in forms that separately show no condensation
572: behaviour, does show something that could be interpreted as a
573: condensation-like phenomenon. The following results were obtained from
574: a random initial configuration.
575: 
576: In Figure~\ref{simfigo} (a), it can be seen that the site occupation
577: distribution (black circles) has a decaying part and a peak at around
578: $n=150$ and is generally reminiscent of a condensed system.  The
579: column distribution (Figure~\ref{simfigo} (a) grey crosses) also shows
580: a large $n$ peak, at around $n=3500$, although it is much broader than
581: one usually associated with a clear condensation.  The typical
582: configuration data in Figure~\ref{simfigoconf} indicate that the high
583: $n$ bump of the
584: column distribution represents more than one highly occupied
585: column, each composed of many site condensates.  Without a direct
586: comparison with theory available, it is difficult to
587: interpret this broad peak: it could be a finite-size effect that will
588: not be seen in larger systems, but on the other hand it could be a
589: true condensation that is being adversely affected by the finite size
590: of the system.  In either case it is interesting from a network point
591: of view as real-world networks are often of finite size.
592: 
593: The degree distributions of this system are also interesting.  They both 
594: have a binomial form for low degree, but the in-degree distribution departs 
595: from this at high degree --- it has a small secondary peak implying that 
596: several sites are highly connected, but a completely connected site to 
597: which all others point, as arises in the column condensation case, is absent.
598: Thus in the network context we have a network with connectivity much 
599: like that of a random graph, except for the existence of several 
600: well-connected `hub' nodes which would not be present in the random case. 
601: These hub nodes also have many high weight links pointing to them, 
602: giving them a large in-strength.
603: 
604: For an initial configuration with all particles located on a single
605: site, we obtain distributions of the same form as above and configurations
606: typically contain more than one highly occupied column.
607: 
608: \begin{figure}%[htb]
609: \begin{center}
610: \includegraphics[width=12cm]{other_sc_idod.eps}
611: \caption{Distributions from simulation of the model with 
612: the hop-rate chosen such that without the site or 
613: column attraction elements no condensation would be 
614: present.
615: (a) Site and column occupation distributions, corresponding 
616: to the distributions of weight among links and in-strength 
617: among nodes in the network context respectively. 
618: (b) In- and out-degree distributions.}
619: \label{simfigo}
620: \end{center}
621: \end{figure}
622: 
623: \begin{figure}%[htb]
624: \begin{center}
625: \includegraphics[width=12cm]{configother.eps}
626: \caption{Typical configuration of the system output from simulation of the 
627: model with the hop-rate chosen such that without the site 
628: or column attraction elements no condensation would be present.}
629: \label{simfigoconf}
630: \end{center}
631: \end{figure}
632: 
633: \section{Theory}\label{theory}
634: In this section we present the exact steady state of the model and
635: exploit this solution in order to understand theoretically the
636: condensation transitions observed in simulations. 
637: 
638: \subsection{Steady state}
639: 
640: First, we give the steady state, the proof of which is given in an
641: appendix. The steady state probabilities
642: $P(\{\underline{n}_{l}\})$, for finding the system in a configuration
643: $\{\underline{n}_{l}\}$, are given by the simple form
644: \begin{equation} \label{P(C)}
645: P(\{\underline{n}_{l}\}) = Z_{L,M}^{-1} \prod_{l=1}^L f(\underline{n}_l)\;,
646: \end{equation}
647: i.e.\ a product over columns. Here, the functions $f(\underline{n}_l)$
648: depend on a single column (i.e.\ the in-links of node $l$), and for
649: the rates (\ref{rate}), they are given by
650: \begin{equation} \label{f(n)}
651: f(\underline{n}_l) = f^c(X_l) \prod_{k=1}^L f^s(n_{kl}) \;,
652: \end{equation}
653: where we are using $X_l = \sum_{k=1}^L n_{kl}$ to represent the total
654: number of particles in a column (the node in-strength), and where
655: \begin{equation} \label{f^a(x)}
656: f^a(x) = \prod_{i=1}^x u^a(i)^{-1}\;,
657: \end{equation}
658: for $a=s,c$, where $f^a(0)=1$.
659: The normalisation $Z_{L,M}$ in (\ref{P(C)}) is given by
660: \begin{equation} \label{Z_{L,M}}
661: Z_{L,M} = \sum_{\{n_{kl}\}} \,\prod_{l=1}^L \left[ f^c(X_l) \prod_{k=1}^L
662: f^s(n_{kl})  \right] \,\prod_{k=1}^L\delta(\sum_{l=1}^L
663: n_{kl} - M)\;. 
664: \end{equation}
665: The $\delta$-function in this normalisation enforces particle
666: conservation within each row (which is the conservation of node
667: out-strength in the network): it is this $\delta$-function that induces
668: correlations between different columns in the steady state.
