Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-27T17:02:28.273Z Has data issue: false hasContentIssue false

Random multi-hooking networks

Published online by Cambridge University Press:  13 February 2023

Kiran R. Bhutani
Affiliation:
Department of Mathematics, The Catholic University of America, Washington, DC 20064, USA
Ravi Kalpathy*
Affiliation:
Department of Mathematics, The Catholic University of America, Washington, DC 20064, USA
Hosam Mahmoud
Affiliation:
Department of Statistics, The George Washington University, Washington, DC 20052, USA
*
*Corresponding author. E-mail: kalpathy@cua.edu
Rights & Permissions [Opens in a new window]

Abstract

We introduce a broad class of multi-hooking networks, wherein multiple copies of a seed are hooked at each step at random locations, and the number of copies follows a predetermined building sequence of numbers. We analyze the degree profile in random multi-hooking networks by tracking two kinds of node degrees—the local average degree of a specific node over time and the global overall average degree in the graph. The former experiences phases and the latter is invariant with respect to the type of building sequence and is somewhat similar to the average degree in the initial seed. We also discuss the expected number of nodes of the smallest degree. Additionally, we study distances in the network through the lens of the average total path length, the average depth of a node, the eccentricity of a node, and the diameter of the graph.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press

1. Introduction

Trees have long been in the focus of research on random graphs. The classic types, such as those that appear in data structures [Reference Brown and Shubert3,Reference Drmota10,Reference Knuth14,Reference Mahmoud15] and digital processing [Reference De La Briandais6,Reference Fredkin12,Reference Szpankowski20], grow incrementally, one node at a time. In more recent times, authors considered more complex types of random graphs grown by adjoining entire graphs to a growing network [Reference Bahrani and Lumbroso1,Reference Bhutani, Kalpathy and Mahmoud2,Reference Chen and Mahmoud5,Reference Desmarais and Holmgren7,Reference Drmota, Gittenberger and Panholzer11,Reference Gopaladesikan, Mahmoud and Ward13,Reference Mahmoud16,Reference van der Hofstad21]. We consider a growing network model in which the number of components attached at a stage follows a predetermined building sequence of numbers.

Societies and social networks grow and change over time in multiple random ways, which include growth patterns that add “components” at each step. Networks grown by adding components reflect these dynamics better than networks evolving on single node additions. One can embed a graph in a predetermined growth structure leading to multiple scenarios of growing networks.

In this paper, we develop a model where networks grow by hooking multiple copies of the seed at multiple nodes of the growing network chosen in a random fashion and study the theoretical and statistical properties of the networks so generated.

2. The building sequence

We assume that a network grows by attaching a number of components at each step to the existing structure, which starts with $\tau _0 \ge 2$ vertices. In the next subsection, we give a formal definition. Here, we only say a word on the number of components added at each step. After $n$ steps of growth, the number of components attached to obtain the next network is $k_n$, a predetermined sequence of nonnegative numbers.

2.1. Regularity conditions

Let $\tau _0\ge 2$. This represents the number of nodes in a building block (a seed). We grow the network by adding a number of copies of the seed at places called latches. At each latch, a designated vertex in the seed (called the hook) is fused with the latch. A formal definition of this process is given in the sequel.

We shall consider adding $k_{n}\ge 1$ copies of the seed to construct the $(n+1)$st network, under the following regularity conditions:

  1. (R1) $k_n \le \tau _0 + (\tau _0 -1) \sum _{i=0}^{n-1} k_i$.

  2. (R2) $\lim _{n \to \infty } {k_n}/{\sum _{i=0}^n k_i} = a \in [0, 1]$.

  3. (R3) $\lim _{n \to \infty } {\tau _0}/{k_n} = b \ge 0$.

A sequence of nonnegative integers $\{k_n\}_{n=0}^\infty$ satisfying (R1)–(R3) is called a building sequence. Condition (R1) is to guarantee the feasibility of choosing latches. At no point in time does the process require more (distinct) latches than the number of nodes existing in the network. Conditions (R2)–(R3) facilitate the existence of limits for properties of interest and expedite finding their values. Note that $a=0$ and $b=0$ are both allowed. For instance, for a constant sequence $k_n = k \in \mathbb {N}$, we have $a=0$, and $b = \tau _0/ k \gt 0$, whereas when $k_n = n+1$, we have both $a= 0$ and $b=0$.

Regularity conditions (R1)–(R3) are not too restrictive and the class covered by the investigation remains very broad. The examples that come up in practice satisfy these regularity conditions. For example, at one extreme, the building sequence $k_n=1$ builds networks of linear growth, including trees. At the other extreme, the case of equality in Condition (R1) builds a deterministic network where the entire vertex set is chosen at each step (a take-all model); such extremal case grows the network exponentially fast.

3. The multi-hooking network

A network grows as follows. We start with a connected seed graph $G_0$ with vertex set of size $\tau _0$ and edge set of size $\eta$. One of the vertices in the seed is designated as a hook (vertex $h$). When a copy of the seed is adjoined to the network, it is the seed's hook that latches into that larger graph. The hooking is accomplished by fusing together the hook and a latch (vertex) chosen from the network.

At step $n$, $k_{n-1}$ copies of the seed are hooked into the graph, $G_{n-1} = (V_{n-1}, {\mathcal {E}}_{n-1})$, with vertex set $V_{n-1}$ and edge set ${\mathcal {E}}_{n-1}$, that exists at time $n-1$. To complete the $n$th hooking step, we sample $k_{n-1}$ latches from the graph $G_{n-1}$. The selection mechanism can take a number of forms, such as choosing distinct hooks as opposed to allowing repetitions.

We use the notation $|A|$ for the cardinality of a set $A$. We consider a uniform model that selects $k_{n-1}$ distinct nodes in the network, with all $|V_{n-1}| \choose k_{n-1}$ subsets being equally likely. In the language of statistics, this boils down to sampling without replacement.

Figure 1 illustrates a seed and a network grown from it in three steps under the building sequence $k_n = n+1$. So, $G_1$ grows by choosing a latch from $G_0$ (the starred node in $G_0$), $G_2$ grows by choosing the two starred nodes from $G_1$, and $G_3$ grows by choosing the three starred nodes from $G_2$. The networks in Figure 1 have loops and multiple edges, as we do not restrict the study to simple graphs.

Figure 1. A seed (top) with a hook and three networks grown from it (second row) under the building sequence $k_n=n+1$. The white vertices in the network $G_1$, $G_2$, and $G_3$ represent the reference vertex.

3.1. Notation

The notation ${\rm Hypergeo}(t, r, s)$ stands for the hypergeometric random variable associated with the random sampling of $s$ objects out of a total of $t$ objects, of which $r$ objects are of a special type. So, the hypergeometric random variable counts the number of special objects in the sample.

It is customary to call the cardinality of the vertex set of a graph the order of the graph and reserve the term size of the graph to the cardinality of the set of edges in the graph. Let $V_n$ be the set of vertices of the graph $G_n$, and ${\mathcal {E}}_n$ be the set of edges of that graph. Thus, the seed $G_0 =(V_0, {\mathcal {E}}_0)$ is a connected graph with the set $V_0$ of vertices and the set ${\mathcal {E}}_0$ of edges.

Let $\tau _n$ be the order of the graph at age $n$. Hence, the cardinality of the vertex set $V_0$ of the seed is $|V_0|=\tau _0$. The $n$th hooking step adds $k_{n-1}$ copies of the seed at $k_{n-1}$ distinct latches chosen uniformly at random from $G_{n-1}$. Each copy contributes $\tau _0 -1$ new vertices to the network. The reason for subtracting 1 is the absorption of the hook. This gives the recurrence

(1) \begin{equation} \tau_n = \tau_{n-1} + k_{n-1} (\tau_0-1) . \end{equation}

Unwinding this recurrence, we obtain

(2) \begin{equation} \tau_n = (\tau_0-1)\sum_{i=1}^n k_{i-1} + \tau_0. \end{equation}

We use the notation $\deg (v)$ to denote the degree of node $v$ in a given graph, and we set $h^* = \deg (h)$.

3.2. Useful limits

By the regularity conditions, we can argue from (2) that

$$\frac {\tau_n} {k_{n-1}} = (\tau_0-1)\frac 1 {k_{n-1}}\sum_{i=0}^{n-1} k_i + \frac {\tau_0} {k_{n-1}} \to \frac {\tau_0-1} a+b =: \gamma.$$

Reorganize (1) as

$$1 = \frac {\tau_{n-1}} {\tau_n} + \frac {k_{n-1}}{\tau_n} (\tau_0-1) .$$

to find the limit

$$\frac {\tau_{n-1}} {\tau_n} = 1 - \frac {k_{n-1}}{\tau_n} (\tau_0-1) \to 1- \frac{(\tau_0-1)} {\gamma}.$$

4. A degree profile of the network

Various aspects of the degrees of nodes in a network are of interest in different contexts. For example, in the language of epidemiology, the degree of a node may be a useful representation of a highly infective person. From a health policy point of view, having knowledge about the degrees in conjunction with other graph parameters may help in identifying hot spots that trigger outbreaks and may be useful in controlling and mitigating the contagion. In the context of a social network, the degree of a node may represent the popularity and social skills of the person represented by the node.

Equally interesting are the global overall average degree in the entire graph (where we look at all the nodes), the local degree of a specific node during its temporal evolution, and the number of nodes of the smallest degree. We deal with the average behavior of each of these in a separate subsection. The different aspects of the degree complete a profile of the graph.

4.1. Evolution of the degree of a specific node

Suppose a node appears for the first time at step $j$. What will become of its degree at step $n$? At step $j$, several copies are added. To avoid a heavy notation identifying the time of appearance $j$, the copy number, which node within the copy to be tracked, and $n$, we use a simpler notation that needs only $j$ and $n$, for after all nodes of the same degree in the seed have the same distribution over time.

Theorem 4.1. Suppose $\{k_n\}_{n=0}^\infty$ is a building sequence of the family of graphs $\{G_n\}_{n=1}^\infty$. Let $X_{j:n}$ be the degree of a node at time $n$ that had appeared for the first time at step $j$. If initially its degree (in the seed) is $\delta$, then we have

$${\mathbb{E}}[X_{j:n}] =\delta + h^* \sum_{i=j}^{n-1} \frac {k_i} {\tau_i} = \frac {h^*a} {(1-a)(\tau_0 -1) + ab}(n-j) + o(n-j) + O(1).$$

Proof. Suppose a node $v$ appears at time $j$ for the first time. So, it belongs to one of the copies adjoined to the graph at that time. As the graph evolves, in any single step, the degree of $v$ can increase, if it is one of the nodes selected as latches in that step; otherwise its degree stays put, and when it does increase, it goes up by $h^* = \deg (h)$, the degree of the hook in the seed. This gives rise to a recurrence:

$$X_{j:n} = X_{j:n-1} + h^* {\mathbb{I}}_{n-1}(v),$$

where ${\mathbb {I}}_{n-1}(v)$ is an indicator of the event of choosing $v$ among the $k_{n-1}$ latches of that step of growth. On average, we have

$${\mathbb{E}}[X_{j:n}] = {\mathbb{E}}[X_{j:n-1}] + h^* \frac {{\tau_{n-1}-1\choose k_{n-1}-1}} {{\tau_{n-1}\choose k_{n-1}}} = {\mathbb{E}}[X_{j:n-1}] +h^* \frac {k_{n-1}} {\tau_{n-1}}.$$

Unwinding the recurrence, we obtain the exact average:

$${\mathbb{E}}[X_{j:n}] =\delta + h^* \sum_{i=j}^{n-1} \frac {k_i} {\tau_i}.$$

By the limits in Subsection 3.2, we obtain

$${\mathbb{E}}[X_{j:n}] = O(1) + o(n-j) + \frac {h^*a} {(1-a)(\tau_0 -1) + ab}(n-j).$$

Remark 4.1. If $a=0$ the ${\mathbb {E}}[X_{j:n}]$ is only $o(n-j)$.

Remark 4.2. Consider the case $a \gt 0$. The average in Theorem 4.1 indicates that the degree of a specific node experiences phases. The degree of a node in the early phase with $j = j(n) = o(n)$ grows linearly with its age in the network. When $j(n)\sim \rho n$, for $0 \lt \rho \lt 1$, we still get a linear growth, but the coefficient of linearity is attenuated to $(1-\rho) ({h^*a}/{[(1-a)(\tau_0 -1) + ab]})$. At $\rho =1$, we have ${\mathbb {E}}[X_{j:n}] = o(n)$.

Remark 4.3. If $a=0$, we can only assert that ${\mathbb {E}}[X_{j:n}] =o(n-j) + \delta$. In this case, a finer analysis is needed to identify the leading order of the average degree of a node that appears at time $j$. For instance, in the case of a tree grown from the complete graph $K_2$, we have $\delta = 1, h^*=1, k_n =1$, and $a=0$. The exact formula in this case yields

$${\mathbb{E}}[X_{j:n}] =1 + \sum_{i=j}^{n-1} \frac 1 {i+2}=H_{n+1} - H_{j+1} + 1.$$

Whence, we have the phases

$${\mathbb{E}}[X_{j:n}]\sim \begin{cases} \ln n, & \text{if} \ j \ \text{is fixed};\\ \ln \dfrac {n} {j(n)}, & \text{if} \ j(n)\to \infty\ \text{and} \ j(n) = o(n);\\ \ln \dfrac 1 {\rho}, & \text{if} \ j(n) \sim \rho n,\ 0 \lt \rho \lt 1;\\ 1, & \text{if} \ j(n) = n - o(n). \end{cases}$$

4.2. The overall average degree

The main result about the overall average degree in the graph is developed in this section. The result is expressed in terms of $\eta$, the number of edges in the seed graph.

Theorem 4.2. Suppose $\{k_n\}_{n=0}^\infty$ is a building sequence of the family of graphs $\{G_n\}_{n=1}^\infty$. Let $Y_n$ be the degree of a randomly chosen node in the graph $G_n$ at age $n$. We have

$$\lim_{n\to\infty}{\mathbb{E}}[Y_n] =\frac {2\eta}{\tau_0-1}.$$

Proof. Upon hooking $k_{n-1}$ copies of the seed to $k_{n-1}$ distinct nodes of $G_{n-1} = (V_{n-1}, {\mathcal {E}}_{n-1})$, we add $\eta k_{n-1}$ edges to the graph. Therefore, we have

$$|{\mathcal{E}}_n| = |{\mathcal{E}}_{n-1}| + \eta k_{n-1}.$$

This recurrence has the solution

$$|{\mathcal{E}}_n| = \eta\left(1+\sum_{i=0}^{n-1} k_i\right).$$

Using the classical First Theorem of Graph Theory, we obtain

$$\sum_{v\in V_n} \deg (v) = 2{|\mathcal{E}}_n| = 2 \eta\left(1+\sum_{i=0}^{n-1} k_i\right).$$

Scaling the equation by $\tau _n$, we get

$$\frac 1 {\tau_n} \sum_{v\in V_n} \deg (v) = \frac {2\eta} {\tau_n}\left(1+\sum_{i=0}^{n-1} k_i\right).$$

Taking limits, and using Eq. (2), we obtain

$$\lim_{n\to\infty} {\mathbb{E}}[Y_n]= 0+\lim_{n\to\infty}2\eta \frac{\sum_{i=0}^{n-1} k_i} {\tau_n} = \frac {2\eta}{\tau_0-1 }.$$

Remark 4.4. The average degree in the seed is $2\eta /\tau _0$. For any building sequence, the asymptotic average degree in the graph is $2\eta / (\tau _0-1)$, only slightly higher than the average degree in the initial seed. This should be anticipated because the additions introduce a number of copies of the seed, each of which has the degree properties of the seed with the hook eliminated.

4.3. Nodes of the smallest degree

We study only the nodes of the smallest degree. Let $d^*$ be the smallest degree in the seed. Note that the smallest admissible degree in the graph is $d^*$. After the network grows, the smallest degree in it may be $d^*$ or higher. Let $X_n$ be the number of nodes of degree $d^*$ at time $n$. Thus, $X_0$ is the number of nodes of degree $d^*$ in the seed. Later graphs can have more nodes of degree $d^*$. The seed in Figure 1 has $d^* = 2$, and $X_0 = 1, X_1 = 2, X_2=3$, and $X_3 =3$.

4.3.1. Stochastic recurrence

In the evolution at step $n$, we hook $k_{n-1}$ copies of the seed to the graph $G_{n-1}$. Let $A_0$ be the event $\deg (h) = d^*$ and $\mathbb {I}_{A_0}$ be an indicator that assumes value 1, if $\deg (h) = d^*$, otherwise, it assumes the value 0. A latch of degree $d^*$ in the sample will have a higher degree (namely, its degree goes up to $d^* + \deg (h)$) in $G_n$. So, we lose such vertices in the count of $X_n$. If the hook degree is $d^*$, every hooked copy contributes only $X_0-1$ vertices of degree $d^*$.

For the case when $k_{n-1} = 1$ and the latch is $\ell$, the change from $X_{n-1}$ to $X_n$ for the four cases can be seen as shown in the table below:

Thus, for any value of $k_{n-1}$, the count $X_n$ therefore satisfies a (conditional) stochastic recurrence:

(3) \begin{align} X_n & = X_{n-1} + (X_0 - 1 -\mathbb{I}_{A_0})\, {\rm Hypergeo}(\tau_{n-1}, X_{n-1}, k_{n-1}) \nonumber\\ & \quad + (X_0 - \mathbb{I}_{A_0})(k_{n-1} -{\rm Hypergeo}(\tau_{n-1} , X_{n-1}, k_{n-1})). \end{align}

4.3.2. The average proportion of nodes of degree $d^*$

Take (conditional) expectation of (3) to get

(4) \begin{align} {\mathbb{E}}[X_n\, \vert \, G_{n-1}] & = X_{n-1} + (X_0 - 1 -\mathbb{I}_{A_0}) \frac {X_{n-1}} {\tau_{n-1}} k_{n-1} \nonumber\\ & \quad + (X_0 - \mathbb{I}_{A_0})\left(k_{n-1} - \frac {X_{n-1}} {\tau_{n-1}} k_{n-1}\right) \nonumber\\ & = \left(1 -\frac {k_{n-1}} {\tau_{n-1}} \right) X_{n-1} + (X_0 - \mathbb{I}_{A_0}) k_{n-1}. \end{align}

Theorem 4.3. Suppose $\{k_n\}_{n=0}^\infty$ is a building sequence of the family of graphs $\{G_n\}_{n=1}^\infty$, starting from a seed with $X_0$ nodes of the smallest degree $d^*$. Let $X_n$ be the number of vertices of this degree in the graph after $n$ steps of evolution according to the building sequence. We have

$${\mathbb{E}}[X_n] = (X_0-\mathbb{I}_{A_0}) \sum_{i=1}^n k_{i-1}\prod_{j=i+1}^n \left(1 - \frac {k_{j-1}} {\tau_{j-1}}\right) +X_0\prod_{j=1}^n \left(1 - \frac {k_{j-1}} {\tau_{j-1}}\right).$$

Subsequently, the average proportion converges to a limit independent of the limits $a$ and $b$; namely we have the convergence

$${\mathbb{E}}\left[ \frac {X_n} {\tau_n}\right] \to \frac {X_0-\mathbb{I}_{A_0}} {\tau_0}.$$

Proof. Taking a double expectation of (4) yields

(5) \begin{equation} {\mathbb{E}}[X_n] = \left(1 - \frac {k_{n-1}} {\tau_{n-1}}\right) {\mathbb{E}}[X_{n-1}] + (X_0 - \mathbb{I}_{A_0}) k_{n-1} . \end{equation}

This recurrence equation is of the standard linear form

(6) \begin{equation} y_n = g_n y_{n-1} + h_n, \end{equation}

with solution

(7) \begin{equation} y_n = \sum_{i=1}^n h_i\prod_{j=i+1}^n g_j + y_0 \prod_{j=1}^n g_j. \end{equation}

So, the sought solution for the average of the number of nodes of degree $d^*$ (for $n\ge 1$) is

$${\mathbb{E}}[X_n] = (X_0-\mathbb{I}_{A_0}) \sum_{i=1}^n k_{i-1} \prod_{j=i+1}^n \left(1 - \frac {k_{j-1}} {\tau_{j-1}}\right) + X_0 \prod_{j=1}^n \left(1 - \frac {k_{j-1}} {\tau_{j-1}}\right).$$

The strategy for the asymptotic part of the statement is two-fold: We prove the existence of a limit (under any building sequence) for the proportion from the exact solution. We then find the value of the limit from the recurrence under the mild regularity conditions imposed on the building sequence.

First, express the expected proportion as

(8) \begin{equation} {\mathbb{E}}\left[\frac {X_n}{\tau_n}\right] = (X_0-\mathbb{I}_{A_0})\sum_{i=1}^n \frac {k_{i-1}} {\tau_n}\prod_{j=i+1}^n \left(1 - \frac {k_{j-1}} {\tau_{j-1}}\right) + \frac {X_0} {\tau_n} \prod_{j=1}^n \left(1 - \frac {k_{j-1}} {\tau_{j-1}}\right); \end{equation}

at $i=n$, the first product does not exist, and is taken to be 1, as usual. Let

$$c_n = \sum_{i=1}^n \frac{k_{i-1}} {\tau_n}\prod_{j=i+1}^n \left( 1 - \frac {k_{j-1}} {\tau_{j-1}}\right) = \sum_{i=1}^{n-1} \frac{k_{i-1}} {\tau_n}\prod_{j=i+1}^n \left( 1 - \frac {k_{j-1}} {\tau_{j-1}}\right)+ \frac {k_{n-1}}{\tau_n};$$

We manipulate this to turn it into a recurrence as follows:

\begin{align*} c_{n+1} & = \sum_{i=1}^n \frac{k_{i-1}} {\tau_{n+1}}\prod_{j=i+1}^{n+1} \left( 1 - \frac {k_{j-1}} {\tau_{j-1}}\right) + \frac {k_n}{\tau_{n+1}}\\ & = \frac {\tau_n}{\tau_{n+1}}\left(1 - \frac {k_n}{\tau_n} \right)\sum_{i=1}^n \frac{k_{i-1}} {\tau_n}\prod_{j=i+1}^n \left( 1 - \frac {k_{j-1}} {\tau_{j-1}}\right) + \frac {k_n}{\tau_{n+1}} \\ & = \left(\frac {\tau_n - k_n}{\tau_{n+1}} \right) c_n+ \frac {k_n}{\tau_{n+1}} \\ & = \left(\frac {\tau_{n+1} - (\tau_0-1)k_n - k_n}{\tau_{n+1}} \right) c_n+ \frac {k_n}{\tau_{n+1}} \\ & = \left(\frac {\tau_{n+1} - \tau_0 k_n }{\tau_{n+1}} \right) c_n+ \frac {k_n}{\tau_{n+1}} \end{align*}

Rearrange the recurrence in the form

$$c_{n+1} - \frac 1 {\tau_0} =c_n - \frac {\tau_0 k_n }{\tau_{n+1}}\, c_n + \frac {k_n}{\tau_{n+1}} -\frac 1 {\tau_0} = \left(c_n - \frac 1 {\tau_0}\right) \left(\frac {\tau_{n+1}-\tau_0 k_n}{\tau_{n+1}}\right),$$

leading to the inequality

\begin{align*} \left|c_{n+1} - \frac 1 {\tau_0} \right| & \le\left|c_n - \frac 1 {\tau_0}\right| \left|\frac {\tau_{n+1}-\tau_0 k_n + k_n}{\tau_{n+1}}\right| \\ & = \left|c_n - \frac 1 {\tau_0} \right| \frac{\tau_n}{\tau_{n+1}}\\ & \le \left|c_{n-1} - \frac 1 {\tau_0} \right| \frac {\tau_n}{\tau_{n+1}} \times \frac {\tau_{n-1}}{\tau_n}\\ & \quad \vdots\\ & \le \left|c_0 - \frac 1 {\tau_0} \right| \frac {\tau_n}{\tau_{n+1}} \times \frac {\tau_{n-1}}{\tau_n} \times \cdots\times \frac {\tau_0}{\tau_{1}}. \end{align*}

Noting that the sum in $c_n$ is empty at $n=0$, we have $c_0 = 0$ and the bounds simplify to $0 \le |c_n - 1 /\tau _0| \le 1/\tau _n$. So, both inferior and superior limits of $|c_n - 1 /\tau _0|$ are equal to $0$, which furnishes the existence of a limit for $c_n$ equal to $1/\tau _0$, too.

As for the remainder part

$$r_n := \frac {X_0}{\tau_n}\prod_{j=1}^n \left( 1 - \frac {k_{j-1}} {\tau_{j-1}}\right),$$

in (8), it clearly converges to 0, as $\tau _n$ is increasing, and the product is bounded from above by 1. Plugging in the limits $\lim _{n\to \infty } c_n= 1/\tau _0$ and $\lim _{n\to \infty } r_n = 0$ in (8), we reach the conclusion that

$$\lim_{n\to \infty} {\mathbb{E}}\left[ \frac {X_n} {\tau_n}\right] = \frac {X_0- \mathbb{I}_{A_0}} {\tau_0}.$$

Remark 4.5. In the case when the hook is not of the smallest degree $d^*$, we have ${\mathbb {E}}[X_n/\tau _n] \to X_0/\tau _0$. The initial proportion of nodes of the smallest degree in the seed is preserved on average in larger subsequent graphs.

Remark 4.6. In the case when the hook is of the smallest degree $d^*$, we have ${\mathbb {E}}[X_n/\tau _n] \to (X_0-1)/\tau _0$. The long-term proportion of nodes of the smallest degree is less than the proportion of nodes of degree $d^*$ in the seed.

Remark 4.7. In the case when the hook is the only node of the smallest degree $d^*$ in the seed, we have $X_0=1$, and ${\mathbb {E}}[X_n/\tau _n] \to 0$, for all $n\ge 1$. Indeed, the degree $d^*$ disappears after the first latching at the initial hook and never reappears.

Remark 4.8. The limit in Theorem 4.3 is more than just an ultimate value in the take-all case. In this case, it is the actual value for each $n\ge 0$, which can be seen from the recurrence. The only term that does not vanish is the last term in sum, yielding $(X_0 - I_{A_0}) k_{n-1}/ \tau _n = (X_0 - I_{A_0}) / \tau _0$.

5. Distances in the network

We measure node distances in $G_n$ relative to a reference point (vertex). We take the reference to be the hook of $G_0$. We look at two (related) kinds of distances: The total path length and the average distance in the graph. Let the nodes of the $n$th graph be labeled with the numbers $1,2, \ldots, \tau _n$, with 1 being reserved for the reference vertex and the rest of the nodes are arbitrarily assigned distinct numbers from the set $\{2, \ldots, \tau _n\}$. The depth of a node in the network is its distance from the reference vertex (i.e., the length of the shortest path from the node to the reference vertex measured in the number of edges). We denote the depth of the $i$th node in the $n$th network by $\Delta _{n,i}$ The total path length is the sum of all depths; namely it is

$$T_n = \sum_{i=1}^{\tau_n} \Delta_{n,i}.$$

For instance, the networks $G_0, G_1, G_2$, and $G_3$ in Figure 1 have total path lengths $T_0=3, T_1 = 6$, $T_2 = 18$, and $T_3 = 45$, respectively.

5.1. Average total path length

As the network grows, at step $n$, a sample of size $k_{n-1}$ latches is chosen from $G_{n-1}$ to grow $G_{n-1}$ into $G_n$. Suppose these latches are $\ell _1, \ldots, \ell _{k_{n -1}}$ at depths $d_1, \ldots, d_{k_{n-1}}$. In view of the absorption of the hooks of the added graphs, a copy's hook fused at the $j$th latch adds $\tau _0-1$ nodes, which appear in $G_n$ at depths equal to their distance from the hook of the copy translated by an additional distance from the latch to the reference vertex. So, collectively, the vertices of the copy hooked to $\ell _j$ increase the total path length by $T_0 + (\tau _0-1) d_j$. We have a conditional recurrence:

$${\mathbb{E}}[T_n \, \vert \, G_{n-1}, d_1, \ldots, d_{k_{n-1}}] = T_{n-1} + \sum_{j=1}^{k_{n-1}} (T_0 + (\tau_0-1) d_j).$$

Averaging over the graphs and the choices of the latches within, we get

(9) \begin{equation} {\mathbb{E}}[T_n] = {\mathbb{E}}[T_{n-1}] + T_0 k_{n-1} + (\tau_0-1) \sum_{j=1}^{k_{n-1}} {\mathbb{E}}[ \Delta_{n-1,\ell_j}]. \end{equation}

Lemma 5.1.

$$\sum_{j=1}^{k_{n-1}} {\mathbb{E}}[ \Delta_{n-1,\ell_j}] = \frac {k_{n-1}} {\tau_{n-1}} \, {\mathbb{E}}[T_{n-1}].$$

Proof. Condition on the event $\ell _1 =i_1, \ldots, \ell _{k_{n-1}} = i_{k_{n-1}}$, to get

\begin{align*} \sum_{j=1}^{k_{n-1}} {\mathbb{E}}[ \Delta_{n-1,\ell_j}] & = \sum_{j=1}^{k_{n-1}} \sum_{1\le i_1 \lt i_2 \lt \cdots \lt i_{k_{n-1}} \le \tau_{n-1}} {\mathbb{E}} [ \Delta_{n-1,\ell_j}\, \vert \, d_1 =i_1, \ldots d_{k_{n-1}} = i_{k_{n-1}}]\\ & \quad \times {\mathbb{P}}(d_1 =i_1, \ldots, d_{k_{n-1}} =i_{k_{n-1}}). \end{align*}

The subsets of size $k_{n-1}$ latches that appear in a sample of vertices from $G_{n-1}$ are all equally likely, and we get

\begin{align*} & \sum_{j=1}^{k_{n-1}} {\mathbb{E}}[\Delta_{n-1,\ell_j}] \\ & \quad =\frac 1 {{{\tau_{n-1}} \choose {k_{n-1}}}}\sum_{j=1}^{k_{n-1}} {\mathbb{E}}\left[ \left.\left(\sum_{1\le i_1 \lt i_2 \lt \cdots \lt i_{k_{n-1}} \le \tau_{n-1}} \Delta_{n-1,\ell_j}\right)\right| \ell_1 =i_1, \ldots \ell_{k_{n-1}}= i_{k_{n-1}}\right]. \end{align*}

Let us write out the inner sum in expanded form:

\begin{align*} & \Delta_{n-1,1} + \Delta_{n-1,2} +\cdots + \Delta_{n-1,k_{n-1}} \\ & \quad + \Delta_{n-1,1} + \Delta_{n-1,2} +\cdots + \Delta_{n-1,k_{n-1}-1} + \Delta_{n-1,k_{n-1}+1}\\ & \quad \enspace \vdots\\ & \quad + \Delta_{n-1,\tau_{n-1} - k_{n-1}+1} + \Delta_{n-1,\tau_{n-1} - k_{n-1}+2}+\cdots + \Delta_{n-1,\tau_{n-1}}. \end{align*}

Upon a reorganization collecting similar terms, we get

$${\tau_{n-1 -1} \choose k_{n-1}-1} (\Delta_{n-1,1} + \Delta_{n-1,2} + \cdots + \Delta_{n-1,\tau_{n-1}}) = {\tau_{n-1 -1} \choose k_{n-1}-1} T_{n-1}.$$

Plugging this expression in the expectation, we proceed to

$$\sum_{j=1}^{k_{n-1}} {\mathbb{E}}[\Delta_{n-1,\ell_j}] = \frac {{\tau_{n-1 -1} \choose k_{n-1}-1}} {{\tau_{n-1} \choose k_{n-1}}} {\mathbb{E}}[T_{n-1}] = \frac {k_{n-1}} {{\tau_{n-1}}} {\mathbb{E}}[T_{n-1}].$$

Theorem 5.1. Suppose $\{k_n\}_{n=0}^\infty$ is a building sequence of the family of graphs $\{G_n\}_{n=1}^\infty$, starting from a seed of total path length $T_0$. Let $T_n$ be the total path length after $n$ steps of evolution according to the building sequence. We have

$${\mathbb{E}}[T_n] = T_0\tau_n\left(\sum_{i=1}^n \frac {k_{i-1}} {\tau_i} + \frac 1 {\tau_0}\right).$$

Proof. By Lemma 5.1 and the recurrence (9), we have a recurrence for the average total path length:

\begin{align*} {\mathbb{E}}[T_n] & = {\mathbb{E}}[T_{n-1}] + T_0 k_{n-1}+ \frac {(\tau_0-1) k_{n-1}} {\tau_{n-1}} {\mathbb{E}}[T_{n-1}] \\ & =\left(1 + \frac {(\tau_0-1) k_{n-1}} {\tau_{n-1}} \right){\mathbb{E}}[T_{n-1}] + T_0 k_{n-1}. \end{align*}

Again, the recurrence is of the standard form (6) with the solution (7). In the specific case at hand, this solution is

$${\mathbb{E}}[T_n] = T_0\sum_{i=1}^n k_{i-1}\prod_{j=i+1}^n \left(1 + \frac {(\tau_0-1){k_{j-1}} } {\tau_{j-1}}\right) +T_0 \prod_{j=1}^n \left(1 + \frac {(\tau_0 -1)k_{j-1}} {\tau_{j-1}}\right).$$

The recurrence (1) on the order of the graph simplifies the solution into telescopic products

$${\mathbb{E}}[T_n] = T_0\sum_{i=1}^n k_{i-1}\prod_{j=i+1}^n \frac {\tau_j} {\tau_{j-1}} +T_0 \prod_{j=1}^n \frac {\tau_j} {\tau_{j-1}} = T_0 \tau_n \left(\sum_{i=1}^n \frac{k_{i-1}}{\tau_i} + \frac 1 {\tau_0}\right).$$

5.2. Average depth

Theorem 5.1 provides a benchmark for the calculation of the average depth. Let the depth of a randomly selected node in the $n$th network be $D_n$.

Corollary 5.1.

$${\mathbb{E}}[D_n] = T_0 \sum_{i=1}^n \frac{k_{i-1}}{\tau_i} + D_0.$$

Proof. Given a specific development leading to $G_n$, the average depth in that graph is

$${\mathbb{E}}[D_n \, \vert \, G_{n}] = \frac {\Delta_{n,1} +\cdots + \Delta_{n,\tau_n}} {\tau_n} = \frac{T_n}{\tau_n}.$$

Upon taking expectation, it follows that ${\mathbb {E}}[D_n] = {\mathbb {E}}[T_n] / \tau _n$. The form given in the statement ensues from Theorem 5.1.

Corollary 5.2. Under the regularity conditions (R1)–(R3), we have the asymptotic equivalent

$${\mathbb{E}}[D_n] = \frac {a T_0}{\tau_0 -1 + ab} n + o(n), \quad \text{as} \ n \to \infty.$$

Remark 5.1. Corollary 5.2 is more useful when

$$\lim_{n\to \infty }\frac {k_n} {\sum_{i=0}^n k_i} = a\not = 0.$$

When $a=0$, as in the case of trees for example, one needs to sharpen the argument to find the leading asymptotic term, as we do in some specific cases below.

5.3. Distances under specific building sequences

At one extreme, there is the sequence of least possible growth ($k_n= 1$). At the other extreme, we have a take-all model ($k_n = \tau _n$) in which all the nodes of $G_n$ are taken as latches for $\tau _{n}$ copies of the seed.

In the case of $k_n = k$, for fixed $k\in \mathbb {N}$, of nearly the least growth, the average depth is

$${\mathbb{E}}[D_n] = T_0 \sum_{i=1}^n \frac k {\tau_i} +D_0.$$

The limit $a$ in regularity condition (R2) is 0, and Corollary 5.2 only tells us that ${\mathbb {E}}[D_n] =o(n)$. However, we can sharpen the asymptotic equivalence from the specific construction of the case.

Here, we have $\tau _i = (\tau _0 - 1) i + \tau _0$, which gives

$${\mathbb{E}}[D_n] = \frac {T_0 k} {\tau_0-1} \sum_{i=1}^n \frac 1{ i + \frac {\tau_0} {\tau_0-1}} + D_0.$$

In terms of the generalized harmonic numbersFootnote 1

$$H_n(x) = \frac 1 {1+x} + \frac 1 {2+x} +\cdots+\frac 1 {n+x},$$

the depth in the near-least-growth is compactly expressed as

$${\mathbb{E}}[D_n] = \frac {T_0 k} {\tau_0-1} H_n\left(\frac {\tau_0}{\tau_0-1}\right) + D_0 \sim \frac {T_0 k} {\tau_0-1} \ln n, \quad \text{as} \ n\to\infty.$$

Remark 5.2. The case $k=1$ and $\tau _0=2$ grows a recursive tree. The seed is a rooted tree on two vertices, in which $T_0=1$ and $D_0 = \frac 1 2$. In this case, the average depth becomes

$${\mathbb{E}}[D_n] =H_n(2) + \frac 1 2 = H_{n+2} - 1 \sim \ln n, \quad \text{as} \ n\to\infty,$$

which recovers a known result [Reference Smythe and Mahmoud19].

Remark 5.3. At the other end of the spectrum, there is the take-all model, in which $k_i = \tau _i$, leading at step $n$ to a graph of order $\tau _n = \tau _0^{n+1}$. Here, the limit $a$ is $(\tau _0-1)/\tau _0$ and the limit $b$ is 0. According to Corollary 5.2, we have ${\mathbb {E}}[D_n] \sim D_0 n$, as $n\to \infty$. This asymptotic estimate can be sharpened as the case is amenable to exact calculation:

$${\mathbb{E}}[D_n] = T_0\sum_{i=1}^n \frac {\tau_{i-1}} {\tau_i} + D_0 = T_0\sum_{i=1}^n \frac 1 {\tau_0} + D_0= D_0 (n+1) .$$

6. Eccentricity

The eccentricity $C(v)$ of a node $v$ in a graph $\mathcal {G}$ is the distance between $v$ and a vertex farthest from $v$ in $\mathcal {G}$. The eccentricity is instrumental in constructing a notion of the diameter of a graph (extreme distances). We use the eccentricity of the hook and the various latches selected in $G_{n-1}$ to determine the diameter of the graph $G_n$.

The eccentricity is technically defined as follows. If $\mathcal {Q}$ is a path in a graph $\mathcal {G}$, we denote its length by $|\mathcal {Q}|$ (the number of edges in it). There can be several paths joining two vertices $u$ and $v$ in $\mathcal {G}$, and the distance between $u$ and $v$, denoted by $d(u,v)$, is the length of the shortest such path. That is, with ${\mathcal {P}}(u,v)$ denoting the collection of paths between $u$ and $v$, the distance between these two nodes is given by

$$d(u,v) = \min_{{\mathcal{Q}}\in {\mathcal{P}}(u,v)} |{\mathcal{Q}}|.$$

The eccentricity $C(v)$ of a vertex $v$ in a graph with vertex set $\mathcal {V}$ is:

$$C(v) = \max_{u\in \mathcal{V}} d(v,u) = \max_{u\in \mathcal{V}} \min_{{\mathcal{Q}}\in {\mathcal{P}}(v,u)} |{\mathcal{Q}}|.$$

For instance, the eccentricity in Figure 1 of the reference vertex of $G_0$ is 2, of the reference vertex in $G_1$ is 2 as well, but of the reference vertex in $G_2$ is 4 and becomes 6 in $G_3$.

6.1. Eccentricity of a node in $G_n$

The $k_{n-1}$ nodes selected as latches from the graph $G_{n-1}= (V_{n-1}, {\mathcal {E}}_{n-1})$ are vertices that play a key role in designing the network at stage $n$ and onward and contribute significantly in determining the diameter of the graph at the next stage.

As a node's eccentricity changes over time, its value at step $n$ in $G_n$ may be different from its value at step $n-1$ in $G_{n-1}$. We need an eccentricity notation reflecting the possible change over time. For that we use $C_n(v)$ to speak of the eccentricity of a vertex $v$ in $G_n$.

If $v \in V_n$ is a vertex in a copy of $G_0$ latched at a vertex $\ell _i \in V_{n-1}$, we express that by saying $v \in V_0^{co_i}$, otherwise we say $v \in V_{n-1}$. We now introduce some notation:

  1. 1. $\mathbb {L}_n = \{\ell _1, \ell _2, \cdots \ell _{k_n}\}$ is the set of latches selected in the graph $G_n$ to produce the graph $G_{n+1}$.

  2. 2. $\widetilde {C}_n(v) = C_n(v) \, \vert \, G_{n-1}, \mathbb {L}_{n-1}$. This is the conditional eccentricity $C_n(v)$ of the node $v$ in the graph $G_n$, given $G_{n-1}$ and the $k_{n-1}$ latches in it.

  3. 3. For any $v \in V_{n-1}$, we define $d_v^{\#} = \max _{\ell _j \in \mathbb {L}_{n-1}} d(v, \ell _j)$. So, $d_v^{\#}$ is the maximum distance from $v$ to the nodes in $\mathbb {L}_{n-1}$.

Also, in what follows we use the notation ${\mathbb {I}}_{\mathcal {C}}$ to indicate a predicate (condition) $\mathcal {C}$. So, it is 1, when $\mathcal {C}$ holds, and is 0, otherwise.

Theorem 6.1. Suppose $\{k_n\}_{n=0}^\infty$ is a building sequence of the family of graphs $\{G_n\}_{n=1}^\infty$. Let $v$ be a node in the graph $G_n$. Conditional upon the choice of the latches $\ell _1, \ldots, \ell _{k_{n-1}}$ in $G_{n-1}$, the eccentricity $C_n(v)$ is given by

$$\widetilde{C}_n(v)= \begin{cases} \max\{C_{n-1}(v), d_v^{\#}+C_0(h)\}, & \text{if} \ v\in V_{n-1};\\ \max\{d(v,h) + C_{n-1}(\ell_i), & \\ \quad \max\{C_0(v) , (d(v,h) + d_{\ell_i}^{\#} + C_0(h)){\mathbb{I}}_{\{k_{n-1} \gt 1\}}\}\}, & \text{if} \ v \in V_0^{co_i}. \end{cases}$$

Proof. The graph $G_n$ is obtained by attaching a copy of the seed $G_0$ at each of the latches $\ell _1, \ell _2, \cdots, \ell _{k_{n-1}}$ selected in the graph $G_{n-1}$.

We denote the vertex set of the $r$th copy of the seed, for $r= 1, \ldots, k_{n-1}$, by $V_0^{{\rm co}_r}$. We now compute the distance from a node $v$ to a vertex $u$ in $G_n$ by considering the four cases:

For a given $v\in V_n$, the maximum (over the range of $u$ and $j$) in each block is

The result now follows.

Remark 6.1. Suppose a vertex $\ell$ is chosen as a latch from $G_{n-1}$. From Theorem 6.1, we pick up the top line and write

$$\widetilde C_n(\ell) = \max \{C_{n-1}(\ell), d_\ell^{\#}+ C_0(h)\}.$$

If $k_{n-1} = 1$, then $d_\ell ^{\#}=0$, in which case we have $\widetilde C_n(\ell ) = \max \{C_{n-1}(\ell ), C_0(h)\}$.

7. Diameter of the graph $G_n$

The diameter of a connected graph with vertex set $\mathcal {V}$ is the longest distance between any two nodes in it [Reference Chartrand and Zhang4]. That is, the diameter is the maximum eccentricity, $\max _{v\in \mathcal {V}} C(v)$. For example, the diameters of the graphs $G_0,G_1,G_2$, and $G_3$ in Figure 1 are, respectively, $2, 4, 6$, and $10$.

We now introduce some additional notation:

  1. 1. $\widetilde {{\unicode{x0414}}_{n}} = {\unicode{x0414}}_{n} \, \vert \, G_{n-1}, \mathbb {L}_{n-1}$. This is the conditional diameter ${\unicode{x0414}}_{n}$ of the graph $G_n$ given $G_{n-1}$ and the $k_{n-1}$ latches in $\mathbb {L}_{n-1}$ (see Subsection 6.1 for the definition of $\mathbb {L}_{n-1}$).

  2. 2. Only for $k_n \gt 1$, we define $q_{n} = \max _{\ell, \tilde \ell \in \mathbb {L}_{n}} d(\ell, \tilde \ell ) = \max _{\ell \in \mathbb {L}_{n}} \ell ^{\#}$. Thus, $q_{n}$ computes the maximum distance between any two latches in $G_{n}$.

  3. 3. $\alpha _{n} = \max _{\ell \in \mathbb {L}_{n}} C_{n}(\ell )$. Thus, $\alpha _{n}$ is the maximum eccentricity of a latch in $G_{n}$.

Theorem 7.1. Suppose $\{k_n\}_{n=0}^\infty$ is a building sequence of the family of graphs $\{G_n\}_{n=1}^\infty$. The conditional diameter $\widetilde {{\unicode{x0414}}_{n}} = {\unicode{x0414}}_n\, \vert \, G_{n-1}, \mathbb {L}_{n-1}$ of a graph of age $n$ is given by

$$\widetilde{{\unicode{x0414}}_{n}} = \max \{{\unicode{x0414}}_{n-1}, (2C_0(h) + q_{n-1}) {\mathbb{I}}_{\{k_{n-1} \gt 1\}},\ C_0(h) + \alpha_{n-1} \}.$$

Proof. The (conditional) diameter $\widetilde {{\unicode{x0414}}_{n}}$ of $G_n$ may remain the same as the diameter ${\unicode{x0414}}_{n-1}$ of $G_{n-1}$,Footnote 2 unless we can find longer paths in $G_n$. The latter case arises, if

  1. (a) There is a pair $x$ and $y$ of latches in $G_{n-1}$, and a pair of vertices (say $u$ in the copy latched at $x$ and $v$ in the copy latched at $y$), such that $d(u,x) + d(x,y) + d(y, v) \gt {\unicode{x0414}}_{n-1}$. The case can be, only if $k_{n-1} \gt 1$. The longest such distance is obtained by maximizing over $x,y,u$ and $v$ to obtain $2C_0(h) + q_{n-1}$.Footnote 3

  2. (b) Or, we can find a vertex $u$ far enough from a latch $\ell$ in $G_{n-1}$ and another vertex $v$ in the copy latched at $\ell$ such that $d(u,\ell ) + d(\ell, v) \gt {\unicode{x0414}}_{n-1}$. The longest such distance is obtained by maximizing over $\ell, u$ and $v$ to obtain $\alpha _{n-1}+C_0(h)$.

The longest distance in the graph is the maximum of the three possibilities discussed.

Remark 7.1. Consider the case where, at stage $n$ (for each $n\ge 1$), we pick among the $k_{n-1}$ latches two, say $\ell, \tilde \ell$ in $G_{n-1}$, such that $d(\ell, \tilde \ell )$ is the diameter of $G_{n-1}$. Note that this selection mechanism is no longer random in the sense discussed in all the preceding sections. Let us call the diameter of the graph so constructed ${\unicode{x0414}}_{n-1}^*$. This is only possible if $k_{n-1} \gt 1$, for each $n$. By arguments similar to what we used in the proof of Theorem 7.1, we get ${\unicode{x0414}}_{n}^* = 2C_0(h) + {\unicode{x0414}}_{n-1}^*$. Unwinding we get ${\unicode{x0414}}_{n}^* = 2n C_0(h) + {\unicode{x0414}}_{0}$.

Remark 7.1 shows that, under this special hooking mechanism, the diameter ${\unicode{x0414}}_{n}$ at step $n$ only requires the knowledge of the seed graph and $n$. It does not take into consideration how many latches were picked at stages $1$ through $n-1$ as long as there are two latches $\ell, \tilde \ell$ picked at each stage such that $d(\ell, \tilde \ell )$ is maximum.

Acknowledgment

The authors would like to sincerely thank Dr. Shaimaa M. Abd-Elaal, Egyptian Ministry of Health and Population, for many fruitful discussions that gave them a perspective on applications of the model proposed.

Competing interest

The authors declare no conflict of interest.

Footnotes

1 Customarily, $H_n (0)$ is denoted by $H_n$.

2 In the graph $G_2$ in Figure 1, if we pick the three latches at distances 2,3,4 from the top vertex, the diameter of the graph so obtained in step 3 will be equal to ${\unicode{x0414}}_2$, the diameter of $G_2$.

3 This situation occurs in the graph $G_3$ in Figure 1.

References

Bahrani, M. & Lumbroso, J. (2018). Split-decomposition trees with prime nodes: Enumeration and random generation of cactus graphs. 2018 Proceedings of the Fifteenth Workshop on Analytic Algorithmics and Combinatorics (ANALCO), New Orleans, Louisiana, pp. 143–157.CrossRefGoogle Scholar
Bhutani, K., Kalpathy, R., & Mahmoud, H. (2021). Average measures in polymer graphs. International Journal of Computer Mathematics: Computer Systems Theory 6(1): 3753. doi:10.1080/23799927.2020.1860134Google Scholar
Brown, G. & Shubert, B. (1984). On random binary trees. Mathematics of Operations Research 9: 4365.CrossRefGoogle Scholar
Chartrand, G. & Zhang, P (2012). A first course in graph theory. Mineola, NY: Dover Publications.Google Scholar
Chen, C. & Mahmoud, H. (2016). Degrees in random self-similar bipolar networks. Journal of Applied Probability 53: 434447.CrossRefGoogle Scholar
De La Briandais, R. (1959). File searching using variable length keys. Proceedings of the Western Joint Computer Conference. San Francisco, CA: AFIPS, pp. 295–298.CrossRefGoogle Scholar
Desmarais, C. & Holmgren, C. (2019). Degree distributions of generalized hooking networks. 2019 Proceedings of the Sixteenth Workshop on Analytic Algorithmics and Combinatorics (ANALCO), San Diego, California, pp. 103–110.CrossRefGoogle Scholar
Desmarais, C. & Holmgren, C. (2020). Normal limit laws for vertex degrees in randomly grown hooking networks and bipolar networks. The Electronic Journal of Combinatorics 27(2): P2.45.CrossRefGoogle Scholar
Desmarais, C. & Mahmoud, H. (2021). Depths in hooking networks. Probability in the Engineering and Informational Sciences 19. doi:10.1017/S0269964821000164Google Scholar
Drmota, M (2009). Random trees: An interplay between combinatorics and probability. Vienna: Springer-Verlag.CrossRefGoogle Scholar
Drmota, M., Gittenberger, B., & Panholzer, A. (2008). The degree distribution of thickened trees. DMTCS Proceedings, Fifth Colloquium on Mathematics and Computer Science, vol. AI, pp. 149–162.CrossRefGoogle Scholar
Fredkin, E. (1960). Trie memory. Communications of the ACM 3: 490499.CrossRefGoogle Scholar
Gopaladesikan, M., Mahmoud, H., & Ward, M. (2014). Building random trees from blocks. Probability in the Engineering and Informational Sciences 28: (1) 6781.CrossRefGoogle Scholar
Knuth, D. (1998). The art of computer programming, vol. 3: Sorting and searching, 2nd ed. Reading, MA: Addison-Wesley.Google Scholar
Mahmoud, H (1992). Evolution of random search trees. New York: John Wiley & Sons.Google Scholar
Mahmoud, H. (2019). Local and global degree profiles of randomly grown self-similar hooking networks under uniform and preferential attachment. Advances in Applied Mathematics 111: 101930.CrossRefGoogle Scholar
Mahmoud, H. (2019). A spectrum of series-parallel graphs with multiple edge evolution. Probability in the Engineering and Informational Sciences 33(4): 487499.CrossRefGoogle Scholar
Resnick, S. & Samorodnitsky, G. (2016). Asymptotic normality of degree counts in a preferential attachment model. Advances in Applied Probability 48(A): 283299.CrossRefGoogle Scholar
Smythe, R. & Mahmoud, H. (1995). A survey of recursive trees. Theory of Probability and Mathematical Statistics 51: 127.Google Scholar
Szpankowski, W (2001). Average case analysis of algorithms on sequences. New York: Wiley-Interscience.CrossRefGoogle Scholar
van der Hofstad, R (2016). Random graphs and complex networks, vol. 1. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Figure 0

Figure 1. A seed (top) with a hook and three networks grown from it (second row) under the building sequence $k_n=n+1$. The white vertices in the network $G_1$, $G_2$, and $G_3$ represent the reference vertex.