key: cord-0005240-vqnh68mo authors: Draief, Moez; Ganesh, Ayalvadi title: A random walk model for infection on graphs: spread of epidemics & rumours with mobile agents date: 2010-08-17 journal: Discret Event Dyn Syst DOI: 10.1007/s10626-010-0092-5 sha: e6022c23ba00765057ef78e895b1a227fa4b0dc3 doc_id: 5240 cord_uid: vqnh68mo We address the question of understanding the effect of the underlying network topology on the spread of a virus and the dissemination of information when users are mobile performing independent random walks on a graph. To this end, we propose a simple model of infection that enables to study the coincidence time of two random walkers on an arbitrary graph. By studying the coincidence time of a susceptible and an infected individual both moving in the graph we obtain estimates of the infection probability. The main result of this paper is to pinpoint the impact of the network topology on the infection probability. More precisely, we prove that for homogeneous graphs including regular graphs and the classical Erdős–Rényi model, the coincidence time is inversely proportional to the number of nodes in the graph. We then study the model on power-law graphs, that exhibit heterogeneous connectivity patterns, and show the existence of a phase transition for the coincidence time depending on the parameter of the power-law of the degree distribution. We finally undertake a preliminary analysis for the case with k random walkers and provide upper bounds on the convergence time for both the complete graph and regular graphs. In recent years, there has been a surge of hand-held wireless computing devices such as PDAs together with the proliferation of new services. These portable computing devices are equipped with a short-range wireless technology such as WiFi or Bluetooth. Despite providing a great deal of flexibility this ability to wirelessly connect to other devices, and to transfer data on the move, attracted the attention of virus writers who exploit such features for launching computer-virus outbreaks that take advantage of human mobility (Wang et al. 2009; Kleinberg 2007; Leavitt 2005) . Over the past couple of years, there have been indeed reports of malicious code that takes advantage of bluetooth vulnerabilities such as the Cabir worm that was detected during the World Athletics Championship (Mobile 2010) and another at a company that has been reported by CommWarrior (http://www.f-secure.com/ v-descs/commwarrior.shtml). Despite their small scale, these incidents bode more threats taking advantage of events and locations where individuals gather in close proximity (Yan et al. 2007; Su et al. 2006) . In much of the literature on mathematical epidemiology, the members of the population are assumed to occupy fixed locations and the probability of infection passing between a pair of them in a fixed time interval is taken to be some function of the distance between them. Mean-field (aka homogeneous mixing) models are a special case where an infected individual can potentially infect a number of random individuals chosen uniformly at random among the population (Daley and Gani 2001) . Recently, there has been an increasing interest in understanding the impact of the network topology on the spread of epidemics in networks with fixed nodes, see Draief (2006) , Draief and Massoulié (2010) . There has however been little analytical work to date on how mobility patterns may affect the outcomes of the processes concerned. In this work, we consider a different model in which the agents are mobile and can only infect each other if they are in sufficiently close proximity. The model is motivated both by certain kinds of biological epidemics, whose transmission may be dominated by sites at which individuals gather in close proximity (e.g. workplaces or public transport for a disease like SARS, cattle markets for foot-and-mouth disease, etc.) and by malware spreading between wireless devices via Bluetooth connections, for example. In what follows, we briefly describe some of the relevant related work on modelling epidemic spreading in mobile environments. To our knowledge, the first attempts to model virus spreading in mobile networks rely on the use of nonrigorous mean-field approximations (similar to the classical Kephart-White model Kephart and White 1991) that incorporate the mobility patterns of users. In Mickens and Noble (2005) , the authors derive a threshold for the persistence of the epidemic by computing the average number of neighbours of a given node. Using a similar approach but with different mobility patterns, Nekovee (2007), Rhodes and Nekovee (2008) explore the evolution of the number of devices that are infected in terms of the contact rate between users. A related line of work studying the dissemination of information in opportunistic networks (Chaintreau et al. 2007 ) focuses on the following analogous problem: Suppose that a set of mobile agents with wireless communication capabilities, forming a temporary network without the aid of a fixed infrastructure, are interested in a piece of information that is initially held by one user. The information is transmitted between users who happen to be in each other's range. As in the case of static networks (Pittel 1987) , one may be interested in the time it takes for the rumour to be known to all users. To this end we need to understand how information is transmitted between an informed and an ignorant user. This is also relevant to patterns of opinion formation and propagation of trends on social networks (Amini et al. 2009 ). Our work gives some insight on the impact of the network structure on the likelihood of successfully transmitting the rumour. Our contribution In contrast to the previous work which has focused on Euclidean models and homogeneous mobility patterns, in this work we consider a model wherein the different locations that a user can reach have varying popularity. More precisely, we consider a simple and stylised mathematical model of the spread of infection as follows. There is a finite, connected, undirected graph G = (V, E) on which the individuals perform independent random walks following a certain rate transition matrix Q. The infection can pass from an infected to a susceptible individual only if they are both at the same vertex. We assume that the probability of the virus being passed over a time interval of length τ is 1 − exp(−βτ ), where β > 0. The parameter β is known the infection rate. We shall focus a single infected and a single susceptible individual and ask what the probability is that the susceptible individual becomes infected by time t. This probability has been studied in the case of a complete graph in Datta and Dorlas (2004) . Here, we extend their results to a much wider class of graphs. It is simplistic to consider just a single infective and a single susceptible individual. Nevertheless, insights gained from this setting are relevant in the "sparse" case, where the number of both infected and susceptible individuals is small and intercontact times are fairly large. In that case, it is not a bad approximation to consider each pair of individuals in isolation. The "dense" setting will require quite different techniques and is not treated here. Random walks on graphs Motivated by applications in Physics, Biology, Social Sciences and Computer Science, there has been an ever increasing interest in analysing the properties of interacting particles or agents moving on a given network (Draief et al. 2005) . In particular, random walks play a central role in computer science, spanning a wide range of areas in both theory and practice, including distributed computing (Bui et al. 2004) . In fact many distributed algorithms use random walks as a building block. Applications in networks include token management (Coppersmith et al. 1993) , load balancing (Karger and Ruhl 2004) , small-world routing (Draief and Ganesh 2006; Kleinberg 2000) , search Gkantsidis et al. (2005) , information propagation and gathering (Kempe et al. 2001) , network topology monitoring (Ganesh et al. 2007 ) and group communication in ad-hoc networks (Dolev et al. 2006) . The paper Lovász (1993) provides a survey of the properties of a random walk on a finite graph. There has been fewer work related to multiple interacting mobile agents on a finite network. In Cooper et al. (2009) the authors propose a number of these dynamics and study their asymptotic properties on regular graphs. Aldous (1991) derives an upper bound for the expected meeting time of two independent copies of Markov chain as a function of the hitting time for a single chain. Coppersmith et al. (1993) provide an upper bound for the expected meeting time of a variant of the problems of two random walks on a general graph whereby an adversary tries to keep the tokens apart. Finally, Dimitriou et al. (2006) addresses a similar problem to ours. The authors analyse the dynamics of interacting random walks k − 1 of which are healthy and one is infectious where the infetion is transimitted as soon as an infectious and a healthy particle meet at given site. This model is related to the broadcast model presented in Cooper et al. (2009) . They derive upper bounds for the time it takes to infect all the healthy particles using standard results on the coalescence time of multiple random walks (Aldous and Fill 2002) . Organisation of the paper The rest of the paper is organised as follows. In Section 2, we present our model and state our main results that relate the coincidence time of the two walkers to the stationary distribution of a (general) random walk on a graph which enables us to upper bound the probability of infection at a given time t. In Section 3, we analyse the case of the standard continuous random walk and instantiate our results for regular graphs, the Erdős-Rényi graph and power-law networks. In Section 4, we present a more detailed analysis for the complete graph and regular graphs in the case of k ≥ 2 walkers providing tighter bounds than the ones derived in Dimitriou et al. (2006) for both models. We now describe the model precisely. We assume that the particles are mobile on a graph G = (V, E) and that their movements are characterised by independent continuous-time Markov chains (CTMCs) on the finite state space V = 1, . . . , n, with the same transition rate matrix given by Q = (q ij ) i, j∈V where j∈V q ij = 0 for all i ∈ V. More precisely, we assume that if a given particle occupies node i then it jumps to node j at rate q i, j ≥ 0. We assume that the graph G is connected and that q i, j > 0 if (i, j) ∈ E (the edge (i, j) is present in the graph G) and 0 otherwise. It is well known that if Q is the transition matrix of an ergodic Markov chain then there exists an invariant distribution π on V such that π Q = (0, . . . , 0) and is the total variation distance between ν and η two probability measures on V. From now on we assume that the Markov chain with transition rate matrix Q is reversible, i.e. π i q ij = π j q ji (for a detailed account on reversible Markov chains see (Aldous and Fill 2002, Chapter 3) . Considering 0 = λ 1 < λ 2 ≤ · · · ≤ λ n the eigenvalues of −Q, and using (Ganesh et al. 2007 , Lemma 1), we have that, We now consider two independent random walkers (X t , t ≥ 0) and (Y t , t ≥ 0) moving between the nodes of the graph according the matrix Q where X t and Y t correspond to the positions of the infected and susceptible individuals respectively at time t. A first natural object to study in this framework is the meeting time between the two walks defined by which is finite regardless of the initial condition. In Aldous and Fill (2002, Proposition 5, Chapter 14) , it is shown that the worst-case mean meeting time where H j is the time is takes the (single) chain (X t ) t to hit node j starting from node i. We define the coincidence time up to time t, denoted τ (t), as the total time up to t during which both walkers are at the same vertex, i.e., ( 2) Let I(t) denote the indicator function that the initial susceptible becomes infected by time t. Then, conditional on τ (t), we have where β > 0 is the infection rate. Let γ (t) = E(I(t)) be the probability that the initial susceptible becomes infected by time t. We are interested in estimating the coincidence time τ (t) and the infection probability γ (t) for different families of graphs. We consider the case when these chains are started independently in the stationary distribution and provide estimates of the coincidence time and an upper bound for the infection probability for general mobility models. Theorem 1 Suppose X 0 and Y 0 are chosen independently according to the invariant distribution π . Then, we have Proof Observe that, for all s ≥ 0, because X s and Y s are independent, and are in stationarity. Hence, it is immediate from Eq. 2 that Next, taking expectations in Eq. 3 with respect to the conditioning random variable τ (t), we have where the inequality follows from Jensen's inequality. We now introduce some terminology and define some examples of graph models that we shall consider. For two functions f (·) and g(·) on the natural numbers, we write f (n) ∼ g(n) to mean that their ratio tends to 1 as n tends to infinity. We write . For a sequence of events A n indexed by n ∈ N, we say that they occur with high probability (w.h.p.) if P(A n ) tends to 1 as n tends to infinity. Example: uniform random walk If we consider the case where the rate transition matrix is given by where d i is the degree of i, i.e. the number of neighbours of node i in the graph G then it is known that the stationary distribution is the uniform distribution over V, π i = 1/n for all i ∈ V, and a direct application of Theorem 1 yields Moreover, the two walks need not start from the stationary distribution. In fact, it is not difficult to see that in this case, if the graph is connected, then the second eigenvalue of Q above is positive, and in fact bounded away from zero, for all n, using the fact that Therefore, after O(log(n)), using Eq. 1, the distribution of the walks coincide with the uniform distribution w.h.p. In this section, we choose a matrix Q in order to allow a non-uniform probability of being at a given node in stationary regime. More precisley, the particles perform independent continuous-time Markov chains (CTMCs) on the finite state space V, with the same transition rate matrix given by Moreover, they both intially start in positions distributed according to the invariant distribution π given by Note that if X t and Y t are not started from the stationary distribution, then by Eq. 1, after c log n λ2 , c constant but large, the distributions of X t and Y t coincide with the uniform distribution on V w.h.p. Moreover, by Cheeger's inequality (see Mohar 1997) , we have that where (G) is the maximal degree of the nodes in G and η(G) is the isoperimetric constant or expansion of a graph G defined by denotes the number of edges having one vertex in U and the other in its complement, U c (i.e., crossing the cut (U, U c )); and |U| the number of vertices or size of U. In particular if the graph is an expander, then the spectral gap λ 2 is large, and the (uniform) random walk is rapidly mixing (Mohar 1997) . In particular, the infection does not take place, w.h.p., before the two walks approach the stationary distribution. In what follows, we will to consider graphs which are expanders as illustrated in Ganesh et al. (2005) through the computation of the isoperimetric constant for Erdős-Rényi and power-law graphs and Friedman's proof Alon's second eigenvalue conjecture for regular graphs (Friedman 2003). We now introduce a number of families of graphs of interest. Complete graphs Consider the complete graph on n nodes, namely the graph in which there is an edge between every pair of nodes. Thus, d v = n − 1 and π v = 1/n for all v ∈ V, so we have by Theorem 1 that E[τ (t)] = t/n. This result should be intuitive by symmetry. Lemma 1 also gives us an upper bound on the infection probability, γ (t) ≤ 1 − exp(−βt/n). Roughly speaking, this says that it takes time of order n/β for the susceptible individual to become infected; for t n/β, the probability of being infected is vanishingly small. Again, this is consistent with intuition. Thus, a complete graph is regular with r = n − 1. It is readily verified that π v = 1/n for all v ∈ V if G is r-regular for any r ≥ 3. Hence, if G is connected, we have the same estimates for τ (t) and γ (t) as for the complete graph, which is a special case corresponding to r = n − 1. The next examples we consider will be families of random graphs widely used in practice to model networks. Erdős-Rényi random graphs The Erdős-Rényi graph G(n, p) is defined as a random graph on n nodes, wherein each edge is present with probability p, independent of all other edges. We consider a family of such random graphs indexed by n, and take p to be a function of n chosen so that np > c log n for some constant c > 1. We also condition on the graph being connected. For p as above, the probability of connectivity tends to 1 as n tends to infinity, so conditioning on connectivity does not alter any of the estimates we shall derive later for the coincidence time on such graphs. In this model, the node degrees are identically distributed Binomial random variables with parameters (n − 1, p). In particular, they concentrate around the mean value of np, and have exponentially decaying tails away from this value. Thus, while Erdős-Rényi graphs are not exactly regular, they exhibit considerable homogeneity in node degrees. Power-law random graphs In contrast to the above graph models, many real-world networks exhibit considerable heterogeneity in node degrees, and have empirical degree distributions whose tails decay polynomially; see, e.g., Barabási and Albert (1999), Faloutsos et al. (1999) . This observation has led to the development of generative models for graphs with power-law tails (Barabási and Albert 1999; Bollobás and Riordan 2004) as well as random-graph models possessing this property (Chung and Lu 2003) . For definiteness, we work with the model proposed in Chung and Lu (2003) , but we believe that similar results will hold for the other models as well. In the model of Chung and Lu (2003) , each node v is associated with a positive weight w v , and edges are present independently with probabilities related to the weights by We assume that W ≥ w 2 max , so that the above defines a probability. It can be verified that E[d v ] = w v and so this model is also referred to as the expected degree model. The model allows self-loops. The Erdős-Rényi graph G(n, p) is a special case corresponding to the choice w v = np for all v ∈ V. If the weights are chosen to have a power-law distribution, then so will the node degrees. The following 3-parameter model for the ordered weight sequence is proposed in Chung and Lu (2003) , parametrised by the mean degree d, the maximum degree m, and the exponent γ > 2 of the weight distribution: where Note that W = n−1 i=0 w i ∼ nd. We consider a sequence of such graphs indexed by n. The maximum expected degree m and the average expected degree d may, and indeed typically will, depend on n. In models of real networks, we can typically expect d to remain bounded or to grow slowly with n, say logarithmically, while m grows more quickly, say as some fractional power of n. In this paper, we only assume the following: Here, δ is a constant that does not depend on n. In other words, the average expected degree is uniformly bounded away from zero. The third assumption simply restates the requirement that w 2 0 ≤ W, so that Eq. 5 defines valid probabilities. The last assumption ensures that i 0 , defined in Eq. 7, tends to infinity. We now describe our results about these models. On each graph, consider two independent random walks with initial condition X 0 , Y 0 chosen independently from the invariant distribution π for the random walk on that graph. We Finally, consider a sequence of power-law random graphs def ined via Eqs. 5 and 6, and satisfying the assumptions in Eq. 8. Then, we have the following: where c > 0 is a constant that may depend on γ , but not on n. The thereom illustrates a discrepency between the dynamics of the system of two particles for regular graphs, Erdős-Rényi random graphs and power-law graphs for γ > 3 on one hand, and power-law graphs γ ≤ 3 on the other hand. For the former, the particles spend O(t/n) amount of time together up to time t. In fact, as we will illustrate it in Section 4 for the complete and regular graphs, the healthy particle is infected w.h.p. before time β+2 β n. As for power-law graphs with γ ≤ 3, the particles spend considerably more time together, at high degree nodes, and we anticipate that the infection occurs much faster. The remainder of this section is dedicated to proving Theorem 2. If the graph G is regular, then, by Eq. 4, π v = 1/n for all v ∈ V. Hence, the claim of the theorem follows from Theorem 1. A more thorough analysis is provided in Section 4 for regular and complete graphs, where we present som eintial findings for the case where there are k particles: one infected and k − 1 healthy. In order to estimate E[τ (t)], we need to compute To this end, define where A uv = 1 ((u,v)∈E) is the indicator that u and v are connected in G, and We will derive the first and second moments of the variables D and D 2 . It then suffices to use Chebyshev's inequality to establish concentration results for both variables D and D 2 . By Eq. 9 and Theorem 1, and the fact that v∈V d 2 v = D 2 + D and D = v∈V d v , we will have an estimate of the coincidence time that holds whp. We begin by computing the mean and variance of D in the expected degree model with arbitrary weight sequence {w i , i = 0, . . . , n − 1}. For notational convenience, we define We obtain Erdős-Rényi graphs G(n, p) by setting w i = np for all i, and so, w k = (np) k for such graphs. Next, consider power-law graphs with weight sequence specified by Eqs. 6 and 7. Since i 0 tends to infinity by assumption, we have for such graphs that Now, straightforward calculations yield that w ∼ d for all γ > 2, whereas, for k ≥ 2, we have We can now compute the mean and variance of D, the sum of node degrees. where n = |V| is the total number of nodes. In particular, for large n, if G is the Erdős-Rényi random graph G(n, p), then whereas, if G is a power law random graph satisfying the assumptions of Theorem 2, then The proof is provided in the Appendix of the paper. The following corollary is now an easy consequence of Chebyshev's inequality. We now establish a similar concentration result for the sum of squared degrees. To this end, recall that We have the following: Proposition 2 Let D 2 be def ined as in Eq. 11. We then have We now specialise the results to Erdős-Rényi and power-law random graphs, showing that D 2 concentrates near its expected value with high probability. Suppose G(n, p) is a sequence of Erdős-Rényi random graphs indexed by n (where p depends on n but this is not made explicit in the notation), and that np is uniformly bounded away from zero. Then Proof We have, by Proposition 2 and the fact that w k = (np) k for the Erdős-Rényi random graph G(n, p), that E[D 2 ] = n 2 (n − 1) p 2 ∼ n 3 p 2 , Var(D 2 ) ≤ 8n 4 p 3 + 2n 3 p 2 . Hence, by Chebyshev's bound, we obtain for all ε > 0 that, Now, by the assumption that np is bounded away from zero, (n − 1) 2 p and n(n − 1) 2 p 2 tend to infinity as n tends to infinity. Thus, P(|D 2 − E[D 2 ]| > εE[D 2 ]) tends to zero for all ε > 0. This establishes the claim of the Proposition. Using similar arguments, albeit with more involved computation, we derive an analogous result for power-law graphs. The proof is provided in the Appendix of the paper. Proposition 4 Suppose G n , n ∈ N is a sequence of power-law random graphs satisfying the assumptions in Theorem 2, with γ > 2. Then, D 2 ∼ E[D 2 ] whp, and To complete the proof of Theorem 2, it suffices to combine the results of Lemmas 1 and 4 with the fact that v π 2 v = D2+D D 2 . In this section, we focus our attention on the complete graph for which the uniform and the standard continuous-time random walks coincide. Due to the absence of topology, i.e. a walk can jump to any other node uniformly at random, we can explore this model in more detail. First, let us analyze the case of two random walks, one being infectious and the other healthy. The dynamic of the system can be described through a three state Markov chain: state 1 corresponds to the case where the two walks are apart, the state 2 corresponds to when the two walks coincide, and state 3 corresponds to the infection of the healthy walk by the infected one. The rate matrix of the corresponding continuous-time Markov chain is then given by and the corresponding transition matrix P t = e tQ . The eigenvalues of Q are 0, −λ 1 and −λ 2 with λ 1 = 1 2 β + 2 + (β + 2) 2 − 8β/(n − 1) The infection probability after time t is therefore given by βt n yielding a tighter bound than the one derived using Theorem 1. Let m 3 (i) be the mean hitting time of state 3, starting from state i, for i = 1, 2. Note that m 3 (1) is the mean time for the healthy node to get infected. By the firststep analysis (Norris 1997, Theorem 3.3 .3), we have From this, it follows, that the average time to infection ET Inf is given by To conclude this section we provide a first analysis of the case with k random particles k ≥ 2 where initially one particle is infectious and the remaining k − 1 are healthy. In what follows we provide an upper bound on the time to infection of the whole population. First it is easy to note that at time t a given particle can occupy any of the n nodes of the graph. Therefore by combining Eq. 17 and the union bound, the average time to infection E T (k) Inf is bounded by We now derive a better bound. First, note that at any given time, on the complete graph, each of the particles is located uniformly at random on the graph. If we assume that there are i infectious particles and (k − i) healthy ones, then for k ≤ √ n, it is not difficult to show that the probability that there is at least an encounter between one healthy and one infectious particles is equivalent to i(k−i) n . To simplify the analysis, we assume that if there are healthy and infectious particles located at the same node then there is at most one infection that occurs before any of the k particles jumps to another node which happens with probability β β+k . The above assumption slows down the infection process and thus yields an upper bound. It is not difficult to note that the time until there is a new particle infected is stochastically dominated by N i=1 E i where E i are i.i.d. r.v. distributed according to an exponential distribution with parameter k and N is a geometric random variable with parameter β β+k i(k−i) n . Hence, The above analysis can be extended to the case of r-regular graphs (r ≥ 3), using Cooper et al. (2009, Theorem 22) , one can derive that the probability of a meeting between a healthy and an infectious particle, at the stage where there are i infectious particles is asymptotically (in n) equal to r−2 r−1 i(k−i) n yielding the following result. The time to infection, starting from 1 infected particle and k − 1 healthy ones, all performing standard continuous-random walks on an r-regular graph, r ≥ 3 and k = o( √ n), satisf ies Proof of Proposition 1 It is immediate from Eq. 10 that which establishes the first equality in Eq. 14. Next, rewrite Eq. 10 as and observe from the independence of the edges that Var(A ii ). , and so, Upon simplifying, this yields the second equality in Eq. 14. Now, using the fact that w k = (np) k for Erdős-Rényi graphs G(n, p), we readily obtain Eq. 15. Next, suppose G is a power-law graph (more precisely, G n is a sequence of power law graphs) satisfying the assumptions of Theorem 2. It follows from Eq. 12 that while and w ∼ d for all γ > 2. By Eq. 14, it suffices to show that in order to show that Var(D) ∼ 2nw ∼ 2nd. Suppose first that γ > 3. Then, by Eq. 19 and the fact that w = nd, where the last equality follows from Eq. 8, and the fact that d ≤ n. Now let γ = 3. Then, by Eq. 19 and the fact that w = nd, 1 nd where the last equality follows by Eq. 8. On the other hand, if 2 < γ < 3, then, by Eq. 19, 1 nd where we have used the inequality m ≤ √ nd from Eq. 8 to obtain the second equality. To obtain the last equality, note that it follows from Eq. 8 that m = o(n) and hence that d = o(n) as well. We have thus established the first equality in Eq. 20 for all γ > 2. The proof of the second equality is similar and is omitted. This completes the proof of the propostion. We first note that Therefore, which is the first part of Eq. 16. Next, for distinct nodes u, v ∈ V, we have Similarly, we obtain Using the fact that distinct edges are independent, we get Var(X u ) ≤ 4w 3 u + 2w 2 u . Now, by Eqs. 11, 21 and 22 Computing the above sums yields the second part of Eq. 16. We will show that Var(D 2 ) = o(E[D 2 ] 2 ), so that the claim follows by Chebyshev's bound, as in the proof of the previous lemma. We will consider separately the parameter ranges γ ≥ 4, 3 ≤ γ < 4 and 2 < γ < 3, where γ is the exponent in the power law describing the degree distribution. In the following, c 1 , c 2 , . . . will denote generic positive constants, not necessarily the same from line to line. Recall that w ∼ d. Suppose first that γ ≥ 4. Then, by Eq. 13, w 3 = O(d 3 log m d ) and w 2 ∼ c 1 d 2 . Therefore, by Lemma 2, where the last equality holds because of the assumption in Eq. 8 that d ≥ δ for some constant δ > 0. Thus, we see that since m ≤ n. Suppose next that 3 ≤ γ < 4. Then, by Eq. 13, w 3 = O(d γ −1 m 4−γ ), while w 2 ∼ c 1 d 2 if 3 < γ < 4 and w 2 ∼ c 2 d 2 log m d if γ = 3. Therefore, by Lemma 2, E[D 2 ] ≥ c 1 nd 2 − c 2 d 2 log 2 m d ≥ c 1 nd 2 − c 2 d 2 log 2 n = (nd 2 ), We have used the assumption that d is uniformly bounded away from zero to obtain the second inequality above. Since we also assumed in Eq. 8 that d = o(m), we have We have used Eq. 8 to obtain the second equality above and the fact that γ ≥ 3 to obtain the last equality. Moreover, E[D 2 ] ∼ cnd 2 for 3 < γ < 4, whereas E[D 2 ] ∼ cnd 2 log(m) for γ = 3. Finally, suppose that 2 < γ < 3. Then, by Eq. 13, w 3 = O(d γ −1 m 4−γ ) and w 2 ∼ c 1 d γ −1 m 3−γ , so that, by Lemma 2, Now, by Eq. 8, and the fact that γ > 2 we have (m/d) γ −1 = o(n). Since the maximum degree m is assumed to grow as a power of n, we have Var(D2) E[D2] 2 = o(1). Note that E[D 2 ] ∼ cnd γ −1 m 3−γ , for 2 < γ < 3. Using Chebyshev's inequality, this establishes the claim of the proposition. Meeting times for independent Markov chains Reversible Markov chains ans random walks on graphs Marketing in a random network Emergence of scaling in random networks The diameter of a scale-free random graph Random walks in distributed computing: a survey Impact of human mobility on opportunistic forwarding algorithms The average distances in random graphs with given expected degrees Multiple random walks in random regular graphs Collisions among random walks on a graph Epidemic modelling: an introduction Random walks on a complete graph: a model for infection The infection time of graphs Random walk for self-stabilizing group communication in ad hoc networks Epidemic processes on complex networks Efficient routing in poisson small-world networks Queues, stores, and tableaux A proof of Alons second eigenvalue conjecture A random walk model for infection on graphs Peer counting and sampling in overlay networks based on random walks The effect of network topology on the spread of epidemics Hybrid search schemes for unstructured peer-to-peer networks Simple efficient load balancing algorithms for peer-to-peer systems Spatial gossip and resource location protocols Directed-graph epidemiological models of computer viruses The wireless epidemic The small-world phenomenon: an algorithm perspective Mobile phones: the next frontier for hackers Random Walks on Graphs: a survey Modeling epidemic spreading in mobile environments Some applications of Laplace eigenvalues of graphs Worm epidemics in wireless ad hoc networks On spreading a rumor The opportunistic transmission of wireless worms between mobile devices A preliminary investigation of worm infections in a bluetooth environment Understanding the spreading patterns of mobile phone viruses Bluetooth worm propagation: mobility pattern matters He then completed a DEA in Probability Theory at the University Paris VI IIT Madras in 1988, MS and PhD in EE from Cornell University in 1991 and 1995 His research interests are in the mathematical modelling of communication and computer networks, and in decentralised algorithms for such networks. Specific interests include large deviations and applications to queueing theory and statistics, random graph models and stochastic processes on graphs, and decentralised algorithms for resource allocation in the Internet and in wireless networks.