669:  
670: \subsection{Condensation theory}
671: Now, we exploit the exact steady state in order to understand
672: theoretically the condensation transitions observed in simulations.
673: Ideally, one would wish to demonstrate condensation in the site- and
674: column-distributions of particles, $p^s(n)$ and $p^c(n)$. The
675: expressions for these distributions involve the normalisation
676: $Z_{L,M}$ given by (\ref{Z_{L,M}}), which can be thought of as a
677: canonical partition function since $M$ is fixed. However, as we now
678: outline, it turns out to be simplest to work within the grand
679: canonical ensemble in which the particle number is allowed to
680: fluctuate.
681: 
682: \subsubsection{Grand canonical ensemble}
683: We introduce the grand canonical partition function in the usual way,
684: by defining fugacities $\{z_k\}$ so that we can replace the canonical
685: partition function (\ref{Z_{L,M}}) by 
686: \begin{equation} \label{GCZ1}
687: \mathcal{Z}_L = \sum_{\{n_{kl}=0\}}^\infty \prod_{k=1}^L
688: z_k^{\sum_l n_{kl}} \prod_{l=1}^L\,\left[
689:  f^c(X_l) \prod_{k=1}^L f^s(n_{kl}) \right] \;,
690: \end{equation} 
691: where the fugacities are chosen to ensure that, on average, each row
692: contains the proper number of particles $M_k$:
693: \begin{equation} \label{fugacity}
694: z_k \frac{\partial \, {\rm ln} \mathcal{Z}_L}{\partial z_k} =
695: \overline{\sum_l n_{kl}} = M_k\;.
696: \end{equation}
697: Here, the overbar indicates an average taken in the grand canonical
698: ensemble. Since we are interested in the case where $M_k = M$ for all
699: values of $k$, the lhs of this equation must be independent of $k$
700: therefore $z_k = z$ for all values of $k$; in this case, the fugacity $z$
701: is chosen to determine the {\it total} number of particles in the
702: system. After a little rearrangement, (\ref{GCZ1}) can be written
703: \begin{eqnarray} \label{GCZ2}
704: \mathcal{Z}_L &=&
705:  \prod_{l=1}^L\,\left[ \left( \prod_{k=1}^L
706:  \sum_{n_{kl}} \right) f^c(X_l) \prod_{k=1}^L
707:  f^s(n_{kl})\, z^{n_{kl}} \right] \\
708: &=&
709:  \left[ 
710:  \sum_{\{ n_{k}\}}  f^c(X) \prod_{k=1}^L
711:  f^s(n_{k})\, z^{n_{k}} \right]^L \;,
712: \end{eqnarray} 
713: where in going from the first to the second line
714: we use the fact
715: that the $l$ subscript on the $n_{kl}$
716: plays no role --- each column makes the same contribution to
717: $\mathcal{Z}_L$ --- and the square bracket can simply be raised to the
718: power of $L$. Now, $X = \sum_k n_k$. By taking
719: derivatives of $\mathcal{Z}_L$ with respect to $z$, one finds that the
720: fugacity must be chosen to satisfy
721: \begin{equation} \label{saddle}
722: M = z \frac{\partial\,{\rm ln} F(z)}{\partial z}\;,
723: \end{equation}
724: where we have defined the function
725: \begin{equation} \label{F(z)}
726: F(z) = \sum_{\{n_k\}} f^c(X) \prod_{k=1}^L
727:  f^s(n_{k})\, z^{n_{k}} \;.
728: \end{equation}
729: 
730: The equation (\ref{saddle}) is the key result of this
731: subsection. Condensation typically arises when one is unable to
732: satisfy (\ref{saddle}) above some critical density; under these
733: circumstances, the grand canonical ensemble is no longer valid and the
734: change of the partition function from the grand canonical form below
735: the critical density to some other form above it signals the phase
736: transition. We now discuss conditions under which (\ref{saddle}) can
737: be satisfied in order to illustrate how one can theoretically
738: understand and predict the condensed phases observed in simulations.
739: 
740: 
741:  
742: 
743: 
744: 
745: \subsubsection{Condensation} \label{Condensation}
746: It is straightforward to show that the function $F(z)$ is a smoothly
747: increasing function of its argument $z$. Let us assume that we have
748: chosen hop rates such that the sums which determine $F(z)$ have a
749: radius of convergence $\alpha$. The convergence properties of $F(z)$
750: and its derivative determine whether or not the system undergoes
751: condensation. These convergence properties are determined by the form
752: of the function $f(\underline n)$ which in turn provides the conditions
753: on the rewiring rates under which the system condenses. We now present
754: two very simple examples to illustrate the two types of condensation.
755: 
756: First, we consider conditions to observe site condensation. We
757: take $f^c(x) = 1$, which is the case when $u^c(x) = 1$ for all $x$, hence
758: \begin{equation}
759: F(z) = \left[ \sum_{n=0}^\infty f^s(n) z^n \right]^L\;.
760: \end{equation}
761: Hence, from (\ref{saddle}), the fugacity is determined by
762: \begin{equation}
763: \rho = \frac{\sum_{n=0}^\infty n f^s(n) z^n}{\sum_{n=0}^\infty f^s(n)
764:   z^n}\;,
765: \end{equation}
766: where $\rho= M/L$ is the density of particles in each row. When both
767: the numerator and denominator approach finite values as $z$ approaches
768: the radius of convergence $\alpha$ there exists a finite critical
769: density $\rho_c$ above which (\ref{saddle}) cannot be satisfied. This
770: signals condensation: for $\rho > \rho_c$, each row contains $(\rho -
771: \rho_c)L$ `excess' particles which condense in a single, randomly
772: located site within each row. In order that $\rho_c$ is finite, one
773: requires that $f^s(n)$ decays asymptotically like $f^s(n) \sim
774: \alpha^{-n} n^{-b}$ for large $n$ where $b>2$. This in turn implies
775: that the hop rates $u^s(n)$ must decay asymptotically as
776: \begin{equation} \label{u^s(n)}
777: u^s(n) \sim \alpha(1+b/n) \;,
778: \end{equation}
779: again, where $b>2$, if one is to observe site condensation in each row
780: at finite density. In the network model, this implies that the rate of
781: rewiring a node must decay more slowly than $2/n$, where $n$ is the
782: weight of the link pointing out of the node being rewired. These
783: considerations guide the choice of hop rates (\ref{site rate})
784: discussed in Section \ref{site condensation}.
785: 
786: Now, we consider conditions to observe column condensation. To deal
787: with the sum over particle configurations in $F(z)$, it turns out to
788: be convenient to sum over the variable $X$ and introduce a
789: $\delta$-function to ensure that $X$ represents the number of
790: particles in a column. Thus $F(z)$ may be written in the form
791: \begin{equation}
792: F(z) = \sum_{\{n_{k}\}} \sum_{X=0}^\infty f^c(X) \,
793:  \delta(\sum_{k=1}^L n_{k} - X) \prod_{k=1}^L f^s(n_{k})
794:  z^{n_{k}}\;,
795: \end{equation} 
796: where the $k$ subscripts refer to the rows within a single column. To
797: proceed, we take $f^s(n) = 1$, which is the case when $u^s(n) =
798: 1$ for all $n$. With this choice, the sum over $\{n_k\}$
799: is straightforward to perform --- the $\delta$-function constraint
800: supplies a combinatorial factor associated with the number of ways of
801: adding $L$ integers to get $X$ --- and one finds
802: \begin{equation}
803: F(z) = \sum_X {L{-}1{+}X \choose L{-}1}\,f^c(X)z^X\;.
804: \label{F(z)X}
805: \end{equation}
806: In this case, the rhs of (\ref{saddle}) represents the average number
807: of particles in a column, $\overline X$. The asymptotics of $f^c(X)$
808: determine the convergence properties of $F(z)$ (\ref{F(z)X}). 
809: In order to deduce the
810: required asymptotics, we expand the binomial factor for $X \gg L$,
811: using the approximation 
812: \begin{equation}
813: {L{-}1{+}X \choose L{-}1} \approx \frac{X^{L-1}}{(L-1)!}\,,
814: \end{equation}
815: for large $X$. To observe column condensation, the rhs of
816: (\ref{saddle}) must converge to a value $X_c=\mathcal{O}(L)$ as $z \to
817: \alpha$. This is the case when $f^c(X)$ decays asymptotically like
818: $f^s(n) \sim \alpha^{-X} X^{-\lambda}$ with $\lambda > 1+L$. When
819: these conditions are satisfied, the system condenses above a critical
820: density $\rho_c = X_c/L$ in such a way that every site in a single,
821: randomly located column typically contains $(X-X_c)$ particles. A
822: hop rate which yields $f^c(X)$ with the required asymptotics is
823: \begin{equation} \label{u^c(n)}
824: u^c(X) \sim \alpha(1+c L/X) \;,
825: \end{equation}
826: where $c>1$, which motivates the choice of hop rates (\ref{col rate})
827: discussed in Section \ref{column condensation}.
828: 
829: Thus these two simple examples, $f^c(x)=1$ and $f^s(n)=1$, illustrate
830: site and column condensation respectively.  We would like to stress
831: that in both site and column condensation, while only the asymptotics
832: determine whether or not the system condenses, the actual value of the
833: critical density depends on the details of the form of the rewiring
834: rates for all values of their arguments, not just the asymptotics.
835: 
836: We now indicate how one can use the grand canonical ensemble to
837: predict the background particle distribution in the condensed
838: phase. We consider the case $f^s(n)=1$, for which $p^c(X)$, the
839: probability that a column contains exactly $X$ particles, is given by
840: \begin{equation} \label{p(X)}
841: p^c(X) = \frac{{L{-}1{+}X \choose L{-}1}\,f^c(X)z^X}{F(z)}\;.
842: \end{equation}
843: This equation holds throughout the low density phase. In the condensed
844: phase, the grand canonical ensemble breaks down, however (\ref{p(X)})
845: with $z=\alpha$ correctly reproduces the form of the background
846: distribution of column occupation numbers. For the rates (\ref{col
847: rate}), (\ref{p(X)}) cannot be written in a convenient form at $z=1$,
848: however it can easily be evaluated numerically for a given value of
849: $b^c$. The result of such a computation, with $b^c = 1.05$, is
850: compared with simulation in Figure~\ref{simfigcc}~(a).
851: 
852: 
853: \section{Generalisations}
854:  \label{generalisations}
855: 
856: In this section we consider some generalisations of the network model
857: which, in the steady state, are still given by a factorised form. The
858: aim is to illustrate that the class of network models with factorised
859: steady states extends to a wide variety of rewiring dynamics and so,
860: by making the connection between network dynamics and interacting
861: particle systems, we are rewarded with a versatile approach to the
862: analysis of network properties.
863: 
864: \subsection{General dependence of the rewiring rates on departure and
865:   destination nodes} 
866: 
867: We begin by considering three generalisations of the rewiring rates: 
868: \begin{enumerate}
869: \item General dependence of the rewiring rates on
870: the weights of links rather than the specialised form (\ref{rate}).
871: 
872: \item Rewiring rates which depend on the
873: weights of the links pointing into both the node from which the link
874: is being removed, and the weights of the links pointing into the 
875: node into which the link is being rewired --- this allows one to
876: consider preferential attachment.
877: 
878: \item Heterogeneity in the rewiring dynamics such that one can
879:   consider rates which differ depending on the source, departure and
880:   destination nodes of the link.
881: 
882: \end{enumerate}
883: With these generalisations, one unit of the weight associated with the
884: link pointing from node $k$ into node $l$ is rewired to point into
885: another randomly selected node $l'$ with a rate
886: $u_{kl}(\underline{n}_l) \, t_{kl'}(\underline{n}_{l'})$. The
887: heterogeneity described in the above point 3 enters through the $k$,
888: $l$, and $l'$ superscripts on $u$ and $t$. The corresponding
889: interacting particle system has been discussed in the literature in
890: the context of `Urn' models (discussed in \cite{GL02}) and
891: `Misanthrope' processes (discussed in \cite{EH05}) in a fully
892: connected geometry, in which particle hop rates depend on the numbers
893: of particles at both the site of departure and the destination site,
894: but now defined with $L$ species of particles.
895: 
896: In the steady state, this model can still be expressed in the
897: factorised form 
898: \begin{equation} \label{P(C)gen}
899: P(\{\underline{n}_{l}\}) = Z_{L,M}^{-1} \prod_{l=1}^L f_l(\underline{n}_l)\;,
900: \end{equation}
901: (though note $f_l(\underline{n}_l)$ now carries a sub-script $l$)
902: provided the hop rates satisfy the constraint
903: \begin{eqnarray} \label{constraint}
904:  \frac{u_{kl}(\underline{n}_{l}; n_{k'l}{-}1) \,
905:  t_{kl}(\underline{n}_{l}; n_{kl}{-}1)}
906: {u_{kl}(\underline{n}_{l}) \,
907:  t_{kl}(\underline{n}_{l}; n_{kl}{-}1, n_{k'l}{-}1)} = 
908:  \frac{u_{k'l}(\underline{n}_{l}; n_{kl}{-}1) \,
909:  t_{k'l}(\underline{n}_{l}; n_{k'l}{-}1)}
910: {u_{k'l}(\underline{n}_{l}) \,
911:  t_{k'l}(\underline{n}_{l}; n_{kl}{-}1, n_{k'l}{-}1)} \;,
912: \end{eqnarray}
913: for all pairs of rows $k$ and $k'$. We have introduced the notation
914: $\underline{n}_{l}; n_{kl}{-}1$ which represents the configuration
915: $\underline{n}_l$ but with the term $n_{kl}$ replaced by
916: $n_{kl}{-}1$. When the hop rates satisfy this constraint, the
917: functions $f_l(\underline{n}_l)$ can be written
918: \begin{equation} \label{f(n)gen}
919: f_l(\underline{n}_l) = \prod_{k=1}^L \left[ \prod_{i=1}^{n_{kl}}
920:   \frac{t_{kl}(0,\ldots,0,i{-}1,n_{k+1\,l},\ldots,n_{Ll})}
921: {u_{kl}(0,\ldots,0,i,n_{k+1\,l},\ldots,n_{Ll})} \right]\;,
922: \end{equation}
923: having chosen $f(0,\ldots,0) = 1$. This can be written in a number of
924: different forms due to (\ref{constraint}) --- the symmetry in
925: $f_l(\underline{n}_l)$ is obscured within this constraint. A proof of
926: this steady state is presented in Appendix B.
