1: \documentclass[twocolumn,aps,floatfix,showpacs]{revtex4}
2:
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5: \usepackage{graphics}
6:
7: \begin{document}
8:
9: \newcommand\UBA{{\sc uba}}
10: \newcommand\RBN{{\sc rbn}}
11: \newcommand\EP{{\sc ep}}
12: \newcommand\SP{{\sc sp}}
13: \newcommand\trm\textrm
14: \newcommand\mbf\mathbf
15: \newcommand\s{\mbf{s}}
16: \newcommand\kb{\mbf{k}}
17: \newcommand\Nt{\tilde N}
18: \newcommand\pt{\tilde p}
19: \newcommand\ut{\tilde u}
20: \newcommand\yt{\tilde y}
21: \newcommand\tti{\tilde t}
22: \newcommand\Nh{\hat N}
23: \newcommand\uh{\hat u}
24: \newcommand\rh{\hat r}
25: \newcommand\ti{\trm i}
26: \newcommand\tn{\trm n}
27: \newcommand\tu{\trm u}
28: \newcommand\tsa{{0*}}
29: \newcommand\tsia{{0,0*}}
30: \newcommand\msa{\trm\tsa}
31: \newcommand\msia{\trm\tsia}
32: \newcommand\Pex{P_{\trm{\sc cc}}}
33: \newcommand\PNex{P_{\trm{\sc cc}}^1}
34: \newcommand\Pn{P_{\tn,N}}
35: \newcommand\Pu{P_{\tu,N}}
36: \newcommand\Oc{{\cal O}}
37:
38: \title{Exhaustive percolation on random networks}
39:
40: \author{Bj\"orn~Samuelsson}
41: \email[]{bjorn.samuelsson@duke.edu}
42: \author{Joshua~E.~S.~Socolar}
43: \email[]{socolar@phy.duke.edu}
44: \affiliation{Physics Department and Center for Nonlinear and Complex
45: Systems, Duke University, Durham, NC, 27514}
46:
47: \date{\today}
48:
49: \begin{abstract}
50: We consider propagation models that describe the spreading of an
51: attribute, called ``damage'', through the nodes of a random network.
52: In some systems, the average fraction of nodes that remain undamaged
53: vanishes in the large system limit, a phenomenon we refer to as {\it
54: exhaustive percolation}. We derive scaling law exponents and
55: exact results for the distribution of the number of undamaged nodes,
56: valid for a broad class of random networks at the exhaustive
57: percolation transition and in the exhaustive percolation regime.
58: This class includes processes that determine the set of frozen nodes
59: in random Boolean networks with indegree distributions that decay
60: sufficiently rapidly with the number of inputs. Connections between
61: our calculational methods and previous studies of percolation
62: beginning from a single initial node are also pointed out. Central
63: to our approach is the observation that key aspects of damage
64: spreading on a random network are fully characterized by a single
65: function specifying the probability that a given node will be
66: damaged as a function of the fraction of damaged nodes. In addition
67: to our analytical investigations of random networks, we present a
68: numerical example of exhaustive percolation on a directed lattice.
69: \end{abstract}
70:
71: \pacs{89.75.Da, 02.50.Ey, 02.10.Ox, 05.50.+q}
72:
73: \maketitle
74:
75: \section{Introduction}
76: \subsection{Overview}
77: Propagation models on lattices or more general graphs describe the
78: spreading of some discrete signal through a set of discrete
79: entities. In the most general terms, the signal corresponds to some
80: qualitative change that causes the entity to interact differently with
81: its neighbors. Examples include the spreading of damage in power
82: grids \cite{Sachtjen:00, Kinney:05}, the spreading of disease through
83: a population \cite{Hethcote:00, Newman:02, Cohen:03}, the spreading of
84: a computer virus on the Internet \cite{Pastor-Satorras:01, Lloyd:01},
85: or the alteration of gene expression patterns in a cell due to a
86: mutation \cite{Hughes:00, Ramo:06}. In the general case, the
87: individual entities are represented as nodes in a graph where the
88: links indicate paths along which the signal can spread
89: \cite{Watts:99,Newman:03, Callaway:00, Moreno:02, Watts:02, Motter:04,
90: Crucitti:04}. Because the signal can be thought of as disrupting the
91: static or dynamical state of the original system, we refer to its
92: propagation as spreading {\it damage}, though in many cases the ``damage''
93: may enhance a desired property or simply represent some natural
94: dynamical process. A single instance of a given spreading process
95: initiated from a particular subset of nodes is often called an
96: avalanche.
97:
98: In analyzing spreading processes, one is often interested in the
99: transition between those that die out quickly and those that spread to
100: a finite fraction of the system in the large-system limit, a
101: transition that may occur as the probability of transmitting damage
102: across links is varied. This percolation transition is relevant for
103: systems in which the fraction of initially damaged nodes tends to
104: zero in the limit of infinite system size. The order parameter for
105: the transition is the average fraction of nodes damaged in a single
106: avalanche, which remains zero for small transmission probabilities and
107: continuously increases when the probability rises above a threshold
108: value. We will refer to this as the {\it sparse percolation} (\SP)
109: transition. The \SP\ transition occurs for spreading processes in
110: which the probability that a node becomes damaged is zero unless at
111: least one of its neighbors is damaged. (If this probability were
112: nonzero, a nonzero fraction of the nodes would always get damaged.)
113:
114: For a certain class of propagation models, there is another transition
115: of interest. When the fraction of initially affected nodes remains
116: fixed as the system size is increased, the fraction of nodes that
117: remain {\it undamaged} can undergo a transition from finite values to
118: zero at transmission probabilities above some threshold. We refer to
119: this as the {\it exhaustive percolation} (\EP) transition. The \EP\
120: transition occurs only for propagation models in which the probability
121: of a node remaining undamaged is zero when all of its neighbors are
122: damaged (all of its inputs in the case of a directed graph). We
123: assume also that there is a nonzero probability for a node to remain
124: undamaged if it has at least one undamaged input. There is then one
125: more condition for the \EP\ transition: the density of directed loops
126: of any specified size must vanish in the large system limit. For any
127: loop there is a finite probability that no member of the loop will be
128: damaged, since no member of the loop can have all of its inputs
129: damaged until one of the members becomes damaged through a
130: probabilistic event. Thus \EP\ is {\it not} observable on spatial
131: lattices of the type generally encountered in statistical mechanics.
132: \EP\ is observable, however, on directed lattices and on graphs in
133: which the nodes serving as inputs to a given node are selected at
134: random.
135:
136: In this paper we derive the probability distribution for the number of
137: undamaged nodes at the \EP\ transition on random graphs for a general
138: class of propagation models exhibiting what we call {\it unordered
139: binary avalanches} (\UBA). This is analogous to finding the
140: distribution of avalanche sizes at the usual percolation transition,
141: but here we are asking for the distribution of the number of nodes
142: {\it not} participating in the avalanche.
143:
144: As an application of our \EP\ results, we consider the problem of
145: identifying unfrozen nodes in a random Boolean network (\RBN). In a
146: \RBN, each node has a binary state that is updated according to a rule
147: that takes the values of some other nodes as inputs. The dynamics of
148: \RBN s has been investigated extensively; see, e.g.,
149: \cite{Kauffman:69, Derrida:86a, Bastolla:97, Socolar:03, Aldana:03a,
150: Aldana:03b, Samuelsson:03, Kauffman:04, Drossel:05b}. A \RBN\ can
151: have several dynamical attractors, but some nodes might have the same
152: value at all times on all attractors. Such nodes are called {\it
153: stable} and the set of stable nodes is important for the dynamics in
154: \RBN s \cite{Flyvbjerg:88b, Bastolla:98a, Bilke:01}.
155:
156: Almost all stable nodes in a broad class of \RBN s can be identified
157: through a dynamic process that was introduced by Flyvbjerg
158: \cite{Flyvbjerg:88b} and formalized to facilitate numeric simulations
159: by Bilke and Sjunnesson \cite{Bilke:01}. We call the stable nodes
160: that can be identified by this dynamic process {\it frozen} (and nodes
161: that are not frozen are called {\it unfrozen}). Provided that the
162: Boolean rule distribution is symmetric with respect to inversion of
163: any subset of inputs, the set of frozen nodes can be identified
164: through an \UBA\ in which frozen inputs cause new nodes to become
165: frozen (damaged). Most rule distributions that have been examined in
166: the literature exhibit this symmetry. The requirement is satisfied,
167: for example, for any model that assigns given probability $p$ for
168: obtaining a 1 in each entry of the truth table for each node.
169:
170: This paper is organized as follows. We first develop the notation and
171: basic definitions required for discussing \UBA s in general. In
172: Section~\ref{sec: basic defs}, we give an introduction to the \UBA\
173: formalism from the perspective of percolation processes. A more
174: formal description is given in Section~\ref{sec: introduction to EP},
175: followed by a numerical illustration of the basic concepts. In
176: Section~\ref{sec: random networks}, we present analytic derivations
177: for \UBA\ in random networks with emphasis on \EP\ and the \EP\
178: transition. We also present explicit results for the special case of
179: Erd\H{o}s--R\'{e}nyi networks with a natural choice for the avalanche
180: rules.
181:
182: In Section~\ref{sec: application} we show how to apply the \UBA\
183: formalism to obtain the statistics of frozen nodes in two-input \RBN
184: s. In the present context, this serves as an illustration of the
185: general theory, but this particular example was also the primary
186: motivation for studying \EP. The results on \RBN s are consistent
187: with those found by Kaufman, Mihaljev, and Drossel.~\cite{Kaufman:05}.
188: The main advantage of using the \EP\ formalism for this problem is
189: that it makes clear how the calculation can be extended to networks
190: with more than two inputs per node, including networks with an
191: in-degree distribution that (with a low probability) allows
192: arbitrarily large in-degrees.
193:
194: \subsection{Basic definitions}
195: \label{sec: basic defs}
196:
197: An unordered binary avalanche (\UBA ) is defined as a spreading process
198: with the following properties:
199:
200: \begin{description}
201:
202: \item[Binary states:] the state of each node can be characterized as a
203: binary variable $s$, with $s=0$ meaning {\it undamaged} and $s=1$
204: meaning {\it damaged};
205:
206: \item[Boolean rules:] the state of each node is determined by a
207: Boolean function of the states of its input nodes;
208:
209: \item[Order independence:] the probability of having a given set of
210: nodes damaged at the end of the process does not depend upon the
211: order in which nodes are chosen for updating.
212:
213: \end{description}
214:
215: Order independence refers to the dynamics of the spreading process or
216: a simulation of it. In such a simulation, one typically chooses a
217: site and updates it according to a rule depending on the states of
218: sites that provide inputs to it, repeating the process until a test of
219: every site yields no change in the state of the system. We are
220: interested in cases where the order in which sites are chosen for
221: possible updating has no bearing on the final state of the system.
222:
223: \UBA\ is a natural extension of site or bond percolation. To determine
224: the avalanche size distribution in site percolation, for example, one
225: identifies an initial subset of damaged sites and then tests neighbors
226: of damaged sites to see whether the damage spreads to them. After a
227: given site is tested for the first time, its value is permanently
228: fixed. The process is iterated until no new damaged sites are
229: generated. See, e.g., Ref.~\cite{Sahimi:03}. This method of
230: investigating site percolation is equivalent to assigning all sites a
231: value, then beginning with a damaged site and determining all of the
232: damaged sites in a connected cluster. Site percolation where each
233: site has the probability $p$ to be occupied can be recast as a \UBA\
234: system as follows. Let each site be associated with a rule that is an
235: {\sc or}-rule of all of its neighbors with probability $p$ and is a constant
236: 0 with probability $1-p$. Then the above described site percolation
237: is achieved by first selecting the rules and clamping the value of a
238: given site to 1, and then repeatedly updating the system according to
239: the Boolean rules. In this situation, the 1s in the final state mark
240: a site percolation cluster. A more practical way of simulating the
241: same \UBA\ is to determine probabilistically the Boolean rule at each
242: site only when that site is first encountered in the percolation
243: process and to update only those nodes where the rules have been
244: determined.
245:
246: To ensure order independence in \UBA, it is sufficient to require that
247: each Boolean function is non-decreasing, meaning that if one of the
248: inputs to the rule changes from 0 to 1, the output is not allowed to
249: change from 1 to 0. For non-decreasing Boolean functions, if a
250: specific node is eventually going to be assigned the value 1 during an
251: avalanche, updating other nodes to 1 first cannot change the outcome.
252:
253: We are particularly interested in \UBA s that are initiated by damage
254: at a set of nodes comprising a nonzero fraction of the total number of
255: nodes. Such a process would be relevant, for example, if the
256: probability that any given node is damaged at the start is independent
257: of the system size.
258:
259: To clarify both the distinction between \EP\ (exhaustive percolation)
260: and \SP\ (sparse percolation) and the similarities between them, we
261: describe a particular case of a propagation model that exhibits both
262: transitions. Consider a graph with a total of $N$ nodes, some of
263: which have three input links each while the others have no input links
264: at all. The graph is random in that the node supplying the input
265: value on any given link is selected at random, but stays fixed
266: throughout the avalanche. Let $\nu_0$ be the fraction of nodes with
267: no inputs. Define a spreading process as follows: The initial
268: condition is that all nodes with no inputs are considered damaged.
269: Each other node is now selected in turn to see whether the damage
270: spreads to it. If a node has one damaged input, the probability that
271: it will be damaged is $p_1$; if it has two damaged inputs, the
272: probability of damage is $p_2$ (with $p_2 \ge p_1$); and nodes with
273: three damaged inputs are guaranteed to become damaged ($p_3 = 1$).
274: These probabilities are realized, for example, by the following
275: Boolean rule distribution: a 3-input {\sc or}-rule with probability
276: $p_1$; a 3-input majority rule with probability $p_2-p_1$; and a
277: 3-input {\sc and}-rule with probability $1-p_2$.
278:
279: As $N$ goes to infinity, the number of initially damaged nodes can be
280: a nonzero number that grows slower than $N$, meaning that
281: $\nu_0$ goes to zero as $N$ goes to infinity. In this limit, the
282: \SP\ transition occurs at $p_1 = 1/3$ and the spreading from each
283: initially damaged node is described by a {\it Galton--Watson process}.
284: In a Galton--Watson process, a tree is created by adding branches to
285: existing nodes, with the number of branches emerging from each node
286: drawn from a fixed probability distribution. Such branching processes
287: have been investigated extensively. (See, e.g.,
288: Ref.~\cite{Harris:63}.) In particular, the correspondence to
289: Galton--Watson processes means that for critical \SP, the probability
290: of finding $n$ damaged nodes scales like $n^{-3/2}$ for $1\ll n\ll N$
291: \cite{Otter:49, Ramo:06}.
292:
293: For any nonzero value of $\nu_0$, the \EP\ transition occurs for $p_2$
294: satisfying $(1 - p_2)(1-\nu_0) = 1/3$ (assuming that this value of
295: $p_2$ is greater than $p_1$.) The analysis described in
296: Section~\ref{sec: random networks} provides a
297: method of calculating the probability $P(u)$ of having $u$ {\it
298: undamaged} nodes in this case. The result in the large $N$ limit is
299: $P(u) \sim P(0)u^{-1/2}$ for large $u$. A difference between \EP\ and
300: \SP\ is that both $P(0)$ and the cutoff on the $u^{-1/2}$ distribution
301: scale with $N$ for \EP, while for \SP\ only the cutoff scales with
302: $N$.
303:
304: \section{Introduction to exhaustive percolation}
305: \label{sec: introduction to EP}
306:
307: \subsection{Formal description of UBA}
308: \label{sec: formal UBA}
309:
310: We now describe a formalism and establish some notation that is
311: suitable for a detailed treatment of \UBA. Let $N$ denote the number
312: of nodes in a network with a specified set of links and let the nodes
313: be indexed by $j=1,\ldots,N$. The network state is described by the
314: vector $\s=\{s_1,\ldots,s_N\}$. Let $K_j$ denote the number of inputs
315: to node $j$, and let $\kb_j$ denote the vector of $K_j$ inputs to node
316: $j$. Furthermore, let $R$ denote a Boolean function and let
317: $\Pi_j(R)$ denote the initial probability that node $j$ has the rule
318: $R$. [It is required that $R$ has precisely $K_j$ inputs for
319: $\Pi_j(R)$ to be nonzero.]
320:
321: To efficiently simulate \UBA, we keep track of the information that is
322: known about each node at each step in the process. In
323: particular, it is important to keep track of whether or not the change
324: from 0 to 1 of a given input has already been accounted for in
325: determining the output. The simplest way to do this is to introduce
326: an extra state \tsa\ that labels a site whose rule $R$ implies an
327: output value of 1 but for which the update to 1 has not yet been
328: implemented. When a node changes its state from 0 to \tsa, it is a
329: silent change in the sense that the Boolean rules at the other nodes
330: treat an input \tsa\ exactly the same as 0. To retrieve the final
331: state of the network, all occurrences of \tsa\ must be updated to 1.
332: When a single update to 1 is made, the information that the given node
333: has value 1 is passed along to all nodes with inputs from it. The
334: values of these nodes may then change from 0 to \tsa. The conditional
335: probability that the value of node $i$ is updated from 0 to \tsa\ when
336: $j$ changes value from \tsa\ to 1, is given by
337: \begin{equation}
338: U_i(\s,j) \equiv \frac{P_1(\kb'_i) - P_1(\kb_i)}{1 - P_1(\kb_i)},
339: \label{eq: Ui}
340: \end{equation}
341: where $\kb'_i$ is the value of $\kb_i$ after $s_j$ has been updated and
342: $P_1(\kb_i)$ is the probability that $R_i(\kb_i) = 1$:
343: \begin{equation}
344: P_1(\kb_i) \equiv \sum_R R(\kb_i)\Pi_i(R).
345: \label{eq: P1}
346: \end{equation}
347: The numerator in Eq.~(\ref{eq: Ui}) is the probability that $R_i$
348: produces a 1 after the update of node $j$ minus the probability that
349: $R_i$ produced a 1 before the update. The denominator is the
350: probability that node $i$ had the value 0 before the update.
351:
352: Let $\Pi_i(1)$ denote the probability that the rule at node $i$ has
353: output 1 regardless of its input values. If some particular nodes are
354: selected for initiation of the \UBA, $\Pi_i(1)$ is set to one for
355: these nodes [which means $\Pi_i(R) = 0$ for all other rules].
356:
357: We are now ready to present a formal algorithm for determining the
358: final state of an instance of \UBA\ on a finite network. We carry out
359: the following procedure (where $:=$ denotes the assignment operator):
360: \begin{enumerate}
361: \item $s_j:=0$ for all $j$;
362: \item $s_j:=\msa$ with probability $\Pi_j(1)$ for each $j$;
363: \item Some $j$ with $s_j=\msa$ is selected; \label{j select}
364: \item $s_i:=\msa$ with probability $U_i(\s,j)$ for each $i$ with $s_i=0$;
365: \item $s_j:=1$; \label{j := 1}
366: \item Steps \ref{j select}--\ref{j := 1} are iterated as long as there
367: exists a node in state \tsa. \label{iterate}
368: \end{enumerate}
369:
370: \UBA\ can also be considered on infinite networks, but that requires a
371: more technical description of the process. First, the choices of $j$ in
372: step 3 for both descriptions must be such that any given $j$ that
373: satisfies the conditions in step~\ref{j select} will be selected in a
374: finite number
375: of iterations. Second, the ensemble of final states needs to be
376: defined in terms of a suitable limit process because the stopping
377: criterion in step~\ref{iterate} can not be applied to an infinite
378: system.
379:
380: Note that the dynamics is only dependent on the probability functions
381: $\{P_1(\kb_i)\}$. That is, the precise rule distributions affect the
382: avalanche results only through their contributions to $P_1$. Because
383: the Boolean rules are non-decreasing functions, $P_1(\kb_i)$ is also a
384: non-decreasing function. In fact, every non-decreasing function,
385: $f(\kb_i)$, with values in the interval $[0,1]$ can be realized by
386: $P_1(\kb_i)$ for a suitable Boolean rule distribution. One such rule
387: distribution can be constructed as follow: for each $i$ and each
388: $\kb_i$, select a random number $y$ from a uniform distribution on the
389: unit interval and set $R_i(\kb_i) = 1$ if and only if $y < f(\kb_i)$.
390:
391: \subsection{An example of EP on a lattice}
392: \label{sec: EP lattice}
393:
394: To illustrate the concepts of \UBA\ and \EP, consider a directed
395: network on a two-dimensional square lattice with periodic boundary
396: conditions. Each node in the lattice has integral coordinates $(i,j)$
397: where $i+j$ is odd and the node at $(i,j)$ receives inputs from the
398: two nodes at $(i-1,j\pm1)$. The rule for propagation of damage to a
399: node is either {\sc or} or {\sc and}, with probabilities
400: $\Pi_{(i,j)}(\trm{\sc or}) = r$ and $\Pi_{(i,j)}(\trm{\sc and}) = 1 -
401: r$, respectively.
402:
403: \begin{figure}[bt]
404: \begin{center}
405: \includegraphics{uba_2d.eps}
406: \end{center}
407: \caption{\label{fig: lattice UBA} An example of \UBA\ on a lattice,
408: displaying undamaged nodes (dots), initially damaged nodes (filled
409: circles), and nodes damaged during the avalanche (empty circles).
410: Each node has either an {\sc or}-rule or an {\sc and}-rule with
411: inputs from its neighbors in the row immediately above the node.
412: The probability for a node to be initially damaged is
413: $\rho = 1/8$ and the probability for obtaining an {\sc
414: or}-rule is $r=0.3$. Periodic boundary conditions are used and the
415: first row and column are repeated in gray after the last row and
416: column to illustrate the periodic boundary conditions. }
417: \end{figure}
418:
419: Figure~\ref{fig: lattice UBA} displays an avalanche
420: that is initiated by letting each node be initially damaged with
421: probability $\rho=1/8$. A node assigned {\sc or} becomes
422: damaged if either of its neighbors one layer above is damaged; a node
423: assigned {\sc and} becomes damaged if and only if both neighbors above
424: it are damaged. This means that
425: \begin{align}
426: P_1(\kb_{(i,j)}) &= \left\{
427: \begin{array}{ll}
428: 0 & \trm{if }\kb_{(i,j)} = (0,0)\\
429: r & \trm{if }\kb_{(i,j)} \in \{(0,1),(1,0)\}\\
430: 1 & \trm{if }\kb_{(i,j)} = (1,1)~.
431: \end{array}\right.
432: \end{align}
433: Note that clusters of damaged nodes formed in an avalanche initiated
434: by a single damaged node cannot contain any holes, as the uppermost
435: undamaged node in the hole would have to have two damaged inputs and
436: hence would become damaged when updated.
437:
438: For localized initial damage, the \SP\ threshold is found at $r =
439: 1/2$. Above this value of $r$, domains of damage tend to widen as the
440: avalanche proceeds. Since the growing cluster has no holes, this is
441: simultaneously an \EP\ transition. The \EP\ transition can be found
442: for smaller values of $r$ in lattices where each node is initially
443: damaged with a given nonzero probability $\rho$. [For every
444: initially damaged node, $\Pi_{(i,j)}(1)$ is set to 1, meaning that
445: $P_1(\kb_{(i,j)})=1$ for every value of $\kb_{(i,j)}$.]
446:
447: \begin{figure}[bt]
448: \begin{center}
449: \includegraphics{EP_2d.eps}
450: \end{center}
451: \caption{\label{fig: lattice EP} The average fraction of undamaged
452: nodes for \UBA\ on a lattice of the type shown in Fig.~\ref{fig:
453: lattice UBA}, as a function of the selection probability $p$ for
454: {\sc or}-rules and the probability $\rho=1/8$ for initial
455: damage. The lattice has periodic boundary conditions and covers a
456: square that has a side of $10$, $10^2$, $10^3$, and $10^4$ lattice
457: points, respectively, with steeper curves for larger systems. The
458: statistical uncertainty in the estimated mean is less than the line
459: width. }
460: \end{figure}
461:
462: Figure~\ref{fig: lattice EP} shows the average number of unaffected
463: nodes as a function of $r$ for $\rho=1/8$ on lattices with
464: periodic boundary conditions. The numerics displayed in
465: Fig.~\ref{fig: lattice EP} clearly suggest that there is a
466: second-order \EP\ phase transition. Furthermore, these numerical
467: results suggest that the avalanche in Fig.~\ref{fig: lattice UBA} is
468: within the parameter regime for \EP\ and that \EP\ does not occur in
469: this case due to finite size effects.
470:
471: For the case $r = 0$, it is possible to map the \EP\ transition onto
472: ordinary, directed, site percolation on the same lattice. When all
473: nodes in the lattice have {\sc and}-rules, the following algorithm may
474: be used to determine whether a given node will be damaged: select a
475: node; put a mark on the selected node unless it is initially damaged;
476: and recursively mark each initially undamaged node that has an output
477: to a marked node. The selected node will get damaged if and only if
478: this recursion ends in a finite number of steps. The algorithm
479: describes ordinary directed site percolation where the initially
480: undamaged (damaged) nodes are considered active (inactive) sites and
481: the process propagates in the opposite direction relative to the \UBA.
482: We therefore expect the \EP\ transition to occur for a value of
483: $\rho$ equal to $1-p_{\trm c}$, where $p_{\trm c} = 0.70549$ is the threshold for
484: directed site percolation \cite{Essam:88} and we have confirmed this
485: with numerical tests. Further study of \EP\ on the lattice is beyond
486: the scope of this paper.
487:
488: \subsection{Suppression of EP by resistant motifs}
489: \label{sec: resistant motifs}
490:
491: In the lattice example above, the fact that the network had no
492: feedback loops smaller than the lattice size was important. In
493: general, \EP\ is suppressed by the presence of short feedback loops.
494: As already noted, for \EP\ to occur, it is required that the output of
495: each rule in the rule distribution is 1 if all of its inputs have the
496: value 1. Otherwise, there would be a finite fraction of nodes that
497: keep the value 0 regardless of the influence from the rest of the
498: network. Generalization of this reasoning allows us to rule out \EP\
499: in other situations, indicating that \EP\ is most likely to occur in
500: directed or highly disordered networks. To pursue this idea, we
501: introduce the notion of {\it resistant motifs}.
502:
503: A {\it motif} is a small network with a particular arrangement of
504: internal links. A given motif may occur many times in a network with
505: different rules assigned to its nodes and with different
506: configurations of external inputs. A motif is {\it resistant} with
507: respect to a given ensemble of rule assignments if the probability of
508: damage entering the motif when all external inputs are damaged is
509: strictly less than unity. For the rule distributions that we consider
510: for the \EP\ transition in random networks, each node has a nonzero
511: probability of being assigned a rule that sets its output to 0 if at
512: least one of its inputs is 0. Thus when all of the nodes in a
513: feedback loop of any length have the value 0, there is a nonzero
514: probability that they will all remain 0 even if all external inputs to
515: the loop are set to 1. Every feedback loop of a given length is
516: therefore a resistant motif.
517:
518: If the number of occurrences of a resistant motif grows linearly with
519: the network size, there will in total be a finite fraction of nodes
520: that remain unaffected with a finite probability. For such networks,
521: \EP\ cannot occur in the limit of large systems. Examples include
522: typically studied regular lattices and small world networks with link
523: directions assigned so that short feedback loops are prevalent.
524:
525: The problem resistant motifs can be avoided in random networks having
526: a mean indegree $\langle K\rangle$ that is well-defined and
527: independent of $N$, in which case the number of feedback loops of a
528: given length approaches a constant. Though the total number of
529: resistant motifs may grow with system size, the larger motifs have a
530: low probability of avoiding damage. For large $N$, the out-degree
531: distribution is a Poisson distribution with a mean value of $\langle
532: K\rangle$. The outputs emerging from a given node form a tree with
533: approximately $\langle K\rangle^m$ nodes at the $m$th level. Thus,
534: the probability for a given node to be part of a cycle of $m$ nodes is
535: approximately $\langle K\rangle^m/N$, which means that the typical
536: number of feedback loops of length $m$ is approximately $\langle
537: K\rangle^m/m$. On the other hand, the loop may contain either
538: initially damaged nodes or some rules that allow damage to enter from
539: external inputs. The probability that this will {\em not} occur
540: decays exponentially with $m$. If the decay is faster than $\langle
541: K\rangle^{-m}$, the density of nodes in undamaged resistant motifs
542: will approach zero.
543:
544: In summary, \EP\ (for the considered type of rule distributions) is excluded on
545: lattices with a high density of feedback loops. For random networks,
546: however, the fraction of nodes in undamaged resistant motifs can go to
547: zero in the large $N$ limit. This property allows \EP\ to occur on
548: random networks as demonstrated in the following section.
549:
550: \section{EP on random networks}
551: \label{sec: random networks}
552: \subsection{Criteria for EP}
553: \label{sec: criteria for EP}
554: Consider a network such that the inputs to each node are chosen
555: randomly and uniformly from all nodes in the network and the
556: probability functions $\{P_1(\kb_i)\}$ are determined from a given
557: distribution of Boolean rules. For such networks, \UBA\ can be handled
558: analytically.
559:
560: Define $g(x)$ as the probability for a rule in the random network to
561: output $1$ if each input has the value $1$ with probability $x$. The
562: function $g$ reflects the probability for propagation of damage to a
563: single node, for the considered instance of \UBA. We refer to $g$ as
564: the {\it damage propagation function}. In random networks,
565: $P_1(\kb_i)$ is independent of $i$ and can be replaced by $P_1(\kb)$.
566: Let $K$ denote the number of components of $\kb$, i.e., the number of
567: inputs to the considered node. $g(x)$ can then be expressed as
568: \begin{align}
569: \label{eq: g definition}
570: g(x) &=
571: \sum_{K=0}^\infty P(K)\!\!\!\!\sum_{\kb\in\{0,1\}^{K}}x^{I}(1-x)^{K-I}P_1(\kb),
572: \end{align}
573: where $I$ is the number of 1s in $\kb$ and $P(K)$ is the probability to
574: draw a rule with $K$ inputs.
575:
576: Let $N$ denote the total number of nodes, and let $n_0$, $n_\msa$, and
577: $n_1$ denote the number of nodes with the values $0$, $\msa$, and $1$,
578: respectively. With these definitions and the fact that $U_i(\s,j)$ is
579: independent of $i$ for the random network, the role of
580: $\{P_1(\kb_i)\}$ is taken over by $g(n_1/N)$ and Eq.~\eqref{eq: Ui} is
581: replaced by
582: \begin{align}
583: U(\s,j) &= \frac{g(n'_1/N)-g(n_1/N)}{1-g(n_1/N)},
584: \label{eq: n-update first}
585: \end{align}
586: where $n'_1=n_1+1$. This means that the size of the network, the
587: number of initially damaged nodes, and the damage propagation function
588: $g$ taken together are sufficient to uniquely determine the stochastic
589: spreading process.
590:
591: After one pass of the update steps \ref{j select}--\ref{j := 1} (from
592: Section~\ref{sec: formal UBA}), the new values $n'_0$ and $n'_\msa$ of
593: $n_0$ and $n_\msa$ are given by
594: \begin{align}
595: n'_0 &= n_0 - \delta
596: \intertext{and}
597: n'_\msa &= n_\msa + \delta - 1
598: \intertext{where}
599: \delta &= B_{n_0}[U(\s,j)],
600: \label{eq: n-update last}
601: \end{align}
602: with $B_n(a)$ being a stochastic function that returns the number of
603: selected items among $n$ items if the selection probability for each
604: of them is $a$. The avalanche ends when $n_\msa=0$.
605:
606: The number of damaged nodes, $n$, in a complete avalanche is the final
607: value of $n_1$, whereas the number of undamaged nodes, $u$, is the
608: final value of $n_0$. An order parameter for the system is $\phi =
609: \lim_{N\rightarrow\infty}\langle n/N\rangle$, where the average is
610: taken over the ensemble of networks. The \SP\ transition is found when
611: $\phi$ changes from zero to a nonzero value, whereas the \EP\ transition
612: is found when $\phi$ reaches 1.
613:
614: To understand the typical development of an avalanche, it is
615: convenient to change from the variables $n_0$, $n_\msa$, and $n_1$,
616: which are constrained to sum to $N$, to the variables $x_1 \equiv
617: n_1/N$ and
618: \begin{align}
619: c &\equiv \frac{n_0}{1-g(x_1)}.
620: \label{eq: c definition}
621: \end{align}
622: As long as $n_\msa>0$, the average value of $c$ after a single update
623: is given by
624: \begin{align}
625: \langle c'\rangle
626: &= \frac{\langle n'_0\rangle}{1-g(x'_1)}\\
627: &= \frac{n_0-c[g(x'_1)-g(x_1)]}{1-g(x'_1)}\\
628: &= c.
629: \label{eq: c constant}
630: \end{align}
631: Hence, as long as $n_\msa>0$ for all members of an ensemble of
632: avalanches, $\langle c\rangle$ (the average of $c$ over the ensemble)
633: is conserved as the avalanche proceeds.
634:
635: From Eqs.~\eqref{eq: n-update first}--\eqref{eq: n-update last} and
636: the definition of $c$, the variance in $c$ can be calculated. We
637: begin by computing the increment of the variance due to one update
638: step, $\sigma^2(c')$. To leading order as $N\rightarrow\infty$,
639: we get
640: \begin{align}
641: \sigma^2(c')
642: &= \frac{\sigma^2(\delta_0)}{[1-g(x'_1)]^2}\\
643: &= \frac{n_0 U(\s,j)[1-U(\s,j)]}{[1-g(x'_1)]^2}\\
644: &= \frac{c\,U(\s,j)}{1-g(x_1)}\\
645: &= \frac{c}{N[1-g(x_1)]^2}\frac{dg(x)}{dx}\bigg|_{x=x_1}.
646: \label{eq: c-variance}
647: \end{align}
648: Eq.~\eqref{eq: c-variance} gives the increment of the variance of $c$
649: from one update step. To get the total variance of $c$, we need to sum
650: over all updates from $n_1=0$ to the desired value of $n_1$. Provided
651: that there is an upper bound $\kappa$ such that $dg(x)/dx<\kappa$ for all
652: $x$, the total variance of $c$ satisfies
653: \begin{align}
654: \sigma_{\trm{tot}}^2(c)
655: &< n_1\frac{c\,\kappa}{N[1-g(x_1)]^2}
656: <\frac{\kappa N}{1-g(x_1)}
657: \label{eq: c tot var upper}
658: \end{align}
659: for $x_1 < 1$. (Note that $1/[1-g(x)]$ is a nondecreasing
660: function because $g(x)$ is nondecreasing.)
661:
662: The avalanche is initiated with $n_\msa \equiv n_\msa^\ti$,
663: $n_0=N-n_\msa^\ti$, and $n_1=0$. The process ends when $n_0+n_1=N$ and
664: we seek the distribution of $n_0$ or $n_1$ when this happens.
665: According to Eq.~\eqref{eq: c tot var upper}, the standard deviation
666: of $c/N$ scales like $1/\sqrt{N}$, which implies that both $n_0/N$ and
667: $n_\msa/N$ have standard deviations that scale like $1/\sqrt{N}$.
668: ($x_1$ has zero standard deviation because $n_1$ is incremented by
669: exactly unity on every update step.) Thus in the large system limit,
670: the probability of any member of the the ensemble of avalanches
671: stopping is negligibly small as long as $n_\msa/N$ is finite, and we
672: may treat $c$ as exactly conserved as long as this condition holds.
673:
674: Using the initial values $x_1 = 0$ and $n_0 = N - n_\msa^\ti$, which
675: determine $c$, Eq.~\eqref{eq: c definition} can be rearranged to give
676: \begin{align}
677: n_0 &= [1-g(x_1)]\frac{N-n_\msa^\ti}{1-g(0)}.
678: \end{align}
679: Noting that $n_0/N = 1 - x_1 - n_\msa/N$, we see that
680: in the large $N$ limit, the process continues as long as the strict inequality
681: \begin{align}
682: 1 - x_1 &> [1-g(x_1)]
683: \frac{\displaystyle 1-\lim_{N\rightarrow\infty}n_\msa^\ti/N}
684: {1-g(0)}~
685: \label{eq: x1 fp}
686: \end{align}
687: holds, since the inequality implies that $n_\msa/N$ remains finite.
688: Moreover, in the large $N$ limit it is impossible to reach values of
689: $x_1$ for which the inequality has the opposite sign, because the
690: process stops when $n_\msa$ reaches zero.
691:
692: Note that because of the zero probability of a node remaining
693: undamaged when all of its neighbors are damaged, we have $g(1)=1$,
694: which in turn implies that Eq.~\eqref{eq: x1 fp} becomes an equality
695: at $x_1 = 1$. If Eq.~\eqref{eq: x1 fp} is satisfied for all $x_1 <
696: 1$, the process will be exhaustive in the sense that it will not end
697: with a finite value of $n_0/N$. If, on the other hand, the inequality
698: changes sign for $x_1$ above some threshold value, then the process
699: will terminate when the threshold is reached. If the left hand side
700: of Eq.~\eqref{eq: x1 fp} forms a tangent line to the right hand side
701: of the expression at some value of $x_1$, the process will exhibit
702: critical scaling laws. The critical case for \EP\ occurs when the
703: when the tangency occurs at $x_1 = 1$. Examples of these behaviors
704: are presented below and in Section~\ref{sec: application}.
705:
706: As an aside, we note that the \SP\ transition is an instance of
707: criticality at $x_1=0$. For the above mentioned criterion of
708: criticality to hold at $x_1=0$, the right hand
709: side of Eq.~\eqref{eq: x1 fp} must have the value $1$ and the slope $-1$ at
710: $x_1=0$. Thus, the system is critical with respect to \SP\ if
711: $\lim_{N\rightarrow\infty}n_\msa^\ti/N=0$ and
712: \begin{align}
713: \label{eq: critical g}
714: \frac{dg(x)}{dx}\bigg|_{x=0} &= 1 - g(0).
715: \end{align}
716:
717: Eqs.~\eqref{eq: g definition} and~\eqref{eq: critical g} immediately
718: give a criterion for critical percolation on graphs in which every
719: possible directed link (including self-inputs) exists with an
720: independent, fixed probability, assuming the conventional choice in
721: which damage spreads to a given node with probability $p$ from each of
722: its damaged neighbors. In this case we have
723: \begin{align}
724: g(x) &=
725: \sum_{K=0}^\infty P(K)\bigl[1 - (1-px)^K\bigr],
726: \end{align}
727: which yields
728: \begin{align}
729: \frac{dg(x)}{dx}\bigg|_{x=0} &= p \sum_{K=0}^\infty P(K) K \\
730: \ &= p \langle K \rangle.
731: \label{eq: ER percolation}
732: \end{align}
733:
734: This result is closely related to the well-known criterion for the
735: presence of a percolating cluster in an Erd\H{o}s--R\'{e}nyi graph:
736: percolation occurs when the probability $p_{\trm {\sc er}}$ for the
737: presence of a link between two randomly selected nodes exceeds $1/N$,
738: where $N$ is the number of nodes.~\cite{Bollabas:85} In the present
739: context, $p_{\trm {\sc er}}$ is mapped to $p_{\trm{link}} p$, where
740: $p_{\trm{link}}$ is the probability that a link exists connecting the
741: two randomly selected nodes and $p$ is the probability that damage
742: spreads across that link. At the same time, we have $\langle K\rangle
743: = p_{\trm{link}} N$. (Recall that $K$ is only the indegree of a node,
744: not the total number of links connected to it.) Thus Eq.~\eqref{eq:
745: ER percolation}, which implies that the critical value of $p$ is
746: $1/\langle K \rangle$, is consistent with the well-known theory of
747: Erd\H{o}s--R\'{e}nyi graphs.~\cite{Bollabas:85}
748:
749: Eq.~\eqref{eq: ER percolation} applies for any distribution of
750: indegrees so long as $\langle K \rangle$ is well-defined and the
751: source of each input is selected at random (so that the outdegrees are
752: Poisson distributed). We note that the latter condition is {\em not}
753: met by random regular graphs (graphs in which all nodes have the same
754: outdegree) because the probabilities of two nodes getting an output
755: from the same node are correlated.
756:
757: \SP\ can also be understood by the theory of Galton--Watson processes.
758: If $\lim_{N\rightarrow\infty}n_\msa^\ti/N=0$, the update described by
759: Eqs.~\eqref{eq: n-update first}--\eqref{eq: n-update last} is
760: consistent with a Galton--Watson processes that has a Poisson
761: out-degree distribution with a mean value
762: \begin{align}
763: \lambda &= \frac1{1 - g(0)}\frac{dg(x)}{dx}\bigg|_{x=0}.
764: \end{align}
765: See References~\cite{Harris:63, Otter:49, Ramo:06}. See
766: Appendix~\ref{app: SP} for more details on \SP\ in relation to known
767: results. Cases of tangencies at intermediate values of $x_1$ are
768: beyond the scope of the present work.
769:
770: Returning to the question of the \EP\ transition, it is convenient to
771: change variables once again. We define $x_\msia \equiv 1-x_1$ and
772: $q(x_\msia) \equiv 1-g(x_1)$. In words, $q(x)$ is the probability
773: that a randomly selected node will output 0 given that each of its
774: inputs has the value 0 with probability $x$. We refer to $q$ as the
775: {\it damage control function} as it characterizes the probability that
776: damage will be prevented from spreading to a single node.
777: Equation~\eqref{eq: x1 fp} is then transformed to
778: \begin{align}
779: x_\msia &> q(x_\msia)
780: \frac{\displaystyle 1-\lim_{N\rightarrow\infty}n_\msa^\ti/N}
781: {q(1)}.
782: \label{eq: x0 gt}
783: \end{align}
784: Critical \EP\ is found when the left hand side of Eq.~\eqref{eq: x0
785: gt} forms a tangent line to the right hand side of the expression at
786: $x_\msia=0$. At criticality, the right hand side of Eq.~\eqref{eq: x0
787: gt} should have the value 0 and the slope 1. Hence, the conditions
788: $q(0)=0$ and
789: \begin{align}
790: \frac{dq(x)}{dx}\bigg|_{x=0} &= \frac{q(1)}{1-n_\msa^\ti/N}
791: \label{eq: EP criticality}
792: \end{align}
793: are required for an \EP\ transition.
794:
795: \subsubsection*{Example: EP on random digraphs}
796: We now consider the special case of graphs in which every possible
797: directed link (including self-inputs) exists with an independent,
798: fixed probability. (We have already discussed \SP\ on such graphs.)
799: If damage spreads along each directed link with probability $p$, there
800: is no \EP\ transition because there is a nonzero probability for a
801: node to remain undamaged when all of its inputs are damaged. A
802: minimal change that allows \EP\ on such graphs is to give a special
803: treatment to nodes whose inputs are all damaged, in which case the
804: considered node should always get damaged. For the same reason, all
805: nodes with no inputs must be initially damaged. Other nodes might
806: also be initially damaged, and we let this happen with a given
807: probability $\rho$ for each node with at least one input. For such a
808: network, we can calculate the damage propagation function according to
809: \begin{align}
810: g(x) &= \sum_{K=0}^\infty P(K)
811: \bigl[1-(1-px)^K+(1-p)^Kx^K\bigr]\\
812: &= 1-e^{-\langle K\rangle px}\bigl(1-e^{-\langle
813: K\rangle(1-x)}\bigr).
814: \end{align}
815:
816: The corresponding damage control function becomes
817: \begin{align}
818: q(x) &= e^{-\langle K\rangle p(1-x)}\bigl(1-e^{-\langle K\rangle x}\bigr).
819: \end{align}
820: A necessary condition for the \EP\ transition is derived from
821: Eq.~\eqref{eq: EP criticality}, yielding
822: \begin{align}
823: \langle K\rangle e^{-p\langle K\rangle} &= \frac1{1-\rho}~.
824: \label{eq: E-R EP slope}
825: \end{align}
826: For the \EP\ transition to occur, it is also required that
827: \begin{align}
828: f(x) \equiv x - q(x)(1-\rho) &\ge 0
829: \label{eq: E-R EP ineq}
830: \end{align}
831: for all $x\in[0,1]$ according to Eq.~\eqref{eq: x0 gt}. If both
832: Eqs.~\eqref{eq: E-R EP slope} and~\eqref{eq: E-R EP ineq} are
833: satisfied, the \EP\ transition occurs at the value of $p$
834: given by Eq.~\eqref{eq: E-R EP slope}:
835: \begin{align}
836: p_{\trm c} &= \frac{\ln\langle K\rangle+\ln(1-\rho)}{\langle K\rangle}~.
837: \end{align}
838:
839: Equation~\eqref{eq: E-R EP slope} turns out to be a sufficient and
840: necessary condition for the \EP\ transition. Provided that
841: Eq.~\eqref{eq: E-R EP slope} holds, the first derivative satisfies
842: $f'(0)=0$. From the observation $f'''(x)<0$, it is then
843: straightforward to show that $f(x)$ has no local minimum on the
844: interval $(0,1)$. Since $f(0)=0$ and $f(1)>0$, Eq.~\eqref{eq: E-R EP
845: ineq} holds for all $x\in[0,1]$.
846:
847: It is instructive to examine the phase diagram at fixed $\rho$. A
848: negative value of $p_{\trm c}$ indicates that the system is always in
849: the \EP\ regime, so for $\langle K\rangle < 1$ the system exhibits
850: \EP\ and it is not possible to observe a transition. For $\langle
851: K\rangle > 1$, an \EP\ transition can be observed at $p=p_{\trm c}$.
852: A curious feature of this system is that $p_{\trm c}$ is not a
853: monotonic function of $\langle K\rangle$, having a maximum value of
854: $(1-\rho)/e$ at $\langle K\rangle = e/(1-\rho)$ and approaching zero
855: as $\langle K\rangle$ approaches infinity. Thus if $p$ is held fixed
856: at any value between zero and $(1-\rho)/e$, the system will undergo
857: two transitions as $\langle K\rangle$ is increased from zero. The
858: system will begin in the \EP\ regime (i.e. $p>p_{\trm c}$), undergo a
859: transition to subcritical behavior at some $\langle K\rangle$, then
860: reenter the \EP\ regime for a higher value of $\langle K\rangle$. The
861: calculated phase diagram is shown in Fig.~\ref{fig: E-R EP phase} and
862: has been verified by direct numerical simulations of avalanches.
863: Roughly speaking, at low $\langle K\rangle$ \EP\ occurs due to the
864: high density of initially damaged nodes with no inputs. At high
865: $\langle K\rangle$, on the other hand, \EP\ occurs due to the high
866: probability of nodes being damaged because of their large number of
867: inputs.
868:
869: \begin{figure}[bt]
870: \begin{center}
871: \includegraphics{E-R_EP_phase.eps}
872: \end{center}
873: \caption{\label{fig: E-R EP phase} Phase diagram for \EP\ on
874: random digraphs, where damage spreads along each directed link with
875: probability $p$ and a node is guaranteed to get damaged in the
876: special case that all of its inputs are connected to damaged nodes.
877: All nodes with zero inputs are initially damaged, and the other
878: nodes are initially damaged with probability $\rho$. The gray area
879: bounded by a solid line shows the region where \EP\ occurs for
880: $\rho=0$ and the dashed lines show the \EP\ transition when $\rho$
881: has the values $1/4$, $1/2$, and $3/4$, respectively.}
882: \end{figure}
883:
884: \subsection{The probability of complete coverage}
885: \label{sec: cc}
886: An important quantity associated with \EP\ is the probability of an
887: avalanche yielding complete coverage of the system; i.e., the
888: probability that all sites are damaged by the \UBA\ so that $u=0$.
889: Let $\Pex(N,q; n_0,n_\msa)$ denote the probability that a \UBA\ on a
890: random network will yield complete coverage for a system with a given
891: network size $N$, a given damage control function $q$, and starting
892: with particular values of $n_0$ and $n_\msa$. For future convenience
893: we also define $\Pex(N,q)$ to be the probability for complete coverage
894: assuming that each node is initially damaged with probability $1-q(1)$
895: and we average over the corresponding probability distribution for
896: $n_\msa$.
897:
898: To calculate $\Pex(N,q; n_0,n_\msa)$, we note that
899: \begin{align}
900: \Pex(N,q;m,0)=0\quad {\rm if\ } m>0,
901: \end{align}
902: since the process stops when $n_\msa = 0$. We also have
903: \begin{align}
904: \Pex(N,q;0,m)=1\quad {\rm for\ any\ } m,
905: \end{align}
906: since updating can never create 0s. These values of $\Pex$ can be
907: used for recursive calculation of $\Pex$. Let $n_\msia$ denote
908: $n_0+n_\msa$, or $Nx_\msia$. Performing steps \ref{j select}--\ref{j
909: := 1} (from Section~\ref{sec: formal UBA}) one time decreases
910: $n_\msia$ by $1$ as described by Eqs.~\eqref{eq: n-update
911: first}--\eqref{eq: n-update last}. This means that $\Pex(N,q;
912: n_0,n_\msa)$ can be calculated for all $n_\msia=m$ if $\Pex(N,q;
913: n_0,n_\msa)$ is known for all $n_\msia=m-1$. The recursion starts at
914: $n_\msia=0$ with $\Pex(N,q;0,0)=1$ and uses the boundary conditions
915: $\Pex(N,q; n_\msia,0)=0$ and $\Pex(N,q; 0,n_\msia)=1$ for $n_\msia>0$.
916:
917: For large $N$, $\Pex$ can be calculated in the framework of a
918: continuous approximation. Let $p(n_\msia,c)$ denote a continuous version
919: of $\Pex(N,q$; $n_0,n_\msa)$. Then, the boundary conditions
920: $\Pex(N,q; n_\msia,0)=0$ and $\Pex(N,q$; $0,n_\msia)=1$ are
921: expressed as
922: \begin{align}
923: p[n_\msia,c_{\trm{max}}(x_\msia)] &= 0,
924: \label{eq: p-diff bound beg}
925: \intertext{and}
926: p(n_\msia,0) &= 1,
927: \intertext{where}
928: c_{\trm{max}}(x_\msia) &= \frac{n_\msia}{q(x_\msia)}.
929: \label{eq: p-diff bound end}
930: \end{align}
931:
932: In the continuous approximation, the recurrence relation that can be
933: derived from Eqs.~\eqref{eq: n-update first}--\eqref{eq: n-update
934: last} is transformed to a partial differential equation. In such an
935: update, the change $n_\msia$ decreases by unity and, for large $N$,
936: the change in $c$ is much less than $c$ itself. In the continuous
937: approximation, this means that $p(n_\msia,c)$ satisfies a partial
938: differential equation of the form
939: \begin{align}
940: \frac{\partial p}{\partial n_\msia}
941: &= h_1(n_\msia,c)\frac{\partial p}{\partial c}
942: + h_2(n_\msia,c)\frac{\partial^2p}{\partial c^2},
943: \label{eq: p-diff 0}
944: \end{align}
945: where $h_1(n_\msia,c)$ and $h_2(n_\msia,c)$ are functions to be
946: determined. This is recognizable as a 1D diffusion equation in which
947: $n_\msia$ plays the role of time and $c$ the role of space. Note that
948: later times in the diffusion equation correspond to earlier stages of
949: the \UBA, since $n_\msia$ decreases as nodes are converted to 1s. The
950: boundary conditions on the diffusion are given by Eqs.~\eqref{eq:
951: p-diff bound beg} and~\eqref{eq: p-diff bound end}. We are
952: interested in computing $p(n_{\msia},c)$ for values of $n_\msia$ and
953: $c$ corresponding to $n_\msa = n_\msa^\ti$ and $n_1 = 0$.
954:
955: The fact that the average of $c$ is constant means that the
956: coefficient of the drift term in the diffusion equation must vanish;
957: i.e., $h_1(n_\msia,c)=0$. The diffusion coefficient,
958: $h_2(n_\msia,c)$, is given by
959: \begin{align}
960: h_2 &= \tfrac12\sigma^2(c'),
961: \label{eq: h2}
962: \end{align}
963: where $\sigma^2(c')$ is the variance of $c'$ when a fixed $c$ is
964: updated.
965:
966: Using Eqs.\ \eqref{eq: c-variance} and \eqref{eq: h2} and converting
967: $g$'s to $q$'s, we find
968: \begin{align}
969: \frac{\partial p}{\partial n_\msia}
970: &= \frac{c}{2N[q(x_\msia)]^2}
971: \frac{dq(x)}{dx}\bigg|_{x=x_\msia}
972: \frac{\partial^2p}{\partial c^2}.
973: \label{eq: p-diff 1}
974: \end{align}
975:
976: The large $N$ behavior of Eq.~\eqref{eq: p-diff 1}, with the boundary
977: conditions in Eqs.\ \eqref{eq: p-diff bound beg} and \eqref{eq: p-diff
978: bound end}, can be found by expanding $q(x)$ around $x=0$. If $q(x)$
979: is well-behaved, such an expansion can be written as
980: \begin{align}
981: q(x) &= \alpha_1x - \alpha_2x^2 + \Oc(x^3).
982: \label{eq: q-Taylor}
983: \end{align}
984: This expansion can always be performed if the probability $P(K)$ for a
985: node to have $K$ inputs decays as least as fast as $K^{-4}$ and in the
986: case that $p_K$ decays slower than $K^{-4}$ but faster than $K^{-3}$,
987: only the residue term can be affected. See Appendix \ref{app: q
988: calc}. In particular, the expansion is always valid if $K$ has a
989: maximal value.
990:
991: The most interesting case in terms of asymptotic behavior is when
992: $\alpha_1$ is close to $1$ and $\alpha_2$ is positive. With suitable
993: $N$-dependent transformations of $p$ and its arguments, described in
994: Appendix \ref{app: PEP}, the large $N$ behavior of Eq.~\eqref{eq:
995: p-diff 1} can be expressed in terms of a function $\pt(\tti,\yt)$
996: determined by the differential equation
997: \begin{align}
998: \frac{\partial\pt}{\partial\tti} &= \frac12
999: \frac{\partial^2\pt}{\partial\yt^2},
1000: \label{eq: diffuse}
1001: \end{align}
1002: with the boundary conditions
1003: \begin{align}
1004: \pt(\tti,1/\tti) &= 0 \quad\trm{for }\tti<0
1005: \label{eq: diffuse bound 0}
1006: \intertext{and}
1007: \lim_{\tti\rightarrow-\infty}\pt(\tti,\yt) &= \yt
1008: \quad \trm{for }\yt\ge0.
1009: \label{eq: diffuse bound 1}
1010: \end{align}
1011: The Crank--Nicholson method can be used to calculate $\pt(\tti,\yt)$
1012: numerically in an efficient way. (See, e.g., \cite{numrecip}.)
1013:
1014: Appendix~\ref{app: PEP} shows that the probability for complete
1015: coverage is given by
1016: \begin{align}
1017: \Pex(N, q) &\approx \Nt^{-1/3}\pt[0,\Nt^{1/3}(1-\alpha_1)],
1018: \label{eq: PexNq}
1019: \end{align}
1020: where $\Nt=\alpha_1N/\alpha_2$. The calculation assumes that the
1021: avalanche is initiated on the nodes whose outputs are independent of
1022: their inputs, as accounted for in $q(1)$.
1023:
1024: To our knowledge, the critical point for \EP\ has not been
1025: investigated previously in its own right. Two special cases have been
1026: studied, however. First, results for numbers of frozen and unfrozen
1027: nodes in critical \RBN s can be mapped to an \EP\ process, as
1028: discussed in Section~\ref{sec: application}. In this context, frozen
1029: nodes in the network are considered to be the damaged nodes of the
1030: \UBA, and the scaling with $N$ of the number of unfrozen nodes at the
1031: phase transition has been investigated for certain class of \RBN s
1032: \cite{Socolar:03,Kaufman:05}.
1033:
1034: Second, in the special case that $q(x)=x$, the exact result
1035: \begin{align}
1036: \Pex(N,x\mapsto x;n_0,n_\msa)
1037: &= \frac{n_\msa}{n_0+n_\msa}
1038: \label{eq: q-ident main}
1039: \end{align}
1040: is obtained. [See Eq.~\eqref{eq: q-ident} in Appendix \ref{app:
1041: PEP}.] This means that the probability for complete coverage is
1042: exactly $n^\ti_\msa/N$. The simplest realization of $q(x)=x$ is
1043: provided by a network of one-input nodes with rules that copy the
1044: input state. Such networks have strong connections to random maps
1045: from a set of $N$ elements into itself. A map $T$ is derived from a
1046: network of one-input nodes by letting each node map to the node from
1047: where its input is taken. In this picture, the damage originating from
1048: one initially damaged node $i$, corresponds to the set of nodes $j$
1049: such that $T^k(j)=i$ for some $k\ge0$ (where $T^k$ denotes the $k$th
1050: iterate of $T$). Such a $j$ is called a {\it predecessor} to $i$. See,
1051: e.g., Ref.~\cite{Bollabas:85} for an overview of the theory of
1052: random maps and see Refs.~\cite{Rubin:54, Harris:60} for results on
1053: predecessors in random maps. See Appendix~\ref{app: exact} for
1054: analytic results that relate \UBA\ to random maps.
1055:
1056: \subsection{On the number of damaged nodes in random networks}
1057: \label{sec: avalanche size}
1058:
1059: In the Sections~\ref{sec: criteria for EP} and~\ref{sec: cc} we
1060: focused on determining the parameters that lead to \EP\ (a vanishing
1061: fraction of undamaged nodes large $N$ limit) and on the probability
1062: that the number of undamaged nodes will be exactly zero (complete
1063: coverage). We now consider the full probability distribution for the
1064: number of nodes damaged in an avalanche in a manner that provides a
1065: suitable base for understanding both \SP\ and \EP\ in random networks.
1066: The calculational strategy involves considering a given set of $n$
1067: nodes to be the damaged set and computing the probability that this is
1068: both consistent with all of the Boolean rules and the probability that
1069: the avalanche will actually cover the whole set. The probability of
1070: consistency is calculated via elementary combinatorics. The
1071: probability of reaching the whole set is precisely the probability of
1072: complete coverage for an avalanche on the sub-network of $n$ candidate
1073: nodes. For this we can directly apply the results of the last
1074: section. For the purposes of explaining the calculation, we refer to
1075: the selected set of $n$ nodes as the {\it candidate set}.
1076:
1077: We let $\Pn(n)$ denote the probability that $n$ nodes will be damaged
1078: in an avalanche, averaged over the ensemble of $N$-node networks with
1079: a rule distribution characterized by a given damage propagation
1080: function $g$ or the corresponding damage control function $q$. We
1081: assume that the avalanche is initiated by randomly selecting $\ell$
1082: nodes to set to $\msa$, regardless of their Boolean rules, then
1083: setting to $\msa$ all nodes with rules that always output 1 for any
1084: inputs. The set of $\ell$ initially damaged nodes must be a subset of
1085: the candidate set. The probability that the candidate set contains
1086: all of the nodes with ``always 1'' rules will be taken into account by
1087: the value of $g(0)$ in the expression below for the consistency
1088: probability. We use the notation $\binom m k$ for the usual binomial
1089: coefficient (the number of combinations of $k$ objects chosen from a
1090: set of $m$ objects).
1091:
1092: The probability $\Pn(n)$ can be expressed as
1093: \begin{align}
1094: \Pn(n) = \binom{N-\ell}{n-\ell} P_{\trm c}(n, \ell; N)\, \PNex(n,\ell;N),
1095: \label{eq: Pnn}
1096: \end{align}
1097: where $P_{\trm c}(n, \ell; N)$ and $\PNex(n,\ell;N)$ are defined below.
1098: The binomial factor counts the number of different sets of $n-\ell$
1099: nodes that could be damaged in a process corresponding to a given set
1100: of $\ell$ nodes that are initially damaged without regard to their
1101: rules. $P_{\trm c}(n, \ell; N)$ is the probability that a given choice of
1102: $n-\ell$ nodes assumed to be damaged by the avalanche will constitute
1103: a final state that is consistent with the Boolean rules for each node,
1104: including the nodes that are initially damaged because their rules
1105: require it. $\PNex(n,\ell;N)$ is the probability that the avalanche
1106: will not die out before damaging all $n$ nodes. This factor is
1107: necessary to avoid counting final states that contain loops of damaged
1108: nodes consistent with the rules but unreachable because damage cannot
1109: spread to the loop from any nodes outside the loop.
1110:
1111: Consistency with the Boolean rules requires that the given set of
1112: $n-\ell$ nodes damaged in the avalanche have inputs that cause them to
1113: be damaged. In a random network, the probability that any single node
1114: will be damaged is $g(x_1)$, where $x_1$ is the fraction of damaged
1115: nodes. Similarly, the probability that any node will {\it not} be
1116: damaged is $1-g(x_1)$. We are considering candidate sets of damaged
1117: nodes with $x_1 = n/N$. Thus we have
1118: \begin{align}
1119: P_{\trm c}(n,\ell;N) = [g(n/N)]^{n-\ell}[1-g(n/N)]^{N-n}.
1120: \label{eq: Pconsistency}
1121: \end{align}
1122:
1123: The computation of $\PNex(n,\ell;N)$ involves the rule distribution on
1124: the restricted network formed by the candidate set with all inputs
1125: from the undamaged nodes removed. This distribution, $g^1(x)$, is
1126: different from $g(x)$ because $P_{\trm c}$ already accounts for rules that
1127: are not consistent with the pattern of damage. Thus the spreading of
1128: damage on the $n$-node network involves $g(nx/N)$, the probability
1129: that a rule outputs $1$ when a fraction $x$ of the $n$-node candidate
1130: set is damaged. The probability must be normalized such that it goes
1131: to unity when $x$ goes to 1. (We know that a node in the $n$-node set
1132: should get damaged if all of its inputs are damaged.) Thus
1133: we have
1134: \begin{align}
1135: g^1_{N,n}(x) = \frac{g(nx/N)}{g(n/N)}
1136: \label{eq: g1}
1137: \end{align}
1138: or, equivalently,
1139: \begin{align}
1140: q^1_{N,u}(x) &= \frac{q[u/N+(1-u/N)x]-q(u/N)}{1-q(u/N)}.
1141: \label{eq: q1}
1142: \end{align}
1143: (Recall that $u = N-n$ is the number of undamaged nodes after an
1144: avalanche.)
1145:
1146: There are two cases of interest for the probability of complete
1147: coverage of the candidate set. For \EP, $g(0)>0$ and the fixed number
1148: $\ell$ of nodes arbitrarily selected for damage is irrelevant compared
1149: to the finite fraction of nodes with rules that produce damage for any
1150: combination of inputs. In this case, we assume $\ell=0$, which allows
1151: reduction of $\Pex$ to its two-argument form defined at the beginning of
1152: Section~\ref{sec: cc}:
1153: \begin{align}
1154: \PNex(n,0;N) & = \Pex(n,q^1_{N,N-n}).
1155: \label{eq: cc on n EP}
1156: \end{align}
1157:
1158: For \SP, we have $g(0)=0$ so the avalanche must be initiated with a
1159: nonzero value of $\ell$. In this case we have
1160: \begin{align}
1161: \PNex(n,\ell;N) = \Pex(n,q^1_{N,N-n};n-\ell,\ell).
1162: \label{eq: cc on n SP}
1163: \end{align}
1164: Note that $\PNex(n,\ell;N)$ depends on $N$ only through $q^1$.
1165:
1166: For notational convenience, we now let $\PNex$ stand for whichever
1167: expression on the right-hand side of Eqs.~\eqref{eq: cc on n EP}
1168: or~\eqref{eq: cc on n SP} is relevant, and we use $u$ where $N-n$
1169: would be the strictly proper form. By combining Eqs.~\eqref{eq: Pnn}
1170: and \eqref{eq: Pconsistency}, we get
1171: \begin{align}
1172: \Pn(n) =\,&\binom{N-\ell}{n-\ell}[g(n/N)]^{n-\ell}[1-g(n/N)]^{u}
1173: \PNex.
1174: \label{eq: Pn0}
1175: \end{align}
1176:
1177: To make some important features of Eq.~\eqref{eq: Pn0} apparent, we
1178: introduce the functions
1179: \begin{align}
1180: \rho(n) &= \frac{n^n}{e^nn!},\\
1181: \tau(n, k) &= \frac{n!}{n^k(n-k)!},
1182: \intertext{and}
1183: G(x) &= \biggl(\frac{g(x)}x\biggr)^{\!x}
1184: \biggl(\frac{1-g(x)}{1-x}\biggr)^{\!1-x}.
1185: \end{align}
1186: Then Eq.~\eqref{eq: Pn0} can be rewritten as
1187: \begin{align}
1188: \Pn(n)
1189: =\,&\frac{\rho(n)\rho(u)}{\rho(N)}
1190: \frac{\tau(n, \ell)}{\tau(N, \ell)}
1191: \biggl(\frac{n/N}{g(n/N)}\biggr)^{\!\ell} \nonumber\\
1192: &\times [G(n/N)]^N \PNex.
1193: \label{eq: Pn1}
1194: \end{align}
1195:
1196: Stirling's formula,
1197: \begin{align}
1198: n! &\approx \sqrt{2\pi n}\,\frac{n^n}{e^n},
1199: \label{eq: Stirling}
1200: \end{align}
1201: yields
1202: \begin{align}
1203: \rho(n)&\approx\frac1{\sqrt{2\pi n}}
1204: \intertext{and}
1205: \frac{\rho(n)\rho(u)}{\rho(N)}&\approx\frac1{\sqrt{2\pi nu/N}}.
1206: \end{align}
1207:
1208: The factor $\tau(n,\ell)/\tau(N,\ell)$ is approximately 1 for large $n$ and
1209: satisfies
1210: \begin{align}
1211: \frac{\tau(n, \ell)}{\tau(N, \ell)}&\le1
1212: \end{align}
1213: for $n\leq N$, with equality if $n=N$ or $\ell=1$ or $\ell=0$. The
1214: only factors in Eq.~\eqref{eq: Pn1} that can show exponential
1215: dependence on $N$ are the $G$ and $\PNex$ factors. Because $\PNex$ is
1216: a probability (and therefore cannot exceed unity) and $G(x)\leq1$ with
1217: equality if and only if $g(x)=x$, $\Pn(n)$ vanishes exponentially as
1218: $N$ goes to infinity for any fixed $n/N$ such that $g(n/N)\ne n/N$.
1219: This is consistent with the above result that the probability of an
1220: avalanche stopping with $x_1 \neq g(x_1)$ is vanishingly small. [See
1221: Eqs.~\eqref{eq: x1 fp} and~\eqref{eq: x0 gt}.]
1222:
1223: For \EP, we are interested in the number of undamaged nodes, $u$. We
1224: let
1225: \begin{align}
1226: \Pu(u) &= \Pn(N-u)
1227: \intertext{and}
1228: Q(x) &= G(1-x) \nonumber \\
1229: &= \biggl(\frac{1-q(x)}{1-x}\biggr)^{\!1-x}
1230: \biggl(\frac{q(x)}x\biggr)^{\!x}.
1231: \end{align}
1232: For \EP, $g(0)>0$ and a fixed $\ell$ is irrelevant when
1233: $N\rightarrow\infty$. Hence, we let $\ell=0$ and rewrite
1234: Eq.~\eqref{eq: Pn1} as
1235: \begin{align}
1236: \Pu(u) =\frac{\rho(n)\rho(u)}{\rho(N)} [Q(u/N)]^N \PNex.
1237: \label{eq: Pu0}
1238: \end{align}
1239:
1240: To some respects, $\Pu$ is similar to $\Pn$: the factor
1241: $[\rho(u)\rho(n)]/\rho(N)$ is fully symmetric with respect to
1242: interchange of $n$ and $u$; and the role of $G(n/N)$ in Eq.~\eqref{eq:
1243: Pn1} is identical to the role of $Q(u/N)$ in Eq.~\eqref{eq: Pu0}.
1244: However, the behavior of $\PNex$ for $n\ll N$ given by Eq.~\eqref{eq:
1245: cc on n SP} is significantly different from the behavior of $\PNex$
1246: for $u\ll N$ given by Eq.~\eqref{eq: cc on n EP}.
1247:
1248: For \EP, we consider damage control functions $q(x)$ that can be
1249: expanded according to Eq.~\eqref{eq: q-Taylor}. For supercritical \EP,
1250: with $\alpha_1<1$, $\Pu(u)$ decays exponentially with $u$. In Appendix
1251: \ref{app: super EP}, we demonstrate that
1252: \begin{align}
1253: \lim_{N\rightarrow\infty}\Pu(u)
1254: &= (1-\alpha_1)\frac{(u\alpha_1)^{u}}{u!}e^{-u\alpha_1}
1255: \label{eq: pu(u) exact asympt}
1256: \\
1257: \ &\approx\frac{1-\alpha_1}{\sqrt{2\pi}}e^{u(1-\alpha_1)}
1258: \alpha_1^uu^{-1/2}.
1259: \label{eq: pu(u) asympt}
1260: \end{align}
1261:
1262: For critical \EP, Eq.~\eqref{eq: PexNq} gives
1263: \begin{align}
1264: \PNex(n,0;N) & = \Pex(n,q^1_{N,N-n}) \nonumber \\
1265: & \approx \tilde{n}^{-1/3}\pt[0,\tilde{n}^{1/3}(1-\alpha_1^1)],
1266: \label{eq: cc on n crit EP}
1267: \end{align}
1268: where $\tilde{n} \equiv \alpha_1^1 n/\alpha_2^1$ and $\alpha_1^1$ and
1269: $\alpha_2^1$ are the first two coefficients of the power series
1270: expansion of $q^1(x)$ about $x=0$. With $\alpha_1=1$ and
1271: $\alpha_2>0$, a Taylor expansion of $\log Q(x)$ about $x=0$ gives
1272: \begin{align}
1273: Q(x) &\approx \exp\biggl(-\frac{\alpha_2^2x^3}2\biggr)
1274: \end{align}
1275: for small $x$. This yields that the typical number of undamaged nodes,
1276: $u$, scales like $N^{2/3}$. In Appendix \ref{app: crit EP}, we derive
1277: the asymptotic distribution of $u$ for large $N$. With
1278: $\ut=\Nt^{-2/3}u = (\alpha_2/N)^{2/3}u$, we find that the large $N$
1279: limit of the probability density for $\ut$ is
1280: \begin{align}
1281: P(\ut) &= \frac{\exp(-\frac12\ut^3)}
1282: {\sqrt{2\pi\ut}}\,\pt(0,2\ut).
1283: \label{eq: p(ut) asympt}
1284: \end{align}
1285:
1286: Eq.~\eqref{eq: Pn1} is suitable for understanding \SP\ as well as \EP.
1287: For \SP, $g(0)=0$ and $\ell>0$. In the large $N$ limit, \SP\ is a
1288: branching process with a Poisson distribution in the number of
1289: branches from each node. The average number of branches per node is
1290: given by the derivative of $g(x)$ at $x=0$, because
1291: $\lim_{x\rightarrow0} g(x)/x$ is the average number of nodes that will
1292: be damaged in one update as a direct consequence of damaging a single
1293: node in the large network limit. In Appendix \ref{app: SP}, we
1294: re-derive known results on \SP\ in the framework of our formalism.
1295:
1296: \section{An application: Frozen nodes in random Boolean networks}
1297: \label{sec: application}
1298:
1299: An important application of our results on \EP\ in random networks is
1300: the determination of the size distribution for the set of unfrozen
1301: nodes in 2-input random Boolean networks, a subject of interest since
1302: the introduction of the Kauffman model in 1969 \cite{Kauffman:69}.
1303: The Kauffman model was originally proposed as a vehicle for studying
1304: aspects of the complex dynamics of transcriptional networks within
1305: cells.
1306:
1307: In a Boolean network, there are usually some nodes that will reach a
1308: fixed final state after a transient time regardless of the initial
1309: state of the network. For most random Boolean networks, nearly all of
1310: these nodes can be found by a procedure introduced in
1311: Ref.~\cite{Flyvbjerg:88b} and applied numerically in
1312: Ref.~\cite{Bilke:01}. We refer to nodes identified by this procedure
1313: as {\it frozen}.
1314:
1315: The nodes that cannot be identified as frozen are labeled {\it
1316: unfrozen}. Their output may switch on and off for all time or
1317: simply have different values on different attractors of the network
1318: dynamics. A frozen node will always reach its fixed final state
1319: regardless of the initial state of the network. The converse is not
1320: true: an unfrozen node can have a fixed final state that is
1321: independent of the initial state due to correlations that are not
1322: accounted for in the identification procedure for frozen nodes. In a
1323: typical random Boolean network, the number of nodes that are
1324: mislabeled in this sense is negligible \cite{Bilke:01}. For the
1325: purposes of investigating dynamics of the network at long times, one
1326: is interested in the size of the unfrozen set.
1327:
1328: The procedure for identification of the frozen nodes starts by marking
1329: all nodes with a constant output function as frozen. There may then
1330: be nodes that, as a consequence of receiving one or many inputs from
1331: frozen nodes, will also produce a constant output. These nodes are
1332: also marked as frozen, and the process continues iteratively until
1333: there are no further nodes that can be identified as frozen.
1334:
1335: We note here that the process of finding frozen nodes in a \RBN\ can
1336: often be framed as a \UBA, where the property of being frozen
1337: corresponds to damage. That is, the process of identifying frozen
1338: nodes involves continually checking all nodes to see whether their
1339: inputs are frozen in such a way that they themselves become frozen, a
1340: process which satisfies the conditions for \UBA. The damage
1341: propagation and damage control functions for the \UBA\ are determined
1342: by the relative weights of different Boolean logic functions in the
1343: \RBN. By changing these weights, one can observe a transition in the
1344: dynamical behavior of \RBN s corresponding precisely to the \EP\
1345: transition in the \UBA. We consider here \RBN s with exactly two
1346: inputs at each node, with some explicit choices for the weights of the
1347: Boolean logic functions that permit observation of both sides of the
1348: transition.
1349:
1350: The only restriction required for mapping the freezing of nodes in a
1351: \RBN\ to a \UBA\ system is that the logic functions in the \RBN\ be
1352: symmetric with respect to the probability of freezing being due to
1353: {\sc true} and {\sc false} inputs. That is, the probability that a
1354: node with a certain set of frozen inputs will itself be frozen should
1355: not depend on the values of the frozen inputs. This condition is
1356: satisfied for the most commonly investigated classes of rule
1357: distributions, where there is a given probability $p$ for obtaining a
1358: 1 at each entry in the truth table for each rule. If the above
1359: mentioned symmetry condition were violated, it would be necessary to
1360: distinguish nodes frozen {\sc true} from nodes frozen {\sc false},
1361: which would mean that the state of a node could not be specified by a
1362: binary variable. For the rest of this section we consider only \RBN s
1363: that respect the symmetry condition.
1364:
1365: It is useful to distinguish different types of Boolean logic
1366: functions. A {\it canalizing} rule is one for which the output is
1367: independent of one of the inputs for at least one value of the other
1368: input. Among the 16 possible 2-input Boolean rules, 2 rules are
1369: constant (``always on'' or ``always off''), 12 rules are non-constant
1370: and canalizing, and 2 rules are non-canalizing ({\sc xor} and not-{\sc
1371: xor}). The original version of the Kauffman model assumes that all
1372: 2-input Boolean rules are equally likely, which turns out to give
1373: critical dynamics.
1374:
1375: Let $p_i$ denote the probability that a randomly selected node's
1376: output is frozen if exactly $i$ of its inputs are frozen. The damage
1377: propagation function $g(x)$ can be expressed directly in terms of
1378: $p_i$:
1379: \begin{align}
1380: g(x) = p_0(1-x)^2 + 2p_1x(1-x) + p_2x^2.
1381: \end{align}
1382: Nodes with constant rules are guaranteed to be frozen. (These nodes
1383: will initiate the \UBA.) Nodes with non-constant canalizing rules are
1384: unfrozen if both inputs are unfrozen, and they are frozen with
1385: probability $1/2$ if exactly one randomly selected input is frozen.
1386: Nodes with rules that are non-canalizing become frozen if and only if
1387: both of their inputs are frozen. Finally, if both inputs are frozen,
1388: the output of any 2-input rule is frozen. Thus for the 2-input
1389: Kauffman model, $p_0=1/8$, $p_1=1/2$, and $p_2=1$.
1390:
1391: \begin{figure}[bt]
1392: \begin{center}
1393: \includegraphics{gq.eps}
1394: \includegraphics{GQ.eps}
1395: \end{center}
1396: \caption{\label{fig: gq} The functions (a) $g(x)\equiv 1-q(1-x)$ and
1397: (b) $G(x)\equiv Q(1-x)$ for three 2-inputs rule distributions. All
1398: three distributions have $p_0=1/8$ and $p_2=1$, whereas $p_1$ takes
1399: the values $7/16$, $1/2$, and $9/16$. The case that has $p_1=1/2$
1400: (marked with $=$) is critical with respect to \EP\ and corresponds
1401: to the propagation of frozen node values in the original Kauffman
1402: model. The other cases $p_1=7/16$ ($<$) and $p_1=9/16$ ($>$) are
1403: subcritical and supercritical, respectively. The dashed line in (a)
1404: shows the identity function $x\mapsto x$.}
1405: \end{figure}
1406:
1407: If the two non-canalizing rules in the 2-input Kauffman model are
1408: replaced by canalizing rules, $p_1$ becomes $9/16$, whereas $p_0$ and
1409: $p_2$ are unchanged. Such networks exhibit supercritical \EP. To get a
1410: subcritical network, we replace two of the canalizing rules with
1411: non-canalizing rules and get $p_1=7/16$. (Note that some care must be
1412: taken to maintain the {\sc true}--{\sc false} symmetry mentioned
1413: above.) The functions $g(x)$ and $G(x)$ for critical, supercritical,
1414: and subcritical rule distributions are shown in Fig.~\ref{fig: gq}.
1415:
1416: \begin{figure}[bt]
1417: \begin{center}
1418: \includegraphics{K2_sub.eps}
1419: \\\vspace*{-42pt}
1420: \includegraphics{K2_cri.eps}
1421: \\\vspace*{-42pt}
1422: \includegraphics{K2_sup.eps}
1423: \end{center}
1424: \caption{\label{fig: K2} The probability density distribution $N\Pn(n)$
1425: with respect to the fraction of nodes ($n/N$) involved in an
1426: avalanche. The rule distributions have the same $g(x)$ as displayed
1427: in Fig.~\ref{fig: gq}, showing rule distributions that are ($<$)
1428: subcritical, ($=$) critical, and ($>$) supercritical with respect to
1429: \EP. The displayed networks sizes, $N$, are 10 (large dots), 100
1430: (small dots), $10^3$ (bold line), $10^4$, $10^5$, and $10^6$
1431: (gradually thinner lines).}
1432: \end{figure}
1433:
1434: \begin{figure}[bt]
1435: \begin{center}
1436: \includegraphics{K2_num_N1000.eps}
1437: \end{center}
1438: \caption{\label{fig: K2_num} A numeric comparison between analytic
1439: calculations (black lines) and explicit reductions of random Boolean
1440: networks (gray lines). For both cases, the probability density
1441: distribution $N\Pn(n)$ is displayed as a function of $n/N$. The rule
1442: distributions have the same $g(x)$ as displayed in Figs.\ \ref{fig:
1443: gq} and \ref{fig: K2}, showing rule distributions that are ($<$)
1444: subcritical, ($=$) critical, and ($>$) supercritical with respect to
1445: \EP. The \UBA\ rule distributions are realized in random Boolean
1446: networks by rule distributions with the following respective
1447: selection probabilities: $1/8$, $1/4$, $5/8-p_{\trm r}$, and
1448: $p_{\trm r}$ for a constant rule, a rule that depends on exactly 1
1449: input, a canalizing rule that depends on 2 inputs, a 2-input
1450: reversible rule. The values of $p_{\trm r}$ are ($<$) 0, ($=$) 1/8,
1451: and ($>$) 1/4. For each rule distribution, $10^6$ networks were
1452: tested.}
1453: \end{figure}
1454:
1455: As can be seen from Fig.~\ref{fig: gq}, a small change in $g(x)$ may
1456: lead to a qualitative change in $G(x)$ for rule distributions close to
1457: criticality. Such changes have a strong impact on the avalanche size
1458: distribution for large $N$. Figure~\ref{fig: K2} shows the probability
1459: density distribution of the fraction, $n/N$, of nodes that are
1460: affected by avalanches in networks with the above mentioned rule
1461: distributions. The probability distributions are obtained by recursive
1462: calculation of the distribution of $n_\msa$ as $n_1$ increases. The
1463: recurrence relations are obtained from Eqs.~\eqref{eq: n-update
1464: first}--\eqref{eq: n-update last} and the result is exact up to
1465: truncation errors. To verify these calculations, we generated
1466: $10^6$ random Boolean networks of size $N=10^3$ for each of the above
1467: described rule distributions. The distributions in the numbers of
1468: frozen nodes in those networks are displayed in Fig.~\ref{fig:
1469: K2_num}.
1470:
1471: \begin{figure}[bt]
1472: \begin{center}
1473: \includegraphics{EP_K2_cri.eps}
1474: \includegraphics{EP_K2_sup.eps}
1475: \end{center}
1476: \caption{\label{fig: EP_K2} Rescaled versions of the probability
1477: distributions displayed in Fig.~\ref{fig: K2}: (a) the probability
1478: density for the critical case, with respect to the rescaled number
1479: of undamaged nodes, $\ut\equiv(\alpha_2/N)^{2/3}u=u/(4N^{2/3})$; (b)
1480: the probability distribution $\Pu(u)$ for the supercritical case
1481: without rescaling. The displayed networks sizes, $N$, are 10 (large
1482: dots), 100 (small dots), $10^3$ (bold line), $10^4$, $10^5$, and
1483: $10^6$ (gradually thinner lines). The analytically derived
1484: asymptotes are shown as dashed lines. In (b), the distributions for
1485: networks of sizes $10^4$, $10^5$, and $10^6$ are not plotted because
1486: they are indistinguishable from the asymptotic curve.}
1487: \end{figure}
1488:
1489: In Fig.~\ref{fig: EP_K2}, the probability distributions of the number
1490: of undamaged nodes, $u$, are shown in comparison to the asymptotic
1491: results in Eqs.~\eqref{eq: pu(u) exact asympt} and~\eqref{eq: p(ut) asympt}.
1492: Our analytic results are strengthened by the data in Fig.~\ref{fig:
1493: EP_K2} as the distributions for finite networks approaches the
1494: predicted asymptotes. Finite size effects are clearly visible in the
1495: critical case even for network sizes as big as $N=10^6$, whereas
1496: convergence in the supercritical case is achieved for $N \gtrsim
1497: 10^3$.
1498:
1499: Kaufman, Mihaljev, and Drossel\ studied distributions of unfrozen nodes in
1500: 2-input critical \RBN s using a method similar to ours in that
1501: differential equations for populations of different types of nodes are
1502: developed from a discrete process in which frozen nodes are identified
1503: by the propagation of information from their inputs~\cite{Kaufman:05}.
1504: Their result for the numbers of unfrozen nodes
1505: in 2-input critical \RBN s corresponds to a particular application of
1506: Eq.~\eqref{eq: p(ut) asympt}. In Ref.~\cite{Kaufman:05}, the function
1507: corresponding to $P(\ut)$ [which they call $G(y)$] is
1508: determined by running a stochastic process and a numerically motivated
1509: approximation is proposed:
1510: \begin{align}
1511: P(\ut) &\approx 0.25\exp(-\tfrac12\ut^3)
1512: \frac{1-0.5\sqrt{\ut}+3\ut}{\sqrt{\ut}}.
1513: \label{eq: p(ut) Kaufman approx}
1514: \end{align}
1515: The scaling law $P(\ut)\propto\ut^{-1/2}$ for small $\ut$ is also
1516: derived analytically in Ref.~\cite{Kaufman:05}.
1517:
1518: For large $x$, Eqs.~\eqref{eq: diffuse}--\eqref{eq: diffuse bound 1}
1519: imply $\pt(0,x)\propto x$ for large positive $x$. This means that
1520: \begin{align}
1521: P(\ut) \approx \sqrt{\frac{2\ut}{\pi}}\exp(-\tfrac12\ut^3)
1522: \end{align}
1523: for large $\ut$. Thus the large $\ut$ limit of Eq.~\eqref{eq: p(ut)
1524: Kaufman approx} differs from the exact result by a factor of
1525: $(3/4)\sqrt{\pi/2}$, an underestimate of about 6\%.
1526:
1527: We are able to improve further on Eq.~\eqref{eq: p(ut) Kaufman approx}
1528: by numerical investigations of $\pt(0,x)$ calculated by the
1529: Crank--Nicholson method (see, e.g., \cite{numrecip}) using
1530: Eqs.~\eqref{eq: diffuse}--\eqref{eq: diffuse bound 1}. We find that
1531: the high-precision numerical results are fit by the function
1532: \begin{align}
1533: P(\ut) \approx \sqrt{\frac{2\ut}{\pi}}\exp(-\tfrac12\ut^3)
1534: \biggl(\!1\!+\frac1{3.248\ut+4.27\ut^2+4.76\ut^3}\!\biggr)
1535: \end{align}
1536: with a relative error that is maximally 0.25\% and vanishing for large
1537: $\ut$.
1538:
1539: By explicitly keeping track of the populations of nodes with each of
1540: the different types of Boolean logic functions as links from frozen
1541: nodes are deleted, Kaufman, Mihaljev, and Drossel~\cite{Kaufman:05}
1542: also derive results for other quantities, such as the number of links
1543: in the sub-network of unfrozen nodes. The \EP\ formalism described
1544: above can be applied once again to investigate these additional
1545: quantities in a broader class of networks. Detailed results for \RBN
1546: s with various degree distributions will be presented elsewhere.
1547:
1548: \section{Summary and discussion}
1549:
1550: Unordered binary avalanches can in some cases lead to damage on every
1551: node or almost every node of a network, a phenomenon we have dubbed
1552: {\it exhaustive percolation}. We have studied a broad class of random
1553: networks that can exhibit \EP. We have shown how to calculate the
1554: probability $\Pex(N)$ that complete coverage occurs (i.e that all
1555: nodes are damaged) and also derived expressions for the probability
1556: distribution $P(u)$ of the number of undamaged nodes, $u$, in the
1557: large $N$ limit when \EP\ does occur. A logical curiosity in our
1558: approach is the fact that the calculation of $P(u)$ involves
1559: application of the $\Pex$ result to subnetworks containing candidate
1560: sets of damaged nodes.
1561:
1562: Our primary results flow from the realization that all of the relevant
1563: information about a \UBA\ defined on a random network is contained in
1564: the damage propagation function $g(x)$ or, equivalently, the damage
1565: control function $q(x)$. We derive scaling law exponents and exact
1566: results for the distribution of $u$ that are valid for a broad class
1567: of random networks and Boolean rule distributions in the \EP\ regime
1568: and for networks at the \EP\ critical point. This class includes the
1569: \UBA s that determine the set of frozen nodes in \RBN s with more than
1570: two inputs per node and therefore constitute a generalization of the
1571: results on the set of unfrozen nodes in \RBN s presented in
1572: Ref.~\cite{Kaufman:05}. Interestingly, the asymptotic behavior found
1573: in Ref.~\cite{Kaufman:05} for the distribution of $u$ at the critical
1574: point is shown to be valid for a broad class of network problems.
1575:
1576: For networks outside the above mentioned class but within the
1577: framework of \UBA, we find connections to previous work on
1578: Galton--Watson processes \cite{Otter:49} and random maps
1579: \cite{Harris:60}. The central result of our investigations is
1580: displayed in Eqs.~\eqref{eq: pu(u) asympt} and~\eqref{eq: p(ut)
1581: asympt}, which provide explicit formulas for the probability of
1582: finding $u$ undamaged nodes after an avalanche runs to completion.
1583: The out-degree distributions of the networks described by our formulas
1584: are all Poissonian, but the in-degree distributions may have different
1585: forms, including power laws, so long as the probability of having
1586: in-degree $K$ decays faster than $K^{-3}$. The exact nature of the
1587: \EP\ transition on networks with broader in-degree distributions is an
1588: interesting issue for future research. Further work is also needed to
1589: handle correlations between input links to different nodes, a
1590: situation that arises, for example, in random regular graphs or
1591: networks with scale free out-degree distributions.
1592:
1593: Our original motivation for studying \EP\ arose from attempts to
1594: understand the dynamical behavior of \RBN s. We have described one
1595: nontrivial example of how the \EP\ formalism is relevant: the
1596: calculation of the probability distribution for the number of unfrozen
1597: nodes in any \RBN\ with a rule distribution that leads to a given
1598: damage control function $q$ for the associated \UBA. The problem of
1599: determining how many of the unfrozen nodes are actually relevant for
1600: determining the attractor structure of the \RBN\ can also be framed as
1601: an \EP\ problem, which will be addressed in a separate publication.
1602:
1603: \section*{Acknowledgment} This work was supported by the National Science
1604: Foundation through Grant No.~PHY-0417372.
1605:
1606: \begin{thebibliography}{38}
1607: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1608: \expandafter\ifx\csname bibnamefont\endcsname\relax
1609: \def\bibnamefont#1{#1}\fi
1610: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1611: \def\bibfnamefont#1{#1}\fi
1612: \expandafter\ifx\csname citenamefont\endcsname\relax
1613: \def\citenamefont#1{#1}\fi
1614: \expandafter\ifx\csname url\endcsname\relax
1615: \def\url#1{\texttt{#1}}\fi
1616: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1617: \providecommand{\bibinfo}[2]{#2}
1618: \providecommand{\eprint}[2][]{\url{#2}}
1619:
1620: \bibitem[{\citenamefont{Kinney et~al.}(2005)\citenamefont{Kinney, Crucitti,
1621: Albert, and Latora}}]{Kinney:05}
1622: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Kinney}},
1623: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Crucitti}},
1624: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Albert}}, \bibnamefont{and}
1625: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Latora}},
1626: \bibinfo{journal}{Eur. Phys. J. B} \textbf{\bibinfo{volume}{46}},
1627: \bibinfo{pages}{101} (\bibinfo{year}{2005}).
1628:
1629: \bibitem[{\citenamefont{Sachtjen et~al.}(2000)\citenamefont{Sachtjen, Carreras,
1630: and Lynch}}]{Sachtjen:00}
1631: \bibinfo{author}{\bibfnamefont{M.~L.} \bibnamefont{Sachtjen}},
1632: \bibinfo{author}{\bibfnamefont{B.~A.} \bibnamefont{Carreras}},
1633: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{V.~E.} \bibnamefont{Lynch}},
1634: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{61}},
1635: \bibinfo{pages}{4877} (\bibinfo{year}{2000}).
1636:
1637: \bibitem[{\citenamefont{Newman}(2002)}]{Newman:02}
1638: \bibinfo{author}{\bibfnamefont{M.~E.~J.} \bibnamefont{Newman}},
1639: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{66}},
1640: \bibinfo{pages}{016128} (\bibinfo{year}{2002}).
1641:
1642: \bibitem[{\citenamefont{Cohen et~al.}(2003)\citenamefont{Cohen, Havlin, and
1643: ben{-}Avraham}}]{Cohen:03}
1644: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cohen}},
1645: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Havlin}}, \bibnamefont{and}
1646: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{ben{-}Avraham}},
1647: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{91}},
1648: \bibinfo{pages}{247901} (\bibinfo{year}{2003}).
1649:
1650: \bibitem[{\citenamefont{Hethcote}(2000)}]{Hethcote:00}
1651: \bibinfo{author}{\bibfnamefont{H.~W.} \bibnamefont{Hethcote}},
1652: \bibinfo{journal}{SIAM Rev.} \textbf{\bibinfo{volume}{42}},
1653: \bibinfo{pages}{599} (\bibinfo{year}{2000}).
1654:
1655: \bibitem[{\citenamefont{Pastor-Satorras and
1656: Vespignani}(2001)}]{Pastor-Satorras:01}
1657: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Pastor-Satorras}}
1658: \bibnamefont{and}
1659: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Vespignani}},
1660: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{86}},
1661: \bibinfo{pages}{3200} (\bibinfo{year}{2001}).
1662:
1663: \bibitem[{\citenamefont{Lloyd and May}(2001)}]{Lloyd:01}
1664: \bibinfo{author}{\bibfnamefont{A.~L.} \bibnamefont{Lloyd}} \bibnamefont{and}
1665: \bibinfo{author}{\bibfnamefont{R.~M.} \bibnamefont{May}},
1666: \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{292}},
1667: \bibinfo{pages}{1316} (\bibinfo{year}{2001}).
1668:
1669: \bibitem[{\citenamefont{Hughes et~al.}(2000)\citenamefont{Hughes, Marton,
1670: Jones, Roberts, Stoughton, Armour, Bennett, Coffey, Dai, He
1671: et~al.}}]{Hughes:00}
1672: \bibinfo{author}{\bibfnamefont{T.~R.} \bibnamefont{Hughes}},
1673: \bibinfo{author}{\bibfnamefont{M.~J.} \bibnamefont{Marton}},
1674: \bibinfo{author}{\bibfnamefont{A.~R.} \bibnamefont{Jones}},
1675: \bibinfo{author}{\bibfnamefont{C.~J.} \bibnamefont{Roberts}},
1676: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Stoughton}},
1677: \bibinfo{author}{\bibfnamefont{C.~D.} \bibnamefont{Armour}},
1678: \bibinfo{author}{\bibfnamefont{H.~A.} \bibnamefont{Bennett}},
1679: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Coffey}},
1680: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Dai}},
1681: \bibinfo{author}{\bibfnamefont{Y.~D.} \bibnamefont{He}},
1682: \bibnamefont{et~al.}, \bibinfo{journal}{Cell} \textbf{\bibinfo{volume}{102}},
1683: \bibinfo{pages}{109} (\bibinfo{year}{2000}).
1684:
1685: \bibitem[{\citenamefont{R{\"a}m{\"o} et~al.}(2006)\citenamefont{R{\"a}m{\"o},
1686: Kesseli, and Yli-Harja}}]{Ramo:06}
1687: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{R{\"a}m{\"o}}},
1688: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kesseli}}, \bibnamefont{and}
1689: \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Yli-Harja}}
1690: (\bibinfo{year}{2006}), \bibinfo{note}{to appear in J.\ Theor.\ Biol.,
1691: doi:10.1016/j.jtbi.2006.02.011}.
1692:
1693: \bibitem[{\citenamefont{Newman}(2003)}]{Newman:03}
1694: \bibinfo{author}{\bibfnamefont{M.~E.~J.} \bibnamefont{Newman}},
1695: \bibinfo{journal}{SIAM Rev.} \textbf{\bibinfo{volume}{45}},
1696: \bibinfo{pages}{167} (\bibinfo{year}{2003}).
1697:
1698: \bibitem[{\citenamefont{Callaway et~al.}(2000)\citenamefont{Callaway, Newman,
1699: Strogatz, and Watts}}]{Callaway:00}
1700: \bibinfo{author}{\bibfnamefont{D.~S.} \bibnamefont{Callaway}},
1701: \bibinfo{author}{\bibfnamefont{M.~E.~J.} \bibnamefont{Newman}},
1702: \bibinfo{author}{\bibfnamefont{S.~H.} \bibnamefont{Strogatz}},
1703: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Watts}},
1704: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{85}},
1705: \bibinfo{pages}{5468} (\bibinfo{year}{2000}).
1706:
1707: \bibitem[{\citenamefont{Moreno et~al.}(2002)\citenamefont{Moreno, G{\'o}mez,
1708: and Pacheco}}]{Moreno:02}
1709: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Moreno}},
1710: \bibinfo{author}{\bibfnamefont{J.~B.} \bibnamefont{G{\'o}mez}},
1711: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.~F.}
1712: \bibnamefont{Pacheco}}, \bibinfo{journal}{Europhys. Lett.}
1713: \textbf{\bibinfo{volume}{58}}, \bibinfo{pages}{630} (\bibinfo{year}{2002}).
1714:
1715: \bibitem[{\citenamefont{Watts}(2002)}]{Watts:02}
1716: \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Watts}},
1717: \bibinfo{journal}{Proc. Natl. Acad. Sci. USA} \textbf{\bibinfo{volume}{99}},
1718: \bibinfo{pages}{5766} (\bibinfo{year}{2002}).
1719:
1720: \bibitem[{\citenamefont{Motter}(2004)}]{Motter:04}
1721: \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Motter}},
1722: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{93}},
1723: \bibinfo{pages}{098701} (\bibinfo{year}{2004}).
1724:
1725: \bibitem[{\citenamefont{Crucitti et~al.}(2004)\citenamefont{Crucitti, Latora,
1726: and Marchiori}}]{Crucitti:04}
1727: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Crucitti}},
1728: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Latora}}, \bibnamefont{and}
1729: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Marchiori}},
1730: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{69}},
1731: \bibinfo{pages}{045104(R)} (\bibinfo{year}{2004}).
1732:
1733: \bibitem[{\citenamefont{Watts}(1999)}]{Watts:99}
1734: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Watts}},
1735: \emph{\bibinfo{title}{Small worlds: the dynamics of networks between order
1736: and randomness}} (\bibinfo{publisher}{Princeton University Press},
1737: \bibinfo{year}{1999}).
1738:
1739: \bibitem[{\citenamefont{Kauffman}(1969)}]{Kauffman:69}
1740: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Kauffman}},
1741: \bibinfo{journal}{J.\ Theor.\ Biol.} \textbf{\bibinfo{volume}{22}},
1742: \bibinfo{pages}{437} (\bibinfo{year}{1969}).
1743:
1744: \bibitem[{\citenamefont{Derrida and Pomeau}(1986)}]{Derrida:86a}
1745: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Derrida}} \bibnamefont{and}
1746: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Pomeau}},
1747: \bibinfo{journal}{Europhys.\ Lett.} \textbf{\bibinfo{volume}{1}},
1748: \bibinfo{pages}{45} (\bibinfo{year}{1986}).
1749:
1750: \bibitem[{\citenamefont{Bastolla and Parisi}(1997)}]{Bastolla:97}
1751: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Bastolla}} \bibnamefont{and}
1752: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Parisi}},
1753: \bibinfo{journal}{J.\ Theor.\ Biol.} \textbf{\bibinfo{volume}{187}},
1754: \bibinfo{pages}{117} (\bibinfo{year}{1997}).
1755:
1756: \bibitem[{\citenamefont{Aldana-Gonzalez
1757: et~al.}(2003)\citenamefont{Aldana-Gonzalez, Coppersmith, and
1758: Kadanoff}}]{Aldana:03a}
1759: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Aldana-Gonzalez}},
1760: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Coppersmith}},
1761: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.~P.}
1762: \bibnamefont{Kadanoff}}, \emph{\bibinfo{title}{Boolean Dynamics with Random
1763: Couplings{\normalfont, in} {P}erspectives and Problems in Nonlinear
1764: Science{\normalfont, edited by E. Kaplan, J. E. Marsden and K. R.
1765: Sreenivasan}}} (\bibinfo{publisher}{Springer}, \bibinfo{year}{2003}),
1766: p.~\bibinfo{pages}{23}.
1767:
1768: \bibitem[{\citenamefont{Aldana and Cluzel}(2003)}]{Aldana:03b}
1769: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Aldana}} \bibnamefont{and}
1770: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Cluzel}},
1771: \bibinfo{journal}{Proc. Natl. Acad. Sci. USA} \textbf{\bibinfo{volume}{100}},
1772: \bibinfo{pages}{8710} (\bibinfo{year}{2003}).
1773:
1774: \bibitem[{\citenamefont{Samuelsson and Troein}(2003)}]{Samuelsson:03}
1775: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Samuelsson}} \bibnamefont{and}
1776: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Troein}},
1777: \bibinfo{journal}{Phys.\ Rev.\ Lett.} \textbf{\bibinfo{volume}{90}},
1778: \bibinfo{pages}{098701} (\bibinfo{year}{2003}).
1779:
1780: \bibitem[{\citenamefont{Kauffman et~al.}(2004)\citenamefont{Kauffman, Peterson,
1781: Samuelsson, and Troein}}]{Kauffman:04}
1782: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Kauffman}},
1783: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Peterson}},
1784: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Samuelsson}},
1785: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Troein}},
1786: \bibinfo{journal}{Proc.\ Natl.\ Acad.\ Sci.\ USA}
1787: \textbf{\bibinfo{volume}{101}}, \bibinfo{pages}{17102}
1788: (\bibinfo{year}{2004}).
1789:
1790: \bibitem[{\citenamefont{Drossel}(2005)}]{Drossel:05b}
1791: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Drossel}},
1792: \bibinfo{journal}{Phys.\ Rev.\ E} \textbf{\bibinfo{volume}{72}},
1793: \bibinfo{pages}{016110} (\bibinfo{year}{2005}).
1794:
1795: \bibitem[{\citenamefont{Socolar and Kauffman}(2003)}]{Socolar:03}
1796: \bibinfo{author}{\bibfnamefont{J.~E.~S.} \bibnamefont{Socolar}}
1797: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~A.}
1798: \bibnamefont{Kauffman}}, \bibinfo{journal}{Phys. Rev. Lett.}
1799: \textbf{\bibinfo{volume}{90}}, \bibinfo{pages}{068702}
1800: (\bibinfo{year}{2003}).
1801:
1802: \bibitem[{\citenamefont{Flyvbjerg}(1988)}]{Flyvbjerg:88b}
1803: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Flyvbjerg}},
1804: \bibinfo{journal}{J.\ Phys.\ A: Math.\ Gen.} \textbf{\bibinfo{volume}{21}},
1805: \bibinfo{pages}{L955} (\bibinfo{year}{1988}).
1806:
1807: \bibitem[{\citenamefont{Bilke and Sjunnesson}(2002)}]{Bilke:01}
1808: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Bilke}} \bibnamefont{and}
1809: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sjunnesson}},
1810: \bibinfo{journal}{Phys.\ Rev.\ E} \textbf{\bibinfo{volume}{65}},
1811: \bibinfo{pages}{016129} (\bibinfo{year}{2002}).
1812:
1813: \bibitem[{\citenamefont{Bastolla and Parisi}(1998)}]{Bastolla:98a}
1814: \bibinfo{author}{\bibfnamefont{U.}~\bibnamefont{Bastolla}} \bibnamefont{and}
1815: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Parisi}},
1816: \bibinfo{journal}{Physica D} \textbf{\bibinfo{volume}{115}},
1817: \bibinfo{pages}{203} (\bibinfo{year}{1998}).
1818:
1819: \bibitem[{\citenamefont{Kaufman et~al.}(2005)\citenamefont{Kaufman, Mihaljev,
1820: and Drossel}}]{Kaufman:05}
1821: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Kaufman}},
1822: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Mihaljev}}, \bibnamefont{and}
1823: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Drossel}},
1824: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{72}},
1825: \bibinfo{pages}{046124} (\bibinfo{year}{2005}).
1826:
1827: \bibitem[{\citenamefont{Sahimi}(2003)}]{Sahimi:03}
1828: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Sahimi}},
1829: \emph{\bibinfo{title}{Heterogenous Materials I}}
1830: (\bibinfo{publisher}{Springer}, \bibinfo{year}{2003}).
1831:
1832: \bibitem[{\citenamefont{Harris}(1963)}]{Harris:63}
1833: \bibinfo{author}{\bibfnamefont{T.~E.} \bibnamefont{Harris}},
1834: \emph{\bibinfo{title}{The Theory of Branching Processes}}
1835: (\bibinfo{publisher}{Springer Verlag}, \bibinfo{year}{1963}).
1836:
1837: \bibitem[{\citenamefont{Otter}(1949)}]{Otter:49}
1838: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Otter}}, \bibinfo{journal}{Ann.
1839: Math. Stat.} \textbf{\bibinfo{volume}{20}}, \bibinfo{pages}{206}
1840: (\bibinfo{year}{1949}).
1841:
1842: \bibitem[{\citenamefont{Essam et~al.}(1988)\citenamefont{Essam, Guttmann, and
1843: De'Bell}}]{Essam:88}
1844: \bibinfo{author}{\bibfnamefont{J.~W.} \bibnamefont{Essam}},
1845: \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Guttmann}},
1846: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{De'Bell}},
1847: \bibinfo{journal}{J.\ Phys.\ A: Math.\ Gen.} \textbf{\bibinfo{volume}{21}},
1848: \bibinfo{pages}{3815} (\bibinfo{year}{1988}).
1849:
1850: \bibitem[{\citenamefont{Bollab\'as}(1985)}]{Bollabas:85}
1851: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Bollab\'as}},
1852: \emph{\bibinfo{title}{Random Graphs}} (\bibinfo{publisher}{Academic Press,
1853: London}, \bibinfo{year}{1985}).
1854:
1855: \bibitem[{\citenamefont{Press et~al.}(1992)\citenamefont{Press, Teukolsky,
1856: Vetterling, and Flannery}}]{numrecip}
1857: \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Press}},
1858: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Teukolsky}},
1859: \bibinfo{author}{\bibfnamefont{W.~T.} \bibnamefont{Vetterling}},
1860: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{B.~P.}
1861: \bibnamefont{Flannery}}, \emph{\bibinfo{title}{Numerical Recipes in C, second
1862: ed.}} (\bibinfo{publisher}{Cambridge University Press},
1863: \bibinfo{year}{1992}).
1864:
1865: \bibitem[{\citenamefont{Harris}(1960)}]{Harris:60}
1866: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Harris}},
1867: \bibinfo{journal}{Ann.\ Math.\ Statist.} \textbf{\bibinfo{volume}{31}},
1868: \bibinfo{pages}{1045} (\bibinfo{year}{1960}).
1869:
1870: \bibitem[{\citenamefont{Rubin and Sitgreaves}(1954)}]{Rubin:54}
1871: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Rubin}} \bibnamefont{and}
1872: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Sitgreaves}}
1873: (\bibinfo{year}{1954}), \emph{\bibinfo{title}{Probability distributions
1874: related to random transformations on a finite set.\ Tech.\ Rep.\ 19A}},
1875: \bibinfo{note}{Applied Mathematics and Statistics Laboratory,
1876: Stanford University}.
1877:
1878: \bibitem[{\citenamefont{Dwass}(1969)}]{Dwass:69}
1879: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Dwass}}, \bibinfo{journal}{J.
1880: Appl. Prob.} \textbf{\bibinfo{volume}{6}}, \bibinfo{pages}{682}
1881: (\bibinfo{year}{1969}).
1882:
1883: \end{thebibliography}
1884:
1885: \appendix
1886:
1887: \section{Calculation of the damage control function}
1888: \label{app: q calc}
1889:
1890: Let $p_K$ denote the probability that a rule has $K$ inputs, and let
1891: $P_0(K,m)$ denote the probability that the output value is zero of a
1892: rule with $K$ inputs fed with $m$ zeros and $K-m$ ones. Then, the
1893: the damage control function is
1894: \begin{align}
1895: q(x) =\,&\sum_{K=0}^\infty p_K\sum^K_{m=0}P_0(K,m) \binom K m
1896: x^m(1-x)^{K-m}.
1897: \label{eq: q explicit 0}
1898: \end{align}
1899: Eq.~\eqref{eq: q explicit 0} can be written as
1900: \begin{align}
1901: q(x) &= a_0 + a_1x + a_2x^2 + \cdots
1902: \label{eq: q explicit 1}
1903: \end{align}
1904: where
1905: \begin{align}
1906: a_i = \sum_{K=i}^\infty\sum_{m=0}^i p_KP_0(K,m)
1907: (-1)^{i-m}\binom K m\binom{K-m}{i-m}.
1908: \label{eq: q expansion terms}
1909: \end{align}
1910: The expansion in Eq.~\eqref{eq: q explicit 1} is well-defined up to
1911: the first term such that the sum in Eq.~\eqref{eq: q expansion terms}
1912: is not absolute convergent. The factor $\binom K m\binom{K-m}{i-m}$
1913: scales like $K^i$ for large $K$ and $P_0(K,m)\le 1$. Hence, $a_i$ is
1914: well-defined if $\sum_{K=0}^\infty K^ip_K$ is convergent and this is
1915: true if $p_K$ decays faster than $K^{-i-1}$.
1916:
1917: In addition, the requirement that the output of each rule in the rule
1918: distribution is 1 if all of its inputs have the value 1, yields that
1919: $a_0=0$. Thus, the expansion
1920: \begin{align}
1921: q(x) &= \alpha_1x - \alpha_2x^2 + \Oc(x^3),
1922: \label{eq: q-Taylor app}
1923: \end{align}
1924: is valid for all rule distributions such that $p_K$ decays faster than
1925: $K^{-4}$. In the case that $p_K$ decays slower than $K^{-4}$ but
1926: faster than $K^{-3}$, only the residue term can be affected.
1927:
1928: \section{Probability for complete coverage}
1929: \label{app: PEP}
1930:
1931: Here, we assume that the expansion in Eq.~\eqref{eq: q-Taylor app} is
1932: well-defined. Then, we get
1933: \begin{align}
1934: \frac{\partial p}{\partial n_\msia}
1935: &= \frac{c}{2\alpha_1Nn_\msia^2}\frac{\partial^2 p}{\partial c^2}
1936: [1+\Oc(n_\msia/N)],
1937: \label{eq: p-diffuse n c}
1938: \intertext{and}
1939: c_{\trm{max}}(x)/N &= \frac{1}{\alpha_1}
1940: + \frac{\alpha_2}{\alpha_1^2}x
1941: + \Oc(x^2).
1942: \end{align}
1943:
1944: To remove the dependence of $n_\msia$ from the leading order term of
1945: the diffusion rate in Eq.~\eqref{eq: p-diffuse n c}, we let
1946: $t=-1/n_\msia$. By also letting $y=1-\alpha_1c/N$, we rewrite
1947: Eq.~\eqref{eq: p-diffuse n c} to a form that easily can be rescaled
1948: as $N$ grows. We get
1949: \begin{align}
1950: \frac{\partial p}{\partial t} &= \frac{1-y}{2}
1951: \frac{\partial^2 p}{\partial y^2}[1+{\cal O}(\tfrac1{Nt})].
1952: \intertext{and}
1953: y_{\trm{min}} &= -\frac{\alpha_2}{\alpha_1Nt}[1+\Oc(\tfrac1{Nt})],
1954: \label{eq: ymin}
1955: \end{align}
1956: where $y=y_{\trm{min}}$ is the transformed value of $c_{\trm{max}}$.
1957: The boundary conditions are $p=0$ for $y=y_{\trm{min}}$ and $p=1$
1958: for $y=1$.
1959:
1960: The $N$ dependence of the leading order term of the boundary condition
1961: in Eq.~\eqref{eq: ymin} can be removed by rescaling of $y$ and $t$.
1962: Typically, $\alpha_2>0$ and this is the case that we will focus on.
1963: [Note that Eq.~\eqref{eq: x0 gt} means that $\alpha_2$ must be
1964: nonnegative at the transition.] If $\alpha_2=0$, either $q(x)=x$ or
1965: $q(x)=x-\alpha_m x^m+\cdots$ with $m>2$ (apart from some pathological
1966: special cases). The first case, $q(x)=x$, is a special case that is
1967: convenient for analytic calculation, whereas the latter case require
1968: calculations analogous to the calculations for $\alpha_2>0$. We will
1969: come back to the case $q(x)=x$.
1970:
1971: For $\alpha_2>0$, we rescale $y$ and $t$ according to
1972: \begin{align}
1973: \yt &= \Nt^{1/3}y
1974: \intertext{and}
1975: \tti&= \Nt^{2/3}t,
1976: \intertext{where}
1977: \Nt &= \frac{\alpha_1}{\alpha_2}N.
1978: \label{eq: Nt}
1979: \end{align}
1980: Then,
1981: \begin{align}
1982: \frac{\partial p}{\partial\tti} &= \frac12
1983: \frac{\partial^2p}{\partial\yt^2}
1984: \bigl(1+\Nt^{-1/3}\yt\bigr)
1985: \bigl[1+{\cal O}\bigl(\tfrac1{\Nt^{1/3}\tti}\bigr)\bigr],
1986: \label{eq: p diff tt yt}
1987: \intertext{where}
1988: \yt_{\trm{min}} &= \tti^{-1}
1989: \bigl[1+{\cal O}\bigl(\tfrac1{\Nt^{1/3}\tti}\bigr)\bigr].
1990: \end{align}
1991:
1992: The boundary conditions are $p=0$ for $\yt_{\trm{min}}$ and $p=1$ for
1993: $y=\Nt^{1/3}$. The only plausible limit of $p$ as
1994: $t\rightarrow-\infty$ is $p=y/\Nt^{1/3}$. To get a motivation that is
1995: mathematically acceptable, one needs to relate the original integer
1996: based formulation of the problem in Eqs.\ \eqref{eq: n-update
1997: first}--\eqref{eq: n-update last}. The large $N$ behavior of
1998: Eq.~\eqref{eq: n-update last},
1999: \begin{align}
2000: \lim_{N\rightarrow\infty}U(\s,j) &= 1/n_\msia,
2001: \label{eq: a lim}
2002: \end{align}
2003: yields
2004: \begin{align}
2005: \lim_{N\rightarrow\infty}\Pex(N,q;n_0,n_\msa)
2006: &= \frac{n_\msa}{n_\msia}.
2007: \label{eq: Pex n fixed N large}
2008: \end{align}
2009: Eq.~\eqref{eq: Pex n fixed N large} can be shown via induction. The
2010: induction is initiated by
2011: \begin{align}
2012: \lim_{N\rightarrow\infty}\Pex(N,q;1,0)&=0
2013: \intertext{and}
2014: \lim_{N\rightarrow\infty}\Pex(N,q;0,1)&=1,
2015: \end{align}
2016: which means that Eq.~\eqref{eq: Pex n fixed N large} is true for
2017: $n_\msia=1$. To obtain the induction step, we assume that
2018: Eq.~\eqref{eq: Pex n fixed N large} is true for
2019: $n'_\msia=n_\msia-1$. Then, we get
2020: \begin{align}
2021: \lim_{N\rightarrow\infty}\Pex(N,q;n'_0,n'_\msa)
2022: &= \frac{n'_\msa}{n_\msia-1}
2023: \end{align}
2024: which leads to
2025: \begin{align}
2026: \lim_{N\rightarrow\infty}\Pex(N,q;n_0,n_\msa)
2027: &= \frac{\langle n'_\msa\rangle}{n_\msia-1}\\
2028: &= \frac{n_\msa + n_0/n_\msia-1}{n_\msia-1}\\
2029: &= \frac{n_\msa}{n_\msia}
2030: \end{align}
2031: that completes the induction step. Eq.~\eqref{eq: Pex n fixed N large}
2032: means that the value of $p$ approaches a linear function of $\yt$ for
2033: $\tti=N^{2/3}/n_0$ as $N\rightarrow\infty$. Hence, the boundary
2034: condition for $t\rightarrow-\infty$ is $p=y/\Nt^{1/3}$.
2035:
2036: Rescaling of $p$ according to
2037: \begin{align}
2038: \pt &= \Nt^{1/3}p
2039: \end{align}
2040: gives the boundary condition
2041: \begin{align}
2042: \lim_{t\rightarrow-\infty}\pt = \yt.
2043: \label{eq: p t neg inf}
2044: \end{align}
2045: If Eq.~\eqref{eq: p t neg inf} is extended to be valid for all
2046: non-negative $\yt$, the boundary condition at $\yt=\Nt^{1/3}$ can be
2047: dropped. In the limit of large $N$, Eq.~\eqref{eq: p diff tt yt}
2048: becomes
2049: \begin{align}
2050: \frac{\partial\pt}{\partial\tti} &= \frac12
2051: \frac{\partial^2\pt}{\partial\yt^2}.
2052: \label{eq: app diffuse}
2053: \end{align}
2054: With $\pt$ is written on the form $\pt(\tti,\yt)$, the boundary
2055: conditions are
2056: \begin{align}
2057: \pt(\tti,1/\tti) &= 0 & \trm{for }&\tti<0
2058: \label{eq: app diffuse bound 0}
2059: \intertext{and}
2060: \lim_{\tti\rightarrow-\infty}\pt(\tti,\yt) &= \yt
2061: & \trm{for }&\yt\ge0,
2062: \label{eq: app diffuse bound 1}
2063: \end{align}
2064: as $N\rightarrow\infty$.
2065:
2066: The solution to Eqs.~\eqref{eq: app diffuse}--\eqref{eq: app diffuse bound 1}
2067: can be calculated numerically. By expressing the transformed variables
2068: $\tti$ and $\yt$ in terms of more fundamental quantities, we get
2069: \begin{align}
2070: \Pex&(N,q; n_0,n_\msa) \nonumber\\
2071: &\approx \Nt^{-1/3}\pt\biggl[-\frac{\Nt^{2/3}}{n_\msia},
2072: \Nt^{1/3}\biggl(1-\frac{\alpha_1n_0}
2073: {Nq(n_\msia/N)}\biggr)\biggr]
2074: \label{eq: Ploopfree rescaled}
2075: \end{align}
2076: and
2077: \begin{align}
2078: \Pex&(N,q; N-n_\msa,n_\msa) \nonumber\\
2079: &\approx \Nt^{-1/3}\pt\biggl[-\frac{\Nt^{2/3}}{N},
2080: \Nt^{1/3}\biggl(1-\frac{\alpha_1n_0}{Nq(1)}\biggr)\biggr]
2081: \end{align}
2082: for large $N$.
2083:
2084: If the avalanche is initiated by letting each node start from 0 with
2085: probability $q(1)$, we get
2086: \begin{align}
2087: \langle n_\msa^\ti\rangle &= N[1-q(1)]
2088: \intertext{and}
2089: \sigma(n_\msa^\ti) &= \sqrt{Nq(1)[1-q(1)]}.
2090: \end{align}
2091:
2092: Provided that $q(x)$ does not depend on $N$, $\Nt$ is fixed and the
2093: spread in $\yt$ that correspond to the initial value of $m$ will go to
2094: zero as $N\rightarrow\infty$. Also, $\Nt^{2/3}/N$ approaches zero as
2095: $N\rightarrow\infty$. In this case, the probability for an avalanche
2096: to yield complete coverage is given by
2097: \begin{align}
2098: \Pex(N, q) &\approx \Nt^{-1/3}\pt[0,\Nt^{1/3}(1-\alpha_1)],
2099: \label{eq: app PexNq}
2100: \end{align}
2101: for large $N$. Only the first two arguments to $\Pex$ are kept in
2102: Eq.~\eqref{eq: app PexNq}, because the process is fully determined by $N$
2103: and $q$.
2104:
2105: In the special case that $q(x)=x$ for all $x\in[0,1]$, Eq.~\eqref{eq:
2106: n-update last} yields $a=1/n_\msia$, which is a strong form of
2107: Eq.~\eqref{eq: a lim}. By using the same induction steps that lead
2108: from Eq.~\eqref{eq: a lim} to Eq.~\eqref{eq: Pex n fixed N large}, we
2109: conclude that
2110: \begin{align}
2111: \Pex(N,x\mapsto x;n_0,n_\msa)
2112: &= \frac{n_\msa}{n_\msia}.
2113: \label{eq: q-ident}
2114: \end{align}
2115:
2116: \section{Asymptotes for sparse percolation}
2117: \label{app: SP}
2118:
2119: Provided that the derivative of $g(x)$ is well defined at $x=0$, we
2120: let $\lambda = g'(0)$, where $g'(x)$ denotes the derivative of
2121: $g(x)$. Then,
2122: \begin{align}
2123: \lim_{N\rightarrow\infty} g^1_{N,n}(x) &= x,
2124: \intertext{which means that}
2125: \lim_{N\rightarrow\infty} q^1_{N,N-n}(x) &= x
2126: \intertext{and}
2127: \lim_{N\rightarrow\infty} \PNex(n,\ell;N) &= \frac\ell n
2128: \end{align}
2129: according to Eq.~\eqref{eq: q-ident}. Thus, the large $N$ limit of
2130: Eq.~\eqref{eq: Pn1} is
2131: \begin{align}
2132: \lim_{N\rightarrow\infty}\Pn(n)
2133: &= \rho(n)\tau(n,\ell)\lambda^{n-\ell}e^{n(1-\lambda)}\frac\ell n\\
2134: &= \frac{\ell(n\lambda)^{n-\ell}}{n(n-\ell)!}e^{-n\lambda}.
2135: \label{eq: lim Pn}
2136: \end{align}
2137: For large $N$, Eq.~\eqref{eq: lim Pn} yields
2138: \begin{align}
2139: \lim_{N\rightarrow\infty}\Pn(n)
2140: &\approx \frac \ell{\sqrt{2\pi}}e^{n(1-\lambda)}\lambda^{n-\ell}n^{-3/2}.
2141: \end{align}
2142:
2143: Due to the correspondence to well investigated branching processes,
2144: Eq.~\eqref{eq: lim Pn} is not a new result. For the special case of
2145: $\ell=1$, Eq.~\eqref{eq: lim Pn} is given explicitly in
2146: Ref.~\cite{Otter:49}, and the general form of Eq.~\eqref{eq: lim Pn}
2147: can easily be obtained by the theorem presented in
2148: Ref.~\cite{Dwass:69}.
2149:
2150: \section{Asymptotes for exhaustive percolation}
2151: \label{app: EP asympt}
2152:
2153: In analogy with our investigation of \SP, we assume that $q(x)$ has a
2154: well-defined derivative at $x=0$ and let $\lambda = q'(0)$. For \EP\ to
2155: be likely in the large $N$ limit, it is required that $q(x)\leq x$ for
2156: all $x$, meaning that $\lambda\le1$. The large $N$ behavior of
2157: Eq.~\eqref{eq: Pu0} is partly explained by
2158: \begin{align}
2159: \lim_{N\rightarrow\infty} \frac{\rho(u)\rho(N-u)}{\rho(N)}[Q(u/N)]^N
2160: &= \rho(u)\lambda^ue^{u(1-\lambda)},
2161: \label{eq: lim EP}
2162: \end{align}
2163: but it remains to investigate the role of $\PNex(N-u,0;N)$. To this end,
2164: we consider the ratio $\PNex(N-u,0;N)/\Pex(N)$. [Here, we have dropped the
2165: argument $q$ from $\Pex(N, q)$.]
2166:
2167: When $N\rightarrow\infty$, there are two processes that influence on
2168: this ratio: $q^1_{N,u}$ approaches $q$ and $N$ increases. The increase
2169: of $N$ makes the involved probabilities more sensitive for the
2170: shrinking differences between $q^1_{N,u}$ and $q$. Thus, there are two
2171: competing processes as $N\rightarrow\infty$. The sensitivity with
2172: respect to $q$ is limited by the variance of the number of nodes with
2173: initial state \tsa, because this variance can be seen as a rescaling
2174: of $q$. The change in $q(x)$ by such a rescaling scales like
2175: $q(x)/\sqrt{N}$ for large $N$. If $q(x)$ has a well-defined nonzero
2176: derivative at $x=0$, the difference $q^1_{N,u}(x)-q(x)$ scales like
2177: $q(x)/N$ for large $N$. Hence, the decrease in the difference between
2178: $q^1_{N,u}$ and $q$ dominates over the increase in sensitivity,
2179: meaning that
2180: \begin{align}
2181: \lim_{N\rightarrow\infty} \frac{\PNex(N-u,0;N)}{\Pex(N)}
2182: &= 1.
2183: \end{align}
2184: Thus,
2185: \begin{align}
2186: \lim_{N\rightarrow\infty}\frac{\Pu(u)}{\Pex(N)}
2187: &= \rho(u)\lambda^ue^{u(1-\lambda)}\\
2188: &= \frac{(u\lambda)^{u}}{u!}e^{-u\lambda},
2189: \label{eq: Pu1}
2190: \end{align}
2191: where $(u\lambda)^u$ should be interpreted with the convention that
2192: $0^0=1$ in order to handle the case $u = 0$ properly.
2193:
2194: \subsection{Limit distributions for supercritical EP}
2195: \label{app: super EP}
2196:
2197: If $0<\lambda<1$ and $x=0$ is the only solution to $q(x)=x$ in the
2198: interval $0\le x\le1$, the exponential decay of $[Q(u/N)]^N$, in
2199: Eq.~\eqref{eq: Pu0}, with increasing $u$ ensures that
2200: \begin{align}
2201: \sum_{u=0}^\infty\lim_{N\rightarrow\infty}\Pu(u) &= 1
2202: \intertext{and}
2203: \Bigl[\lim_{N\rightarrow\infty}\Pex(N)\Bigr]\sum_{u=0}^\infty
2204: \frac{(u\lambda)^{u}}{u!}e^{-u\lambda} &= 1.
2205: \end{align}
2206: Thus, $\lim_{N\rightarrow\infty}$ has a unique value for each
2207: $\lambda$. This value can be calculated by considering the simplest
2208: case, $q(x)=\lambda x$. From the definition of the spreading
2209: process, we get
2210: \begin{align}
2211: \Pex(N,x\mapsto\lambda x;n_0,n_\msa) &= \Pex(N,x\mapsto x;n_0,n_\msa).
2212: \end{align}
2213: Then, Eq.~\eqref{eq: q-ident} and averaging over initial configurations
2214: yield
2215: \begin{align}
2216: \Pex(N,x\mapsto\lambda x) &= 1-\lambda,
2217: \end{align}
2218: which means that
2219: \begin{align}
2220: \lim_{N\rightarrow\infty}\Pex(N, q) &= 1-\lambda~
2221: \end{align}
2222: for all $q$ that satisfy the above mentioned criteria. We get,
2223: \begin{align}
2224: \lim_{N\rightarrow\infty}\Pu(u)
2225: &= (1-\lambda)\frac{(u\lambda)^{u}}{u!}e^{-u\lambda},
2226: \label{eq: P(u) asympt}
2227: \intertext{which for large $u$ means that}
2228: \lim_{N\rightarrow\infty}\Pu(u)
2229: &\approx\frac{1-\lambda}{\sqrt{2\pi}}e^{u(1-\lambda)}
2230: \lambda^uu^{-1/2}.
2231: \end{align}
2232:
2233: \subsection{Scaling at the EP transition}
2234: \label{app: crit EP}
2235:
2236: This section aims to derive the asymptotic distribution of $u$ for
2237: large $N$ for critical \EP\ with $\alpha_2>0$ and $\ell=0$. Define
2238: $\alpha_1^1$, $\alpha_2^1$ and $\Nt^1$ analogous to the definitions of
2239: $\alpha_1$, $\alpha_2$ and $\Nt$ in Eqs.~\eqref{eq: q-Taylor} and
2240: \eqref{eq: Nt}. The derivatives of $q^1(x)$ in Eq.~\eqref{eq: q1} at
2241: $x=0$ are given by
2242: \begin{align}
2243: (q^1)'(0) &= q'(u/N)\frac{1-u/N}{1-q(u/N)}
2244: \intertext{and}
2245: (q^1)''(0) &= q''(u/N)\frac{(1-u/N)^2}{1-q(u/N)}
2246: \end{align}
2247: and we get
2248: \begin{align}
2249: \Nt^1 &=\Nt+\Oc(u/N)
2250: \intertext{and}
2251: \alpha_1^1 &= \alpha_1 + (\alpha_1-1-2\alpha_2)u/N
2252: +\Oc(u^2/N^2).
2253: \end{align}
2254: For critical networks, with $\alpha_1=1$, we get $\Nt = N/\alpha_2$ and
2255: \begin{align}
2256: \alpha_1^1 &= 1 - 2u/\Nt + \Oc(u^2/N^2).
2257: \end{align}
2258: Insertion into Eq.~\eqref{eq: app PexNq} yields
2259: \begin{align}
2260: \PNex(N-u,0;N)
2261: &\approx \Nt^{-1/3}\pt(0,2\Nt^{-2/3}u)
2262: \end{align}
2263: for $u\ll N$.
2264:
2265: Because $Q(x)\leq1$ with equality if and only if $q(x)=x$, $\Pu$
2266: that vanishes exponentially, as $N$ goes to infinity, for any fixed
2267: $u/N$ such that $q(u/N)\ne u/N$. For a typical network with
2268: $\alpha_2>0$ at the \EP\ transition, the only solution to $q(x)=x$ is
2269: $x=0$. For such a network, the large $N$ behavior of $\Pu$ is found
2270: by expanding $Q(x)$ around $x=0$. To the leading
2271: non-vanishing order, we get
2272: \begin{align}
2273: Q(x) &\approx \exp\biggl(-\frac{\alpha_2^2x^3}2\biggr),
2274: \end{align}
2275: which yields
2276: \begin{align}
2277: \Pu(u)
2278: &\approx \rho(u)\exp\biggl(-\frac{u^3}{2\Nt^2}\biggr)
2279: \PNex.
2280: \end{align}
2281:
2282: Hence,
2283: \begin{align}
2284: \Pu(u) &\approx \Nt^{-1/3}\rho(u)
2285: \exp\biggl(-\frac{u^3}{2\Nt^2}\biggr)
2286: \pt(0,2\Nt^{-2/3}u),
2287: \end{align}
2288: with asymptotic equality for large $N$. The probability density,
2289: $P(\ut)$, for the distribution of $\ut=\Nt^{-2/3}u$ as
2290: $N\rightarrow\infty$ approaches
2291: \begin{align}
2292: P(\ut) &= \frac{\exp(-\frac12\ut^3)}{\sqrt{2\pi\ut}}
2293: \,\pt(0,2\ut).
2294: %\label{eq: p(ut) asympt}
2295: \end{align}\\
2296:
2297: \section{Exact results}
2298: \label{app: exact}
2299:
2300: A network with $g(x)=x$ is critical for all $x$. For such a network,
2301: $G(x)=1$ for all $x$ and $\PNex(n,\ell;N)=\ell/n$. Hence, Eq.~\eqref{eq: Pn1}
2302: yields
2303: \begin{align}
2304: \Pn(n)
2305: &= \frac{\rho(n)\rho(N-n)}{\rho(N)}\frac{\tau(n, \ell)}
2306: {\tau(N, \ell)}\frac\ell n\\
2307: &= \frac\ell n\binom{N-\ell}{n-\ell}\frac{n^{n-\ell}
2308: (N-n)^{N-n}}{N^{N-\ell}}.
2309: \label{eq: Pn ident}
2310: \end{align}
2311: For $n$ and $N$ satisfying $n\gg1$ and $N-n\gg1$, we get
2312: \begin{align}
2313: \Pn(n) &\approx \frac \ell{\sqrt{2\pi}}\frac{\sqrt{N}}{n\sqrt{n(N-n)}}\,.
2314: \end{align}
2315:
2316: In the special case $\ell=1$, $\Pn(n)$ is the distribution of the number
2317: of predecessors to an element in a random map. This distribution,
2318: which is consistent with eq.~\eqref{eq: Pn ident} for $\ell=1$, was
2319: obtained in Ref.\ \cite{Rubin:54} and restated in Ref.\
2320: \cite{Harris:60}.
2321:
2322: For completeness, we provide an explicit expression for the
2323: distribution of avalanche sizes in the case that $g(x)$ is a first
2324: order polynomial. From Eq.~\eqref{eq: Pn1}, we get
2325: \begin{align}
2326: \Pn(n)
2327: =\,& \Pu(u) \nonumber\\
2328: =\,&
2329: \frac{n^{n-\ell}u^u}{N^{N-\ell}}\binom{N-\ell}{n-\ell}
2330: \biggl(g(0)+\frac\ell n\biggr)
2331: \nonumber\\&\times
2332: \biggl(\frac{ug(0)+ng(1)}n\biggr)^{\!n-\ell}
2333: \biggl(\frac{nq(0)+uq(1)}u\biggr)^{\!u}.
2334: \label{eq: explicit Pn}
2335: \end{align}
2336:
2337: \end{document}
2338:
2339: