cs0702156/main.tex
1: \documentclass{amsart}
2: 
3: \usepackage{bbm}
4: \usepackage{graphicx}
5: \usepackage{verbatim}
6: %\usepackage{changebar}
7: \def\cbstart{ } \def\cbend{ } 
8: 
9: \usepackage[latin1]{inputenc}
10: 
11: \newtheorem{proposition}{Proposition}
12: \newtheorem{theorem}{Theorem}
13: \newtheorem{lemma}{Lemma}
14: \newtheorem{definition}{Definition}
15: \newtheorem{corollary}{Corollary}
16: 
17: \newcommand\ind[1]{\mathbbm{1}_{\{#1\}}}
18: 
19: \def\cal{\mathcal}
20: \def\C{{\mathbb C}}
21: \def\N{{\mathbb N}}
22: \def\R{{\mathbb R}}
23: \def\Z{{\mathbb Z}}
24: \def\P{{\mathbb P}}
25: \def\E{{\mathbb E}}
26: \def\Var{\mathrm{Var}}
27: \def\Cov{\mathrm{Cov}}
28: \def\eps{\varepsilon}
29: \def\Wdir{./}
30: \def\etal{{\em et al.}}
31: \def\T{\mathcal{T}}
32: \setcounter{tocdepth}{1}
33: 
34: \renewcommand\labelitemi{---}
35: 
36: \title[Analysis of  Steiner subtrees]{Analysis of  Steiner subtrees of Random Trees for Traceroute Algorithms}
37: 
38: \author{Fabrice  Guillemin}
39: \address[F.~Guillemin]{Orange Labs, 2, Avenue Pierre Marzin, F-22300 Lannion}
40: \email{Fabrice.Guillemin@orange-ftgroup.com}
41: \author{Philippe Robert}
42: \address[Ph.~Robert]{INRIA-Rocquencourt,  RAP project, Domaine de Voluceau, 78153 Le Chesnay, France}
43: \email{Philippe.Robert@inria.fr}
44: \urladdr{http://www-rocq.inria.fr/\~{}robert}
45: \date{\today}
46: 
47: 
48: 
49: \begin{document}
50: 
51: \begin{abstract}
52: We consider  in this  paper the problem  of discovering,  via a traceroute  algorithm, the
53: topology of a network, whose graph is  spanned by an infinite branching process.  A subset
54: of nodes  is selected  according to  some criterion.  As  a measure  of efficiency  of the
55: algorithm,  the Steiner distance  of the  selected nodes,  i.e. the  size of  the spanning
56: sub-tree of  these nodes, is investigated.  For  the selection of nodes,  two criteria are
57: considered: A node is randomly selected with a probability, which is either independent of
58: the depth of the node (uniform model)  or else in the depth biased model, is exponentially
59: decaying with  respect to  its depth.  The  limiting behavior  the size of  the discovered
60: subtree is investigated for both models.
61: \end{abstract}
62: 
63: \keywords{Traceroute Algorithm. Steiner Distance. Branching Processes.  Oscillating Behavior. Asymptotic  Expansions.}
64: 
65: \maketitle 
66: 
67: 
68: \hrule
69: 
70:  \tableofcontents 
71: 
72: \vspace{-5mm}
73: 
74: \hrule
75: 
76: \vspace{5mm}
77: 
78: \section{Introduction}
79: In  the past  ten  years, the  Internet has  known  an extraordinary  expansion and  still
80: experiences  today a  sustained  growth.  The  counterpart  of this  success  is that  the
81: different  autonomous  systems  composing  the  global Internet  have  been  independently
82: developed  by different  operators.   This raises  some  issue since  the  Internet is  by
83: construction a flat network, where the different components are interdependent in terms of
84: connectivity availability, security,  quality of service etc.  It thus  turns out that the
85: knowledge of the physical layout of a  network is of prime interest for network operators. The
86: physical topology of a component of the Internet is in general very difficult to describe.
87: To  establish a  representation of  the whole  or a  part of  the Internet,  some topology
88: exploration methods have to be  devised.  Various topology discovery experiments have been
89: initiated  by  different organizations  in  order  to infer  the  topology  of the  global
90: Internet,   notably    the   Skitter   project   by   CAIDA    \cite{caida},   the   DIMES
91: project~\cite{Dimes}  and  many other  initiatives.   The  method  generally proposed  for
92: analyzing  the topology  of  a network  is based  on  the traceroute  facility offered  by
93: routers.  Roughly speaking,  a  traceroute  procedure  consists of  sending
94: traceroute messages between hosts as follows:
95: 
96: \vspace{4mm}
97: 
98: \hrule
99: 
100: \vspace{3mm}
101: 
102: \noindent
103: {\sc Traceroute Algorithm} \\
104: If $H$ and $G$ are  hosts participating in the topology discovery experiment, 
105: $H$ sends to $G$ a traceroute message so that all the hosts/routers on the path 
106: $(H,G)$ are identified.
107: 
108: \vspace{3mm}
109: 
110: \hrule
111: 
112: \vspace{4mm}
113: 
114: The purpose  of this paper is to  investigate the efficiency of  the traceroute algorithm.
115: While a large  number of experimental papers are available in  the technical literature on
116: the  analysis of  the topology  of the  Internet, a  very few  studies  provide analytical
117: insight  into  the  efficiency  of   these  topology  discovery  methods;  see  Vespignani
118: \etal~\cite{Vespignani} for a discussion and Azzana \etal~\cite{Azzana:01} for an analysis
119: in the case of specific deterministic trees.
120: 
121: In this paper, a more realistic model is proposed to include some randomness in the degree
122: of  the nodes  of  the graph  representing the  topology  of a  network. One  specifically
123: considers a network  with a random tree architecture spanned  by a Galton-Watson branching
124: process.  \cbstart{We shall restrict the analysis to the case of offspring distributions, which have a finite second momenmt. This notably precludes the case of power law distributions with infinite second moments, typically distributions $G$ such that $\P(G \geq n) \sim C n^{-\alpha} $ with $\alpha \in (0,2)$.}\cbend
125: 
126: 
127: The Internet graph is definitely not a tree, since many studies (see the Skitter
128: project)  show that  there  is a  core  of highly  connected  routers. Nevertheless,  some
129: components of the Internet  have a  topology close to a tree structure. This is notably
130: the case  of access or collect  networks, which play  the role of capillarity  networks in
131: charge  of collecting  and distributing  traffic  between customers  and the  core of  the
132: Internet. This  latter component is not  critical for the  problem we study in  this paper
133: since core routers are easy to discover by traceroute procedures.  This is why we focus on
134: collect networks, which can be represented  by a tree architecture, spanned by a branching
135: process. In addition, to get more insight  into the topology discovery process in the case
136: of a large network, it is assumed that the underlying branching process does not terminate
137: with probability $1$; in particular the depth of the tree is infinite.
138: 
139: The discovery process is as follows: a random number of nodes are selected among the nodes
140: of the tree.  After the selected nodes have performed the traceroute algorithm, the set of
141: the nodes  discovered is  the spanning  tree of the selected  nodes. The performance
142: criterion used in  this paper is simply the  size of this sub-tree. In graph  theory it is
143: known as the  {\em Steiner distance} of the  selected nodes (with the  slight difference that
144: the selected  nodes are not  counted).  It has  been the subject  of a recent  interest by
145: Mahmoud  and  Neininger~\cite{Mahmoud2}   and  Christophi  and  Mahmoud~\cite{ChMa}  which
146: considered the asymptotic  behavior of the distance between two random  nodes of the tree.
147: Panholzer~\cite{Panholzer}, Panholzer and Prodinger~\cite{Panholzer2} proved central limit
148: theorems  when multiple  points are  considered.   The asymptotics  investigated in  these
149: papers concern the  size of the random tree.  In our paper, we will  study two situations:
150: when the size  of the tree and the number  of selected nodes go to  infinity and also when
151: the   infinite  tree   is  fixed   and   the  number   of  selected   nodes  grows.    See
152: Panholzer~\cite{Panholzer} for a thorough discussion of the literature in this domain.
153: 
154: Two stochastic models for selecting the nodes  in the network are considered. In the first
155: model,  the uniform model,  we adopt  the point  of view  of an  external observer  to the
156: network; a set of  nodes is chosen at random and a  traceroute algorithm is performed.  In
157: the second model,  the depth biased model, the  observer is located at one  node (the root
158: node) and  it chooses more  likely nodes  not too far  away.  As it  will be seen,  in the
159: uniform model , the selected nodes are  basically in the ``bottom'' of the tree where most
160: of the nodes  are, while in the second  model they are more concentrated at  the ``top'' of
161: the tree.
162: 
163: In the first model, referred to as the uniform model, nodes whose depth is less than $N>0$
164: are   randomly  chosen  with   probability  $1-\exp(-\lambda)$   for  some   $\lambda  >0$
165: independently of the position of the node  in the tree.  The quantity analyzed here is the
166: ratio $\rho_N(\lambda)$ of the mean size $\E(R_N)$ of the sub-tree discovered and the mean
167: number  $\E(T_N)$ of  nodes  of the  tree  whose depth  is less  than  $N$.  The  quantity
168: $\rho_N(\lambda)$ denotes the fraction of  the tree discovered.  The asymptotic results of
169: this paper first determine the limit  $\rho(\lambda)$ of $\rho_N(\lambda)$ as $N$ tends to
170: infinity.  In a  second step, the asymptotic behavior of  $\rho(\lambda)$ for $\lambda \to
171: 0$  is investigated.   This  last  point gives  an  indication of  the  efficiency of  the
172: algorithm when only a few nodes are selected in the topology discovery experiment.
173: 
174: For the  uniform model, it is shown  in Theorem~\ref{rho1} that, for  small $\lambda$, the
175: exploration rate $\rho(\lambda)/\lambda$ is equivalent to $\log_m\lambda$ where $m$ is the
176: mean  value of  the offspring  distribution of  a node,  so that  at the  first  order the
177: algorithm is very efficient.   A second order analysis, Proposition~\ref{vartree}, reveals
178: that the standard deviation  of the size of the discovered tree  scales with the mean size
179: of the tree, except when the  offspring distribution is deterministic. This latter case is
180: degenerate in  the sense that  the standard deviation  is negligible when compared  to the
181: mean value.
182: 
183: In the second model, referred to as the depth biased model, the probability of selecting a
184: node depends on its  depth in the tree so that the mean number  of selected nodes at depth
185: $n$ is $\alpha^n$ for some $\alpha >0$.  It is shown in Theorem~\ref{oscil} that the ratio
186: of the average of the size  $R(\alpha)$ of the sub-tree discovered and the
187: average number  of selected nodes is equivalent to  $1/(1-\alpha)$.
188: 
189: \begin{comment}
190: In 
191: general they  can be  expressed as $F(\phi(x))$  where $F$  is some periodic  function and
192: $\phi$ is a  slowly increasing function.  Technically this phenomenon  comes from the fact
193: that  a Mellin  transform has  an infinite  number of  poles on  a vertical  axis  (for an
194: analytic approach,  see~\cite{Flajolet:14}) or the  fact that a  discrete renewal theorem
195: has  to be used  (for a  probabilistic approach,  see~\cite{Mohamed:01}).  In  the present 
196: case, the corresponding oscillating term cannot be represented in such a way; this unusual
197: asymptotic behavior is obtained through an integral representation of the main quantity of
198: interest.   
199: \end{comment}
200: The  paper is  organized  as follows:  In  Section~\ref{formulation}, the  models for  the
201: selection of the nodes  of the tree are introduced.  The uniform  model is investigated in
202: Section~\ref{uniform}  and  the depth  biased  model  in  Section~\ref{biased}.  The  main
203: ingredients for the analysis of these models are Kesten-Stigum Theorem and some results on
204: the rates of convergence for Galton-Watson branching processes and a general limit theorem
205: proved in Section~\ref{Convsec}.
206: 
207: \subsection*{Acknowledgments}
208: The authors wish to thank two anonymous referees for their work,  their detailed comments 
209: have helped us a lot to improve and correct mistakes in the  first version of the paper. 
210: 
211: 
212: 
213: \section{Problem Formulation}\label{formulation}
214: Throughout this  paper, we consider  a Galton-Watson branching  process, whose graph  is a
215: tree denoted by ${\cal T}$. Each element of the $n$th generation (or $n$th level) gives birth
216: to $G$ nodes at the $(n+1)$th generation  independently of the other elements of the $n$th
217: level, where the offspring $G$ is some {\em integrable} random variable.  (See Athreya and
218: Ney~\cite{Athreya:04} and  Lyons and Peres~\cite{Lyons:07}  for an introduction  to random
219: trees.)
220: 
221:  It is  assumed  that $\P(G{=}0){=}0$ and $P(G\geq 2)>0$, in particular the  tree  is supercritical, i.e. $m=\E(G)>1$.
222: For $n\geq 0$, the variable $Z_n$ denotes  the number of nodes at level $n$, in particular
223: $Z_0=1$.  For  $1\leq\ell\leq Z_n$,  a node  of the tree  can be  represented as  a pair
224: $(n,\ell)$, where  $n$ is its  generation and $\ell$  its rank within the  generation. (For notational conventions, see
225: Neveu~\cite{Neveu:20}  for example.)  Let  ${\cal T}_{k}^{n,\ell}$  denote  the sub-tree  of
226: ${\cal T}$ with depth less than or equal to  $k$ and with root at node $(n,\ell)$. The size of
227: ${\cal T}_{k}^{n,\ell}$ is denoted by $T_{k}^{n,\ell}$.  When $(n,\ell)$ is the root node,
228: i.e. $(n,\ell)=(0,1)$, the  upper index $(0,1)$ is omitted.   With the above notation,
229: one gets easily that for all $N>1$ and $n= 1,\dots, N$
230: \begin{equation}\label{eq1st}
231: T_N=\sum_{i=0}^{n-1}Z_i+ \sum_{\ell=1}^{Z_{n}} T_{N-n}^{n,\ell}.
232: \end{equation}
233: 
234: Let us consider a counting measure $\mathcal{N}$ on the tree representing the distribution
235: of the  points selected in  the tree: For a  subset $A$ of  the nodes of the  tree, ${\cal
236: N}(A)$ denotes  the total  number of  points in $A$.  By selecting  nodes, a  sub-tree from
237: $\mathcal{T}$ is obtained  through the traceroute algorithm; this sub-tree  is  referred to as sampled
238: tree. See Figure~\ref{trafig}.
239: 
240: \begin{figure}[ht]
241: \begin{center}
242: \scalebox{.5}{\includegraphics{tree}}
243: \end{center}
244: \caption{Traceroute Algorithm.}\label{trafig}
245: \end{figure}
246: 
247: To complete  the description of the  problem, it remains to  specify how the  nodes of the
248: original tree are selected.  In the following, we shall consider two selection criteria:
249: \begin{description}
250: \item[Uniform model] Nodes are  chosen at random on all the nodes  of the tree whose depth
251: is  less than  or equal  to $N$,  $N$  being a  fixed integer.   A node  is selected  with
252: probability $1-\exp(-\lambda)$ independently of his depth  in the tree. The mean number of
253: nodes involved in the discovery experiment is then $(1-\exp(-\lambda)) (m^{N+1}-1)/(m-1)$.
254: (Recall that the mean size of the  $n$th generation is $m^n$, $n\geq 0$, where $m =\E(G)$,
255: the mean of the offspring variable $G$.)
256: 
257: To investigate the topology discovery process, we shall consider for a fixed
258: $N>0$ the  $N$ first levels  of the  original tree $\mathcal{T}$  and count the  number of
259: nodes which are discovered, given by
260: \begin{equation}\label{relfond}
261: R_N=\sum_{n=0}^N \sum_{\ell=1}^{Z_{N-n}}\ind{{\cal N}(\T_{n}^{N-n,\ell})\not=0}.
262: \end{equation}
263: In the following, we shall be particularly interested in the quantity
264: \begin{equation}
265: \label{defrhoN}
266: \rho_N(\lambda)= \frac{\E(R_N)}{\E(T_N)},
267: \end{equation}
268: i.e., the ratio of the mean number of  discovered nodes to the mean number of nodes in the
269: tree,  when the analysis  is restricted  to the  $N$ first  levels of  the tree.  Then the
270: behavior of this ratio when the number $N$ of levels tends to infinity is investigated.
271: 
272: \item[Depth biased model] Nodes at given level $n$ are selected with probability
273: $1-\exp[-(\alpha/m)^n]$ for some $\alpha\in(0,1)$.  The mean number of nodes selected at
274: level $n$ is $m^n(1-\exp[-(\alpha/m)^n]) \sim \alpha^n$ and therefore is 
275: exponentially decreasing with respect to the depth. The rational behind that is the fact
276: that, for this model, the traceroute procedure will rarely  select nodes ``far away'' from the root
277: node, in contrary to the uniform case where geometric aspects are completely ignored for the
278: selections of the hosts. 
279: 
280: By denoting by $R(\alpha)$ the total number
281: of nodes discovered, the efficiency of the traceroute algorithm is measured in this case 
282: through the ratio  of the mean $\E(R(\alpha))$ to  the average number of
283: selected nodes. The limiting behavior when the  average number of selected  nodes becomes
284: large, i.e. when $\alpha\nearrow 1$,  is investigated. 
285: 
286: \end{description}
287: Additionally it is  assumed that the root node  of the tree is always selected;  it is not
288: difficult to show that for both models described above, the root node belongs to
289: the sample tree  with a very high probability and then the above assumption is not really restrictive.  This implies that a node $(n,\ell)$ of
290: the tree  ${\cal T}$ at  level $n$  belongs to the  sampled tree whenever  ${\cal N}({\cal
291:   T}_{N-n}^{n,\ell})$ is not  $0$.  In other words, a node of  the original tree belongs
292: to the discovered tree if at least one of his descendants has been selected. \cbstart In the following, we shall use the following notation: for a subtree $\T^{N-n,\ell}_{n}$ rooted at a vertex $(N-n,\ell)$ of the $(N-n)$th generation of the tree $\T$ and with depth $n$, the quantity $\P(\mathcal{N}(\T^{N-n,\ell}_n)\neq 0)$ is the probability that a least one vertex of the subtree $\T^{N-n,\ell}_{n}$ is marked and $(N-n, \ell)\in \T$.\cbend
293: 
294: Before proceeding to the analysis of the  topology discovery process, we prove in the next
295: section  a technical  result, which  is important  in  the analysis  of the  speed of  the
296: exploration process.
297: 
298: 
299: 
300: \section{A Convergence Result}\label{Convsec}
301: To prove asymptotic expansions in the following sections, the
302: following proposition will repeatedly be used. Its proof is based on integral
303: representations and Fubini's Theorem instead of complex analysis techniques as it is
304: usually the case in the context of harmonic series.  See Robert~\cite{Robert:09} for a presentation of
305: these methods.  
306: \begin{proposition}\label{asympprop}
307: Let $V$  be a positive  random variable with  $\E(V^2)<+\infty$ and $h$ be  a non-negative
308: twice  differentiable function  on $\R_+$  such that  $h(0)=0$. In  addition,  it is
309:   assumed that  the function  $h'$ is integrable  with $h'(0)\not=0$  and \cbstart that there exists some constant $K>0$ such that $|h''(x)|<K$  for all $x\in [0,\infty)$. \cbend
310: 
311: The function $\Psi(h)(x)$ defined by
312: \begin{equation}\label{Branaire}
313: \Psi(h)(x)=\sum_{n=0}^{+\infty} \frac{1}{m^n}\E\left(h\left(x Vm^n\right)\right),\quad x\geq 0,
314: \end{equation}
315: is such that 
316: \[
317: \lim_{x\to 0} \frac{\Psi(h)(x)}{x\log_m(1/x)} = \E(V)h'(0).
318: \]
319: \end{proposition}
320: \begin{proof}
321: Since $h$ is non-negative and $|h'|$ integrable with respect to Lebesgue measure on $\R_+$,
322: Fubini's Theorem applied twice  shows that $\Psi(h)$ can be expressed as 
323: \begin{align}
324: \Psi(h)(x)&=\sum_{n=0}^{+\infty} \frac{1}{m^n}\E\left(h\left(x Vm^n\right)\right)=
325: \E\left(\sum_{n=0}^{+\infty} \frac{1}{m^n} h\left(x Vm^n\right)\right)\notag\\
326: &= \E\left(\sum_{n=0}^{+\infty} \frac{1}{m^n} \int_0^{+\infty} h'(u)\ind{u\leq x
327:   Vm^n}\,du \right)\notag \\
328: &= \E\left(  \int_0^{+\infty} h'(u)\sum_{n=0}^{+\infty} \frac{1}{m^n} \ind{u\leq x
329:   Vm^n}\,du \right).\label{Lang}
330: \end{align}
331: The function $\Psi(h)$ is thus well defined. 
332: 
333: Since $h'(0)>0$,  Fatou's Lemma applied successively gives the relation
334: \begin{multline*}
335: \liminf_{x\to 0} \frac{\Psi(h)(x)}{x}\geq 
336: \sum_{n=0}^{+\infty} \liminf_{x\to 0} \frac{m-1}{m^n}\E\left(\frac{h\left(x
337:   Vm^n\right)}{x}\right)\\
338: \geq \sum_{n=0}^{+\infty} \frac{m-1}{m^n}\E\left(\liminf_{x\to 0}\frac{h\left(x
339:   Vm^n\right)}{x}\right)
340: =\sum_{n=0}^{+\infty} (m-1)\E\left(V\right)h'(0)=+\infty,
341: \end{multline*}
342: therefore the ratio $\Psi(h)(x)/x$ diverges as $x\to 0$. 
343: 
344: By using representation~\eqref{Lang} of $\Psi(h)$, we have
345: \begin{multline*}
346: (m-1)\Psi(h)(x)
347: = m\E\left(\int_{0}^{xV} h'(u)\,du \right)\\+\E\left(\int_{xV}^{V}
348: \frac{1}{m^{\lfloor \log_m(u/xV)\rfloor}}h'(u)\,du \right)+ \E\left(\int_{V}^{+\infty}
349: \frac{1}{m^{\lfloor \log_m(u/xV)\rfloor}}h'(u)\,du \right),
350: \end{multline*}
351: where $\lfloor y \rfloor$ is the integer part of $y\in\R$. One first shows that only the
352: central term of the right hand side plays a role in the asymptotic behavior of $\Psi(h)$
353: at the first order. 
354: 
355: For the first term, note that, if $\|h''\|_{\infty}$ is the $L_\infty$ norm of $h''$,
356: \begin{multline*}
357: \left|\frac{1}{x}\E\left(\int_{0}^{xV} h'(u)\,du \right)\right|
358: \\\leq \frac{1}{x}\E\left(\int_{0}^{xV} \left(h'(0)+u \|h''\|_{\infty}\right)\,du \right)
359: \leq h'(0)\E(V)+\frac{x}{2}\E(V^2)\|h''\|_{\infty}
360: \end{multline*}
361: 
362: For $u\geq V$, one has 
363: \[
364: x m^{\lfloor \log_m(u/xV)\rfloor}\geq x m^{\lfloor \log_m(1/x)\rfloor}\geq 
365: x m^{\log_m(1/x)-1}=m^{-1},
366: \]
367: and hence,
368: \[
369: \frac{1}{x}\E\left(\int_{V}^{+\infty} 
370: \frac{1}{ m^{\lfloor \log_m(u/xV)\rfloor}}|h'(u)|\,du \right)
371: \leq m \int_0^{+\infty} |h'(u)|\, du.
372: \]
373: By gathering these estimations, it follows that the following equivalence 
374: \begin{multline*}
375: \frac{\Psi(h)(x)}{x}\sim \E\left(\int_{xV}^{V}  \frac{1}{x m^{\lfloor \log_m(u/xV)\rfloor}}h'(u)\,du \right) 
376: \\ = \E\left(V\int_{xV}^{V} m^{\{ \log_m(u/xV)\}}\frac{h'(u)}{u}\,du\right),
377: \end{multline*}
378: holds as $x\to 0$, 
379: with $\{y\}=y-\lfloor y \rfloor$, the fractional value of $y\in\R$. The above equivalence can be rewritten as
380: \begin{multline*}
381: \frac{\Psi(h)(x)}{x}\sim\E\left(V\int_{xV}^{V} m^{\{ \log_m(u/xV)\}}\frac{h'(u)-h'(0)}{u}\,du\right)
382: \\ +h'(0)\E\left(V\int_{xV}^{V} \frac{m^{\{ \log_m(u/xV)\}}}{u}\,du\right)
383: \end{multline*}
384: Due to  the boundedness of  $h''$ and the  integrability of $V^2$,  the first term  in  the
385: right hand side of the above equation is  bounded as $x$ goes to $0$. Hence, only the second term
386: has to be considered. For $x<1$, we have
387: \begin{align*}
388: \int_{xV}^{V} &\frac{m^{\{ \log_m(u/xV)\}}}{u}\,du
389: =\int_1^{1/x} \frac{m^{\{\log_m(u)\}}}{u}\,du\\
390: &=\sum_{{k\geq 0:m^k\leq 1/x}}
391: \int_{m^k}^{m^{k+1}} \frac{m^{\{\log_m(u)\}}}{u}\,du+O(1)
392: =(m-1)\lfloor -\log_m(x)\rfloor +O(1)
393: \end{align*}
394: and the result follows.
395: \end{proof}
396: 
397: Asymptotic behavior of algorithms with an underlying   tree   structure has been
398: extensively investigated,   see   Flajolet  \etal~\cite{Flajolet:14},   Mohamed   and
399: Robert~\cite{Mohamed:01}  and Mahmoud~\cite{Mahmoud:02}  for a  general  presentation.  
400: By using the terminology of Flajolet \etal~\cite{Flajolet:14}, for non-negative sequences
401: $(\lambda_n)$ and $(\mu_n)$, a series like
402: \begin{equation}\label{eqaus}
403: G(x)=\sum_{n\geq 0} \lambda_n g(\mu_n x),
404: \end{equation}
405: for some function $g$ is defined as an {\em harmonic sum}. Because of the integration of the
406: random variable $V$ and given that one wants the weakest assumptions on this random variable,
407: series~\eqref{Branaire} could be seen as a \cbstart special case \cbend  of harmonic sums. 
408: The fact that the sequences $(\lambda_n)$ and $(\mu_n)$ are specific in Expression~\eqref{Branaire} 
409: is not a real restriction, see Robert~\cite{Robert:09}.  
410: 
411: Flajolet \etal~\cite{Flajolet:14} derives the asymptotic expansion of $G(x)$  when
412: $x$ goes to $0$ or $+\infty$ by using Mellin transform techniques. For $s\in\C$, if
413: $h^*(s)$ is the Mellin transform of $h$, i.e. for $s$ in some vertical strip of $\C$,
414: \[
415: h^*(s)=\int_0^{+\infty} h(x) \,x^{s-1}\,dx,
416: \]
417: it is easy to check that the Mellin transform of $\Psi(h)$ is given by
418: \[
419: \Psi(h)^*(s)=\frac{1}{1-m^{-(s+1)}}\E\left(V^{-s}\right) h^*(s).
420: \]
421: Following the methods of Flajolet \etal~\cite{Flajolet:14}, to derive  the asymptotic behavior
422: of $\Psi(h)(x)$ as $x$ goes to infinity, one has to identify the first singularity of
423: $\Psi(h)^*$ on the right of the maximal vertical strip where it is defined. In particular, 
424: some conditions on the finiteness of some fractional moments of the random variable $V$ have to be
425: assumed (as well as growth conditions on $h^*$). From this point of view, our approach is
426: minimal since only the finiteness of $\E(V^2)$ and differentiability  conditions on $h$ are assumed. It turns out that it is important
427: as it will be seen in the following sections, since in practice little is known on the
428: fractional moments of the corresponding variable $V$. 
429: 
430: \section{The Exploration Rate in the Uniform Model}\label{uniform}
431: In this section, nodes are selected at  random with uniform probability in
432: the tree  with depth  less than $N$.   The variable  $R_N$ is the  size of  the underlying
433: sub-tree  (or sampled tree)  containing the  selected nodes.   The asymptotic  behavior of
434: $\rho_N(\lambda)=\E(R_N)/\E(T_N)$,  the fraction of  discovered nodes,  when $N$  tends to
435: infinity   is  investigated.    In   the  second   part   of  this   section,  the   ratio
436: $\mathrm{var}(R_N)/\E(T_N)$ is analyzed.
437: 
438: \subsection{First Order Asymptotics}
439: 
440: In the uniform case, the limiting behavior of the ratio $\rho_N(\lambda)$ when $N$ tends to infinity is given by the following result.
441: 
442: \begin{theorem}\label{ThFirst}\label{rho1}
443: The  ratio of the average size $R_N$ of the sampled tree  to
444: the total average size of the tree $\E(T_N)$ satisfies the relation
445: \begin{equation}
446: \label{defrholambda}
447: \rho(\lambda)\stackrel{\text{def.}}{=}\lim_{N\to+\infty}\rho_N(\lambda) =\sum_{n=0}^{+\infty}    \frac{m-1}{m^{n+1}}\left(1-\E\left(\exp\left(-\lambda
448:   \sum_{i= 0}^{n}Z_i\right)\right)\right).
449: \end{equation}
450: If additionally the condition $\E\left(G^2\right)<+\infty$ holds then
451: \begin{equation}\label{rate}
452: \lim_{\lambda \to 0} \frac{\rho(\lambda)}{\lambda\log_m(1/\lambda)} = 1.
453: \end{equation}
454: \end{theorem}
455: 
456: Relation~\eqref{rate} shows that the rate of increase of the discovery process is infinite
457: near the origin. This implies that with only  a few selected nodes one has the impression of rapidly discovering the whole network.
458: 
459: \begin{proof}
460: By conditioning on the tree, the conditional probability that node $(N-n,\ell)$ does not
461: belong to the sampled tree is 
462: \[
463: \left.\P\left( \rule{0mm}{4mm}{\cal N}\left({\cal T}^{N-n,\ell}_{n}\right)\not=0\right| {\cal T}\right)=
464: 1-\exp\left(-\lambda T_{n}^{N-n,\ell}\right).
465: \]
466: By summing-up these relations, one obtains  that the expected value of $R_N$, i.e., the average
467: number of nodes in the sampled tree, is given by
468: \begin{align*}
469: \E(R_N)&=\sum_{n=0}^{N} \E(Z_{N-n})
470: \left(1-\E\left(\exp\left(-\lambda T_{n}\right)\right)\right)\\
471: &=\sum_{n=0}^{N} m^{N-n}\left(1-\E\left(\exp\left(-\lambda   \sum_{i = 0}^{n}Z_i\right)\right)\right).
472: \end{align*}
473: The limit when $N \to \infty$ of the ratio $\rho_N(\lambda)$ is then given by
474: $$
475: \rho(\lambda)\stackrel{\text{def.}}{=}\lim_{N\to +\infty} \rho_N(\lambda) =\sum_{n=0}^{+\infty}
476:   \frac{m-1}{m^{n+1}}\left(1-\E\left(\exp\left(-\lambda  \sum_{i=0}^{n}Z_i\right)\right)\right)
477: $$
478: since $\E(T_N) \sim m^{N+1}/(m-1)$ for large $N$. This proves the first equality stated in Theorem~\ref{ThFirst}.
479: 
480: We now study the behavior of $\rho(\lambda)$ when $\lambda$ goes to $0$.  Since $\E(G^2)<+\infty$, 
481: Kesten-Stigum's Theorem ensures the existence of a
482: random variable $W$ such that $\P(W>0)=1$ (because of the assumption on the
483: distribution of $G$) and $\E(W)=1$ (See Lyons and Peres~\cite{Lyons:07}) and that, almost surely,
484: \begin{equation}
485: \label{defW}
486: \lim_{n\to+\infty} \frac{Z_n}{m^n}=W.
487: \end{equation}
488:  Let us define 
489: $$
490: f(\lambda)\stackrel{\text{def.}}{=}\sum_{n=0}^{+\infty}
491:  \frac{m-1}{m^{n+1}}\left(1-\E\left(\exp\left(-\lambda W \frac{m^{n+1} -1}{m-1}\right)\right)\right).
492: $$
493: Then,
494: \begin{multline}\label{aux:1}
495: \frac{|\rho(\lambda)-f(\lambda)|}{\lambda}  \\ \leq 
496: \sum_{n=0}^{+\infty}
497:   \frac{m-1}{\lambda m^{n+1}}\left|\E\left(\exp\left(-\lambda
498:   \sum_{i=0}^{n}Z_i\right)\right)-\E\left(\exp\left(-\lambda W
499:     \frac{m^{n+1}-1}{m-1}\right)\right)\right|.
500: \end{multline}
501: 
502: Since $W$ is integrable, Lebesgue's dominated convergence Theorem gives that
503: \begin{multline}\label{eqaux:1}
504: \lim_{\lambda\to 0} \frac{1}{\lambda}
505: \E\left(\exp\left(-\lambda
506:   \sum_{i=0}^{n}Z_i\right)-\exp\left(-\lambda W \frac{m^{n+1}-1}{m-1}\right) \right)\\=
507: \E\left(\sum_{i=0}^{n}Z_i- W \frac{m^{n+1}-1}{m-1}\right)=0.
508: \end{multline}
509: We have
510: \begin{multline*}
511: \frac{1}{m^{n+1}\lambda}
512: \left|\E\left(\exp\left(-\lambda\sum_{i=0}^{n}Z_i\right)\right)-\E\left(\exp\left(-\lambda W \frac{m^{n+1}-1}{m-1}\right)\right)\right|
513: \\ \leq \frac{1}{m^{n+1}} \sum_{i=0}^{n}\E |Z_i -m^i W|.
514: \end{multline*}
515: From Athreya and Ney~\cite[Theorem~1, page~54]{Athreya:04},  for $n\geq 1$, there
516: exists a sequence $(W^i)$ of i.i.d. random variables with the same distribution as $W$
517: such that
518: \begin{equation}\label{CV}
519: Z_n-m^n W=\sum_{i=1}^{Z_n} (1-W^{i}).
520: \end{equation}
521: By using Cauchy-Shwartz's Inequality, we obtain
522: \begin{align}
523: \E\left(\left|Z_n-m^n W\right|\right)|&
524: \leq  \sqrt{\E\left((Z_n -m^n   W)^2\right)}\notag \\
525: &=\Var(1-W)\sqrt{\E(Z_n)}\notag \\
526: &=\Var(1-W) m^{n/2}.\label{CS}
527: \end{align}
528: From the above inequality, we deduce that 
529: \begin{multline*}
530: \frac{1}{m^{n+1}\lambda}
531: \left|\E\left(\exp\left(-\lambda\sum_{i=0}^{n}Z_i\right)\right)-\E\left(\exp\left(-\lambda W \frac{m^{n+1}-1}{m-1}\right)\right)\right|
532: \\\leq \frac{\Var(1-W)}{\sqrt{m}-1} \frac{1}{m^{(n+1)/2}}.
533: \end{multline*}
534: Relation~\eqref{eqaux:1} and Lebesgue's Theorem then imply that
535: \[
536: \lim_{\lambda\to 0}\frac{\rho(\lambda)-f(\lambda)}{\lambda}=0.
537: \]
538: Hence, up to an expression which is of the order of $o(\lambda)$, the behavior at $0$ of
539: $\rho(\lambda)$ is equivalent to  the behavior of $f(\lambda)$ as $\lambda$ becomes  small. 
540: 
541: 
542: By using Proposition~\ref{asympprop}, we have by taking $h(u)=1-e^{-u}$  and $V=W m/(m-1)$,
543: \[
544: \sum_{n=0}^{+\infty} \frac{m-1}{m^{n+1}}\left(1-\E\left(\exp\left(-x V m^{n}\right)\right)\right) =\Psi(h)(x)\sim x \log_m (1/x)
545: \]
546: as $x \to 0$. To conclude the proof, we note that
547: $$
548: \lim_{x \to 0} \frac{\Psi(h)(x) -f(x)}{x\log_m x}=0
549: $$
550: and the result follows.
551: \end{proof}
552: 
553: \subsection{Second Order Properties}
554: 
555: The results obtained in the previous section show that the size of the sampled tree is of
556: the same order of magnitude as the original tree when the probability of selecting a node is fixed. When this probability  is very small (i.e., for small $\lambda$), the
557: speed of the discovery process is even very fast. In this section, we evaluate the second
558: moment of the random variable $R_N$ in order to estimate the dispersion of the size of the
559: sampled tree around the mean value. 
560: 
561: \cbstart In the rest of this section, we use the following notation: If  $(n,\ell)$  and  $(n',\ell')$  are  two  nodes of  the
562: tree, the relation  $(n',\ell') < (n,\ell)$ indicates that  the nodes are distinct and
563: that $(n',\ell')$ is a node of the sub-tree whose root is $(n,\ell)$. \cbend
564: 
565: \begin{proposition}[Asymptotic behavior of the variance]\label{vartree}
566: When the size $N$ of the original tree goes to infinity, the variance of the size of the
567: sampled tree is such that
568: \begin{enumerate}
569: \item If the random variable $G$ is not deterministic and $\E(G^2)<+\infty$, then
570: \begin{equation}\label{rho21}
571: \rho_2^{(1)}(\lambda)\stackrel{\mathrm{def}}{=}  \lim_{N\to \infty}\frac{\Var(R_N)}{\E(T_N)^2} = \frac{\Var(G)}{m^2-m} \rho(\lambda)^2
572: \end{equation}
573: where $\rho(\lambda)$ is defined by Equation~\eqref{defrholambda}.
574: \item If $G\equiv m$ almost surely, then
575: \begin{multline} \label{rho22}
576: \rho^{(2)}_2(\lambda)\stackrel{\mathrm{def}}{=}  \lim_{N\to \infty}\frac{\Var(R_N)}{\E(T_N)} \
577: \ =
578: \frac{m-1}{m}\sum_{n=1}^{+\infty}
579: \frac{1}{m^n}\left[
580: \E\left(e^{-\lambda T_n}\right)\left(1-\E\left(e^{-\lambda T_n}\right)\right) \right. \\
581: \left.  +2\sum_{k=0}^{n-1}\E\left(e^{-\lambda T_{n{-}k{-}1}} Z_{n-k} \E\left(e^{-\lambda T_{k}
582: }\right)^{Z_{n-k}-1}
583: \E\left(e^{-\lambda T_{k}}\left(1-e^{-\lambda T_{k}}\right)\right)\right)\right].
584: \end{multline}
585: where $T_n = (m^{n+1}-1)/(m-1)$.
586: \end{enumerate}
587: \end{proposition}
588: It is worth  noting that the case of a deterministic  offspring distribution is degenerate
589: in the sense  that the standard deviation of  the size of the tree discovered  by means of
590: the traceroute  algorithm does not  scale with  the size of  the tree (and  the discovered
591: tree).  The  coefficient of variation of  the random variables  $R_N$ tends to 0  when $N$
592: goes to infinity.
593: 
594: \begin{proof}
595: \cbstart Using Representation~\eqref{relfond} for the size of the sampled tree, one obtains by conditioning on the tree the relation 
596: $$
597: R_N-\E(R_N) =A_{N,1}+A_{N,2}+A_{N,3},
598: $$
599: where 
600: \begin{align*}
601: A_{N,1} &=  \sum_{n=0}^N\sum_{\ell=1}^{Z_{N-n}} \Delta^\ell_n,\\
602: A_{N,2} &= \sum_{n=0}^N (Z_{N-n} - \E(Z_{N-n})) \P\left({\cal N}(\T_{n})\not=0\right),\\
603: A_{N,3} &= \sum_{n=0}^N \sum_{\ell=1}^{Z_{N-n}} \left(1-\exp(-\lambda T^{N-n,\ell}_n)-\P(\cal{N}(\T_n)\neq 0)\right)
604: \end{align*}
605: with
606: $
607: \Delta^\ell_n =\ind{{\cal N}(\T_{n}^{N-n,\ell})\not=0}  - (1-\exp(-\lambda
608: T_n^{N-n,\ell})).
609: $
610: Note that if distinct nodes $(n,l)$ and $(n',l')$ cannot be compared with the relation
611: $''{<}''$ then, conditionally on the tree ${\cal T}$, the corresponding random variables
612: $\Delta^\ell_n$ and $\Delta^{\ell'}_{n'}$ are {\em centered} and {\em independent}. 
613: In addition, note that $ A_{N,1} = R_N-\E(R_N~|~\T)$. 
614: 
615: To study the variance of the random variable $R_N$, we separately consider the second
616: moments of the terms $A_{N,1}$, $A_{N,2}$ and $A_{N,3}$. Of course, the terms $A_{N,2}$ and $A_{N,3}$ are  non null if
617: and only if the variable $G$ is not deterministic.  
618: 
619: \cbend
620: 
621: \subsection*{The second moment of $A_{N,1}$}
622: \cbstart 
623: It is shown that the second moment of $A_{N,1}$ is of the order of $m^N$.
624: By using the independence in the
625: selection of nodes in the tree and the fact that the random variables $\Delta_n^\ell$ are 
626: centered conditionally on $\T$, we have  the identity 
627: \[
628: \E(A_{N,1}^2~|~\T) = \sum_{(n,\ell)\in \T_N}\Var(\Delta^\ell_n~|~\T) + 2\sum_{\substack{(n,\ell),    (n',\ell')\in{\cal T}_N \\(n',\ell')<(n,\ell)}} \E\left(\Delta^\ell_n\Delta^{\ell'}_{n'}~|~\T\right).
629: \]
630: Conditioning on the state of the tree, when $(n',\ell')<(n,\ell)$, one has the identity
631: \[
632: \E\left(\Delta^\ell_n\Delta^{\ell'}_{n'}\mid {\cal T}\right)=
633: \exp\left(-\lambda T^{N-n,\ell}_{n}\right)\left(
634: 1-\exp\left(-\lambda T^{N-n',\ell'}_{n'}\right)\right).
635: \]
636: By symmetry, the above computations yield the following relation for the second moment $\E(A_{N,1}^2)$
637: \begin{align*}
638: U_N&\stackrel{\text{def.}}{=} \E(A_{N,1}^2)-\sum_{n=0}^N\E(Z_{N-n})\Var(\Delta_n^1)\\
639: &= 2\E\left(\sum_{n=0}^N  \E(Z_{N-n}) \exp\left(-\lambda T_{n}^{N-n,1}\right)\sum_{\substack{(N-n',l')\in{\cal T}_N\\ (N-n',l')<(N-n,1)}} 
640: \left(1-\exp\left(-\lambda T_{n'}^{N-n',\ell'}\right)\right)\right),
641: \end{align*}
642: where $\Var(\Delta_n)$ is the variance of the random variable $\ind{\mathcal{N}(\T_n)\neq 0}-\P(\mathcal{N}(\T_n)\neq 0)$.
643: \cbend
644: 
645: For two nodes of the tree such that $(N{-}n',l')<(N{-}n,1)$, Equation~\eqref{eq1st} gives
646: the relation 
647: \[
648: T_n\stackrel{\text{dist.}}{=}
649: \sum_{k=0}^{n-n'-1}\widetilde{Z}_{k}+ 
650: \sum_{\ell'=1}^{\widetilde{Z}_{n-n'}} {T}_{n'}^{N-n',\ell'},
651: \]
652: where $(\widetilde{Z}_{k}, k\geq 0)$ denotes another independent Galton-Watson process
653: independent of $(Z_n, n \geq 0)$ with the same offspring distribution. By using this relation, we have
654: \begin{multline*}
655: \E\left(\exp\left(-\lambda T_{n}^{N-n,1}\right)\sum_{\substack{(N-n',l')\in{\cal T}_N\\ (N-n',l')<(N-n,1)}}\left(1-\exp\left(-\lambda T_{n'}^{N-n',\ell'}\right)\right)\right) \\ = \E\left(\sum_{n'=0}^{n-1}    \sum_{\ell'=1}^{Z_{n-n'}} \exp\left(-\lambda T_{n}^{N-n,\ell}\right)\left(1-\exp\left(-\lambda T_{n'}^{N-n',\ell'}\right)\right)\right)\\
656: = \E\left(\sum_{n'=0}^{n-1} \exp\left(-\lambda\sum_{k=0}^{n-n'-1} Z_k\right)  \E(V_{n-n'})\right),
657: \end{multline*}
658: where
659: $$
660: V_{n-n'} =  \sum_{\ell'=1}^{Z_{n-n'}} \exp\left(-\lambda\sum_{\ell''=1}^{Z_{n-n'}} T_{n'}^{N-n',\ell''}\right)\left(1-\exp\left(-\lambda T_{n'}^{N-n',\ell'}\right)\right).
661: $$
662: By using the independence of the different trees $\mathcal{T}_{n'}^{N-n',\ell'}$ for $\ell =1, \ldots, Z_{n-n'}$, we have
663: \begin{multline*}
664: \E(V_{n-n'}\mid Z_0,\ldots, Z_{n-n'-1}) \\ = \E\left(Z_{n-n'}\left(\E\left(
665: e^{-\lambda T_{n'}}\right)\right)^{Z_{n-n'}-1}\mid Z_0,\ldots,
666: Z_{n-n'-1}\right)\E\left(  e^{-\lambda T_{n'}} \left(1-e^{-\lambda T_{n'}}\right)\right). 
667: \end{multline*}
668: It follows that by using the above  expression for $U_N$, one obtains
669: \begin{multline*}
670: U_N=2\sum_{n=0}^N  \E(Z_{N-n})\sum_{n'=0}^{n-1}\E\left(\exp\left(-\lambda \sum_{k=0}^{n-n'-1}{Z}_{k}\right) \right.\\   
671: \left.\rule{0mm}{7mm} Z_{n-n'} \E\left(\rule{0mm}{4mm}\exp\left(-\lambda T_{n'}\right)\right)^{Z_{n-n'}-1}
672: \E\left(\rule{0mm}{4mm}\exp\left(-\lambda T_{n'}\right)\left(1-\exp\left(-\lambda
673: T_{n'}\right)\right)\right) \right).
674: \end{multline*}
675: Dividing by $\E(T_N)$, we have
676: \begin{multline*}
677: \frac{U_N}{\E(T_N)}=\frac{2(m-1)}{m-1/m^n}\sum_{n=0}^N  \frac{1}{m^n}\sum_{n'=0}^{n-1}\E\left(
678: \exp\left(-\lambda \sum_{k=0}^{n-n'-1}{Z}_{k}\right) \right.\\
679: \left.\rule{0mm}{7mm} Z_{n-n'}\E\left(\rule{0mm}{4mm}\exp\left(-\lambda T_{n'}\right)\right)^{Z_{n-n'}-1}
680: \E\left(\rule{0mm}{4mm}\exp\left(-\lambda T_{n'}\right)\left(1-\exp\left(-\lambda T_{n'}\right)\right)\right)\right).
681: \end{multline*}
682: By letting $N$ go to infinity, we finally obtain the relation for the second moment of the random variable $A_{N,1}$
683: \begin{multline}\label{A1}
684: \lim_{N\to \infty}\frac{\E(A_{N,1}^2)}{\E(T_N)} = 
685: \frac{m-1}{m}\sum_{n=1}^{+\infty}
686: \frac{1}{m^n}\left[
687: \E\left(e^{-\lambda T_n}\right)\left(1-\E\left(e^{-\lambda T_n}\right)\right) \right. \\
688: \left.  +2\sum_{k=0}^{n-1}\E\left(e^{-\lambda T_{n{-}k{-}1}} Z_{n-k} \E\left(e^{-\lambda T_{k}}\right)^{Z_{n-k}-1}
689: \E\left(e^{-\lambda T_{k}}\left(1-e^{-\lambda T_{k}}\right)\right)\right)\right].
690: \end{multline}
691: 
692: \subsection*{The second moment of $A_{N,2}$}
693: We have
694: \begin{multline*}
695: \frac{A_{N,2}}{m^{N}}  = \sum_{n=0}^{N} \frac{\left(Z_{N-n} - \E(Z_{N-n})\right)}{m^{N-n}}\frac{(1-\E(e^{-\lambda T_n}))}{m^n}
696: \\= \sum_{n=0}^{N} \left(\frac{Z_{N-n}}{m^{N-n}} - 1\right) \frac{(1-\E(e^{-\lambda T_n}))}{m^n}.
697: \end{multline*}
698: If $\|H\|_2=\sqrt{\E(H^2)}$ for some random variable $H$, then  we have
699: \[
700: \left\|\frac{A_{N,2}}{m^{N}}-(W - 1)\sum_{n=0}^{N} \frac{(1-\E(e^{-\lambda T_n}))}{m^n}\right\|_2 \leq \sum_{n=0}^{N} \left\|W-\frac{Z_{N-n}}{m^{N-n}}\right\|_2\frac{(1-\E(e^{-\lambda T_n}))}{m^n},
701: \]
702: where $W$ is defined by Equation~\eqref{defW}. \cbstart Athreya and Ney~\cite[Theorem~2, page~9]{Athreya:04} gives that the sequence
703: $\left(\left\|W-{Z_{n}}/{m^{n}}\right\|_2\right)$ converges to $0$. This implies
704: \begin{equation}\label{A2}
705: \lim_{N\to+\infty} \frac{\E(A_{N,2}^2)}{\E(T_N)^2}=\frac{(m-1)\Var(G)}{m^3} 
706: \left(\sum_{n=0}^{+\infty} \frac{(1-\E(e^{-\lambda T_n}))}{m^n}\right)^2.
707: \end{equation}
708: since
709: $\E\left((1-W)^2\right)={\Var(G)}/{m(m-1)}$. 
710: 
711: \subsection{Second moment of $A_{N,3}$} 
712: Clearly
713: $$
714: \|A_{N,3}\|_2 \leq \sum_{n=0}^N\left\|  \sum_{\ell=1}^{Z_{N-n}} \exp(-\lambda T^{N-n,\ell}_n) - \E(\exp(-\lambda T_n))\right\|_2 ,
715: $$
716: and since conditionally on $Z_{N-n}$, the random variables $\exp(-\lambda T^{N-n,\ell}_n)$
717: for $\ell = 1, \ldots, Z_{N-n}$ are independent and identically distributed with mean
718: $\E(\exp(- \lambda T_n))$,  we then have 
719: \begin{align*}
720: \left\|\sum_{\ell=1}^{Z_{N-n}} \exp(-\lambda T^{N-n,\ell}_n) - \E(\exp(-\lambda T_n))  \right\|_2^2 &= \E(Z_{N-n})\Var(\exp(-\lambda T_n))\\ & \leq m^{(N-n)}.
721: \end{align*}
722: It follows that 
723: $$
724: \|A_{N,3}\|_2 \leq \frac{m^{(N+1)/2}}{\sqrt{m}-1}.
725: $$
726: and then
727: \begin{equation}
728: \label{limA3N}
729: \limsup_{N\to\infty}\frac{\E(A_{N,3}^2)}{m^N} \leq \frac{m}{(\sqrt{m}-1)^2}.
730: \end{equation}
731: 
732: Since $R_N-\E(R_N)=A_{N,1}+A_{N,2}+A_{N,3}$, 
733: %%it is easily checked that $$\Var(R_N) = E(A_{N,1}^2) + \Var(A_{N,2}+A_{N,3}).$$
734: Relations~\eqref{A1}, \eqref{A2} and \eqref{limA3N} then imply  that  
735: \begin{enumerate}
736: \item When $G$ is non-deterministic, the expression $A_{N,2}$ dominates in $R_N-\E(R_N)$
737: so that $\Var(R_N)/\E(T_N)^2$ is converging to   the right hand side of Equation~\eqref{A2}. 
738: \item If $G\equiv m$, the term $A_{N,3}$ vanishes so that  $\Var(R_N)/\E(T_N)$ is converging to
739:   the right hand side of Equation~\eqref{A1}. 
740: \end{enumerate}
741: Equations~\eqref{rho21} and \eqref{rho22} are established.
742: \end{proof}
743: 
744: \cbend
745: 
746: As  for the  first  moment of  $R_N$, we  turn  now to  the  analysis of  the behavior  of
747: of the second order characteristics defined by Equations~\eqref{rho21} and \eqref{rho22}
748: when $\lambda$ is in the neighborhood of $0$. For the non deterministic case, we have from Proposition~\ref{rho1}
749: \[
750: \lim_{\lambda\to 0} \frac{\rho^{(1)}_2(\lambda)}{(\lambda\log_m(1/\lambda))^2} =\frac{\Var(G)}{(m^2-m)}.
751: \]
752: In  Proposition~\ref{vartree},  the expression  of  $\rho^{(2)}_2(\lambda)$  is defined  a
753: priori  only for  a  deterministic offspring  distribution,  but can  be  extended to  any  offspring   distribution  by   using  the   right  hand   side  of
754: Equation~\eqref{rho22}.     In    the    following,    we   study    the    behavior    of
755: $\rho_2^{(2)}(\lambda)$ for an arbitrary offspring distribution.
756: 
757: \begin{lemma}[Asymptotic Behavior of $\lambda{\to}\rho^{(2)}_2(\lambda)$ at $0$]\label{Rho22}
758: Provided that the random variable $G$ has a finite second moment, the function $\rho^{(2)}_2(\lambda)$ defined by
759: Equation~\eqref{rho22} is such that  
760: \begin{equation}\label{rate2}
761: \lim_{\lambda\searrow 0}\frac{\rho^{(2)}_2(\lambda)}{\lambda(\log_m\lambda)^2}=1.
762: \end{equation}
763: \end{lemma}
764: 
765: \begin{proof}
766: Define
767: $$
768: f_a(\lambda)\stackrel{\text{def.}}{=} \sum_{n=1}^{+\infty} \frac{m-1}{m^n}\left(\E\left(e^{-\lambda T_n}\right)\left(1-\E\left(e^{-\lambda T_n}\right)\right)\right)
769: $$
770: and 
771: \begin{multline*}
772: f_b(\lambda)\stackrel{\text{def.}}{=} \\ \sum_{n=1}^{+\infty} \frac{m-1}{m^n} \sum_{k=0}^{n-1}\E\left(e^{-\lambda T_{n{-}k{-}1}} Z_{n-k}\E\left(e^{-\lambda T_{k}}\right)^{Z_{n-k}-1}\right)\E\left(e^{-\lambda T_{k}}\left(1-e^{-\lambda T_{k}}\right)\right).
773: \end{multline*}
774: Equation~\eqref{rho22} gives that
775: ${m}\rho^{(2)}_2(\lambda)=2f_b(\lambda)+f_a(\lambda)$. 
776: 
777: \medskip
778: 
779: \paragraph{\em Asymptotic behavior of $f_a$}
780: With similar arguments as in the proof of
781: Theorem~\ref{ThFirst} the asymptotic behavior of  $f_a(\lambda)$ when $\lambda$ goes to
782: $0$  is equivalent to the  asymptotic behavior of
783: \[
784: \sum_{n=1}^{+\infty}
785: \frac{m-1}{m^n}\left(\E\left(\exp\left(-\lambda \frac{W m^{n+1}}{m-1} \right)\right)\left(1-\E\left(\exp\left(-\lambda \frac{W m^{n+1}}{m-1} \right)\right)\right)\right).
786: \]
787: If $W_1$ and $W_2$ are two independent random variables with the same distribution as $W$,
788: the above series can be rewritten as
789: \begin{multline*}
790: \sum_{n=1}^{+\infty}
791: \frac{m-1}{m^n}
792: \E\left(\exp\left(-\lambda \frac{W_1 m^{n+1}}{m-1} \right)
793: \left(1-\E\left(\exp\left(-\lambda \frac{W_2 m^{n+1}}{m-1} \right)\right)\right)\right)
794: \\=
795: (m-1)\sum_{n=1}^{+\infty}
796: \left(\frac{1}{m^n} \E(h(\lambda(W_1+W_2)m^n/(m-1)))-\E(h(\lambda W_1m^n/(m-1)))\right),
797: \end{multline*}
798: with $h(u)=1-e^{-u}$. Consequently,
799: \begin{equation}\label{G1}
800: \lim_{\lambda\to 0} \frac{f_a(\lambda)}{-\lambda\log_m \lambda}=m
801: \end{equation}
802: by Proposition~\ref{asympprop}.
803: 
804: \medskip
805: \paragraph{\em Asymptotic behavior of $f_b$} Let us fix some $\varepsilon >0$ and assume that $\lambda <\varepsilon$.  The function $f_b(\lambda)$ can be rewritten as
806: \begin{equation}
807: \label{decompofb}
808: f_b(\lambda) = \sum_{n=1}^{\lfloor \log_m(\varepsilon/\lambda)\rfloor} \frac{m-1}{m^n} S(n;\lambda)  +  \sum_{n=\lfloor \log_m(\varepsilon/\lambda)\rfloor+1}^\infty  \frac{m-1}{m^n} S(n;\lambda) 
809: \end{equation}
810: where
811: $$
812: S(n;\lambda) =  \sum_{k=0}^{n-1}\E\left(e^{-\lambda T_{n{-}k{-}1}} Z_{n-k}\E\left(e^{-\lambda T_{k}}\right)^{Z_{n-k}-1}\right)\E\left(e^{-\lambda T_{k}}\left(1-e^{-\lambda T_{k}}\right)\right)
813: $$
814: Since for $x\geq 0$, $e^{-x}(1-e^{-x}) \leq x$, we easily deduce that for all $n \geq 1$
815: $$
816: S(n;\lambda) \leq  \sum_{k=0}^{n-1}\E\left( Z_{n-k}\right) \E\left(\lambda T_{k}\right) \leq \frac{n \lambda m^{n+1}}{m-1}
817: $$
818: and then
819: $$
820: \sum_{n=1}^{\lfloor \log_m(\varepsilon/\lambda)\rfloor} \frac{m-1}{m^n} S(n;\lambda) \leq \frac{m}{2}\lambda \log_m(\varepsilon/\lambda) (\log_m(\varepsilon/\lambda)+1).
821: $$
822: 
823: The second term in the right hand side of Equation~\eqref{decompofb} can be written as
824: \begin{multline}
825: \label{termtech2}
826: \sum_{n=\lfloor \log_m(\varepsilon/\lambda)\rfloor+1}^\infty  \frac{m-1}{m^n}      S( \lfloor \log_m(\varepsilon/\lambda)\rfloor +1;\lambda)  
827:  \\ 
828: + \sum_{n=\lfloor \log_m(\varepsilon/\lambda)\rfloor+1}^\infty  \frac{m-1}{m^n}  \left(S( n;\lambda)  - S( \lfloor \log_m(\varepsilon/\lambda)\rfloor +1;\lambda)  \right).
829: \end{multline}
830: By using the fact that for $x>0$ and $\alpha>0$, $xe^{-\alpha x}\leq 1/\alpha$, we get that 
831: $$
832:  S( \lfloor \log_m(\varepsilon/\lambda)\rfloor +1;\lambda)  \leq \frac{1}{\E(e^{-\lambda T_k})}  \sum_{k=0}^{ \lfloor \log_m(\varepsilon/\lambda)\rfloor} \frac{\lambda\E(T_k)}{-\log\E(e^{-\lambda T_k})}. 
833: $$
834: The relation $\E\left(\exp\left(-\lambda T_k\right)\right) \geq 
835: \exp\left(-\lambda \E(T_k)\right)
836: \geq \exp(- m\varepsilon/(m-1))$  holds by Jensen's Inequality under the condition that $k \leq \lfloor
837: \log_m(\varepsilon/\lambda)\rfloor $. In addition,  
838: $$
839: \E(T_n^2) \leq \left(\sum_{i=0}^n \sqrt{\E(Z_i^2)}\right)^2 \leq \frac{m^{2n}}{(m-1)^2} \left(\frac{\sigma^2}{m-1}+1\right) ,
840: $$
841: where $\sigma^2$ is the variance of the random variable $G$, so that for $k \leq  \lfloor \log_m(\varepsilon/\lambda)\rfloor $
842: \begin{equation}
843: \label{rapportmoment}
844: \frac{\lambda\E(T_k^2)}{\E(T_k)} \leq \frac{\lambda m^k m^k}{(m-1)(m m^k-1)} \left(\frac{\sigma^2}{m-1}+1\right)  \leq \frac{\varepsilon}{(m-1)^2} \left(\frac{\sigma^2}{m-1}+1\right). 
845: \end{equation}
846: Since for $x \geq 0$, $e^{-x} \leq 1-x+{x^2}/{2}$, we have
847: $$
848: \E(e^{-\lambda T_k}) \leq 1 -\lambda\E(T_k) +\frac{\E\left((\lambda T_k)^2\right)}{2} \leq 1-\lambda \E(T_k) \left(1-  \frac{\varepsilon}{(m-1)^2} \left(\frac{\sigma^2}{m-1}+1\right) \right).
849: $$
850: and then, for $k \leq  \lfloor \log_m(\varepsilon/\lambda)\rfloor $, 
851: \begin{multline}\label{eqaux}
852: \frac{\lambda\E(T_k)}{-\log\E(e^{-\lambda T_k})} \leq
853: \frac{\lambda\E(T_k)}{1- \E(e^{-\lambda T_k}) }\\ \leq
854: \left(1-  \frac{\varepsilon}{(m-1)^2} \left(\frac{\sigma^2}{m-1}+1\right)\right)^{-1}
855:  \stackrel{\mathrm{def.}}{=}\kappa(\varepsilon)
856: \end{multline}
857: as long as 
858: $$
859: \varepsilon< (m-1)^2\left/\left(\frac{\sigma^2}{m-1}+1\right)\right. \stackrel{\mathrm{def.}}{=}\varepsilon_1.
860: $$
861: 
862: It follows that for $\varepsilon<\varepsilon_1$,
863: $$
864: S( \lfloor \log_m(\varepsilon/\lambda)\rfloor +1;\lambda)  \leq  (1+\log_m(\varepsilon/\lambda))  e^{m\varepsilon/(m-1)}\kappa(\varepsilon).
865: $$
866: and therefore,
867: $$
868: \sum_{n=\lfloor \log_m(\varepsilon/\lambda)\rfloor+1}^\infty  \frac{m-1}{m^n}      S( \lfloor \log_m(\varepsilon/\lambda)\rfloor +1;\lambda)  \leq \frac{\lambda}{\varepsilon}  \log_m(\varepsilon/\lambda)  e^{m\varepsilon/(m-1)}\kappa(\varepsilon),
869: $$
870: which is $o(\lambda(\log\lambda)^2)$ when $\lambda \to 0$. In addition, the second term in the right hand side of Equation~\eqref{termtech2} can be rewritten as
871: $$
872:   \sum_{k= \lfloor \log_m(\varepsilon/\lambda)\rfloor +1}^\infty \frac{m-1}{m^k} \sum_{n=1}^\infty \frac{1}{m^n} \E\left(e^{-\lambda T_{n-1}} Z_n \E(e^{-\lambda T_k})^{Z_n-1}\right)  \E\left(e^{-\lambda T_{k}}\left(1-e^{-\lambda T_{k}}\right)\right).
873: $$
874: We first note that
875: \begin{multline*}
876: \sum_{n=1}^\infty \frac{1}{m^n} \E\left(e^{-\lambda T_{n-1}} Z_n \E(e^{-\lambda T_k})^{Z_n-1}\right) =\\
877: \sum_{n=1}^{\lfloor \log_m(\varepsilon/\lambda)\rfloor} \frac{1}{m^n} \E\left(e^{-\lambda T_{n-1}} Z_n \E(e^{-\lambda T_k})^{Z_n-1}\right) \\ 
878: + \sum_{n = \lfloor \log_m(\varepsilon/\lambda)\rfloor+1}^\infty \frac{1}{m^n} \E\left(e^{-\lambda T_{n-1}} Z_n \E(e^{-\lambda T_k})^{Z_n-1}\right).
879: \end{multline*}
880: The first term in the right hand side of the above equation is less than or equal to the
881: quantity $\log_m(\varepsilon/\lambda)$ since $\E(Z_n)=m^n$. The second term can be upper bounded as
882: \begin{multline*}
883: \sum_{n = \lfloor \log_m(\varepsilon/\lambda)\rfloor+1}^\infty \frac{1}{m^n} \E\left(e^{-\lambda T_{n-1}} Z_n \E(e^{-\lambda T_{k}})^{Z_n-1}\right) \\ \leq \sum_{n=\lfloor \log_m(\varepsilon/\lambda)\rfloor+1}^\infty \frac{1}{m^n}\E\left(Z_n  e^{-\lambda Z_n} \right) 
884: \leq \frac{1}{(m-1)\varepsilon},
885: \end{multline*}
886: where we have used the fact that $T_k \geq 1$ for all $k \geq 0$ and $xe^{-\lambda x}\leq 1/(e\lambda)$ for all $x>0$. It follows that the second term in the right hand side of Equation~\eqref{termtech2} is upper bounded by the quantity 
887: $$
888: \frac{\lambda}{\varepsilon}\left(\log_m(\varepsilon/\lambda)+ \frac{1}{(m-1)\varepsilon}\right),
889: $$
890: which is $o(\lambda(\log_m\lambda)^2)$ when $\lambda \to 0$. 
891: 
892: By using the above inequalities, we come up with the conclusion that for every $\varepsilon >0$,
893: \begin{equation}
894: \label{limsupfb}
895: \limsup_{\lambda \to 0} \frac{f_b(\lambda)}{\lambda(\log_m\lambda)^2} \leq \frac{m}{2}.
896: \end{equation}
897: 
898: 
899: For establishing a lower bound for $f_b(\lambda)$, we introduce the size-biased Galton-Watson branching process.  The sequence of random variables $(Z_n/m^n)$ being a positive  martingale, it induces a probability distribution $\widetilde{\P}$ such that, for any $n\geq 1$  and any random variable $Y$ measurable with respect to the random variables $Z_1, \ldots, Z_n$, 
900:  \[
901:  \int Y d\widetilde{\P}=\E\left(Y\frac{Z_n}{m^n}\right).
902:  \]
903:  It is known, see Lyons and Peres~\cite{Lyons:07}, that under the probability
904:  $\widetilde{\P}$, the sequence $(Z_n)$ as the same distribution as  a branching process with
905:  immigration $(\widetilde{Z}_n)$ where the number of children has the same distribution as
906:  $G$ and the number of new immigrants is distributed as $\widetilde{G}$ such that
907:  $\P(\widetilde{G}=n)=n\P(G=n)/m$. If $\widetilde{Z}_0=1$, it is easy to check that
908:  \[
909:  \widetilde{\E}\left(\widetilde{Z}_n\right)=m^n+\frac{m^n-1}{m(m-1)}\E(G^2).
910:  \]
911:  If $\widetilde{T}_n=\widetilde{Z}_0+\widetilde{Z}_1+\cdots+\widetilde{Z}_n$, we have by Jensen inequality
912: \begin{align*}
913: \E\left(\rule{0mm}{5mm}e^{-\lambda T_{n{-}k{-}1}}\right.&\left. \frac{Z_{n-k}}{m^{n-k}} \E\left(e^{-\lambda T_{k}}\right)^{Z_{n-k}-1}\right) =\widetilde{\E}\left(e^{-\lambda \widetilde{T}_{n{-}k{-}1}} \E\left(e^{-\lambda T_{k}}\right)^{\widetilde{Z}_{n-k}-1}\right) \\
914: &\geq  \widetilde{\E}\left(e^{-\lambda \widetilde{T}_{n{-}k}} \E\left(e^{-\lambda T_{k}}\right)^{\widetilde{T}_{n-k}}\right) \\
915: &\geq \exp\left(-\lambda(1+\E(T_k)) \widetilde{\E}( \widetilde{T}_{n{-}k})\right)\\
916: &\geq \exp\left(-\frac{\lambda m}{m-1} \left(m^{n-k}+\frac{m^{n+1}}{(m-1)}\right)\left(1+\frac{g_2}{m}\right)\right)
917: \end{align*}
918: since 
919: $$
920: \widetilde{\E}( \widetilde{T}_{n}) \leq \frac{m^{n+1}}{m-1}\left(1+\frac{g_2}{m}  \right),
921: $$
922: where $g_2 = \E(G^2)$. In addition, by using the fact that $e^{-x}(1-e^{-x}) \geq x-2x^2$
923: holds for $x>0$, we have
924: \begin{multline*}
925:  \sum_{n=1}^{\lfloor \log_m(\varepsilon/\lambda)\rfloor} \frac{m-1}{m^n} S(n;\lambda)  
926: \geq  \sum_{n=1}^{\lfloor \log_m(\varepsilon/\lambda)\rfloor} (m-1) \\\times
927: \sum_{k=0}^{n-1} \frac{1}{m^k} 
928: \exp\left(-\frac{\lambda m}{m-1} \left(m^{n-k}+\frac{m^{n+1}}{(m-1)}\right)
929: \left(1+\frac{g_2}{m}\right)\right)
930: \E\left(\lambda T_k-2(\lambda T_k)^2\right) \\
931: \geq \exp\left(-\frac{\varepsilon m}{(m-1)}
932:   \left(1+\frac{m}{(m-1)}\right)\left(1+\frac{g_2}{m}\right)\right)\\ \sum_{n=1}^{\lfloor
933:   \log_m(\varepsilon/\lambda)\rfloor} (m-1) \sum_{k=0}^{n-1}\frac{\lambda\E(T_k)}{m^k}
934: \left(1-\frac{\lambda\E(T_k^2)}{\E(T_k)}\right). 