927: 
928: \subsubsection{Preferential attachment}
929: To illustrate how one might exploit this steady state, we consider a
930: network model with preferential attachment rewiring dynamics. Thus we
931: choose  $u_{kl}(\underline{n}_{l}) = 1$ for all $k$ and $l$, and 
932: \begin{equation}
933:  t_{kl'}(\underline{n}_{l'}) = t^s(n_{kl'}) t^c (\sum_{k=1}^L n_{kl'})\;,
934: \end{equation}
935: which defines the rate a link pointing from node $k$ into node $l$ is
936: rewired to point into node $l'$. This rate therefore is a function of
937: the weight of the link pointing from the source node $k$ into the
938: destination node $l'$ before the rewiring event, and a function of the
939: total in-strength of the destination node. It satisfies the
940: constraint (\ref{constraint}) and one finds that the functions
941: $f(\underline{n}_l)$ are given by (\ref{f(n)}), where
942: \begin{equation}
943: f^a(x) = \prod_{i=1}^x t^a(i-1)\;,
944: \end{equation}
945: for $a=s,c$, where $f^a(0)=1$. In order that this model will exhibit
946: the behaviour observed in simulations, one can follow the analysis of
947: Section \ref{Condensation} to deduce that if the hop rates decay
948: asymptotically like
949: \begin{equation}
950: t^s(n) \sim 1-b/n\;,
951: \end{equation}
952: with $b>2$, then one observes site condensation above some finite
953: critical system density, and if
954: \begin{eqnarray}
955: t^c(X) \sim 1-\lambda/X \;,
956: \end{eqnarray}
957: for $X \gg L$, with $\lambda>L+1$, then one observes row condensation
958: above some finite critical system density.
959: 
960: 
961: \subsection{Further generalisations}
962: There are a number of further ways one can choose to redefine the
963: model and still obtain a factorised steady state. One way is to
964: consider geometries other than the fully connected one, such as
965: rewiring dynamics in which the weight being rewired from node $l$ is
966: always rewired to a `neighbouring' node $l+1$. All the models
967: considered so far also have factorised steady states in this geometry
968: although there may exist extra constraints in certain
969: cases. Generalisations to more complicated geometries are also
970: possible, and have been discussed in the context of the ZRP in
971: \cite{Evans00}.
972: 
973: 
974: We can also relax the property of integer weights and consider the
975: case where the weights are continuous variables and an arbitrary
976: amount of the weight is allowed to be rewired.  The approach is
977: similar to the way one generalises the zero-range process to
978: continuous masses \cite{EMZ}.
979: 
980: 
981: It is straightforward to allow each node to have a different total
982: out-strength, but more complicated generalisations whereby these
983: out-strengths are not conserved may also be possible. Non-conservation
984: in the context of the ZRP has been discussed in \cite{AELM05, EH05}. Another
985: straightforward modification of the model is to prohibit links which
986: point into the same node they point out of, i.e., restrict to
987: $n_{kk}=0$.  A further possibility is to allow negative weights, as
988: well as positive, which corresponds to a generalisation of the
989: so-called `Bricklayers' model' \cite{B03} --- positive and negative
990: weights have been considered in a rewiring social network in
991: \cite{HJDXW05}.
992: 
993: A point about the factorised steady states we would like to emphasise
994: is that one is free to choose any form for $f(\underline{n}_l)$ and
995: then infer the rewiring rates from a recursion such as
996: (\ref{recursion}). 
997: 
998: \section{Conclusion} \label{conclusion}
999: 
1000: In summary we have introduced a dynamical model of a weighted, directed
1001: network wherein the out-strength of each node is conserved but the
1002: in-strength is not.  The rewiring dynamics of this model can be mapped
1003: onto a multi-species zero-range process for which the steady state is
1004: exactly solvable.
1005: 
1006: Through numerical simulations and theoretical analysis we have
1007: identified two condensation transitions in the steady state.  In the
1008: network context they correspond to a disparity transition where for
1009: each node a single link contains a finite fraction of the
1010: out-strength, and the more familiar condensation transition where the
1011: in-strength of a single node captures a finite fraction of the
1012: out-strengths for all nodes in the system.  These transitions
1013: demonstrate some of the varied behaviour that is possible in weighted
1014: directed networks.
1015: 
1016: Within the multi-species ZRP picture these transitions correspond to,
1017: in the first case, condensation of all species at independent sites
1018: or, in the second case, a collective condensation of all species onto
1019: the same site.
1020: From our knowledge of the multi-species ZRP we are able to identify
1021: a number of generalisations of the model and dynamics which preserve the 
1022: exactly solvable steady state. For example
1023: these correspond in the network context to continuous weights,
1024: preferential attachment dynamics and node fitness.
1025: It would be interesting to explore further  the 
1026: different possible behaviours afforded by these and other
1027: generalisations.
1028: 
1029: 
1030: \subsection*{Acknowledgements}
1031: This work was supported by the Scottish Universities Physics Alliance
1032: (SUPA) and EPSRC programme grant GR/S10377/01.  AGA thanks the
1033: Carnegie Trust for the Universities of Scotland for financial support.
1034: 
1035: 
1036: 
1037: 
1038: \appendix
1039: 
1040: \section*{Appendices}
1041: 
1042: \setcounter{equation}{0}
1043: \def\theequation{A.\arabic{equation}}
1044: \subsection*{Appendix A. Proof of steady state (\ref{P(C)}) to
1045:   (\ref{f^a(x)})}
1046: 
1047: The steady state (\ref{P(C)}-\ref{f^a(x)}) can be demonstrated by
1048: noting that, since the network model is defined in a fully connected
1049: geometry (i.e.\ a link can be rewired from a node to any other node in
1050: the network), the steady state probabilities must satisfy detailed
1051: balance with respect to the rewiring dynamics. (In Section
1052: \ref{generalisations} we comment on how one generalises to other
1053: geometries.)  The detailed balance condition implies that
1054: \begin{eqnarray} \label{balance}
1055: u^s(n_{kl}{+}1)u^c(X_l{+}1) P(\{\underline{n}_l\};n_{kl}{+}1,n_{km}{-}1)  =
1056: %\nonumber \\ 
1057: u^s(n_{km})u^c(X_m)  P(\{\underline{n}_l\}) \;,
1058: \end{eqnarray}
1059: where we have introduced the notation
1060: $\{\underline{n}_l\};n_{kl}{+}1,n_{km}{-}1$ to represent a particle
1061: configuration with $n_{kl}$ replaced by $n_{kl}{+}1$ and $n_{km}$
1062: replaced by $n_{km}{-}1$. In this balance equation, the lhs represents
1063: a particle hop from site $kl$ to site $km$ and the rhs represents the
1064: reverse process. We now regard a solution of the form
1065: (\ref{P(C)},\ref{f(n)}) as an ansatz, substitution of which into the
1066: balance equation (\ref{balance})  yields
1067: \begin{eqnarray}
1068: u^s(n_{kl}{+}1) \frac{f^s(n_{kl}{+}1)}{f^s(n_{kl})} u^c(X_l{+}1)
1069:   \frac{f^c(X_l{+}1)}{f^c(X_l)} = 
1070: %\\ \nonumber 
1071: u^s(n_{km}) \frac{f^s(n_{km})}{f^s(n_{km}{-}1)}
1072:   u^c(X_m) 
1073:   \frac{f^c(X_m)}{f^c(X_m{-}1)}\;,
1074: \end{eqnarray}
1075: after cancelling common factors and rearranging slightly. This is
1076: satisfied for 
1077: \begin{equation}
1078: u^a(x{+}1)f^a(x{+}1)=f^a(x)\;,
1079: \end{equation} 
1080: for $a=s$, $c$, which is iterated to give (\ref{f^a(x)}). This proves
1081: the steady state (\ref{P(C)}) to (\ref{f^a(x)}).
1082: 
1083: \setcounter{equation}{0}
1084: \def\theequation{B.\arabic{equation}}
1085: \subsection*{Appendix B. Proof of steady state (\ref{P(C)gen}) to
1086:   (\ref{f(n)gen})}   
1087: 
1088: The steady state (\ref{P(C)gen}) to
1089:   (\ref{f(n)gen}) can be demonstrated by asking that the steady state
1090: probabilities satisfy detailed balance with respect to the rewiring
1091: dynamics, hence
1092: \begin{eqnarray}
1093: u_{kl}(\underline{n}_l; n_{kl}{+}1) \,
1094: t_{kl'}(\underline{n}_{l'}; n_{kl'}{-}1)
1095: P(\{\underline{n}_l\}; n_{kl}{+}1, n_{kl'}{-}1) =
1096: %\nonumber \\ 
1097: u_{kl'}(\underline{n}_{l'})\,
1098: t_{kl}(\underline{n}_l)  P(\{\underline{n}_l\}) \;, 
1099: \end{eqnarray}
1100: for each $k=1,\ldots,L$ and all pairs $l\neq l'$. Inserting the steady
1101: state (\ref{P(C)gen}), cancelling common factors and rearranging
1102: yields
1103: \begin{eqnarray}
1104: \frac{u_{kl}(\underline{n}_l; n_{kl}{+}1)}{t_{kl}(\underline{n}_l)}
1105: \frac{f_l(\underline{n}_l; n_{kl}{+}1)}{f_l(\underline{n}_l)}  =
1106: \frac{u_{kl'}(\underline{n}_{l'})}{t_{kl'}(\underline{n}_{l'}; n_{kl'}{-}1)}
1107: \frac{f_{l'}(\underline{n}_{l'})}{f_{l'}(\underline{n}_{l'}; n_{kl'}{-}1)}\;,
1108: \end{eqnarray}
1109: again, for each $k=1,\ldots,L$. Both sides of this equation must be
1110: equal to a constant which we choose, without loss of generality, to be
1111: equal to one, whereby we obtain the recursion
1112: \begin{equation} \label{recursion}
1113: f_l(\underline{n}_l) =
1114: \frac{t_{kl}(\underline{n}_l; n_{kl}{-}1)}
1115: {u_{kl}(\underline{n}_l)}
1116: f_l(\underline{n}_l; n_{kl}{-}1)\;, 
1117: \end{equation}
1118: which can be iterated to obtain (\ref{f(n)gen}). This recursion also
1119: implies the constraint on the choice of the hop rates: one can obtain
1120: two expressions for $f_l(\underline{n}_l)$ in terms of
1121: $f_l(\underline{n}_l; n_{kl}{-}1, n_{k'l}{-}1)$; one comes by applying
1122: the recursion (\ref{recursion}) first to $k$, then to $k'$; the other
1123: comes by applying the recursion to $k'$ before $k$. Performing these
1124: operations yields the constraint (\ref{constraint}).
1125: 
1126: 
1127: 
1128: 
1129: \begin{thebibliography}{00} 
1130: 
1131: \bibitem{AB02}
1132: R.~Albert and A.-L.~Barab\'asi, Rev.~Mod.~Phys.\ {\bf 47}, 74 (2002)
1133: 
1134: \bibitem{DM03}
1135: S.~N.~Dorogovtsev and J.~F.~F.~Mendes 2003 
1136: {\it Evolution of Networks} (OUP, Oxford) 
1137: 
1138: \bibitem{BA99}
1139: A.-L.~Barab\'asi and R.~Albert,
1140: Science {\bf 286}, 509 (1999) 
1141: 
1142: \bibitem{Newman01a}
1143: M.~E.~J.~Newman,
1144: %    The structure of scientific collaboration networks
1145:     Proc.~Natl.~Acad.~Sci.~USA {\bf 98}, 404 (2001)
1146: 
1147: \bibitem{Newman04}
1148: M.~E.~J.~Newman, Phys.~Rev.~E {\bf 70}, 056131 (2004)
1149: 
1150: \bibitem{GA04}
1151: R.~Guimera and L.~A.~N.~Amaral, 
1152: Eur.~Phys.~J. {\bf 38}, 381 (2004)
1153: 
1154: \bibitem{BBPV05}
1155: M.~Barth\'elemy, A.~Barrat, R.~Pastor-Satorras and A.~Vespignani, 
1156: %    Characterization and modeling of weighted networks
1157:     Physica~A {\bf 346}, 34 (2005) 
1158: 
1159: \bibitem{NR02}
1160: J.~D.~Noh and H.~Rieger, 
1161: Phys.~Rev.~E {\bf 66}, 066127 (2002)
1162: 
1163: \bibitem{MAB04}
1164: P.~J.~Macdonald, E.~Almaas, A.-L.~Barab\'asi, 
1165: cond-mat/0405688
1166: %    Title: Minimum spanning trees on weighted scale-free networks
1167: 
1168: \bibitem{YJBT01}
1169: S.~H.~Yook, H.~Jeong, A.-L.~Barab\'asi and Y.~Tu,
1170: Phys.~Rev.~Lett.\ {\bf 86}, 5835 (2001)
1171: 
1172: \bibitem{AK05}
1173: T.~Antal and P.~L.~Krapivsky, 
1174: %    Weight-driven growing networks
1175:     Phys.~Rev.~E {\bf 71},  026103 (2005) 
1176: 
1177: \bibitem{AKR05} E.~Almaas, P.~L.~Krapivsky and S.~Redner, Phys. Rev. E
1178: {\bf 71}, 036124 (2005)
1179: 
1180: \bibitem{BBV04}
1181: A.~Barrat, M.~Barth\'elemy and A.~Vespignani, 
1182: %    Weighted evolving networks: Coupling topology and weight dynamics
1183:     Phys.~Rev.~Lett.\ {\bf 92}, 228701 (2004)
1184: 
1185: \bibitem{BBV04a}
1186: A.~Barrat, M.~Barth\'elemy and A.~Vespignani,
1187: %    Modeling the evolution of weighted networks
1188:     Phys.~Rev.~E {\bf 70}, 066149 (2004)
1189: 
1190: \bibitem{BBV04b}
1191: A.~Barrat, M.~Barth\'elemy and A.~Vespignani,
1192: %    Modeling the evolution of directed weighted networks
1193:     Lect. Notes Comput. Sci. {\bf 56}, 3243 (2004)
1194: 
1195: \bibitem{DM04}
1196: S.~N.~Dorogotsev and J.~F.~F.~Mendes,
1197: cond-mat/0408343
1198: 
1199: %Statistical ensemble of scale-free random graphs
1200: \bibitem{BCK01}
1201: Z.~Burda, J.~D.~Correia and A.~Krzywicki, 
1202: Phys.~Rev.~E {\bf 64}, 046118 (2001) 
1203: 
1204: % Principles of statistical mechanics of random networks
1205: \bibitem{DMS03}
1206: S.~N.~Dorogovtsev, J.~F.~F.~Mendes and A.~N.~Samukhin, 
1207: Nucl.~Phys.~B {\bf 666}, 396 (2003)
1208: 
1209: \bibitem{AELM05}
1210: A.~G.~Angel, M.~R.~Evans, E.~Levine and D.~Mukamel,
1211: cond-mat/0503487
1212: 
1213: \bibitem{EH05}
1214: M.~R.~Evans and T.~Hanney, J.~Phys.~A {\bf 38}, R195 (2005) 
1215: 
1216: 
1217: \bibitem{KRL00}
1218: P.~L.~Krapivsky, S.~Redner and F.~Leyvraz,
1219: Phys.~Rev.~Lett.\ {\bf 85}, 4629 (2000)
1220: 
1221: 
1222: \bibitem{KR01}
1223: %Organization of growing random networks
1224: P.~L.~Krapivsky and S.~Redner, 
1225: Phys.~Rev.~E {\bf 63}, 066123 (2001)  
1226: 
1227: 
1228: \bibitem{BB01}
1229: %	Bose-Einstein Condensation in Complex Networks
1230: G. Bianconi and A.-L. Barab\'asi, 
1231: Phys. Rev. Lett. {\bf 86}, 5632 (2001)
1232: 
1233: 
1234: \bibitem{Evans00}
1235: M.~R.~Evans, Braz.~J.~Phys.\ {\bf 30}, 42 (2000)                           
1236: 
1237: \bibitem{HJDXW05}
1238: B.~Hu, X.-Y.~Jiang, J.-F.~Ding, Y.-B.~Xie and B.-H.~Wang,
1239: cond-mat/0408125
1240: 
1241: \bibitem{EH03}
1242: M.~R.~Evans and T.~Hanney, J.~Phys.~A {\bf 36}, L441 (2003)
1243: 
1244: \bibitem{HE04}
1245: T.~Hanney and M.~R.~Evans, Phys.~Rev.~E {\bf 69}, 016107 (2004)
1246: 
1247: \bibitem{BGG03}
1248: M.~Barth\'elemy, B.~Gondran and E.~Guichard, 
1249:  Physica A {\bf 319}, 633 (2003)
1250: 
1251: \bibitem{B05}
1252: G.~Bianconi, cond-mat/0412399
1253: 
1254: \bibitem{MBCV05}
1255: A. De Montis, M. Barth\'elemy, A. Chessa and A. Vespignani,
1256: cond-mat/0507106
1257: 
1258: \bibitem{GL02} 
1259: C. Godr\`eche and J. M. Luck, J. Phys.:Condens. Matter {\bf 14}, 1601
1260: (2002) 
1261: 
1262: \bibitem{EMZ}  
1263: M. R. Evans, S. N. Majumdar and R. K. P. Zia 
1264: J. Phys. A: Math. Gen. {\bf 37}, L275 (2004)
1265: 
1266: \bibitem{B03} 
1267: M. Bal\'azs, Annales de l'Institut Henri Poincar\'e - PR {\bf 39}, 639
1268: (2003)
1269:  
1270: \end{thebibliography}
1271: 
1272: \end{document}
1273: 
1274: 
1275: 
1276: