Hostname: page-component-745bb68f8f-cphqk Total loading time: 0 Render date: 2025-02-11T13:42:41.318Z Has data issue: false hasContentIssue false

Giant descendant trees, matchings, and independent sets in age-biased attachment graphs

Published online by Cambridge University Press:  25 April 2022

Hüseyin Acan*
Affiliation:
Drexel University
Alan Frieze*
Affiliation:
Carnegie Mellon University
Boris Pittel*
Affiliation:
Ohio State University
*
*Postal address: Department of Mathematics, Drexel University, Philadelphia, PA 19104, USA. Email: huseyin.acan@drexel.edu
**Postal address: Department of Mathematical Sciences, Carnegie Mellon University, Pittsburgh, PA 15213, USA. Email: alan@random.math.cmu.edu
***Postal address: Department of Mathematics, Ohio State University, Columbus, OH 43210, USA. Email: bgp@math.ohio-state.edu
Rights & Permissions [Opens in a new window]

Abstract

We study two models of an age-biased graph process: the $\delta$ -version of the preferential attachment graph model (PAM) and the uniform attachment graph model (UAM), with m attachments for each of the incoming vertices. We show that almost surely the scaled size of a breadth-first (descendant) tree rooted at a fixed vertex converges, for $m=1$ , to a limit whose distribution is a mixture of two beta distributions and a single beta distribution respectively, and that for $m>1$ the limit is 1. We also analyze the likely performance of two greedy (online) algorithms, for a large matching set and a large independent set, and determine – for each model and each greedy algorithm – both a limiting fraction of vertices involved and an almost sure convergence rate.

Type
Original Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Applied Probability Trust

1. Introduction

It is widely accepted that graphs/networks are an inherent feature of life today. The classical models $G_{n,m}$ and $G_{n,p}$ of Erdős and Rényi [Reference Erdős and Rényi17] and Gilbert [Reference Gilbert22], respectively, lacked some salient features of observed networks. In particular, they failed to have a degree distribution that decays polynomially. Barabási and Albert [Reference Barabási and Albert3] suggested the preferential attachment model (PAM) as a more realistic model of a ‘real-world’ network. There was a certain lack of rigor in [Reference Barabási and Albert3], and later Bollobás, Riordan, Spencer, and Tusnády [Reference Bollobás, Riordan, Spencer and Tusnády11] gave a rigorous definition.

Many properties of this model have been studied. Bollobás and Riordan [Reference Bollobás and Riordan9] studied the diameter and proved that with high probability (w.h.p.) PAM with n vertices and $m>1$ attachments for every incoming vertex has diameter $\approx \log n/\log\log n$ . An earlier result by Pittel [Reference Pittel34] implied that for $m=1$ w.h.p. the diameter of PAM is of exact order $\log n$ . Bollobás and Riordan [Reference Bollobás and Riordan7, Reference Bollobás and Riordan8] studied the effect on component size of deleting random edges from PAM and showed that it is quite robust w.h.p. The degree distribution was studied by Móri [Reference Móri30, Reference Móri31], Flaxman, Frieze, and Fenner [Reference Flaxman, Frieze and Fenner18], and Berger, Borgs, Chayes, and Saberi [Reference Berger, Borgs, Chayes and Saberi4]. Peköz, Röllin, and Ross [Reference Peköz, Röllin and Ross32] established convergence, with rate, of the joint distribution of the degrees of finitely many vertices. Acan and Hitczenko [Reference Acan and Hitczenko1] found an alternative proof, without rate, via a memory game. Pittel [Reference Pittel36] used the Bollobás–Riordan pairing model to approximate, with explicit error estimate, the degree sequence of the first $n^{m/(m+2)}$ vertices, $m\ge 1$ , and proved that, for $m>1$ , PAM is connected with probability $\approx 1-{\textrm{O}}((\!\log n)^{-(m-1)/3})$ . Random walks on PAM have been considered in the work of Cooper and Frieze [Reference Cooper and Frieze14, Reference Cooper and Frieze15]. In [Reference Cooper and Frieze14] there are results on the proportion of vertices seen by a random walk on an evolving PAM and [Reference Cooper and Frieze15] determines the asymptotic cover time of a fully evolved PAM. Frieze and Pegden [Reference Frieze and Pegden20] used random walk in a ‘local algorithm’ to find vertex 1, improving results of Borgs et al. [Reference Borgs, Brautbar, Chayes, Khanna and Lucier12]. The mixing time of such a walk was analyzed by Mihail, Papadimitriou, and Saberi [Reference Mihail, Papadimitriou and Saberi28], who showed rapid mixing. Interpolating between Erdős–Rényi and preferential attachment, Pittel [Reference Pittel35] considered the birth of a giant component in a graph process $G_M$ on a fixed vertex set, when $G_{M+1}$ is obtained by inserting a new edge between vertices i and j with probability proportional to $[\deg (i)+\delta]\cdot[\deg (j)+\delta]$ , with $\delta>0$ being fixed. Confirming a conjecture of Pittel [Reference Pittel35], Janson and Warnke [Reference Janson and Warnke25] recently determined the asymptotic size of the giant component in the supercritical phase in this graph model.

The paragraph above gives a small sample of results on PAM that can be related to its role as a model of a real-world network. It is safe to say that PAM has now been accepted into the pantheon of random graph models that can be studied from a purely combinatorial perspective. For example, Cooper, Klasing, and Zito [Reference Cooper, Klasing and Zito16] studied the size of the smallest dominating set and Frieze, Pérez-Giménez, Prałat, and Reiniger [Reference Frieze, Pérez-Giménez, Prałat and Reiniger21] studied the existence of perfect matchings and Hamilton cycles.

One source of our inspiration was the work of Móri [Reference Móri30, Reference Móri31]; see also Katona and Móri [Reference Katona and Móri27] and van der Hofstad [Reference van der Hofstad23]. They were able to construct a family of martingales in the form of factorial products, with arguments being the degrees of individual vertices. This allowed them to analyze the limiting behavior of vertex degrees in a $\delta$ -version of PAM. In this paper we construct a new factorial-type martingale with one argument being the total size of a ‘descendant’ subtree. This is a generalization of the martingale for $\delta=0$ , found by Pittel [Reference Pittel36].

2. Our results

For each of the two models, PAM and UAM, we study the descendant tree of a given vertex v; it is a maximal subtree rooted at v and formed by increasing paths starting at v. The number of vertices in this subtree is a natural influence measure of vertex v. We also analyze the performance of two online greedy algorithms, for finding a large matching set and for a large independent set. Both algorithms are well known, and the classical random graphs are a gold mine for problems on the expected efficiency of these and similar algorithms, thanks to their homogeneity and independence. The PAM and UAM models are inherently more difficult, due to the rather limited scope of these properties. We carry out this analysis in the context of the PAM graph process described in [Reference Bollobás, Riordan, Spencer and Tusnády11] and its extension taken from [Reference van der Hofstad23, Chapter 8], and the UAM graph process; see [Reference Acan and Pittel2] and [Reference Frieze, Pérez-Giménez, Prałat and Reiniger21].

The PAM graph process, $\boldsymbol\delta$ -extension. Vertex 1 has m loops, so its degree is 2m initially. Recursively, vertex $t+1$ has m edges, and it uses them one at a time either to connect to a vertex $x\in [t]$ or to loop back on itself.

To describe the transition probabilities, let $d_{t,i-1}(x)$ denote the degree of vertex x just before the ith edge of vertex $t+1$ arrives. Let w denote the random receiving end of the ith edge emanating from vertex $t+1$ . Conditioned on $G_{m,\delta}(t)$ and the previous $i-1$ edges emanating from vertex $t+1$ , we have

(2.1) \begin{equation}\mathbb P(w=x) =\begin{cases} \dfrac{d_{t, i-1}(x) +\delta}{(2m + \delta)t + 2i-1 + i\delta/m} & \text{if $ x\in[t]$,}\\[15pt]\dfrac{d_{t,i-1}(t+1) +1+i\delta/m}{(2m + \delta)t + 2i-1 + i\delta/m} & \text{if $ x=t+1$.}\end{cases}\end{equation}

Thus one-step transition from $G_{m,\delta}(t)$ to $G_{m,\delta}(t+1)$ is the m-long sequence of choices made by the m edges emanating from vertex $t+1$ , and given $G_{m,\delta}(t)$ , these choices form a Markov sequence.

For $m=1$ , writing $d_t(x)$ for $d_{t,0}(x)$ , the probabilities above can be written a little more simply in the form

(2.2) \begin{equation}\mathbb{P}({w=x})=\begin{cases} \dfrac{d_t(x)+\delta}{(2+\delta)t+(1+\delta)} & \text{if $ x \in [t]$,}\\[14pt] \dfrac{1+\delta}{(2+\delta)t+(1+\delta)} & \text{if $ x = t+1$.}\end{cases}\end{equation}

Bollobás and Riordan [Reference Bollobás and Riordan9] discovered the following coupling between $\{G_{m,0}(t)\}_t$ for $m>1$ and $\{G_{1,0}(mt)\}_t$ . Start with the $\{G_{1,0}(t)\}$ random process and let the vertices be $v_1,v_2,\dots$ . To obtain $\{G_{m,0}(t)\}$ from $\{G_{1,0}(mt)\}$ :

  1. (1) collapse the first m vertices $v_1,\dots,v_m$ into the first vertex $w_1$ of $G_{m,0}(t)$ , the next m vertices $v_{m+1},\dots,v_{2m}$ into the second vertex $w_2$ of $G_{m,0}(t)$ , and so on;

  2. (2) keep the full record of the multiple edges and loops formed by collapsing the blocks $\{v_{(i-1)m+1}, \dots, v_{im}\}$ for each i.

By doing this collapsing indefinitely we get the jointly defined Bollobás–Riordan graph processes $\{G_{m,0}(t)\}$ and $\{G_{1,0}(mt)\}$ . The beauty of the $\delta$ -extended Bollobás–Riordan model is that, similarly, this collapsing operation applied to the process $\{G_{1,\delta/m}(mt)\}$ delivers the process $\{G_{m,\delta}(t)\}$ [Reference van der Hofstad23].

(For the reader’s convenience we present the explanation in Appendix A.)

Remark 2.1. Note that the process is well-defined for $\delta\ge -m$ , since for such $\delta$ all the probabilities defined in (2.1) are non-negative and add up to 1. For $m=1$ and $\delta=-1$ , it is easy to see from (2.2) that there is no loop in the graph except the loop on the first vertex. Hence a vertex $u>1$ starts with degree 1 and then its degree does not change, since as long as $d_t(u) + \delta = 0$ the vertex cannot attract any neighbors (again from (2.2)). As a result, in this case the graph is a star centered at vertex 1. It follows from the above coupling that $G_{m,-m}(t)$ is also a star centered at vertex 1, and the key problems we want to solve have trivial solutions in this extreme case.

The UAM graph process. Conceptually close to the preferential attachment model is the uniform attachment model (UAM). In this model, vertex $t+1$ selects uniformly at random (repetitions allowed) m vertices from the set [t] and attaches itself to these vertices. (See [Reference Acan and Pittel2] for connectivity and bootstrap percolation results.) This model can be thought of the limit of the PAM model as $\delta\to \infty$ except that loops are not allowed in this case.

2.1. Number of descendants

Fix a positive integer r and let X(t) denote the number of descendants of r at time t. Here r is a descendant of r, and x is a descendant of r if and only if x chooses to attach itself to at least one descendant of r in step x. In other words, if we think of the graph as a directed graph with edges oriented towards the smaller vertices, vertex x is a descendant of r if and only if there is a directed decreasing path from x to r. In [Reference Pittel36] X(t) was proposed as an influence measure of vertex r at time t.

We prove two theorems.

Theorem 2.1. Suppose that r is fixed, $m=1$ and $\delta>-1$ , and set $p_X(t)\,:\!=\, X(t)/t$ . Then almost surely (i.e. with probability $\textrm{1}$ ) $\lim p_X(t)$ exists, and its distribution is the mixture of two beta distributions, with parameters

\[a=1,\ b=r-\dfrac{1}{2+\delta}\quad {and}\quad a=\dfrac{1+\delta}{2+\delta},\ b=r,\]

weighted by

\[\dfrac{1+\delta}{(2+\delta)r-1}\quad {and}\quad \dfrac{(2+\delta)(r-1)}{(2+\delta)r-1}\]

respectively. Consequently almost surely $\lim_{t\to\infty}p(t)>0$ .

Remark 2.2. The two beta distributions in the theorem result from two possible scenarios for vertex r; if there is a loop on r we get the first beta distribution, otherwise we get the second one.

Note.

  1. (i) The proof is based on a new family of martingales

    \[M_{\ell}(t)\,:\!=\, \dfrac{(X(t)+{{\gamma}/{(2+\delta)}})^{(\ell)}}{(t+\beta)^{(\ell)}},\]
    where $(z)^{(\ell)}$ stands for the rising factorial. This family definitely resembles the martingales used in [Reference Móri30, Reference Móri31] for the individual vertices’ degrees. For instance, if $D_j(t)$ is the degree of vertex j at time $t\ge j$ , then for some deterministic $\gamma_k(t)$ , $Z_{j,k}(t)\,:\!=\, \gamma_k(t)(D_j(t)+\delta)^{(k)}$ is a martingale [Reference van der Hofstad23]. However, $M_{\ell}(t)$ depends on X(t), a global parameter of the PAM graph. Our initial proof was quite technical. We owe a debt of gratitude to an anonymous referee who was able to condense our argument to a few lines.
  2. (ii) With high probability $G_{1,\delta}$ is a forest of $\Theta(\!\log t)$ trees rooted at vertices with loops. For the preferential attachment tree (no loops), Janson [Reference Janson24] recently proved that the scaled sizes of the principal subtrees, those rooted at the root’s children and ordered chronologically, converge almost surely (a.s.) to the GEM-distributed random variables. His techniques differ significantly from ours.

For $m>1$ we use Theorem 2.1 to prove a somewhat surprising result that, for fixed r, almost surely all but a vanishingly small fraction of vertices are descendants of the vertex r (see [Reference Janson24]).

Theorem 2.2. Let r be fixed, $m>1$ and $\delta>-m$ , and let $p_X(t)=X(t)/t$ , $p_Y(t)=Y(t)/(2mt)$ , where $Y\big(t\big)$ is the total degree of the descendants of r at time t. Then almost surely $\lim_{t\to\infty} p_X(t)=\lim_{t\to\infty} p_Y(t)=1$ .

For the case of UAM we have the following result.

Theorem 2.3. Consider the UAM graph process $G_m(t)$ . Given a fixed $r>1$ , let $X\big(t\big)$ be the cardinality of the descendant tree rooted at vertex r, and let $p_X(t)\,:\!=\, X(t)/t$ .

  1. (i) For $m=1$ , almost surely $\lim p_X(t)$ exists, and it has the beta distribution with parameters 1 and r, which is the distribution of the minimum of r independent $[0, 1]$ -uniforms. Consequently a.s. $\liminf_{{t}\to\infty}p_X(t)>0$ .

  2. (ii) For $m>1$ , almost surely $\lim_{t\to\infty} p_X(t)=1$ .

2.2. Greedy matching algorithm

We analyze a greedy matching algorithm; a.s. it delivers a surprisingly large matching set even for relatively small m. This algorithm generates the increasing sequence $\{ M(t)\}$ of partial matchings on the sets [t], with $ M(1)=\emptyset$ . Suppose that X(t) is the set of unmatched vertices in [t] at time t. If $t+1$ attaches itself to a vertex $u\in X(t)$ , then $M(t+1)=M(t)\cup\{\{u,t+1\}\}$ , otherwise $M(t+1)=M(t)$ . (If $t+1$ chooses multiple vertices from X(t), then we pick one of those as u arbitrarily.)

First consider the PAM graph. Let

\[h(z)=h_{m,\delta}(z)\,:\!=\, 2\biggl[1-\biggl(\dfrac{m+\delta}{2m+\delta}\biggr)z\biggr]^m-z-1,\]

and let $\rho=\rho_{m,\delta}$ be the unique solution of $h(z)=0$ in the interval [0, 1]. (We have $\rho_{m,\delta}\in (0,1)$ if $\delta>-m$ .)

Theorem 2.4. Let $M\big(t\big)$ and $X\big(t\big)$ , respectively, be the set of greedy matchings and the set of uncovered vertices at time t, and let $x(t)=X(t)/t$ . For any $\delta>-m$ and $\alpha<1/3$ , almost surely

\[\lim_{t\to\infty} t^{\alpha}\max\{0, x(t)-\rho_{m,\delta}\}=0.\]

In consequence, the greedy matching algorithm a.s. finds a sequence of nested matchings $\{M(t)\}$ , where the number of vertices in $M\big(t\big)$ is asymptotically at least $(1-\rho_{m,\delta})t$ .

Remark 2.3. Observe that $\rho_{m,-m}=1$ , which makes it plausible that the maximum matching size is minuscule compared to t. In fact, by Remark 2.1, $G_{m,-m}(t)$ is the star centered at vertex 1 and hence the maximum matching size is 1.

Remark 2.4. Consider the case $\delta=0$ . Let $r_m\,:\!=\, 1-\rho_{m,0}$ ; some values of $r_m$ are

(2.3) \begin{equation} \begin{aligned}r_1&=0.5000, \ &r_2&=0.6458, \ &r_5&=0.8044,\\r_{10}&=0.8863,&r_{20}&=0.9377, &r_{70}&=0.9803. \end{aligned} \end{equation}

With a bit of calculus, we obtain $r_m=1- 2m^{-1}\log2 +{\textrm{O}}\big(m^{-2}\big)$ .

Theorem 2.5. Let $M\big(t\big)$ denote the greedy matching set after t steps of the UAM process. Let $r_m$ denote the unique positive root of $2(1-z^m) -z=0$ , i.e. $r_m=1-m^{-1}\log 2+{\textrm{O}}\big(m^{-2}\big)$ . Then, for any $\alpha<1/3$ , almost surely

\[\lim_{t\to\infty} t^{\alpha}\bigg| \dfrac{2|M(t)|}{t}-r_m\bigg|=0.\]

Some values of $r_m$ in this case are

\begin{alignat*}{5}r_1&=0.6667, \quad &r_2&=0.7808, \quad &r_5&=0.8891,\\*r_{10}&=0.9386, &r_{20}&=0.9674, &r_{35}&=0.9809.\end{alignat*}

2.3. Greedy independent set algorithm

The algorithm generates an increasing sequence of independent sets $\{I(t)\}$ on vertex sets [t]. Namely, $I(1)=\{1\}$ , and $I(t+1)=I(t)\cup\{t+1\}$ if $t+1$ does not select any of the vertices in I(t); if it does, then $I(t+1)=I(t)$ . I(t) is also a dominating set for the PAM/UAM graph with vertex set [t]; indeed, if a vertex $\tau\in [t]\setminus I(t)$ did not have any neighbor in I(t), then vertex $\tau$ would have been added to $I(\tau-1)$ at step $\tau$ . (Pittel [Reference Pittel33] analyzed the performance of this algorithm applied to the Erdős–Rényi random graph with a large but fixed vertex set.)

For the PAM case we prove the following.

Theorem 2.6. Let $w_m$ denote the unique root of $(1-w)^m-w$ in $ (0,1)$ . For any

\[ \chi\in \biggl(0,\min\biggl\{\dfrac{1}{3},\dfrac{2m+2\delta}{3(2m+\delta)}\biggr\}\biggr), \]

almost surely

\begin{equation*} \lim_{t\to\infty}t^{\chi}\bigg|\dfrac{|I(t)|}{t}-w_m\bigg|=0.\end{equation*}

Remark 2.5. Thus the limiting scaled size of the greedy independent set does not depend on $\delta$ , but the convergence rate does.

For the UAM case we prove an almost identical result.

Theorem 2.7. Let $w_m$ be the unique positive root of $-w+(1-w)^m$ in $ (0,1)$ . Then, for any $\alpha<1/3$ , almost surely

\[\lim_{t\to\infty}t^{\alpha}\bigg|\dfrac{|I(t)|}{t} - w_m\bigg|=0.\]

Remark 2.6. Let $w_m$ be the unique positive root of $(1-w)^m-w$ in (0, 1). As $m\to\infty$ ,

\[w_m= \dfrac{\log m}{m}+{\textrm{O}}\left(m^{-1}\log\log m\right).\]

So, for m large and both models, a.s. for all large enough t the greedy algorithm delivers an independent set containing a fraction $\sim{{\log m}/{m}}$ of all vertices in [t]. It was proved in [Reference Frieze, Pérez-Giménez, Prałat and Reiniger21] that for each large t with probability $1-{\textrm{o}}(1)$ , the fraction of vertices in the largest independent set in the PAM graph process and in the UAM graph process is at most $(4+{\textrm{o}}(1))\,{{\log m}/{m}}$ and $(2+{\textrm{o}}(1))\,{{\log m}/{m}}$ , respectively. Since I(t) is dominating, our results prove a.s. existence for all large t of relatively small dominating sets, of cardinality $\le (1+\varepsilon_m) t\,{{\log m}/{m}}$ , ( $\varepsilon_m\to 0$ ).

Remark 2.7. Let $\mathcal I(t)$ denote the cardinality of the largest independent set. We conjecture that for each of the processes there exists a corresponding constant c(m) such that a.s. $\lim_{t\to\infty}t^{-1}\mathcal I(t)=c(m)$ ; of course, $c(m)\ge w_m$ .

Notice that, for the UAM process, the existence of a deterministic function c(m,t), bounded away from 0 and 1 as $t\to\infty$ , such that the distribution of $\mathcal I(t)$ is sharply concentrated around tc(m,t), is not difficult to prove. Indeed, $\mathcal I(t)$ is determined by $t-1$ independent selections of m neighbors among the older vertices made by vertices $2,\dots, t$ . Obviously, $\mathcal I(t)$ meets the following two conditions: (1) changing the outcome of any such selection can affect $\mathcal I(t)$ by at most 1; (2) if $\mathcal I(t)\ge s$ , then there are s vertices that form an independent set, which depends entirely on selections made by these vertices. Using Talagrand’s general inequality [Reference Talagrand37] (see also [Reference Molloy29]), we have, for $z\le \mathbb{E}[\mathcal I(t)]$ ,

\[\mathbb{P}\bigl(|\mathcal I(t)-\mathbb{E}[\mathcal I(t)]|>z+3\sqrt{\mathbb{E}[\mathcal I(t)]}\bigr)\le 2\exp\biggl(-\dfrac{z^2}{16 \mathbb{E}[\mathcal I(t)]}\biggr).\]

That does it, since by Theorem 2.7 $\mathbb{E}[\mathcal I(t)]\ge \mathbb{E}[I(t)]\ge 0.9 w_mt$ when t is large enough and, with exponentially high probability, we have $\mathcal I(t)\le (\sigma+{\textrm{o}}(1))t$ , where $\sigma=\sigma(m)\in (0,1)$ is the root of $\sigma\log\sigma+(1-\sigma)\log(1-\sigma)+\sigma^2m/2$ ; see Lemma 9 of [Reference Frieze, Pérez-Giménez, Prałat and Reiniger21].

3. Descendant trees

Instead of referring the reader back to Section 2, we start this, and other proof sections, by formulating each claim in full.

3.1. Proof of Theorem 2.1

In this subsection we prove Theorem 2.1 restated below.

Theorem 2.1. Suppose that $m=1$ and $\delta>-1$ , and set $p_X(t)\,:\!=\, X(t)/t$ . Then almost surely (i.e. with probability $\textrm{1}$ ) $\lim p_X(t)$ exists, and its distribution is the mixture of two beta distributions, with parameters

\[a=1,\ b=r-\dfrac{1}{2+\delta}\quad {and}\quad a=\dfrac{1+\delta}{2+\delta},\ b=r,\]

weighted by

\[\dfrac{1+\delta}{(2+\delta)r-1}\quad {and}\quad \dfrac{(2+\delta)(r-1)}{(2+\delta)r-1}\]

respectively. Consequently a.s. $\lim_{t\to\infty}p_X(t)>0$ .

Proof. For $t\ge r>1$ , let $X(t)=X_{m,\delta}(t)=X_{m,\delta}(t,r)$ and $Y(t)=Y_{m,\delta}(t)=Y_{m,\delta}(t,r)$ denote the size and total degree of the vertices in the descendant set rooted at r; so $X(r)=1$ and $Y(r)\in [m,2m]$ , where m (resp. 2m) is attained when vertex r forms no loops (resp. forms m loops) at itself. Introduce $p_Y(t)={{Y(t)}/{(2mt)}}$ . This notation will be used in the proof of Theorem 2.2 as well, but of course $m=1$ in Theorem 2.1.

Here

\[Y(t)=\begin{cases}2X(t) &\text{if}\ {r}\ \text{looped on itself,}\\[4pt]2X(t)-1 &\text{if}\ {r}\ \text{selected a vertex in}\ [r-1].\end{cases}\]

$\big($ In particular, $p_Y(t) =p_X(t)+{\textrm{O}}\big(t^{-1}\big)$ . $\big)$ So, by (2.2),

\begin{align*} \mathbb P(X(t+1)&=X(t)+1\mid \circ)\\ &=\dfrac{Y(t)+\delta X(t)}{(2+\delta)t +(1+\delta)}\\[5pt] &= \begin{cases}\dfrac{(2+\delta)X(t)}{(2+\delta)t+(1+\delta)} &\text{if}\ {r}\ \text{looped on itself,}\\[12pt] \dfrac{(2+\delta)X(t)-1}{(2+\delta)t+(1+\delta)} &\text{if}\ {r}\ \text{selected a vertex in}\ [r-1],\end{cases}\end{align*}

where ‘ $\circ$ ’ indicates conditioning on prehistory (particularly X(t) and Y(t)). Thus we are led to consider the process X(t) such that

\begin{align*}\mathbb P(X(t+1)=X(t)+1 \mid \circ) &= \dfrac{(2+\delta)X(t)+\gamma}{(2+\delta)t+(1+\delta)},\\\mathbb P(X(t+1)=X(t) \mid \circ) &= 1 - \mathbb P(X(t+1)=X(t)+1 \mid \circ) ,\end{align*}

$\gamma=0$ if r looped on itself, $\gamma=-1$ if r selected a vertex in $[r-1]$ . So $\gamma$ is determined by $G_{1,\delta}(t)$ . Letting $\beta={{(1+\delta)}/{(2+\delta)}}$ , the above equation can be written as

(3.1) \begin{equation}\begin{aligned}\mathbb P(X(t+1)=X(t)+1 \mid \circ) &= \dfrac{X(t) + \gamma/(2+\delta)}{t+\beta},\\\mathbb P(X(t+1)=X(t) \mid \circ) &= 1 - \dfrac{X(t) + \gamma/(2+\delta)}{t+\beta}.\end{aligned}\end{equation}

For $\delta=0$ the following claim was proved in [Reference Pittel36]. We let $z^{(\ell)}$ denote the rising factorial $\prod_{j=0}^{\ell-1}(z+j)$ .

Lemma 3.1. Let $\beta={{(1+\delta)}/{(2+\delta)}}$ and $Z(t) = X(t) + {{\gamma}/{(2+\delta)}}$ . Then, conditioned on the attachment record during the time interval $\big[r, t\big]$ , i.e. starting with the attachment decision by vertex r, we have

\[{\mathbb{E}[{Z(t+1)^{(\ell)}\mid \circ}]}= \biggl(\dfrac{t+\beta+\ell}{t+\beta}\biggr) Z(t)^{(\ell)}.\]

Consequently

\[M_{\ell}(t)\,:\!=\, \dfrac{Z(t)^{(\ell)}}{(t+\beta)^{(\ell)}}\]

is a martingale.

We thank an anonymous referee for the following proof, which is much shorter than our original proof.

Proof. We have

\begin{align*}\mathbb{E}[(Z(t+1)^{(\ell)} \mid \circ ]&= (Z(t)+1)^{(\ell)} \dfrac{Z(t)}{t+\beta} + Z(t)^{(\ell)} \biggl( 1- \dfrac{Z(t)}{t+\beta}\biggr)\\&= (Z(t)+\ell)\, Z(t)^{(\ell)} \dfrac{1}{t+\beta} + Z(t)^{(\ell)} - Z(t)^{(\ell)} \dfrac{Z(t)}{t+\beta}\\&= Z(t)^{(\ell)} \biggl(\dfrac{\ell+t+\beta}{t+\beta}\biggr).\end{align*}

The second part follows from this immediately.

To identify the $\lim_{t\to\infty} p(t)$ , recall that the classical beta probability distribution has density

\[f(x;a,b)=\dfrac{\Gamma(a+b)}{\Gamma(a)\Gamma(b)}x^{a-1}(1-x)^{b-1},\quad x\in (0,1),\]

parametrized by two parameters $a>0$ , $b>0$ , and moments

\begin{equation*} \int_0^1 x^{\ell} f(x;a, b)\,{\textrm{d}} x=\prod_{j=0}^{\ell-1}\dfrac{a+j}{a+b+j}.\end{equation*}

We can now complete the proof of Theorem 2.1. By Lemma 3.1, we have, for all $t\ge r$ , $\mathbb E [M_{\ell}(t)\mid \gamma]=M_{\ell}(r)$ , or explicitly

\begin{equation*} \mathbb E\biggl[\dfrac{(X(t)+{{\gamma}/{(2+\delta)}})^{(\ell)}}{(t+\beta)^{(\ell)}}\Bigm| \gamma\ \biggr]=\dfrac{(1+{{\gamma}/{(2+\delta)}})^{(\ell)}}{(r+\beta)^{(\ell)}}.\end{equation*}

Since $|\gamma/(2+\delta)|\le 1$ and $X(t)\le t$ , we obtain

\[|M_{\ell}(t)|\le \biggl[\max\biggl(\dfrac{t+1}{t+\beta},\,\dfrac{t+\ell}{t+\beta+\ell-1}\biggr)\biggr]^{\ell}\Longrightarrow \sup_{t\ge 0}|M_{\ell}(t)|\le\max(1,\beta^{-\ell}).\]

Therefore, for every $\ell\ge 1$ , by the martingale convergence theorem, conditioned on $\gamma$ , a.s. there exists

(3.2) \begin{equation}\lim_{t\to\infty}\dfrac{(X(t)+{{\gamma}/{(2+\delta)}})^{(\ell)}}{(t+\beta)^{(\ell)}}\,=\!:\, \mathcal M_{\gamma,\ell}\le \max(1,\beta^{-\ell}),\quad\mathbb E[\mathcal M_{\gamma,\ell}]=\dfrac{(1+{{\gamma}/{(2+\delta)}})^{(\ell)}}{(r+\beta)^{(\ell)}}.\end{equation}

Since $X(t)\le t$ , it follows from (3.2) that a.s. there exists $\lim_{t\to\infty}X(t)/t=\lim_{t\to\infty}p_X(t)$ , and

\begin{equation*} \lim_{t\to\infty}\mathbb{E}[p_X(t)^{\ell}]=\mathbb{E}[\mathcal M_{\gamma,\ell}]=\dfrac{(1+{{\gamma}/{(2+\delta)}})^{(\ell)}}{(r+\beta)^{(\ell)}}=\prod_{j=0}^{\ell-1}\dfrac{1+{{\gamma}/{(2+\delta)}}+j}{r+\beta+j}.\end{equation*}

The sequence of the right-hand side products is the sequence of moments of a unique distribution, which is the beta distribution with parameters $1+{{\gamma}/{(2+\delta)}}$ and $r+\beta-1-{{\gamma}/{(2+\delta)}}$ . By the definition of $\gamma$ and (2.2), we have

\[\mathbb P(\gamma=0)=\dfrac{1+\delta}{(2+\delta)(r-1)+(1+\delta)}=\dfrac{1+\delta}{2r-1+\delta r}.\]

We conclude that $\lim_{t\to\infty}p_X(t)$ has the distribution which is the mixture of the two beta distributions, with parameters

\[a=1,\ b=r-\dfrac{1}{2+\delta} \quad\text{and}\quad a=\dfrac{1+\delta}{2+\delta},\ b=r,\]

weighted by

\[\dfrac{1+\delta}{(2+\delta)r-1}\quad\text{and}\quad\dfrac{(2+\delta)(r-1)}{(2+\delta)r-1}\]

respectively. This completes the proof of Theorem 2.1.

3.2. Proof of Theorem 2.2

In this subsection we prove Theorem 2.2 restated below.

Theorem 2.2. Let $m>1$ and $\delta>-m$ , and let $p_X(t)=X(t)/t$ , $p_Y(t)=Y(t)/(2mt)$ , where $Y\big(t\big)$ is the total degree of the descendants of r at time t. Then almost surely $\lim_{t\to\infty} p_X(t)=\lim_{t\to\infty} p_Y(t)=1$ .

For the proof of this theorem, we need tractable formulas/bounds for the conditional distribution of $Y(t+1)-Y(t)$ . The existence of loops in a preferential attachment graph is a minor nuisance for stating exact conditional probabilities. Hence we will start with the scenario that we do not get a loop on vertex $t+1$ , in which case we can write the exact equations easily and we will then argue that the effect of loops is negligible. Let $\mathcal E_m(t)$ be the event that vertex $t+1$ has no loop. Conditioned on $G_{m,\delta}(t)$ and $\mathcal E_m(t)$ , by the transition probabilities given in (2.1), vertex $t+1$ chooses a vertex $x\in [t]$ with probability

\begin{equation*}\dfrac{d_{t, i-1}(x) +\delta}{(2m + \delta)t + i-1}.\end{equation*}

In the following lemma we use ‘ $ \mid \circ)$ ’ to denote conditioning on $G_{m,\delta}(t)$ and $\mathcal E_m(t)$ .

Lemma 3.2. For $a\in [m]$ ,

\begin{align*}\mathbb{P}({Y(t+1) =Y(t)+m+a \mid \circ})=\binom{m}{a} \dfrac{(Y(t)+\delta X(t))^{(a)} \cdot (2mt-Y(t)+\delta(t-X(t))^{(m-a)}} {((2m+\delta)t)^{(m)}}\end{align*}

and

\[\mathbb{P}({Y(t+1) =Y(t)\mid \circ}) = \dfrac{ (2mt-Y(t)+\delta(t-X(t))^{(m)}} {((2m+\delta)t)^{(m)}}.\]

Proof. Let V(t) denote the descendants of r at time t. Vertex $t+1$ selects, in m steps, a sequence $\{v_1,\dots, v_m\}$ of m vertices from [t], with t choices for every selection. Introduce $\textbf{I}=(\mathbb{I}_1,\dots,\mathbb{I}_m)$ , where $\mathbb{I}_i$ is the indicator of the event $\{v_i \in V(t)\}$ . The total vertex degree of [t] (resp. V(t)) right before step i is $2mt + i-1$ (resp. $Y(t)+\mu_i$ , $\mu_i\,:\!=\, |\{j<i\colon \mathbb{I}_j=1\}|$ ). Conditioned on $G_{m,\delta}(t)$ , $\mathcal E_m(t)$ , and the outcomes of the previous $i-1$ steps, we have

\begin{align*}\mathbb P(\mathbb{I}_i=1)&=\dfrac{Y(t)+\delta X(t)+\mu_i}{2mt+\delta t+i-1},\\[4pt] \mathbb P(\mathbb{I}_i=0)&=\dfrac{2mt-Y(t)+\delta(t-X(t))+i-1-\mu_i}{2mt+\delta t+ i -1}.\end{align*}

Therefore a sequence $(j_1,\dots,j_m)$ will be the outcome of the loopless m-step selection with probability

\begin{align*}& \mathbb P(\textbf{I} =(j_1,\dots,j_m)) \\&\quad=\prod_{i\in [m]} ((2m+\delta)t +i-1)^{-1}\\&\quad \quad \times\prod_{i\colon j_i=1}(Y(t)+\delta X(t)+\mu_i)\times \prod_{i\colon j_i=0}(2mt-Y(t)+\delta(t-X(t))+i-1-\mu_i).\end{align*}

For $j_1+\cdots+j_m = a \in [m]$ , the second product above equals $(Y(t)+\delta X(t))^{(a)}$ and the third product equals $(2mt-Y(t)+\delta(t-X(t))^{(m-a)}$ so that

\[\mathbb P(\textbf{I} =(j_1,\dots,j_m))= \dfrac{(Y(t)+\delta X(t))^{(a)} (2mt-Y(t)+\delta(t-X(t))^{(m-a)}}{((2m+\delta)t)^{(m)}}.\]

Since the total number of admissible sequences $(j_1,\dots,j_m)$ with $j_1+\cdots+j_m = a$ is $\binom{m}{a}$ , we obtain the first formula in Lemma 3.2. The second formula is the case of $\mathbb P(\textbf{I} = (0,\dots,0))$ .

Proof of Theorem 2.2. Lemma 3.2 gives the conditional probabilities for $Y(t+1)-Y(t)$ in the event that there is no loop on vertex $t+1$ . Now let us evaluate the probability that there is no loop on vertex $t+1$ . For $0\le i \le m$ , let $\mathcal E_{i}(t)$ denote the event that no loop has been formed on vertex $t+1$ in the first i steps when vertex $t+1$ is choosing its neighbors. On the event $\mathcal E_{i-1}(t)$ , as the ith edge incident to $t+1$ is about to attach its second end to a vertex in $[t]\cup\{t+1\}$ , the total degree of vertices in [t] is $2mt+i-1$ ( $1\le i\le m$ ). So, by the transition probabilities given in (2.1),

\[\mathbb P(\mathcal E_i(t) \mid \mathcal E_{i-1}(t))= \dfrac{2mt +i-1+t\delta}{2mt+2(i-1)+t\delta +1+{{i\delta}/{m}}} .\]

Hence

(3.3) \begin{equation}\mathbb P(\mathcal E_m(t)) =\prod_{i=1}^m\dfrac{2mt +i-1+t\delta}{2mt+2(i-1)+t\delta +1+{{i\delta}/{m}}} = \prod_{i=1}^m (1 - {\textrm{O}}(1/t))= 1-{\textrm{O}}\big(t^{-1}\big).\end{equation}

It is clear from the proof of Lemma 3.2 that, given X(t) and Y(t),

\[\mathbb P_m(a)\,:\!=\, \binom{m}{a}\dfrac{(Y(t)+\delta X(t))^{(a)} (2mt-Y(t)+\delta(t-X(t))^{(m-a)}}{((2m+\delta)t)^{(m)}}\]

is a probability distribution of a random variable D, a ‘rising-factorial’ counterpart of the binomial distribution $\mathcal D=\text{Bin}(m,p=Y(t)/2mt)$ . (We have not seen this distribution in the literature.) Define the falling factorial $(x)_{\ell}=x(x-1)\cdots (x-\ell+1)$ . It is well known that $\mathbb E[(\mathcal D)_{\mu}]=(m)_{\mu} p^{\mu}$ , $(\mu\le m)$ . For D we have

(3.4) \begin{align} \mathbb E[(D)_{\mu}]&=\sum_a(a)_{\mu} \mathbb P_m(a) \notag \\[3pt] &=\dfrac{(m)_{\mu}\,(Y(t)+\delta X(t))^{(\mu)}}{((2m+\delta)t)^{(\mu)}}\cdot \sum_{a\ge \mu}\binom{m-\mu}{a-\mu} \notag \\[3pt] &\quad \times \dfrac{(Y(t)+\delta X(t)+\mu)^{(a-\mu)}((2m+\delta)t+\mu-(Y(t)+\delta X(t)+\mu))^{((m-\mu)-(a-\mu))}}{(2mt+\mu)^{(m-\mu)}} \notag \\[3pt] &=\dfrac{(m)_{\mu}\,(Y(t)+\delta X(t))^{(\mu)}}{((2m+\delta)t)^{(\mu)}},\end{align}

since the sum over $a\ge \mu$ is $\sum_{\nu\ge 0}\mathbb P_{m-\mu}(\nu)=1$ .

By Lemma 3.2, (3.3), and (3.4),

(3.5) \begin{align}\mathbb E[Y(t+1)-Y(t)\mid\circ] &= \sum_{a=1}^m (a+m) \mathbb P_m(a) + {\textrm{O}}\big(t^{-1}\big) \notag \\[4pt]&= \biggl( \sum_{a=1}^m a \mathbb P_m(a) + \sum_{a=1}^m m \mathbb P_m(a) \biggr) + {\textrm{O}}\big(t^{-1}\big) \notag \\[4pt]&= \dfrac{m(Y(t)+\delta X(t))}{(2m+\delta)t}+ m\biggl(1-\dfrac{((2m+\delta)t-Y(t)-\delta X(t))^{(m)}}{((2m+\delta)t)^{(m)}}\biggr) +{\textrm{O}}\big(t^{-1}\big)\end{align}

and

\begin{equation*} \mathbb E[X(t+1)-X(t)\mid\circ]= 1-\dfrac{((2m+\delta)t-Y(t)-\delta X(t))^{(m)}}{((2m+\delta)t)^{(m)}} + {\textrm{O}}\big(t^{-1}\big),\end{equation*}

where ‘ $\circ$ ’ indicates conditioning on $G_{m,\delta}(t)$ . (Note that $X(t+1)=X(t)$ if and only if $Y(t+1)=Y(t)$ .)

To continue the proof of Theorem 2.2, we note first that $mX(t)\le Y(t)\le 2mX(t)$ . These inequalities follow from the fact that every time a vertex joins the descendant set, the total degree of the descendant set increases by an amount $m+a$ for some $a\in [m]$ . Let

\[p(t)\,:\!=\, \dfrac{2m}{2m+\delta}\,p_Y(t)+\dfrac{\delta}{2m+\delta}\,p_X(t),\]

where $p_Y(t)=Y(t)/(2mt)$ and $p_X(t)=X(t)/t$ as defined before. By the definition of p(t) and the above inequalities, we have

\begin{equation*} \dfrac{p_X(t)}{2}\le p_Y(t)\le p_X(t) \Longrightarrow \dfrac{m+\delta}{2m+\delta}\,p_X(t)\le p(t) \le p_X(t);\end{equation*}

in particular, $p(t)\in [0,1]$ since $\delta\ge -m$ . We will also need

\[\dfrac{((2m+\delta)t-Y(t)-\delta X(t))^{(m)}}{((2m+\delta)t)^{(m)}} = (1-p(t))^m +{\textrm{O}}\big(t^{-1}\big).\]

So, using (3.5), we compute

(3.6) \begin{equation} \begin{aligned}\mathbb E[p_Y(t+1) \mid \circ]&=\mathbb E\biggl[\dfrac{Y(t+1)}{2mt}\cdot \dfrac{t}{t+1}\Bigm| \circ\biggr] \\[3pt]&=\dfrac{t}{t+1} \biggl(p_Y(t)+\dfrac{1}{2t}\bigl[1+p(t)-(1-p(t))^m\bigr]+{\textrm{O}}\big(t^{-2}\big)\biggr) \\[3pt]&=p_Y(t)+q_Y(t), \\[3pt]q_Y(t)&\,:\!=\, \dfrac{1}{2(t+1)}\bigl[1+p(t)-2p_Y(t)-(1-p(t))^m\bigr]+{\textrm{O}}\big(t^{-2}\big).\end{aligned}\end{equation}

Likewise

(3.7) \begin{equation}\begin{aligned}\mathbb E[p_X(t+1)\mid\circ]&=p_X(t)+q_X(t),\\[3pt]q_X(t)&=\dfrac{1}{t+1}\bigl[1-p_X(t)-(1-p(t))^m\bigr]+{\textrm{O}}\big(t^{-2}\big).\end{aligned}\end{equation}

Multiplying (3.6) by ${{2m}/{(2m+\delta)}}$ and (3.7) by ${{\delta}/{(2m+\delta)}}$ , and adding them, we obtain

(3.8) \begin{equation}\begin{aligned}\mathbb E[p(t+1) \mid \circ]&=p(t) +q(t),\\[3pt]q(t)& \,:\!=\, \dfrac{m+\delta}{(2m+\delta)(t+1)}\bigl[1-p(t)-(1-p(t))^m\bigr]+{\textrm{O}}\big(t^{-2}\big).\end{aligned}\end{equation}

From the first line in (3.8) it follows that

\[\sum_{t=1}^{\tau}\mathbb{E}[p(t+1)]=\sum_{t=1}^{\tau}\mathbb{E}[p(t)]+\sum_{t=1}^{\tau}\mathbb{E}[q(t)],\]

implying that

\[\limsup_{\tau\to\infty}\sum_{t\le\tau} \mathbb E[q(t)]\le \limsup_{\tau\to\infty}\mathbb{E}[p(\tau+1)]\le 1.\]

Since $1-z-(1-z)^m\ge 0$ on [0, 1], the second line in (3.8) implies that $|q(t)|\le q(t)+{\textrm{O}}\big(t^{-2}\big)$ . Since $\sum_t t^{-2}<\infty$ , we see that $\sum_t\mathbb E[|q(t)|]<\infty$ .

So a.s. there exists $Q\,:\!=\, \lim_{\tau\to\infty}\sum_{1\le t\le \tau} q(t)$ , with $\mathbb E[|Q|]\le \sum_t\mathbb E[|q(t)|]<\infty$ , that is, a.s. $|Q|<\infty$ . Introducing $Q(t+1)=\sum_{\tau\le t}q(\tau) $ , we see from (3.8) that $\{p(t+1)-Q(t+1)\}_{t\ge 1}$ is a martingale with $\sup_t |p(t+1)-Q(t+1)|\le 1 + \sum_{\tau\ge 1} |q(\tau)|$ . By the martingale convergence theorem we obtain that there exists an integrable $\lim_{t\to\infty}(p(t)-Q(t))$ , implying that a.s. there exists a random $p(\infty)=\lim_{t\to\infty}p(t)$ . Equation (3.8) also implies that

\[1\ge \mathbb E[p(\infty)]=\dfrac{m+\delta}{2m+\delta}\sum_{t\ge 1}\dfrac{1}{t+1}\mathbb E\bigl[1-p(t)-(1-p(t))^m\bigr]+ {\textrm{O}}(1).\]

Since $m+\delta>0$ and

\[\lim_{t\to\infty} \mathbb E\bigl[1-p(t)-(1-p(t))^m\bigr] =\mathbb E\bigl[1-p(\infty)-(1-p(\infty))^m\bigr],\]

and the series $\sum_{t\ge 1} t^{-1}$ diverges, we obtain that $\mathbb P(p(\infty)\in \{0,1\})=1$ .

Recall that

\[p(t)\ge \frac{m+\delta}{2m+\delta}\, p_X(t).\]

If we show that a.s. $\liminf_{t\to\infty}p_X(t)>0$ , it will follow that a.s. $p(\infty)>0$ , whence a.s. $p(\infty)=1$ , implying (by $p(t)\le p_X(t)$ ) that a.s. $p_X(\infty)$ exists, and is 1, and consequently (by the formula for p(t)) a.s. $p_Y(\infty)$ exists, and is 1.

So let us prove that a.s. $\liminf_{t\to\infty}p_X(t)>0$ . Recall that we did prove the latter for $m=1$ . To transfer this earlier result to $m>1$ we need to establish some kind of monotonicity with respect to m. The coupling described in Section 2 comes to the rescue!

Lemma 3.3. For the coupled processes $\{G_{m,\delta}(t)\}$ and $\{G_{1,\delta/m}(mt)\}$ , we have $X_{m,\delta}(t,r) \ge m^{-1} X_{1,\delta/m}(mt,mr)$ .

Proof. Let us simply write $G_1$ and $G_m$ for the two graphs $G_{1,\delta/m}(mt)$ and $G_{m,\delta}(t)$ , respectively. Similarly, write $T_1$ and $T_m$ , respectively, for the descendant tree in $G_{1,\delta/m}(mt)$ rooted at mr and the descendant tree in $G_{m,\delta}(t)$ rooted at r. If $v_a\in T_1$ , i.e. $v_a$ is a descendant of mr, then for $b=\lceil a/m \rceil $ we have $w_b=\{v_{m(b-1)+i}\}_{i\in [m]}\ni v_a$ , implying that $w_b$ is a descendant of r in $G_m$ , i.e. $w_b\in T_m$ . (The converse is generally false: if $w_b$ is a descendant of r, it does not mean that every $v_{m(b-1)+i}$ , ( $i\in [m]$ ), is a descendant of mr.) Therefore

\begin{equation*}\qquad\quad\qquad\qquad X_{m,\delta}(t,r)=|V(T_m)|\ge m^{-1} |V(T_1)|=m^{-1}X_{1,\delta/m}(mt,mr).\qquad\qquad\qquad\quad\end{equation*}

Thus, to complete the proof of the theorem, i.e. for $\delta>-m$ , we (a) use Theorem 2.1, to assert that for the process $\{G_{1,\delta/m}(t)\}$ , a.s. $\lim_{t\to\infty} p_X (t)>0$ , (b) use Lemma 3.3, to assert that a.s. $\liminf_{t\to\infty} p_X (t)>0$ for $\{G_{m,\delta}(t)\}$ as well. The proof of Theorem 2.2 is complete.

3.3. Proof of Theorem 2.3

Theorem 2.3. Consider the UAM graph process $G_m(t)$ . Given $r>1$ , let $X\big(t\big)$ be the cardinality of the descendant tree rooted at vertex r, and let $p_X(t)\,:\!=\, X(t)/t$ .

  1. (i) For $m=1$ , almost surely $\lim p_X(t)$ exists and it has the same distribution as the minimum of $(r-1)$ independent $[0, 1]$ -uniforms. Consequently a.s. $\liminf_{t\to\infty}p_X(t)>0$ .

  2. (ii) For $m>1$ , almost surely $\lim_{t\to\infty} p_X(t)=1$ .

Proof. By the definition of the UAM process, we have

(3.9) \begin{equation}\mathbb{P}(X(t+1)=X(t)+1 \mid \circ)= 1 - (1-p_X(t))^m.\end{equation}

Case (i). Consider $m=1$ . For $r=1$ , we have $p(t)\equiv 1$ . Consider $r\ge 2$ . Equation (3.9) is the case $\delta=-1$ , $\gamma=0$ of (3.1). By Lemma 3.1, we claim that

\[M(t)\,:\!=\, \dfrac{(X(t))^{(\ell)}}{t^{(\ell)}}\]

is a martingale. So arguing as in the proof of Theorem 2.1, we obtain that almost surely (a.s.) $\lim p_X(t)=p_X(\infty)$ exists, and the limiting distribution of $p_X(\infty)$ is a beta distribution with parameters $a=1$ and $b=r$ . That is, the limiting density is $r(1-x)^{r-1}$ , $x\in [0,1]$ . Therefore a.s. $\lim p_X(t)>0$ .

Case (ii). Consider $m>1$ . Clearly $G_1(t)\subset G_m(t)$ . Therefore a.s. $\liminf p_X(t)>0$ as well. Furthermore, it follows from (3.9) that

\[\mathbb{E}[p_X(t+1)\mid\circ]= p_X(t)+\dfrac{1}{t+1}\bigl[1-p_X(t)-(1-p_X(t))^m\bigr]\]

(as before, ‘ $ \mid \circ)$ ’ means conditioning on $G_m(t)$ ), which is a special case of (3.8), with ${\textrm{O}}\big(t^{-2}\big)$ dropped. So we obtain that $\mathbb{P}(p(\infty)\in \{0,1\})=1$ , which in combination with $\mathbb{P}(p_X(\infty)>0)=1$ implies that $\mathbb{P}(p_X(\infty)=1)=1$ .

4. A technical lemma

In this section we will prove Lemma 4.1. We need the following Chernoff bound for its proof (see e.g. [Reference Janson, Łuczak and Ruciński26, Theorem 2.8]).

Theorem 4.1. If $X_1,\dots,X_n$ are independent Bernoulli random variables, $X=\sum_{i=1}^n X_i$ , and $\lambda = \mathbb{E}[X]$ , then

\[\mathbb P(|X-\lambda| > \varepsilon\lambda) < 2\exp\left(-\varepsilon^2\lambda/3\right) \quad {for\ all\ } \varepsilon\in (0, 3/2).\]

Lemma 4.1. Let $\{X(t)\}_{t\ge 0}$ be a sequence of random variables such that $X(0)=0$ and $X(t+1)-X(t) \in \{0,1\}$ . Let $x(t)=X(t)/t$ and, using ‘ $ \mid \circ)$ ’ to denote conditioning on $\{x(s)\colon s\le t\}$ , let us assume

(4.1) \begin{equation}\mathbb{E}[x(t+1) - x(t) \mid \circ]\le \dfrac{h(x(t))}{t} +{\textrm{O}}\big(t^{-2}\big),\end{equation}

where h is a continuous, strictly decreasing function with $h(0)>0$ and $h(1)<0$ , so that $h\big(x\big)$ has a unique root $\rho \in(0,1)$ . Assume also that $h'(x) < -1$ in $(0,1)$ . Then, for any $\gamma<1/3$ , almost surely

\[\lim_{t\to\infty} t^{\gamma}\max\{0,x(t)-\rho\}=0.\]

Lemma 4.2. (Extensions of Lemma 4.1.)Lemma 4.1 can be extended in a couple of ways as follows.

  1. (a) If the hypothesis $X(t+1)-X(t) \in \{0,1\}$ in Lemma 4.1 is replaced with $X(t+1)-X(t) \in \{-1,1\}$ , then the conclusion of Lemma 4.1 still holds. This follows from minor modifications in the proof of Lemma 4.1.

  2. (b) If the inequality sign in (4.1)is replaced with an equality sign, i.e. under the condition

    \[ {\mathbb{E}[{x(t+1) - x(t) \mid \circ}]} = \dfrac{h(x(t))}{t} +{\textrm{O}}\big(t^{-2}\big), \]
    we have the following conclusion: for any $\gamma<1/3$ , almost surely
    \[\lim_{t\to\infty} t^{\gamma} (x(t)-\rho)=0.\]

Proof of Lemma 4.2(b). First of all, by Lemma 4.1, we have $\lim_{t\to\infty} t^{\gamma}\max\{0,x(t)-\rho\}=0$ almost surely. Second, let $g(z) = -h(1-z)$ , so that $g(0) >0$ and $g(1)<0$ , and in (0, 1), we have $g'(z) = h'(1-z)<-1$ . Letting $y(t) = 1-x(t)$ ,

\begin{align*}{\mathbb{E}[{y(t+1) -y (t) \mid \circ}]} &= {\mathbb{E}[{(1-x(t+1)) -(1- x(t)) \mid \circ}]} \\&= -\dfrac{h(x(t))}{t} +{\textrm{O}}\big(t^{-2}\big)\\&= \dfrac{g(y(t))}{t} +{\textrm{O}}\big(t^{-2}\big).\end{align*}

Applying Lemma 4.1 with $X_1(t) = t-X(t)$ , and then switching back to X(t), we see that $\lim_{t\to\infty} t^{\gamma}\max\{0, \rho-x(t)\}=0$ almost surely, as well.

Proof of Lemma 4.1. Let $\varepsilon=\varepsilon_t\,:\!=\, t^{-1/3}\log t$ . We will show

(4.2) \begin{equation}\mathbb P(x(t)>\rho+\varepsilon) \le \exp\left({-}\Theta\left(\log^3t\right)\right).\end{equation}

Once we show (4.2), the Borel–Cantelli lemma gives

\[\mathbb P(x(t)-\rho > t^{-1/3}\log t \ \text{infinitely often})=0,\]

which proves what we want. Let us prove (4.2).

For $T\in [0,t)$ , let $\mathcal E_T$ be the event that { $x(t)> \rho+\varepsilon$ and T is the last time such that $X(\tau) \le (\rho+\varepsilon/2)\tau$ }, that is,

\[X(T)\le (\rho+\varepsilon/2)T, \quad x(\tau)> \rho+\varepsilon/2\tau \ \text{for all}\ \tau\in (T,t), \quad x(t) >\rho+\varepsilon.\]

Since $X(t+1)-X(t) \in \{0,1\}$ , we have

\begin{align*}X(T)+t-T\ge X(t)>t(\rho+\varepsilon).\end{align*}

Using $X(t) = tx(t)$ above, we get

\[T(\rho+\varepsilon/2)+t-T>t(\rho+\varepsilon),\]

implying

(4.3) \begin{equation}t-T > \dfrac{t\varepsilon}{2(1-\rho)}.\end{equation}

We conclude that

\[\{x(t)>\rho+\varepsilon\}\subseteq \bigcup_{T=1}^s \mathcal E_T,\quad s=s(t)\,:\!=\, t-\biggl\lceil\dfrac{t\varepsilon}{2(1-\rho)}\biggr\rceil.\]

Now let us fix a $T\in [0,s]$ and bound $\mathbb P(\mathcal E_T)$ . The main idea of the proof is that, as long as $x(\tau)>\rho$ , by (4.1), the process $\{x(\tau)\}$ has a negative drift.

Let $\xi_\tau$ denote the indicator of the event $\{x(\tau-1)>\rho+\varepsilon/2 \text{ and } X(\tau)=X(\tau-1)+1\}$ and let $\mathcal Z_T\,:\!=\, \xi_{T+2}+\cdots+\xi_t$ . On the event $\mathcal E_T$ , the sum $\mathcal Z_T$ counts the total number of upward unit jumps ( $X(\tau)-X(\tau-1)=1$ , $\tau\in [T+2,t]$ ) and therefore

\[X(T+1) + \mathcal Z_T = X(t)\ge t(\rho+\varepsilon).\]

Since $X(T+1) \le X(T)+1\le T(\rho+\varepsilon/2)+1$ , we must have

\begin{equation*}Z_T> (\rho+\varepsilon)(t-T),\quad Z_T\,:\!=\, 1+\mathcal Z_T.\end{equation*}

Writing $p(x(\tau))\,:\!=\, \mathbb{P}({X(\tau+1) = X(\tau)+1 \mid \circ})$ ,

\begin{align*}{\mathbb{E}[{x(\tau+1) \mid \circ}]} &= p(x(\tau)) \dfrac{X(\tau)+1}{\tau+1} + (1-p(x(\tau))) \dfrac{X(\tau)}{\tau+1} \\&= \dfrac{p(x(\tau))}{\tau+1} + \dfrac{\tau x(\tau)}{\tau+1} \\& = x(\tau) + \dfrac{p(x(\tau)) - x(\tau)}{\tau+1},\end{align*}

so that $p(x(\tau)) = h(x(\tau)) + x(\tau) + {\textrm{O}}\left(\tau^{-1}\right)$ .

Recall that for $\tau \ge T+1$ we have $x(\tau) > \rho +\varepsilon/2$ . Since $h'(x)<-1$ in (0, 1), the sum $h(x) + x$ is decreasing in (0, 1). Hence, conditioning on the full record (up to and including time $\tau$ ),

\begin{align*}\mathbb P(\xi_{\tau+1}=1 \mid \circ) &=\mathbb P(X(\tau+1)=X(\tau)+1\mid \circ)\\&= h(x(\tau)) + x(\tau) +{\textrm{O}}\left(\tau^{-1}\right) \\&< h(\rho +\varepsilon/2) + \rho + \varepsilon/2 + {\textrm{O}}\left(\tau^{-1}\right)\\&= h(\rho) + (\varepsilon/2)\cdot h'(y) +\rho +\varepsilon/2 + {\textrm{O}}\left(\tau^{-1}\right) \quad \text{for some }y\in (\rho,\rho+\varepsilon/2)\\& < \rho + {\textrm{O}}\left(\tau^{-1}\right).\end{align*}

Hence the sequence $\{\xi_\tau\}$ is stochastically dominated by the sequence of independent Bernoulli random variables $B_\tau$ with parameters $\min\big(\rho+{\textrm{O}}\left(\tau^{-1}\right),1\big)$ . Consequently $Z_T$ is stochastically dominated by $1+\sum_{j=T+2}^t B_j$ , and

\[\lambda\,:\!=\, \sum_{j=T+2}^t \mathbb E[B_j]=\rho(t-T) +{\textrm{O}}(\!\log t).\]

For the choice of $\varepsilon$ we have, (4.3) gives

\[(\rho+\varepsilon)(t-T) \ge (1+\varepsilon/2)\lambda.\]

Thus, by the Chernoff bound in Theorem 4.1 and using (4.3),

\begin{align*}\mathbb{P}({\mathcal{E}_T})&\le \mathbb P(Z_T>(t-T)(\rho+\varepsilon)) \\&\le \mathbb P(1+B_{T+2}+\cdots+B_t>(t-T)(\rho+\varepsilon)) \\&\le \mathbb P(1+B_{T+2}+\cdots+B_t> (1+\varepsilon/2)\lambda) \\&\le \exp({-}\Theta(\varepsilon^2(t-T))) \le {\textrm{e}}^{-\Theta\left(\log^3 t\right)}.\end{align*}

Using the union bound on T, we complete the proof of (4.2) and of the lemma.

Note. In the next sections we turn to the two greedy algorithms for the PAM and UAM Markov graph processes. We will continue using the symbol ‘ $ \mid \circ)$ ’ to denote conditioning on a current graph G(t).

5. Greedy matching algorithm

Recall that the greedy matching algorithm (for either of two graph models) generates the increasing sequence $\{ M(t)\}$ of partial matchings on the sets [t], with $ M(1)=\emptyset$ . GivenM(t), let

\begin{align*}X(t)&\,:\!=\, \text{number of unmatched vertices at time }t,\\Y(t)&\,:\!=\, \text{total degree of unmatched vertices at time }t,\\U(t)&\,:\!=\, \text{number of unmatched vertices selected by $t+1$ from $[t]\setminus M(t)$},\\x(t)&\,:\!=\, X(t)/t,\\y(t)&\,:\!=\, Y(t)/(2mt).\end{align*}

5.1. The PAM case

Theorem 2.4. Let $X\big(t\big)$ be the number of unmatched vertices at time t in the greedy matching algorithm. For $\delta>-m$ , let $\rho_{m,\delta}$ be the unique root in $ (0,1)$ of

(5.1) \begin{equation}h(z)=h_{m,\delta}(z)\,:\!=\, 2\biggl[1-\biggl(\dfrac{m+\delta}{2m+\delta}\biggr) z\biggr]^m-z-1.\end{equation}

Then, for any $\alpha<1/3$ , almost surely

\begin{equation*} \lim_{t\to\infty} t^{\alpha}\max\{0, x(t)-\rho_{m,\delta}\}=0.\end{equation*}

In consequence, the greedy matching algorithm a.s. finds a sequence of nested matchings $\{M(t)\}$ , where the number of vertices in $M\big(t\big)$ is asymptotically at least $(1-\rho_{m,\delta})t$ .

Proof. Notice first that for $\delta>-m$ the function h(z) is decreasing on (0, 1) and $h(z)=0$ does have a unique solution in the same interval.

We will prove our claim first for a slightly different model that does not allow any loops other than at the first vertex. In this model, vertex 1 has m loops, and the ith edge of vertex $t+1$ attaches to $u\in [t]$ with probability

\[\dfrac{d_{t,i-1}(u)+\delta}{2mt+2(i-1)+t\delta}.\]

Loops not allowed except at vertex 1. In this case, since each degree is at least m, we have $Y(t)\ge mX(t)$ and hence $y(t)\ge x(t)/2$ . Also, since

\[X(t+1)=\begin{cases}X(t)+1 & \text{if $ U(t)=0$,}\\X(t)-1 & \text{if $ U(t)>0$,}\end{cases}\]

we have

(5.2) \begin{equation}\mathbb E[X(t+1) \mid \circ]= X(t)+ \mathbb P(U(t)=0 \mid \circ)-\mathbb P(U(t)>0 \mid \circ).\end{equation}

Since $\mathbb P(\text{vertex}\ t + 1\ \textrm{has some loop})= {\textrm{O}}\big(t^{-1}\big)$ , using $Y(t)\ge mX(t)$ in the last step below, by Lemma 3.2 we get

(5.3) \begin{align}\mathbb P(U(t)=0 \mid \circ)&= \mathbb P(U(t)=0 \text{ and vertex $t+1$ has no loop} \mid \circ) +{\textrm{O}}\big(t^{-1}\big) \notag\\&= (1-{\textrm{O}}\big(t^{-1}\big))\dfrac{(2mt-Y(t)+\delta t-\delta X(t))^{(m)}}{(2mt+\delta t)^{(m)}}+{\textrm{O}}\big(t^{-1}\big) \notag\\&=\dfrac{(2mt+\delta t-Y(t)-\delta X(t))^{m}}{(2mt+\delta t)^{m}}+{\textrm{O}}\big(t^{-1}\big) \notag\\&=\biggl( 1-\dfrac{2m}{2m+\delta}\,y(t)-\dfrac{\delta}{2m+\delta}\,x(t)\biggr)^m+{\textrm{O}}\big(t^{-1}\big) \notag \\&\le \biggl(1-\dfrac{m+\delta}{2m+\delta}\,x(t)\biggr)^m+ {\textrm{O}}\big(t^{-1}\big).\end{align}

Using (5.2) and (5.3) gives

(5.4) \begin{align}\mathbb E[x(t+1) \mid \circ] &\le x(t)+ \dfrac1t\biggl[2\biggl( 1-\dfrac{m+\delta}{2m+\delta}\,x(t)\biggr)^m-x(t)-1\biggr] +{\textrm{O}}\big(t^{-2}\big) \notag \\[3pt]&=x(t)+\dfrac{1}{t}h(x(t))+{\textrm{O}}\big(t^{-2}\big),\end{align}

where h(z) is as defined in (5.1). Note that $X(0) = 0$ and h(z) satisfies the conditions given in Lemmas 4.1 and 4.2, namely $h(0)>0$ , $h(1)<0$ , and $h'(z) <- 1$ for $z\in (0,1)$ . The conclusion of the theorem follows from the first part of Lemma 4.2 in this case.

Loops allowed everywhere. The above analysis is carried over to this more complicated case via an argument similar to that for the descendant trees in Section 2.1.Here is a proof sketch. First, the counterpart of (5.3) is

\begin{align*}&\mathbb P(\{U(t)=0\}\cap \{\text{no loops at }t+1\} \mid \circ)\\&\quad =\Pi_m(t)\prod_{j=0}^{m-1}\biggl(\dfrac{2mt-Y(t)+\delta t-\delta X(t)+j}{2mt+2j+1+\delta t+(j+1)\,\delta/m}\biggr)\\&\quad \le \Pi_m(t)\biggl[\biggl(1-\dfrac{m+\delta}{2m+\delta}\,x(t)\biggr)^m+ {\textrm{O}}\big(t^{-1}\big)\biggr]\\&\quad =\big(1-{\textrm{O}}\big(t^{-1}\big)\big)\biggl[\biggl(1-\dfrac{m+\delta}{2m+\delta}\,x(t)\biggr)^m+ {\textrm{O}}\big(t^{-1}\big)\biggr]\\&\quad = \biggl(1-\dfrac{m+\delta}{2m+\delta}\,x(t)\biggr)^m+{\textrm{O}}\big(t^{-1}\big);\end{align*}

see (3.3) for $\Pi_m(t)$ . Therefore we again obtain (5.4). The rest of the proof remains the same.

Remark 5.1. Let $r=r_{m,\delta}\,:\!=\, 1-\rho_{m,\delta}$ , where $\rho_{m,\delta}$ is the unique root in (0, 1) of

\[h(z)=h_{m,\delta}(z)\,:\!=\, 2\biggl[1-\biggl(\dfrac{m+\delta}{2m+\delta}\biggr)z\biggr]^m-z-1.\]

Then r is the unique root in (0, 1) of

\[f(z)=f_{m,\delta}(z)\,:\!=\, 2-z-2\biggl(\dfrac{m}{2m+\delta}+\dfrac{m+\delta}{2m+\delta}\,z\biggr)^m.\]

Thus, by Theorem 2.4, we have

\[\liminf (1-x(t)) \ge r\]

almost surely, where $1-x(t)$ is the fraction of the vertices in M(t). See (2.3) for various r values when $\delta=0$ .

Remark 5.2. When $\delta \to \infty$ , the function $f_{m,\delta}(z)$ as defined above converges to $2-z-2z^m$ in (0, 1). So it is plausible that for the case of uniform attachment model, the number of vertices in M(t) is asymptotically rt, where r is the unique root of $2-z-2z^m$ . This is in fact the case, as shown in the next theorem.

5.2. The UAM case

Theorem 2.5. Let $M\big(t\big)$ denote the greedy matching set after t steps of the UAM process. Let $r_m$ denote the unique positive root of $2(1-z^m) -z=0$ , i.e. $r_m=1-m^{-1}\log 2+{\textrm{O}}\!\left(m^{-2}\right)$ . Then, for any $\alpha<1/3$ , almost surely

\[\lim_{t\to\infty} t^{\alpha}\bigg| \dfrac{2|M(t)|}{t}-r_m\bigg|=0.\]

Proof. Let $X(t) = t-2|M(t)|$ as before. In particular, we have $X(0)=0$ and $X(1)=1$ . At each step $t\ge 2$ we check the edges incident to vertex t. If some of the edges end at vertices that do not belong to M(t), then we choose the largest (youngest) of those vertices, say w, and set $M(t)\,:\!=\, M(t-1)\cup \{(t,w)\}$ and $X(t)=X(t-1)-1$ . Otherwise $M(t)=M(t-1)$ and $X(t)=X(t-1)+1$ . Let $x(t)=X(t)/t$ be the fraction of unmatched vertices after step t. Then X(t) is a Markov chain with

\[\mathbb{P}({X(t+1)-X(t) =1 \mid X(t)}) = (1-x(t))^m ,\]

since for $X(t+1)-X(t) =1$ to happen, each of the m choices made by vertex $t+1$ must lie outside of M(t), the probability of each such choice is $1-x(t)$ , and the choices are independent of each other. With the remaining probability vertex $t+1$ chooses at least one of m vertices from X(t), in which case X(t) decreases by 1. Consequently

\[{\mathbb{E}[{X(t+1)\mid \circ}]} = X(t) + (1-x(t))^m - \left(1-(1-x(t))^m\right) = X(t) +2(1-x(t))^m-1.\]

Dividing both sides by $t+1$ , we obtain

\[{\mathbb{E}[{x(t+1) \mid \circ}]} = x(t) + \dfrac{h(x(t))}{t+1} - {\textrm{O}}\!\left(1/t^2\right),\]

where $h(z) = 2(1-z)^m-z-1$ . The function h meets the conditions of the second part of Lemma 4.2. Hence $\lim_{t\to\infty} t^{\gamma}|x(t)-\rho_m| = 0$ a.s. On the other hand, if $\rho_m$ is the unique root of h in (0, 1), then $r_m\,:\!=\, 1-\rho_m$ is the unique root of $2(1-z^m)-z$ in (0, 1). Since $x(t)-\rho_m = (1 -2 |M(t)|/t) - (1-\rho_m) = \rho_m -2 |M(t)|/t$ , we also have a.s.

\[\lim_{t\to\infty} t^{\gamma}\bigl|2|M(t)|/t-r_m\bigr| = 0.\]

This completes the proof.

6. Analysis of greedy independent set algorithm

The algorithm, for both the PAM and UAM cases, generates the increasing sequence of independent sets $\{I(t)\}$ on the sets [t], with $I(1)\,:\!=\, \{1\}$ . If vertex $t+1$ does not select a single vertex from t, we set $I(t+1)=I(t)\cup \{t+1\}$ ; otherwise $I(t+1)\,:\!=\, I(t)$ . Given I(t), let

\begin{align*}X(t)&\,:\!=\, \text{number of vertices} \le t\ \text{outside of the current independent set } I(t),\\Y(t)&\,:\!=\, \text{total degree of these outsiders} ,\\Z(t)&\,:\!=\, \text{number of insiders selected by outsiders by time}\ {t},\\U(t)&\,:\!=\, \text{number of insiders selected by vertex}\ t+1,\\x(t) &\,:\!=\, \dfrac{X(t)}{t},\quad y(t)\,:\!=\, \dfrac{Y(t)}{2mt},\quad z(t)=\dfrac{Z(t)}{mt}, \quad i(t)=\dfrac{|I(t)|}{t}.\end{align*}

Since each insider selects only among outsiders, the total degree of insiders is $m|I(t)| +Z(t)$ and

\[Y(t)=2mt- m|I(t)|-Z(t)\Longrightarrow y(t)=1- (i(t)+z(t))/2.\]

6.1. The PAM case

Theorem 2.6. Let $w_m$ denote the unique root of $-w+(1-w)^m$ in $ (0, 1)$ . For any

\[ \chi\in \biggl(0,\min\biggl\{\dfrac{1}{3},\dfrac{2m+2\delta}{3(2m+\delta)}\biggr\}\biggr), \]

almost surely

\begin{equation*} \lim_{t\to\infty}t^{\chi}\bigg|\dfrac{|I(t)|}{t}-w_m\bigg|=0.\end{equation*}

Proof. By the definition of the algorithm, we have

\[|I(t+1)|=\begin{cases}|I(t)|+1 &\text{with probability $\mathbb{P}(U(t)=0 \mid \circ)$,}\\|I(t)| &\text{with probability $\mathbb{P}(U(t)>0 \mid \circ)$.}\end{cases}\]

Now $U(t)=0$ means that vertex $t+1$ selects all m vertices from the outsiders set, so that

\begin{align*}\mathbb{P}(U(t)=0 \mid \circ)&=\dfrac{(Y(t)+\delta X(t))^{(m)}}{((2m+\delta)t)^{(m)}}+{\textrm{O}}\big(t^{-1}\big)\\&=\biggl(\dfrac{2my(t)+\delta x(t)}{2m+\delta}\biggr)^m +{\textrm{O}}\big(t^{-1}\big)\\&=\biggl(1-\dfrac{m+\delta}{2m+\delta}\, i(t)-\dfrac{m}{2m+\delta}\,z(t)\biggr)^m +{\textrm{O}}\big(t^{-1}\big).\end{align*}

The leading term in the first equality is the exact (conditional) probability of $\{U(t)=0\}$ when no loops at vertices other than the first vertex are allowed, and the extra ${\textrm{O}}\big(t^{-1}\big)$ is for our more general case when the loops are admissible. So, using $i(t)=|I(t)|/t$ , we have

(6.1) \begin{multline}\mathbb{E}[i(t+1) \mid \circ]=i(t)+\dfrac{1}{t+1}\biggl[-i(t)+\biggl(1-\dfrac{m+\delta}{2m+\delta} i(t)-\dfrac{m}{2m+\delta}z(t)\biggr)^m\biggr]+{\textrm{O}}\big(t^{-2}\big).\end{multline}

Now $Z(t+1)=Z(t)+U(t)$ , so that

\begin{align*}\mathbb{E}[Z(t+1) \mid \circ]&= Z(t) +\mathbb{E}[U(t) \mid \circ]\\&=Z(t)+m\dfrac{I(t)(m+\delta)+Z(t)}{(2m+\delta)t}+{\textrm{O}}\big(t^{-1}\big),\end{align*}

and using $z(t)=Z(t)/mt$ , we have

(6.2) \begin{align}\mathbb{E}[z(t+1) \mid \circ]=z(t) +\dfrac{1}{t+1}\biggl[-z(t) + \biggl(i(t)\dfrac{m+\delta}{2m+\delta}+z(t)\dfrac{m}{2m+\delta}\biggr)\biggr] +{\textrm{O}}\big(t^{-2}\big).\end{align}

Introduce

\[w(t)=i(t)\dfrac{m+\delta}{2m+\delta}+z(t)\dfrac{m}{2m+\delta}.\]

Multiplying (6.1) and (6.2) by ${{(m+\delta)}/{(2m+\delta)}}$ and ${{m}/{(2m+\delta)}}$ and adding the products, we obtain

(6.3) \begin{equation}\begin{aligned}\mathbb{E}[w(t+1) \mid \circ]&=w(t)+\dfrac{f(w(t))}{t+1}+{\textrm{O}}\big(t^{-2}\big),\\f(w)&\,:\!=\, \dfrac{m+\delta}{2m+\delta}\bigl[-w+(1-w)^m\bigr].\end{aligned}\end{equation}

The function f is qualitatively similar to the function h in the proof of Theorem 2.4. Indeed, f(w) is strictly decreasing, with $f(0)=1$ and $f(1)=-1$ . Therefore f(w) has a unique root $w_m\in (0,1)$ ; it is not difficult to see that

\[w_m=\dfrac{\log m}{m}[1+{\textrm{O}}(\!\log\log m/\log m)],\quad m\to\infty.\]

Let us prove that for any

\[\alpha< c(m,\delta)\,:\!=\, \min\biggl\{1,\dfrac{2m+2\delta}{2m+\delta}\biggr\}\quad (\subseteq (0,1])\]

and $A\ge A(\alpha)$ we have

(6.4) \begin{equation}\mathbb{E}\bigl[(w(t)-w_m)^2\bigr]\le A t^{-\alpha},\end{equation}

First

(6.5) \begin{align}|i(t+1)-i(t)|&=\bigg|\dfrac{|I(t+1)|}{t+1}-\dfrac{|I(t)|}{t}\bigg| \notag \\&\le \dfrac{1}{t+1}(|I(t+1)|-|I(t)|) +\dfrac{|I(t+1)|}{t+1}\notag \\&\le \dfrac{2}{t+1},\end{align}

and similarly

\begin{equation*}|z(t+1)-z(t)|\le \dfrac{2}{t+1}.\end{equation*}

Therefore $|w(t+1)-w(t)|\le 2/(t+1)$ , and consequently

\[(w(t+1)-w_m)^2\le (w(t)-w_m)^2+\dfrac{4}{t^2}+2(w(t)-w_m) (w(t+1)-w(t)).\]

So, conditioning on prehistory, we have

\begin{align*}\mathbb{E}\bigl[(w(t+1)-w_m)^2 \mid \circ\bigr]&\le (w(t)-w_m)^2+2(w(t)-w_m) \mathbb{E}[w(t+1)-w(t) \mid \circ] +{\textrm{O}}\big(t^{-2}\big)\\&=(w(t)-w_m)^2 +\dfrac{2(w(t)-w_m)}{t+1} f(w(t))+{\textrm{O}}\big(t^{-2}\big).\end{align*}

By (6.3),

\[f'(w)\le -\dfrac{m+\delta}{2m+\delta}.\]

Therefore, with

\[c(m,\delta)=\dfrac{2m+2\delta}{2m+\delta},\]

the last inequality gives

\[\mathbb{E}\bigl[(w(t+1)-w_m)^2 \mid \circ\bigr]\le \biggl(1-\dfrac{c(m,\delta)}{t+1}\biggr) (w(t)-w_m)^2 +{\textrm{O}}\big(t^{-2}\big),\]

leading to a recursive inequality

(6.6) \begin{equation}\mathbb{E}\bigl[(w(t+1)-w_m)^2\bigr]\le \biggl(1-\dfrac{c(m,\delta)}{t+1}\biggr)\mathbb{E}\bigl[(w(t)-w_m)^2\bigr] +{\textrm{O}}\big(t^{-2}\big).\end{equation}

The bound (6.4) follows from (6.6) by a straightforward induction on t. Next we use (6.4) to prove that, for $A_1$ large enough,

(6.7) \begin{equation}\mathbb{E}\bigl[(i(t+1)-w_m)^2 \mid \circ\bigr]\le A_1 t^{-\alpha}.\end{equation}

Using (6.5), we have

\[(i(t+1)-w_m)^2\le (i(t)-w_m)^2+\dfrac{4}{t^2}+2(i(t)-w_m) (i(t+1)-i(t)).\]

Consequently

\begin{align*}\mathbb{E}\bigl[(i(t+1)-w_m)^2 \mid \circ\bigr]&\le (i(t)-w_m)^2+2(i(t)-w_m) \mathbb{E}[i(t+1)-i(t) \mid \circ] +{\textrm{O}}\big(t^{-2}\big)\\&=(i(t)-w_m)^2 +\dfrac{2(i(t)-w_m)}{t+1} [-i(t)+(1-w(t))^m]+{\textrm{O}}\big(t^{-2}\big).\end{align*}

Taking expectations of both sides, and using Cauchy’s inequality and (6.4), we obtain

\begin{align*}\mathbb{E}\bigl[(i(t+1)-w_m)^2\bigr]&\le \biggl(1-\dfrac{2}{t+1}\biggr)\mathbb{E}\bigl[(i(t)-w_m)^2\bigr]\\&\quad\, +\dfrac{2}{t+1}\mathbb{E}^{1/2}\bigl[(i(t)-w_m)^2\bigr]\,\mathbb{E}^{1/2}\bigl[(w_m-(1-w(t))^m)^2\bigr]+{\textrm{O}}\big(t^{-2}\big)\\&\le \biggl(1-\dfrac{2}{t+1}\biggr)\mathbb{E}\bigl[(i(t)-w_m)^2\bigr] +\dfrac{2mA^{1/2}}{t^{1+\alpha/2}}\mathbb{E}^{1/2}\bigl[(i(t)-w_m)^2\bigr]+{\textrm{O}}\big(t^{-2}\big),\end{align*}

since $w_m=(1-w_m)^m$ , and

\[|(1-w(t))^m - (1-w_m)^m|\le m|w(t)-w_m|.\]

So it suffices to show the existence of $A_1$ such that

\[\biggl(1-\dfrac{2}{t+1}\biggr) A_1 t^{-\alpha}+\dfrac{2mA^{1/2} A_1^{1/2}t^{-\alpha/2}}{t^{1+\alpha/2}}+{\textrm{O}}\big(t^{-2}\big)\le A_1 (t+1)^{-\alpha}\]

holds for $t>t_0$ , where $t_0$ depends only on A and $\alpha$ . For large t, the above inequality becomes

\[2mA^{1/2}A_1^{1/2}+{\textrm{O}}(t^{-1-\alpha})\le A_1[(2-\alpha)+{\textrm{O}}\big(t^{-1}\big)]+{\textrm{O}}(t^{-1+\alpha}),\]

and

\[A_1> A\biggl(\dfrac{2m}{2-\alpha}\biggr)^2\]

does the job. So (6.7) is proved. Therefore, by Markov’s inequality,

(6.8) \begin{equation}\mathbb{P}(|i(t)-w_m|\ge t^{-\chi})\le At^{-\alpha+2\chi}\to 0,\quad\chi\in (0,\alpha/2).\end{equation}

This inequality already means that $i(t)\to w_m$ in probability. Let us show the considerably stronger statement that $i(t)\to w_m$ with probability 1, at least as fast as $t^{-\chi}$ , for any given $\chi<1/3$ . Pick $\beta>1$ and introduce a sequence $\{t_{\nu}\}$ , $t_{\nu}=\lfloor \nu^{\beta}\rfloor$ . By (6.8), we have

\begin{align*}\sum_{\nu\ge 1}\mathbb{P}(|i(t_{\nu})-w_m|\ge t_{\nu}^{-\chi})&\le \sum_{\nu\ge 1}A_1t_{\nu}^{-\alpha+2\chi}={\textrm{O}}\biggl(\,\sum_{\nu\ge 1}\nu^{-\beta (\alpha-2\chi)}\biggr)<\infty,\end{align*}

provided that $\beta>(\alpha-2\chi)^{-1}$ , which we assume from now. For such choice of $\beta$ , by the Borel–Cantelli lemma with probability 1 for all but finitely many $\nu$ , we have $|i(t_{\nu})-\rho|\le t_{\nu}^{-\chi}$ . Let $t\in [t_{\nu}, t_{\nu+1}]$ . By (6.5), we have

\[|i(t)-i(t_{\nu}))|={\textrm{O}}\biggl(\dfrac{t_{\nu+1}-t_{\nu}}{t_{\nu}}\biggr)\]

uniformly for all $\nu$ . So if $|i(t_{\nu})-\rho|\le t_{\nu}^{-\chi}$ , then (using $t_{\nu}=\Theta\big(\nu^{\beta}\big)$ ) we have, for $t\in [t_{\nu},t_{\nu+1}]$ ,

\begin{align*}|i(t)-w_m|&\le t_{\nu}^{-\chi}+{\textrm{O}}\biggl(\dfrac{t_{\nu+1}-t_{\nu}}{t_{\nu}}\biggr)\\&={\textrm{O}}\left(\nu^{-\beta\chi}+\nu^{-1}\right)\\&={\textrm{O}}\left(\nu^{-\min(\beta\chi,1)}\right)\\&={\textrm{O}}\left(t^{-\min(\chi,\beta^{-1})}\right).\end{align*}

Since $|i(t_{\nu})-w_m|\le t_{\nu}^{-\chi}$ holds almost surely (a.s.) for all but finitely many $\nu$ , we see then that a.s. so does the bound $|i(t)-w_m|={\textrm{O}}(t^{-\min(\chi,\beta^{-1})})$ for all but finitely many t. Now, by taking $\beta$ sufficiently close to $(\alpha-2\chi)^{-1}$ from above, we can make $\min(\chi,\beta^{-1})$ arbitrarily close to $\min(\chi, \alpha-2\chi)$ from below. It remains to notice that $\min(\chi, \alpha-2\chi)$ attains its maximum $\alpha/3$ at $\chi=\alpha/3$ . The proof of Theorem 2.6 is complete.

6.2. The UAM case

Theorem 2.7. Let $w_m$ be the unique root of $-w+(1-w)^m$ in $ (0, 1)$ . Then, for any $\alpha<1/3$ , almost surely

\[\lim_{t\to\infty}t^{\alpha}\bigg|\dfrac{|I(t)|}{t} - w_m\bigg|=0.\]

The following remark is already stated as Remark 2.3.

Remark 6.1. Thus, the convergence rate aside, almost surely the greedy independent algorithm delivers a sequence of independent sets of asymptotically the same size as for the PAM case.

Proof. Let $i(t)=|I(t)|/t$ . From the definition of the UAM process and the greedy independent set algorithm, we obtain

\[|I(t+1)|=\begin{cases}|I(t)|+1 &\text{with conditional probability $ (1-i(t))^m$,}\\|I(t)| &\text{with conditional probability $ 1-(1-i(t))^m$.}\end{cases}\]

Therefore

\[\mathbb{E} [|I(t+1)|\mid \circ ]=|I(t)| +(1-i(t))^m,\]

or equivalently

\[\mathbb{E}[i(t+1) \mid \circ]=i(t)+\dfrac{1}{t+1} [-i(t)+(1-i(t))^m ].\]

The function $-x+(1-x)^m$ differs by a constant positive factor from the function f in (6.3). The function f meets the conditions of Lemma 4.1, and $r_m$ is a unique root of f. Therefore, for any $\alpha<1/3$ , a.s. $\lim_{t\to\infty} t^{\alpha}|i(t)-r_m|=0$ .

Appendix A. Coupling $\{{\boldsymbol{G}}_{{\boldsymbol{m,}}\textbf{0}}{(\boldsymbol{t})}\}_{{\boldsymbol{t}}}$ and $\{{\boldsymbol{G}}_{1,0}({\boldsymbol{mt}})\}_{{\boldsymbol{t}}}$

In order to show that the coupling described in Section 2 really works, we can compute the probability that the ith edge of vertex $w_{t+1}$ connects to vertex $w_x$ in the coupling and compare it with the probability in (2.1). Let $\{G^{\prime}_{m,\delta}(t)\}$ denote the process obtained by collapsing the vertices of $\{G_{1,\delta/m}(mt)\}$ . Note that the $(mt+i)$ th edge of the $\{G_{1,\delta/m}(mt)\}$ process becomes the ith edge of $w_{t+1}$ after collapsing. Hence the ith edge of vertex $w_{t+1}$ connects to $w_x$ ( $x\le t$ ) if and only if the $(mt+i)$ th edge of the $\{G_{1,\delta/m}(mt)\}$ process connects $v_{mt+i}$ with one of the vertices $v_{m(x-1)+1},\dots,v_{mx}$ . Let $d_{mt+i-1}(v_y)$ denote the degree of $v_y$ ( $y\le mt+i$ ) just before the $(mt+i)$ th edge of the $\{G_{1,\delta/m}\}$ process is drawn. Also, let $D_{t,i-1}(w_x)$ denote the degree of $w_x$ at the exact same time. Hence, by definition,

\[D_{t,i-1}(w_x)=\begin{cases}\displaystyle \sum_{y=mx-m+1}^{mx} \ d_{mt+i-1}(v_y) & x\le t,\\[17pt]\displaystyle \sum_{y=mt+1}^{mt+i}\ d_{mt+i-1}(v_y) & x=t+1.\end{cases}\]

By (2.2), for $x\le t$ , the probability that $v_{mt+i}$ connects to one of the vertices $v_{m(x-1)+1},\dots,v_{mx}$ (equivalently, the probability that the ith edge of $w_{t+1}$ connects to $w_x$ ) is

\begin{align*}\dfrac {\sum_{y=mx-m+1}^{mx} \ (d_{mt+i-1}(v_y)+\delta/m)} {(2+\delta/m)(mt+i-1)+1+\delta/m} &= \dfrac{\delta+ \sum_{y=mx-m+1}^{mx}\ d_{mt+i-1}(v_y)}{\delta(t+i/m)+2mt+2i-1}\\[5pt]&=\dfrac{\delta+ D_{t,i-1}(w_x)}{\delta(t+i/m)+2mt+2i-1}.\end{align*}

Similarly, the probability that $v_{mt+i}$ selects one of $v_{mt+1},\dots,v_{mt+i}$ (equivalently, the probability that the ith edge of $w_{t+1}$ is a loop) is

\[ \dfrac {1+i\delta/m+\sum_{j=1}^{i-1} d_{mt+i-1}(v_{mt+j})} {\delta(t+i/m)+2mt+2i-1}= \dfrac {1+i\delta/m+D_{t,i-1}(w_{t+1})} {\delta(t+i/m)+2mt+2i-1}.\]

Note that the two probabilities above are the same as those in (2.1) if we replace $D_{t,i-1}(w_x)$ with $d_{t,i-1}(x)$ . Moreover, the two processes, $\{G^{\prime}_{m,\delta}(t)\}$ and $\{G_{m,\delta}(t)\}$ as defined by (2.1), both start with m loops on the first vertex, which implies $d_{1,0}(\cdot)=D_{1,0}(\cdot)$ . This gives us that $\{G_{m,\delta}(t)\}$ and $\{G^{\prime}_{m,\delta}(t)\}$ are equivalent processes, that is, at every stage they produce the same random graph.

Acknowledgements

We thank the referees for numerous critical comments that genuinely helped us to improve the presentation of our results. We are sincerely grateful to the Editors for an efficient refereeing/editorial process. We dedicate this work to the memory of Irina Pittel, the wife of the third coauthor.

Funding information

During the time of this research, the first coauthor was supported by NSF Fellowship (award no. 1502650). Research of the second author is supported in part by NSF Grant DMS 1661063.

Competing interests

There were no competing interests to declare which arose during the preparation or publication process of this article.

References

Acan, H. and Hitczenko, P. (2016). On a memory game and preferential attachment graphs. Adv. Appl. Prob. 48, 585609.CrossRefGoogle Scholar
Acan, H. and Pittel, B. (2019). On connectivity, conductance and bootstrap percolation for the uniformly random k-out, age-biased graph. Random Structures Algorithms 56, 3762.CrossRefGoogle Scholar
Barabási, I. and Albert, R. (1999). Emergence of scaling in random networks. Science 286, 509512.CrossRefGoogle ScholarPubMed
Berger, N., Borgs, C., Chayes, J. T. and Saberi, A. (2014). Asymptotic behavior and distributional limits of preferential attachment graphs. Ann. Prob. 42, 140.CrossRefGoogle Scholar
Bollobás, B. (2001). Random Graphs, 2nd edn. Cambridge University Press.CrossRefGoogle Scholar
Bollobás, B. and Riordan, O. (2000). Linearized chord diagrams and an upper bound for the Vassiliev invariants. J. Knot Theory Ramifications 9, 847853.CrossRefGoogle Scholar
Bollobás, B. and Riordan, O. (2002). Mathematical results on scale-free random graphs. In Handbook of Graphs and Networks: From the Genome to the Internet, eds S. Bornholdt and H. G. Schuster, pp. 1–34. Wiley-VCH.CrossRefGoogle Scholar
Bollobás, B. and Riordan, O. (2003). Robustness and vulnerability of scale-free random graphs. Internet Math. 1, 135.CrossRefGoogle Scholar
Bollobás, B. and Riordan, O. (2004). The diameter of a scale-free random graph. Combinatorica 4, 534.CrossRefGoogle Scholar
Bollobás, B., Borgs, C., Chayes, J. and Riordan, O. (2003). Directed scale-free graphs. In Proceedings of the 14th Annual ACM–SIAM SODA Conference, pp. 132139. ACM Press.Google Scholar
Bollobás, B., Riordan, O., Spencer, J. and Tusnády, G. (2001). The degree sequence of a scale-free random graph process. Random Structures Algorithms 18, 279290.CrossRefGoogle Scholar
Borgs, C., Brautbar, M., Chayes, J., Khanna, S. and Lucier, B. (2012). The power of local information in social networks. In International Workshop on Internet and Network Economics WINE 2012: Internet and Network Economics (Lecture Notes Comput. Sci. 7695), pp. 406–419. Springer, Berlin and Heidelberg.CrossRefGoogle Scholar
Comtet, L. (1974). Advanced Combinatorics. D. Reidel, Dordrecht.CrossRefGoogle Scholar
Cooper, C. and Frieze, A. M. (2003). Crawling on web graphs. Internet Math. 1, 5790.CrossRefGoogle Scholar
Cooper, C. and Frieze, A. M. (2007). The cover time of the preferential attachment graph. J. Combinatorial Theory B 97, 269290.CrossRefGoogle Scholar
Cooper, C., Klasing, R. and Zito, M. (2005). Lower bounds and algorithms for dominating sets in web graphs. Internet Math. 2, 275300.CrossRefGoogle Scholar
Erdős, P. and Rényi, A. (1959). On random graphs I. Publ. Math. Debrecen 6, 290297.Google Scholar
Flaxman, A., Frieze, A. M. and Fenner, T. (2005). High degree vertices and eigenvalues in the preferential attachment graph. Internet Math. 2, 120.CrossRefGoogle Scholar
Frieze, A. M. and Karoński, M. (2016). Introduction to Random Graphs. Cambridge University Press.Google Scholar
Frieze, A. M. and Pegden, W. (2017). Looking for vertex number one. Ann. Appl. Prob. 27, 582630.CrossRefGoogle Scholar
Frieze, A. M., Pérez-Giménez, X., Prałat, P. and Reiniger, B. (2019). Perfect matchings and Hamiltonian cycles in the preferential attachment model. Random Structures Algorithms 54, 258288.CrossRefGoogle Scholar
Gilbert, E. N. (1959). Random graphs. Ann. Math. Statist. 30, 11411144.CrossRefGoogle Scholar
van der Hofstad, R. (2017). Random Graphs and Complex Networks, Vol. 1. Cambridge University Press.Google Scholar
Janson, S. (2019). Random recursive trees and preferential attachment trees are random split trees. Combinatorics Prob. Comput. 28, 8199.CrossRefGoogle Scholar
Janson, S. and Warnke, L. (2021). Preferential attachment without vertex growth: emergence of the giant component. Ann. Appl. Prob. 31, 1523–1547.CrossRefGoogle Scholar
Janson, S., Łuczak, T. and Ruciński, A. (2000). Random Graphs. Wiley, New York.CrossRefGoogle Scholar
Katona, Z. and Móri, T. (2006). A new class of scale free random graphs. Statist. Prob. Lett. 76, 15871593.CrossRefGoogle Scholar
Mihail, M., Papadimitriou, C. and Saberi, A. (2006). On certain connectivity properties of the internet topology. J. Comput. System Sci. 72, 239251.CrossRefGoogle Scholar
Molloy, M. (1998). The probabilistic method. In Probabilistic Methods for Algorithmic Discrete Mathematics (Algorithms and Combinatorics 16), pp. 1–35. Springer, Berlin and Heidelberg.CrossRefGoogle Scholar
Móri, T. (2003). On random trees. Studia Sci. Math. Hungar. 39, 143155.Google Scholar
Móri, T. (2005). The maximum degree of the Barabási–Albert random tree. Combinatorics Prob. Comput. 14, 339348.CrossRefGoogle Scholar
Peköz, E., Röllin, A. and Ross, N. (2017). Joint degree distributions of preferential attachment random graphs. Adv. Appl. Prob. 49, 368387.CrossRefGoogle Scholar
Pittel, B. (1982). On the probable behaviour of some algorithms for finding the stability number of a graph. Math. Proc. Camb. Phil. Soc. 92, 511526.CrossRefGoogle Scholar
Pittel, B. (1994). Note on the heights of random recursive trees and random m-ary search trees. Random Structures Algorithms 5, 337347.CrossRefGoogle Scholar
Pittel, B. (2010). On a random graph evolving by degrees. Adv. Math. 223, 619671.CrossRefGoogle Scholar
Pittel, B. (2021) On Bollobás–Riordan random pairing model of preferential attachment graph. Random Structures Algorithms 58, 691725.CrossRefGoogle Scholar
Talagrand, M. (1995) Concentration of measure and isoperimetric inequalities in product spaces. Institut des Hautes Études Scientifiques, Publications Mathématiques 81, 73205.CrossRefGoogle Scholar