Modern Discrete Probability Branching processes

Similar documents
Branching processes. Chapter Background Basic definitions

Sharpness of second moment criteria for branching and tree-indexed processes

14 Branching processes

The range of tree-indexed random walk

BRANCHING PROCESSES 1. GALTON-WATSON PROCESSES

Notes 15 : UI Martingales

Notes 18 : Optional Sampling Theorem

UPPER DEVIATIONS FOR SPLIT TIMES OF BRANCHING PROCESSES

Erdős-Renyi random graphs basics

Problem Points S C O R E Total: 120

Hitting Probabilities

Lecture 17 Brownian motion as a Markov process

Lecture 06 01/31/ Proofs for emergence of giant component

Notes 6 : First and second moment methods

Homework 6: Solutions Sid Banerjee Problem 1: (The Flajolet-Martin Counter) ORIE 4520: Stochastics at Scale Fall 2015

Selected Exercises on Expectations and Some Probability Inequalities

Lecture 2. We now introduce some fundamental tools in martingale theory, which are useful in controlling the fluctuation of martingales.

Optional Stopping Theorem Let X be a martingale and T be a stopping time such

Problem Sheet 1. You may assume that both F and F are σ-fields. (a) Show that F F is not a σ-field. (b) Let X : Ω R be defined by 1 if n = 1

Positive and null recurrent-branching Process

Stochastic Process (ENPC) Monday, 22nd of January 2018 (2h30)

Itô s excursion theory and random trees

Hard-Core Model on Random Graphs

Branching within branching: a general model for host-parasite co-evolution

Population growth: Galton-Watson process. Population growth: Galton-Watson process. A simple branching process. A simple branching process

Math 6810 (Probability) Fall Lecture notes

ECE534, Spring 2018: Solutions for Problem Set #4 Due Friday April 6, 2018

Biased random walks on subcritical percolation clusters with an infinite open path

Mean-field dual of cooperative reproduction

Part IA Probability. Definitions. Based on lectures by R. Weber Notes taken by Dexter Chua. Lent 2015

4 Sums of Independent Random Variables

µ X (A) = P ( X 1 (A) )

Statistics 253/317 Introduction to Probability Models. Winter Midterm Exam Monday, Feb 10, 2014

Exam Stochastic Processes 2WB08 - March 10, 2009,

Interlude: Practice Final

On trees with a given degree sequence

arxiv: v1 [math.pr] 13 Nov 2018

Point Process Control

6.207/14.15: Networks Lecture 3: Erdös-Renyi graphs and Branching processes

THE CONTACT PROCESS ON TREES

Resistance Growth of Branching Random Networks

1. Stochastic Processes and filtrations

THE x log x CONDITION FOR GENERAL BRANCHING PROCESSES

A simple branching process approach to the phase transition in G n,p

4 Branching Processes

(6, 4) Is there arbitrage in this market? If so, find all arbitrages. If not, find all pricing kernels.

MATH 56A: STOCHASTIC PROCESSES CHAPTER 6

Lecture 5. 1 Chung-Fuchs Theorem. Tel Aviv University Spring 2011

Solutions For Stochastic Process Final Exam

Universal examples. Chapter The Bernoulli process

arxiv: v3 [math.pr] 10 Nov 2017

Part IA Probability. Theorems. Based on lectures by R. Weber Notes taken by Dexter Chua. Lent 2015

k-protected VERTICES IN BINARY SEARCH TREES

n! (k 1)!(n k)! = F (X) U(0, 1). (x, y) = n(n 1) ( F (y) F (x) ) n 2

Problem Points S C O R E Total: 120

Scaling limits for random trees and graphs

Fundamental Inequalities, Convergence and the Optional Stopping Theorem for Continuous-Time Martingales

arxiv: v2 [math.pr] 25 Feb 2017

T. Liggett Mathematics 171 Final Exam June 8, 2011

Lecture 9 Classification of States

Survival Probabilities for N-ary Subtrees on a Galton-Watson Family Tree

The Minesweeper game: Percolation and Complexity

Chapter 3. Erdős Rényi random graphs

Random trees and branching processes

March 1, Florida State University. Concentration Inequalities: Martingale. Approach and Entropy Method. Lizhe Sun and Boning Yang.

Discrete random structures whose limits are described by a PDE: 3 open problems

P (A G) dp G P (A G)

MATH 56A: STOCHASTIC PROCESSES CHAPTER 2

Markov Decision Processes

Almost sure asymptotics for the random binary search tree

Let (Ω, F) be a measureable space. A filtration in discrete time is a sequence of. F s F t

A Conceptual Proof of the Kesten-Stigum Theorem for Multi-type Branching Processes

MS&E 321 Spring Stochastic Systems June 1, 2013 Prof. Peter W. Glynn Page 1 of 10

Random trees and applications

Lecture 22 Girsanov s Theorem

1 Gambler s Ruin Problem

(A n + B n + 1) A n + B n

6. Brownian Motion. Q(A) = P [ ω : x(, ω) A )

Counting independent sets up to the tree threshold

Other properties of M M 1

Preliminary Exam: Probability 9:00am 2:00pm, Friday, January 6, 2012

MA3H2 Markov Processes and Percolation theory. Stefan Adams

Almost giant clusters for percolation on large trees

FRINGE TREES, CRUMP MODE JAGERS BRANCHING PROCESSES AND m-ary SEARCH TREES

arxiv: v2 [math.pr] 19 Jun 2015

9.1 Branching Process

Cutting edges at random in large recursive trees

Galton-Watson trees and probabilities of local neighbourhoods of the root

arxiv: v1 [math.pr] 25 Jul 2013

Chapter 2 Galton Watson Trees

Theory of Stochastic Processes 3. Generating functions and their applications

Limiting distribution for subcritical controlled branching processes with random control function

< k 2n. 2 1 (n 2). + (1 p) s) N (n < 1

Stein s Method: Distributional Approximation and Concentration of Measure

Bootstrap Percolation on Periodic Trees

THE SECOND LARGEST COMPONENT IN THE SUPERCRITICAL 2D HAMMING GRAPH

Lecture 2: Tipping Point and Branching Processes

RECENT RESULTS FOR SUPERCRITICAL CONTROLLED BRANCHING PROCESSES WITH CONTROL RANDOM FUNCTIONS

4 Expectation & the Lebesgue Theorems

Random transceiver networks

Transcription:

Modern Discrete Probability IV - Branching processes Review Sébastien Roch UW Madison Mathematics November 15, 2014

1 Basic definitions 2 3 4

Galton-Watson branching processes I Definition A Galton-Watson branching process is a Markov chain of the following form: Let Z 0 := 1. Let X(i, t), i 1, t 1, be an array of i.i.d. Z + -valued random variables with finite mean m = E[X(1, 1)] < +, and define inductively, Z t := X(i, t). 1 i Z t 1

Galton-Watson branching processes II Further remarks: 1 The random variable Z t models the size of a population at time (or generation) t. The random variable X(i, t) corresponds to the number of offspring of the i-th individual (if there is one) in generation t 1. Generation t is formed of all offspring of the individuals in generation t 1. 2 We denote by {p k } k 0 the law of X(1, 1). We also let f (s) := E[s X(1,1) ] be the corresponding probability generating function. 3 By tracking genealogical relationships, i.e. who is whose child, we obtain a tree T rooted at the single individual in generation 0 with a vertex for each individual in the progeny and an edge for each parent-child relationship. We refer to T as a Galton-Watson tree.

Exponential growth I Lemma Let M t := m t Z t. Then (M t ) is a nonnegative martingale with respect to the filtration F t = σ(z 0,..., Z t ). In particular, E[Z t ] = m t. Proof: Recall the following lemma: Lemma: Let (Ω, F, P) be a probability space. If Y 1 = Y 2 a.s. on B F then E[Y 1 F] = E[Y 2 F] a.s. on B. On {Z t 1 = k}, E[Z t F t 1 ] = E X(j, t) F t 1 = mk = mz t 1. 1 j k This is true for all k. Rearranging shows that (M t) is a martingale. For the second claim, note that E[M t] = E[M 0 ] = 1.

Exponential growth II Theorem We have M t M < + a.s. for some nonnegative random variable M σ( t F t ) with E[M ] 1. Proof: This follows immediately from the martingale convergence theorem for nonnegative martingales and Fatou s lemma.

1 Basic definitions 2 3 4

: some observations I Observe that 0 is a fixed point of the process. The event {Z t 0} = { t : Z t = 0}, is called extinction. Establishing when extinction occurs is a central question in branching process theory. We let η be the probability of extinction. Throughout, we assume that p 0 > 0 and p 1 < 1. Here is a first result: Theorem A.s. either Z t 0 or Z t +. Proof: The process (Z t) is integer-valued and 0 is the only fixed point of the process under the assumption that p 1 < 1. From any state k, the probability of never coming back to k > 0 is at least p k 0 > 0, so every state k > 0 is transient. The claim follows.

: some observations II Theorem (Critical branching process) Assume m = 1. Then Z t 0 a.s., i.e., η = 1. Proof: When m = 1, (Z t) itself is a martingale. Hence (Z t) must converge to 0 by the corollaries above.

Main result I Let f t (s) = E[s Z t ]. Note that, by monotonicity, η = P[ t 0 : Z t = 0] = lim P[Z t = 0] = lim f t(0), t + t + Moreover, by the Markov property, f t as a natural recursive form: f t (s) = E[s Z t ] = E[E[s Z t F t 1 ]] = E[f (s) Z t 1 ] where f (t) is the t-th iterate of f. = f t 1 (f (s)) = = f (t) (s),

Main result II Theorem ( probability) The probability of extinction η is given by the smallest fixed point of f in [0, 1]. Moreover: (Subcritical regime) If m < 1 then η = 1. (Supercritical regime) If m > 1 then η < 1. Proof: The case p 0 + p 1 = 1 is straightforward: the process dies almost surely after a geometrically distributed time. So we assume p 0 + p 1 < 1 for the rest of the proof.

Main result: proof I Lemma: On [0, 1], the function f satisfies: (a) f (0) = p 0, f (1) = 1; (b) f is indefinitely differentiable on [0, 1); (c) f is strictly convex and increasing; (d) lim s 1 f (s) = m < +. Proof: (a) is clear by definition. The function f is a power series with radius of convergence R 1. This implies (b). In particular, f (s) = i 1 ip i s i 1 0, and f (s) = i 2 i(i 1)p i s i 2 > 0, because we must have p i > 0 for some i > 1 by assumption. This proves (c). Since m < +, f (1) = m is well defined and f is continuous on [0, 1], which implies (d).

Main result: proof II Lemma: We have: If m > 1 then f has a unique fixed point η 0 [0, 1). If m < 1 then f (t) > t for t [0, 1). (Let η 0 := 1 in that case.) Proof: Assume m > 1. Since f (1) = m > 1, there is δ > 0 s.t. f (1 δ) < 1 δ. On the other hand f (0) = p 0 > 0 so by continuity of f there must be a fixed point in (0, 1 δ). Moreover, by strict convexity and the fact that f (1) = 1, if x (0, 1) is a fixed point then f (y) < y for y (x, 1), proving uniqueness. The second part follows by strict convexity and monotonicity.

Main result: proof III

Main result: proof IV Lemma: We have: If x [0, η 0 ), then f (t) (x) η 0 If x (η 0, 1) then f (t) (x) η 0 Proof: By monotonicity, for x [0, η 0 ), we have x < f (x) < f (η 0 ) = η 0. Iterating x < f (1) (x) < < f (t) (x) < f (t) (η 0 ) = η 0. So f (t) (x) L η 0. By continuity of f we can take the limit inside of f (t) (x) = f (f (t 1) (x)), to get L = f (L). So by definition of η 0 we must have L = η 0.

Main result: proof V

Example: Poisson branching process Example Consider the offspring distribution X(1, 1) Poi(λ) with λ > 0. We refer to this case as the Poisson branching process. Then f (s) = E[s X(1,1) ] = i 0 λ λi e i! si = e λ(s 1). So the process goes extinct with probability 1 when λ 1. For λ > 1, the probability of extinction η λ is the smallest solution in [0, 1] to the equation e λ(1 x) = x. The survival probability ζ λ := 1 η λ satisfies 1 e λζ λ = ζ λ.

: back to exponential growth I Conditioned on extinction, M = 0 a.s. Theorem Conditioned on nonextinction, either M = 0 a.s. or M > 0 a.s. In particular, P[M = 0] {η, 1}. Proof: A property of rooted trees is said to be inherited if all finite trees satisfy this property and whenever a tree satisfies the property then so do all the descendant trees of the children of the root. The property {M = 0} is inherited. The result then follows from the following 0-1 law. Lemma: For a Galton-Watson tree T, an inherited property A has, conditioned on nonextinction, probability 0 or 1. Proof of lemma: Let T (1),..., T (Z 1) be the descendant subtrees of the children of the root. Then, by independence, P[A] = E[P[T A Z 1 ]] E[P[T (i) A, i Z 1 Z 1 ]] = E[P[A] Z 1 ] = f (P[A]), so P[A] [0, η] {1}. Also P[A] η because A holds for finite trees.

: back to exponential growth II Theorem Let (Z t ) be a branching process with m = E[X(1, 1)] > 1 and σ 2 = Var[X(1, 1)] < +. Then, (M t ) converges in L 2 and, in particular, E[M ] = 1. Proof: From the orthogonality of increments On {Z t 1 = k} E[M 2 t ] = E[M 2 t 1] + E[(M t M t 1 ) 2 ]. E[(M t M t 1 ) 2 F t 1 ] = m 2t E[(Z t mz t 1 ) 2 F t 1 ] ( k 2 = m 2t E X(i, t) mk) F t 1 = m 2t kσ 2 i=1 = m 2t Z t 1 σ 2.

: back to exponential growth III Hence Since E[M 2 0 ] = 1, E[M 2 t ] = E[M 2 t 1] + m t 1 σ 2. t+1 E[Mt 2 ] = 1 + σ 2 m i, which is uniformly bounded when m > 1. So (M t) converges in L 2. Finally by Fatou s lemma i=2 E M sup M t 1 sup M t 2 < + and E[M t] E[M ] M t M 1 M t M 2, implies the convergence of expectations.

1 Basic definitions 2 3 4

Exploration process I We consider an exploration process of the Galton-Watson tree T. The exploration process, started at the root 0, has 3 types of vertices: - A t : active, E t : explored, N t : neutral. We start with A 0 := {0}, E 0 :=, and N 0 contains all other vertices in T. At time t, if A t 1 = we let (A t, E t, N t ) := (A t 1, E t 1, N t 1 ). Otherwise, we pick an element, a t, from A t 1 and set: - A t := A t 1 {x N t 1 : {x, a t } T }\{a t }, - E t := E t 1 {a t }, - N t := N t 1 \{x N t 1 : {x, a t } T }. To be concrete, we choose a t in breadth-first search (or first-come-first-serve) manner: we exhaust all vertices in generation t before considering vertices in generation t + 1.

Exploration process II We imagine revealing the edges of T as they are encountered in the exploration process and we let (F t ) be the corresponding filtration. In words, starting with 0, the Galton-Watson tree T is progressively grown by adding to it at each time a child of one of the previously explored vertices and uncovering its children in T. In this process, E t is the set of previously explored vertices and A t is the set of vertices who are known to belong to T but whose full neighborhood is waiting to be uncovered. The rest of the vertices form the set N t.

Exploration process III Let A t := A t, E t := E t, and N t := N t. Note that (E t ) is non-decreasing while (N t ) is non-increasing. Let τ 0 := inf{t 0 : A t = 0}, (which by convention is + if there is no such t). The process is fixed for all t > τ 0. Notice that E t = t for all t τ 0, as exactly one vertex is explored at each time until the set of active vertices is empty. Lemma Let W be the total progeny. Then W = τ 0.

Random walk representation I The process (A t ) admits a simple recursive form. Recall that A 0 := 1. Conditioning on F t 1 : - If A t 1 = 0, the exploration process has finished its course and A t = 0. Otherwise, (a) one active vertex becomes an explored vertex and (b) its neutral neighbors become active vertices. That is, A t 1 + [ ] }{{} 1 + X t, t 1 < τ0, }{{} A t = (a) (b) 0, o.w. where X t is distributed according to the offspring distribution.

Random walk representation II We let Y t = X t 1 1 and with S 0 := 1. Then t S t := 1 + Y i, i=1 τ 0 = inf{t 0 : S t = 0} = inf{t 0 : 1 + [X 1 1] + + [X t 1] = 0} = inf{t 0 : X 1 + + X t = t 1}, and (A t ) is a random walk started at 1 with steps (Y t ) stopped when it hits 0 for the first time: A t = (S t τ0 ).

Duality principle I Theorem Let (Z t ) be a branching process with offspring distribution {p k } k 0 and extinction probability η < 1. Let (Z t ) be a branching process with offspring distribution {p k } k 0 where p k = ηk 1 p k. Then (Z t ) conditioned on extinction has the same distribution as (Z t ), which is referred to as the dual branching process.

Duality principle II Some remarks: Note that k 0 p k = k 0 η k 1 p k = η 1 f (η) = 1, because η is a fixed point of f. So {p k } k 0 is indeed a probability distribution. Note further that k 0 kp k = k 0 kη k 1 p k = f (η) < 1, since f is strictly increasing, f (η) = η < 1 and f (1) = 1. So the dual branching process is subcritical.

Duality principle III Proof: We use the random walk representation. Let H = (X 1,..., X τ0 ) and H = (X 1,..., X τ 0 ) be the histories of the processes (Z t) and (Z t ) respectively. (Under breadth-first search, the process (Z t) can be reconstructed from H.) In the case of extinction, the history of (Z t) has finite length. We call (x 1,..., x t) a valid history if x 1 + + x i (i 1) > 0 for all i < t and x 1 + + x t (t 1) = 0. By definition of the conditional probability, for a valid history (x 1,..., x t) with a finite t, P[H = (x 1,..., x t) τ 0 < + ] = P[H = (x 1,..., x t)] P[τ 0 < + ] Because x 1 + + x t = t 1, η 1 t t p xi = η 1 η 1 x i p x i = i=1 i=1 t i=1 p x i t = η 1 p xi. i=1 = P[H = (x 1,..., x t)].

Duality principle: example Example (Poisson branching process) Let (Z t ) be a Galton-Watson branching process with offspring distribution Poi(λ) where λ > 1. Then the dual probability distribution is given by p k = ηk 1 p k = η k 1 λ λk e k! = η 1 λ (λη)k e k! where recall that e λ(1 η) = η, so p k = eλ(1 η) λ (λη)k e k! λη (λη)k = e. k! That is, the dual branching process has offspring distribution Poi(λη).,

Hitting-time theorem Theorem Let (Z t ) be a Galton-Watson branching process with total progeny W. In the random walk representation of (Z t ), for all t 1. P[W = t] = 1 t P[X 1 + + X t = t 1], Note that this formula is rather remarkable as the probability on the l.h.s. is P[S i > 0, i < t and S t = 0] while the probability on the r.h.s. is P[S t = 0].

Spitzer s combinatorial lemma I We start with a lemma of independent interest. Let u 1,..., u t R and define r 0 := 0 and r i := u 1 + + u i for 1 i t. We say that j is a ladder index if r j > r 0 r j 1. Consider the cyclic permutations of u = (u 1,..., u t ): u (0) = u, u (1) = (u 2,..., u t, u 1 ),..., u (t 1) = (u t, u 1,..., u t 1 ). Define the corresponding partial sums r (β) j := u (β) 1 + + u (β) j for j = 1,..., t and β = 0,..., t 1. Observe that (r (β) 1,..., r (β) t ) = (r β+1 r β, r β+2 r β,..., r t r β, [r t r β ] + r 1, [r t r β ] + r 2,..., [r t r β ] + r β ) = (r β+1 r β, r β+2 r β,..., r t r β, r t [r β r 1 ], r t [r β r 2 ],..., r t [r β r β 1 ], r t ) (1)

Spitzer s combinatorial lemma II Lemma Assume r t > 0. Let l be the number of cyclic permutations such that t is a ladder index. Then l 1. Moreover, each such cyclic permutation has exactly l ladder indices. Proof: We first show that l 1, i.e., there is at least one cyclic permutation where t is a ladder index. Let β be the smallest index achieving the maximum of r 1,..., r t, i.e., r β > r 1 r β 1 and r β r β+1 r t. From (1), and r β+i r β 0 < r t, i = 1,..., t β, r t [r β r j ] < r t, j = 1,..., β 1. Moreover, r t > 0 = r 0 by assumption. So, in u (β), t is a ladder index.

Spitzer s combinatorial lemma III Since l 1, we can assume w.l.o.g. that u is such that t is a ladder index. Then β is a ladder index in u if and only if if and only if r β > r 0 r β 1, r t > r t r β and r t [r β r j ] < r t, j = 1,..., β 1. Moreover, because r t > r j for all j, we have r t [r β+i r β ] = (r t r β+i ) + r β and the last equation is equivalent to r t > r t [r β+i r β ], i = 1,..., t β and r t [r β r j ] < r t, j = 1,..., β 1. That is, t is a ladder index in the β-th cyclic permutation.

Back to the hitting-time theorem: proof I Proof: Let R i := 1 S i and U i := 1 X i for all i = 1,..., t and let R 0 := 0. Then {X 1 + + X t = t 1} = {R t = 1}, and By symmetry, for all β {W = t} = {t is the first ladder index in R 1,..., R t}. P[t is the first ladder index in R 1,..., R t] = P[t is the first ladder index in R (β) 1,..., R (β) t ]. Let E β be the event on the last line. Hence P[W = t] = E[1 E1 ] = 1 t t E β=1 1 Eβ

Back to the hitting-time theorem: proof II Proof: By Spitzer s combinatorial lemma, there is at most one cyclic permutation where t is the first ladder index. In particular, tβ=1 1E β {0, 1}. So P[W = t] = 1 [ ] t P t β=1e β. Finally observe that, because R 0 = 0 and U i 1 for all i, the partial sum at the j-th ladder index must take value j. So the event { t β=1e β } implies that {R t = 1} because the last partial sum of all cyclic permutations is R t. Similarly, because there is at least one cyclic permutation such that t is a ladder index, the event {R t = 1} implies { t β=1e β }. Therefore, which concludes the proof. P[W = t] = 1 P [Rt = 1], t

Hitting-time theorem: example Example (Poisson branching process) Let (Z t ) be a Galton-Watson branching process with offspring distribution Poi(λ) where λ > 0. Let W be its total progeny. By the hitting-time theorem, for t 1, P[W = t] = 1 t P[X 1 + + X t = t 1] = 1 t e λt (λt)t 1 (t 1)! λt (λt)t 1 = e, t! where we used that a sum of independent Poisson is Poisson.

1 Basic definitions 2 3 4

Bond percolation on Galton-Watson trees I Let T be a Galton-Watson tree for an offspring distribution with mean m > 1. Perform bond percolation on T with density p. Theorem Conditioned on nonextinction, p c (T ) = 1 m a.s. Proof: Let C 0 be the cluster of the root in T with density p. We can think of C 0 as being generated by a Galton-Watson branching process where the offspring distribution is the law of X(1,1) i=1 I i where the I i s are i.i.d. Ber(p) and X(1, 1) is distributed according to the offspring distribution of T. In particular, by conditioning on X(1, 1), the offspring mean under C 0 is mp. If mp 1 then 1 = P p[ C 0 < + ] = E[P p[ C 0 < + T ]], and we must have P p[ C 0 < + T ] = 1 a.s. In other words, p c(t ) 1 m a.s.

Bond percolation on Galton-Watson trees II On the other hand, the property of trees {P p[ C 0 < + T ] = 1} is inherited. So by our previous lemma, conditioned on nonextinction, it has probability 0 or 1. That probability is of course 1 on extinction. So by P p[ C 0 < + ] = E[P p[ C 0 < + T ]], if the probability is 1 conditioned on nonextinction then it must be that mp 1. In other words, for any fixed p such that mp > 1, conditioned on nonextinction P p[ C 0 < + T ] = 0 a.s. By monotonicity of P p[ C 0 < + T ] in p, taking a limit p n 1/m proves the result.