935: \end{multline*}
936: By using Inequality~\eqref{rapportmoment} and Definition~\eqref{eqaux}, we have, for
937: $\eps<\eps_1$, 
938: \begin{multline*}
939: \sum_{n=1}^{\lfloor \log_m(\varepsilon/\lambda)\rfloor} (m-1) \sum_{k=0}^{n-1}\frac{\lambda\E(T_k)}{m^k} \left(1-\frac{\lambda\E(T_k^2)}{\E(T_k)}\right)  \geq \frac{\lambda}{\kappa(\eps)}\sum_{n=1}^{\lfloor \log_m(\varepsilon/\lambda)\rfloor} \sum_{k=0}^{n-1}\frac{m^{k+1}-1}{m^k}
940: \end{multline*}
941: Since
942: \begin{multline*}
943: \sum_{n=1}^{\lfloor \log_m(\varepsilon/\lambda)\rfloor} \sum_{k=0}^{n-1}\frac{m^{k+1}-1}{m^k} = \frac{m}{2}\lfloor \log_m(\varepsilon/\lambda)\rfloor(\lfloor \log_m(\varepsilon/\lambda)\rfloor+1)+\frac{m\lfloor \log_m(\varepsilon/\lambda)\rfloor}{m-1} \\ -\frac{m}{(m-1)^2}\left(\frac{1}{m^{\lfloor \log_m(\varepsilon/\lambda)\rfloor}}-1\right)
944: \end{multline*}
945: and since we already know that the second term in the right hand side of Equation~\eqref{decompofb} is $o(\lambda(\log_m(\lambda))^2$ when $\lambda\to 0$, we then deduce that for all $\varepsilon \in (0,\varepsilon_1)$
946: $$
947: \liminf_{\lambda \to 0} \frac{f_b(\lambda)}{\lambda(\log_m\lambda)^2} \geq
948: \frac{m}{2\kappa(\eps)} \exp\left(-\frac{\varepsilon m}{(m-1)} \left(1+\frac{m}{(m-1)}\right)\left(1+\frac{g_2}{m}\right)\right)
949: $$
950: and hence,
951: \begin{equation}
952: \label{liminffb}
953: \liminf_{\lambda \to 0} \frac{f_b(\lambda)}{\lambda(\log_m\lambda)^2} \geq \frac{m}{2}.
954: \end{equation}
955: Combining Equations~\eqref{G1}, \eqref{limsupfb} and \eqref{liminffb}, Equation~\eqref{rate2} follows.
956:  \end{proof}
957: 
958: 
959: \begin{proposition}
960: The functions $\rho^{(1)}_2(\lambda)$ and $\rho^{(2)}_2(\lambda)$ are such that
961: \begin{eqnarray}
962: \lim_{\lambda \to 0}\frac{\rho^{(1)}_2(\lambda)}{(\lambda\log_m\lambda)^2} &=& \frac{1}{m^2-m}\Var(G),\\ 
963: \lim_{\lambda \to 0} \frac{\rho^{(2)}_2(\lambda)}{\lambda(\log_m\lambda)^2} &=& 1.
964: \end{eqnarray}
965: where $G$ is the random variable  describing the offspring of a node.
966: \end{proposition}
967: 
968: From Theorem~\ref{rho1},  we observe  that the size  of the sampled tree  scales with the
969: size  of  the  original  tree.  The   same  phenomenon  is  true  for  the  squared coefficient of variation  of the size of the sampled tree if an only if the offspring distribution is not deterministic as shown by Proposition~\ref{vartree}. In the case of a deterministic offspring distribution, when $\lambda \to 0$,  the squared coefficient of variation is approximately equal to $1/(\lambda \E(T_N))$ for large $N$. The quantity $\lambda\E(T_N)$ is precisely the mean number of selected points. This indicates that the distribution of the random variable $R_N$ is concentrated around its mean value.  There is almost no randomness in the discovered tree.
970: 
971: 
972: % \begin{proof}
973: % The first term of the right hand side of Relation~\eqref{vartree} defining
974: % $\rho_2(\lambda)$ is 
975: % \[
976: % \frac{m-1}{m}\sum_{n=1}^{+\infty}
977: % \frac{1}{m^n} \E\left(e^{-\lambda T_n}\right)\left(1-\E\left(e^{-\lambda T_n}\right)\right).
978: % \]
979: % By proceeding as in the proof of Theorem~\ref{ThFirst}, it is not difficult to prove that
980: % this expression is equivalent  to $-C\lambda\log_m(\lambda)$ as $\lambda\to 0$ for some
981: % $C>0$. This term will be therefore neglible provided that one can prove that the
982: % complementary term in $\rho_2(\lambda)$ is of the order of $\lambda(\log_m \lambda)^2$.
983: 
984: % The second term of the right hand side of
985: % Relation~\eqref{vartree} that is, up to the multiplicative constant $2(m-1)/m$, of  the quantity 
986: % \begin{multline}\label{eqche}
987: % \Phi(\lambda)=
988: % \sum_{n=1}^{+\infty}
989: % \frac{1}{m^n}
990: % \sum_{k=0}^{n-1}\E\left(e^{-\lambda T_{n{-}k{-}1}} Z_{n-k} \E\left(\rule{0mm}{4mm}e^{-\lambda T_{k}}\right)^{Z_{n-k}-1}
991: % \right. \\ \left.\E\left(\rule{0mm}{4mm}e^{-\lambda T_{k}}\left(1-e^{-\lambda T_{k}}\right)\right)\rule{0mm}{6mm}\right)
992: % \end{multline}
993: % by using Fubini's theorem, one gets that 
994: % \[
995: % \Phi(\lambda)= 
996: % \sum_{k=1}^{+\infty}
997: % \frac{1}{m^k} \E\left(\rule{0mm}{4mm}e^{-\lambda T_{k}}\left(1-e^{-\lambda  T_{k}}\right)\right)\phi_k(\lambda),
998: % \]
999: % with
1000: % \[
1001: % \phi_k(\lambda)=\sum_{n>0}\E\left(e^{-\lambda T_{n{-}1}} \frac{Z_{n}}{m^{n}} \E\left(\rule{0mm}{4mm}e^{-\lambda T_{k}}\right)^{Z_{n}-1}\right).
1002: % \]
1003: % \paragraph{\em Lower Bound}
1004: % By using Jensen's Inequality, one gets that
1005: % it is clear that $\bar{\phi}(\lambda)$
1006: % \[
1007: % \phi_k(\lambda)\geq 
1008: % \underline{\phi}(\lambda)
1009: % \stackrel{\text{def.}}{=}
1010: % \sum_{n>0}\E\left(e^{-\lambda T_{n{-}1}} \frac{Z_{n}}{m^{n}} e^{-\lambda m^{k+1}Z_{n}/(m-1)}\right).
1011: % \]
1012: % The sequence of random variables $(Z_n/m^n)$ being a positive
1013: % martingale, it induces a probability distribution $\widehat{\P}$ such that, for any $n\geq 1$
1014: % and any random variable $Y$ measurable with respect to the random variables $Z_1$, \ldots,
1015: % $Z_n$, then 
1016: % \[
1017: % \int Y d\widehat{\P}=\E\left(Y\frac{Z_n}{m^n}\right).
1018: % \]
1019: % It is known, see Lyons and Peres~\cite{Lyons:07}, that under the probability
1020: % $\widehat{\P}$, the sequence $(Z_n)$ as the same distribution as  a branching process with
1021: % immigration $(\widehat{Z}_n)$ where the number of children has the same distribution as
1022: % $G$ and the number of new immigrants is distributed as $\widehat{G}$ such that
1023: % $\P(\widehat{G}=n)=n\P(G=n)/m$. If $\widehat{Z}_0=1$, it is easy to check that
1024: % \[
1025: % \E\left(\widehat{Z}_n\right)=m^n+\frac{m^n-1}{m(m-1)}\E(G^2).
1026: % \]
1027: % If $\widehat{T}_n=\widehat{Z}_0+\widehat{Z}_1+\cdots+\widehat{Z}_n$, the lower bound
1028: % $\underline{\phi}(\lambda)$ can be written as
1029: % \begin{align*}
1030: % \underline{\phi}(\lambda)
1031: % &=\sum_{n>0}\E\left(e^{-\lambda \widehat{T}_{n{-}1}} e^{-\lambda
1032: %   m^{k+1}\widehat{Z}_{n}/(m-1)}\right)\\
1033: % &\geq \sum_{n>0}e^{-\lambda \E(\widehat{T}_{n-1})} e^{-\lambda
1034: %   m^{k+1}\E(\widehat{Z}_{n})/(m-1)}\\
1035: % &\geq \sum_{n>0}\exp\left(-\frac{\lambda}{m-1}\left(1+\frac{m_2}{m(m-1)}\right)(1+m^{k+1})m^n\right)
1036: % \end{align*}
1037: % where $m_2=\E(G^2)$
1038: 
1039: % With this estimation, one gets that
1040: % \begin{equation}\label{eqaux}
1041: % \sum_{k=1}^{+\infty}
1042: % \frac{1}{m^k} g_k(\lambda) 
1043: % H\left[\frac{\lambda}{m-1}\left(1{+}\frac{m_2}{m(m-1)}\right)\left(1{+}m^{k+1}\right)\right]
1044: % \leq \Phi(\lambda)
1045: % %\\ \leq \Phi(\lambda) \leq  \sum_{k=1}^{+\infty}\frac{1}{m^k} g_k(\lambda)H(\lambda/(m-1))
1046: % \end{equation}
1047: % with
1048: % \[
1049: % g_k(\lambda) \stackrel{\text{def.}}{=}\E\left(\rule{0mm}{4mm}e^{-\lambda T_{k}}\left(1{-}e^{-\lambda   T_{k}}\right)\right).
1050: % \]
1051: % and, for $x>0$,
1052: % \begin{align*}
1053: % H(x)&\stackrel{\text{def.}}{=}\sum_{n>0} e^{-xm^n}=\int_{xm}^{+\infty} \sum_{n>0} \ind{u>xm^n} e^{-u}\,du
1054: % =\int_{xm}^{+\infty} \lfloor\log_m (u/x)\rfloor e^{-u}\,du\\
1055: % &=\int_{xm}^{+\infty} \{\log_m (u/x)\}e^{-u}\,du+\int_{xm}^{+\infty} \log_m(u) e^{-u}\,du
1056: % -\log_m(x) e^{-xm},
1057: % \end{align*}
1058: % by plugging this decomposition into Equation~\eqref{eqaux}, by using again
1059: % Proposition~\ref{asympprop}, one gets that, for $\lambda$ converging to $0$, $\Phi(\lambda)/-\log_m\lambda$
1060: % is lower bounded by 
1061: % \[
1062: % \sum_{k=1}^{+\infty}
1063: % \frac{1}{m^k} \E\left(\rule{0mm}{4mm}e^{-\lambda T_{k}}\left(1{-}e^{-\lambda   T_{k}}\right)\right).
1064: % \]
1065: % By using Inequality~\eqref{CS}, the above
1066: % expressions $T_k$ can be replaced by $W m^{k+1}/(m-1)$ so that with the help of
1067: % Proposition~\ref{asympprop} one gets the relation
1068: % \begin{equation}\label{eq3}
1069: % \liminf_{\lambda\searrow 0}\frac{\rho_2(\lambda)}{\lambda(\log_m\lambda)^2}\geq 2.
1070: % \end{equation}
1071: 
1072: % \paragraph{\em Upper Bound}
1073: % Since, for $k\geq 1$, $T_k\geq 1$,  one has
1074: % \[
1075: % \phi_k(\lambda)\leq \overline{\phi}(\lambda) \stackrel{\text{def.}}{=}
1076: % e^{\lambda}\sum_{n>0}\E\left(\frac{Z_{n}}{m^{n}} e^{-\lambda Z_{n}} \right),
1077: % \]
1078: % by using the fact that  the derivative of $x\to xe^{-\lambda x}$ is
1079: % bounded by $1$, from  Inequality~\eqref{CS} one gets that
1080: % \[
1081: % \sum_{n>0}\frac{1}{m^n}\E\left(\left|Z_{n} e^{-\lambda Z_{n}}-m^nWe^{-\lambda m^nW}\right| \right)
1082: % \leq \frac{\Var(1-W)}{\sqrt{m}-1} \sum_{n>0}\frac{1}{m^{(n-1)/2}}<+\infty.
1083: % \]
1084: % Since $\overline{\phi}(\lambda) $ is diverging as $\lambda$ goes to $0$, it has the same
1085: % asymptotic behavior as
1086: % \[
1087: % \sum_{n>0} \E\left(We^{-m^n \lambda W}\right)
1088: % \]
1089: % which is equivalent to $-\log_m(\lambda)$ by Proposition~\ref{asympprop}. This estimation and
1090: % Relation~\eqref{eq3} give the asymptotic expansion~\eqref{rate2}. The proposition is
1091: % proved. 
1092: % \end{proof}
1093: 
1094: 
1095: \section{The Depth Biased Model}
1096: \label{biased}
1097: In this section,  it is assumed that conditionally  on the tree, for $n\geq 0$,  a node at
1098: depth   $n$    is   chosen    with   probability   $(1-\exp(-(\alpha/m)^n))$    for   some
1099: $\alpha\in[0,1)$. The mean number  of selected nodes at depth $n$ in  the tree is equal to
1100: $m^n (1-\exp(-(\alpha/m)^n))\sim \alpha^n$ and the total number of selected nodes in the
1101: whole tree $\mathcal{N}(\mathcal{T})$ is such that
1102: $$
1103: \frac{1}{1-\alpha} -\frac{1}{2(1-\alpha^2/m)} \leq \E(\mathcal{N}(\mathcal{T})) = \sum_{n=0}^\infty m^n (1-e^{-(\alpha/m)^n})\leq \frac{1}{1-\alpha},
1104: $$  in  particular  the  mean  number of  selected  nodes  $\mathcal{N}(\mathcal{T})  \sim
1105: 1/(1-\alpha)$ when  $\alpha \to 1$.  The behavior of  the size $R(\alpha)$ of  the sampled
1106: tree is used to estimate the speed of the exploration process, when the number of selected
1107: nodes becomes large. We first give the expression of the mean value $\E(R(\alpha))$ of the
1108: size of the sampled tree.
1109: 
1110: \begin{lemma}
1111: The mean value of the size of the sampled tree in the depth biased model is given by
1112: \begin{equation}
1113: \label{relfond1}
1114: \E(R(\alpha)) = \sum_{n=0}^\infty m^n \left(1-\E\left(\exp\left(-\left(\frac{\alpha}{m}\right)^n\sum_{i=0}^\infty \alpha^i \frac{Z_i}{m^i}\right)\right)\right).
1115: \end{equation}
1116: \end{lemma}
1117: 
1118: \begin{proof}
1119: As in the previous section, for $n\geq 0$ and $1\leq \ell\leq Z_n$, the symbol  ${\cal T}^{n,\ell}$
1120: denotes the sub-tree of ${\cal T}$ whose root is $(n,\ell)$. The node $(n,\ell)$  is in the sampled tree if ${\cal N}({\cal T}^{n,\ell})\neq 0$. 
1121: \cbstart Since nodes at a given depth are selected independently one of  each other, we have
1122: $$ 
1123: \P\left({\cal N}({\cal T}^{n,\ell})\not=0 \mid {\cal T}, (n,\ell) \in {\cal T}\right) =1 - \exp\left(-\left(\frac{\alpha}{m}\right)^n\sum_{i=0}^\infty \alpha^i \frac{Z^{(n,\ell)}_i}{m^i}\right),
1124: $$
1125: where $Z_i^{(n,\ell)}$ is the number of descendants of $(n,\ell)$ at generation $i$ and
1126: where we have used the fact that the sub-tree ${\cal T}^{n,\ell}$ has the same offspring
1127: distribution as the original tree $\mathcal{T}$. Hence,  
1128: $$
1129: \P\left({\cal N}({\cal T}^{n,\ell})\not=0  \mid (n,\ell) \in {\cal T}\right)   = 1 - \E\left(\exp\left(-\left(\frac{\alpha}{m}\right)^n\sum_{i=0}^\infty \alpha^i \frac{Z_i}{m^i}\right)\right).
1130: $$ 
1131: \cbend
1132: It follows that the size of the sampled tree given by 
1133: \begin{equation} 
1134: R(\alpha) =\sum_{n=0}^{+\infty} \sum_{\ell=1}^{Z_{n}}\ind{{\cal N}({\cal T}^{\ell,n})\not=0}
1135: \end{equation}
1136: and its mean value is, by using the independence in the selection of nodes, \cbstart
1137: $$
1138: \E(R(\alpha)) =  \sum_{n=0}^{+\infty} \E(Z_n)\P\left({\cal N}({\cal T}^{n,1})\not=0\right),
1139: $$
1140: %%where $\mathcal{T}^n$ is a subtree of $\T$ rooted at a node of the $n$th generation of $\T$, 
1141: Equation~\eqref{relfond1} follows.\cbend
1142: \end{proof}
1143: 
1144: The growth rate of the exploration process is defined  by the ratio
1145: \[
1146: \frac{\E\left(R(\alpha)\right)}{\E({\cal N}({\cal T}))} =\frac{1}{\eta(\alpha)} \sum_{n=0}^{+\infty} (1-\alpha)m^n
1147: \left(1-\E\left(\exp\left(-\left(\frac{\alpha}{m}\right)^n\sum_{i=0}^\infty \alpha^i \frac{Z_i}{m^i}\right)\right)\right),
1148: \]
1149: where $\eta(\alpha)=(1-\alpha)\E({\cal N}({\cal T})) \to 1$ when $\alpha \to 1$.
1150: 
1151: \begin{theorem}\label{oscil}
1152: If $\E(G^2)<+\infty$, as $\alpha\nearrow 1$, the following limit relation holds 
1153: \[
1154: \lim_{\alpha \to 1}\frac{\E\left(R(\alpha)\right)}{\E({\cal N}({\cal T}))^2} =1.
1155: \]
1156: \end{theorem}
1157: 
1158: \begin{proof}
1159: Let us first introduce the function
1160: $$
1161: H(\alpha) = \sum_{n=0}^{+\infty} (1-\alpha)m^n \left(1-\E\left(\exp\left(-\left(\frac{\alpha}{m}\right)^n\frac{W}{1-\alpha}\right)\right)\right),
1162: $$
1163: where $W$ is defined by Equation~\eqref{defW}. We have
1164: \begin{eqnarray}
1165: \left|H(\alpha) - \eta(\alpha)\frac{\E\left(R(\alpha)\right)}{\E({\cal N}({\cal T}))}
1166: \right| &\leq& (1-\alpha) \sum_{n=0}^\infty \alpha^n \left|\sum_{i=0}^\infty \alpha^i
1167: \E\left(\frac{Z_i}{m^i} - W\right) \right|\label{eq0}\\  
1168: & \leq &\sum_{i=0}^\infty \alpha^i \E\left(\left| \frac{Z_i}{m^i} - W  \right|\right)
1169: \label{eqtech} \\
1170: &\leq& \mathrm{Var}(W)  \sum_{i=0}^\infty \left(\frac{\alpha}{\sqrt{m}}\right)^i =\frac{\mathrm{Var}(W) }{1-\frac{\alpha}{\sqrt{m}}},\notag
1171: \end{eqnarray}
1172: where we have used Inequality~\eqref{CS} in the last step.
1173: 
1174: Let us define the family of non-negative random variables $\mathcal{H}_\alpha$, $0<\alpha<1$ by
1175: $$
1176: \mathcal{H}_\alpha = \sum_{n=0}^{+\infty} (1-\alpha)^2m^n \left(1-\exp\left(-\left(\frac{\alpha}{m}\right)^n\frac{W}{1-\alpha}\right)\right).
1177: $$
1178: We have $(1-\alpha)H(\alpha)= \E(\mathcal{H}_\alpha )$. 
1179: 
1180: Let us fix some $\varepsilon >0$. Since $W>0$ a.s., we can define the quantity 
1181: $$n(W,\alpha)=\max\left(\left\lceil
1182: \log_{m/\alpha}\left(\frac{W}{(\varepsilon(1-\alpha))}\right)\right\rceil, 0\right).$$
1183: For $n\geq n(W,\alpha)$, 
1184: $$
1185: \left( \frac{\alpha}{m} \right)^n\frac{W}{1-\alpha} <\varepsilon.
1186: $$
1187: By using the fact that for $x\geq 0$, $1-e^{-x} \geq x - x^2/2$, we have
1188: $$
1189: \mathcal{H}_\alpha \geq  \sum_{n=n(W,\alpha)}^{+\infty} 
1190: (1-\alpha)^2m^n
1191: \left(1-\exp\left(-\left(\frac{\alpha}{m}\right)^n\frac{W}{1-\alpha}\right)\right)  \geq \alpha^{n(W,\alpha)} W(1-\varepsilon).
1192: $$   Since  the   above  inequality   is  valid   for  all   $\varepsilon  >0$   and  
1193: $\alpha^{n(W,\alpha)}$  converges   to  $1$  as   $\alpha\nearrow  1$,  it   follows  that
1194: $\liminf_{\alpha\to 1}\mathcal{H}_\alpha \geq W$ a.s.
1195: 
1196: Since $1-e^{-x} \leq x$ for $x\geq 0$, we have
1197: $$
1198: \mathcal{H}_\alpha \leq   W \quad \mbox{a.s.}
1199: $$
1200: and then $\limsup_{\alpha\to 1}\mathcal{H}_\alpha \leq W$ a.s. Hence, $\limsup_{\alpha\to 1}\mathcal{H}_\alpha = W$ a.s. Since the family $(\mathcal{H}_\alpha)$ is non negative and bounded by $W$, which is integrable, we have
1201: $$
1202: \lim_{\alpha\to 1} \E(\mathcal{H}_\alpha )= \E(W)=1
1203: $$
1204: and the result follows by using Inequality~\eqref{eqtech}.
1205: \end{proof}
1206: 
1207: When $\alpha <1$ the selected node are closed to the root and only a small fraction of the
1208: whole is  discovered. When  $\alpha \nearrow 1$,  we can  select nodes deeper  in the  tree but
1209: roughly only  one node is selected  in average at  each level. The above  result indicates
1210: that the average  size of the discovered  tree grows as the square  of the average
1211: number of selected nodes.
1212: 
1213: 
1214: \bibliographystyle{amsplain}
1215: \bibliography{main}
1216: 
1217: \end{document}
1218: 
1219: