Hostname: page-component-745bb68f8f-5r2nc Total loading time: 0 Render date: 2025-02-11T04:42:02.338Z Has data issue: false hasContentIssue false

A restaurant process with cocktail bar and relations to the three-parameter Mittag–Leffler distribution

Published online by Cambridge University Press:  22 November 2021

Martin Möhle*
Affiliation:
Eberhard Karls Universität Tübingen
*
*Postal address: Mathematisches Institut, Eberhard Karls Universität Tübingen, Auf der Morgenstelle 10, 72076 Tübingen, Germany. Email address: martin.moehle@uni-tuebingen.de
Rights & Permissions [Opens in a new window]

Abstract

In addition to the features of the two-parameter Chinese restaurant process (CRP), the restaurant under consideration has a cocktail bar and hence allows for a wider range of (bar and table) occupancy mechanisms. The model depends on three real parameters, $\alpha$ , $\theta_1$ , and $\theta_2$ , fulfilling certain conditions. Results known for the two-parameter CRP are carried over to this model. We study the number of customers at the cocktail bar, the number of customers at each table, and the number of occupied tables after n customers have entered the restaurant. For $\alpha>0$ the number of occupied tables, properly scaled, is asymptotically three-parameter Mittag–Leffler distributed as n tends to infinity. We provide representations for the two- and three-parameter Mittag–Leffler distribution leading to efficient random number generators for these distributions. The proofs draw heavily from methods known for exchangeable random partitions, martingale methods known for generalized Pólya urns, and results known for the two-parameter CRP.

Type
Original Article
Copyright
© The Author(s), 2021. Published by Cambridge University Press on behalf of Applied Probability Trust

1. Introduction

The Chinese restaurant process (CRP) is a discrete-time stochastic process which generates exchangeable random partitions. This process is first mentioned in David Aldous’s lecture notes [Reference Aldous1, p. 92] on exchangeability and goes back to ideas of Jim Pitman and Lester Dubins. The one-parameter CRP can be generated [Reference James16,Reference Lo, Brunner and Chan23] from a Dirichlet process, being strongly related to the Blackwell and MacQueen urn [Reference Blackwell and MacQueen4]. Similarly, the two-parameter CRP can be constructed from the two-parameter Poisson–Dirichlet process, also called the Pitman–Yor process [Reference Pitman and Yor31]. Constructions of this form have been generalized using normalized completely random measures [Reference James16,Reference James17] and the Poisson–Kingman process (see also [Reference Pitman, Speed and Goldstein29]). Pitman [Reference Pitman27,Reference Pitman28,Reference Pitman30] further developed the related theory on (partially) exchangeable random partitions. The CRP and the associated random measures have applications in several disciplines, for example in Bayesian statistics [Reference Green, Bingham and Goldie11,Reference James16,Reference James17] and in mathematical population genetics. We mention here Kingman’s work [Reference Kingman22] on random partitions in population genetics, and that the Ewens sampling formula [Reference Ewens8] arises naturally from the CRP or from the Dirichlet process (see [Reference Antoniak2]).

We are interested in the following extension of the CRP. Fix three real parameters $\alpha$ , $\theta_1$ , and $\theta_2$ . The precise range of these parameters will be provided shortly. Imagine a restaurant having an infinite number of tables with labels $i\in\mathbb{N}\,:\!=\,\{1,2,\ldots\}$ and a cocktail bar. All tables and the bar are assumed to have infinite capacity. At the beginning (opening) the restaurant is empty. If at stage $n\in\mathbb{N}_0\,:\!=\,\{0,1,\ldots\}$ there are $n_0\in\mathbb{N}_0$ customers at the bar and $k\in\mathbb{N}_0$ occupied tables with $n_i\in\mathbb{N}$ customers at the ith table, $i\in\{1,\ldots,k\}$ , the next customer is seated

  • at the bar with probability $(n_0+\theta_2-\theta_1)/(n+\theta_2)$ ,

  • at occupied table $i\in\{1,\ldots,k\}$ with probability $(n_i-\alpha)/(n+\theta_2)$ , and

  • at the new table $k+1$ with probability $(\alpha k+\theta_1)/(n+\theta_2)$ ,

where $n\,:\!=\,\sum_{i=0}^kn_i$ . In particular, customer 1 is seated at table 1 with probability $\theta_1/\theta_2$ and at the bar with probability $1-\theta_1/\theta_2$ . In order to fulfill the axioms of probability it is assumed that either $0\le\alpha\le 1$ and $0<\theta_1\le\theta_2<\infty$ or that $\alpha=-\kappa<0$ for some $\kappa>0$ , $\theta_1=m\kappa$ for some $m\in\mathbb{N}$ , and $\theta_1\le \theta_2<\infty$ . The case $\theta_1=0$ is excluded, since in this case all customers would sit at the bar. For $\theta_1=\theta_2$ the bar can be neglected (stays empty) and the model coincides with Pitman’s [Reference Pitman30] two-parameter $(\alpha,\theta)$ -Chinese restaurant model with $\theta\,:\!=\,\theta_1>0$ . Green [Reference Green, Bingham and Goldie11] studied mixed Dirichlet processes in which not all clusters have equal standing. For example (see [Reference Green, Bingham and Goldie11, Section 5.3]), we may think about a distinguished ‘background’ cluster behaving differently from all other clusters. The cocktail bar can be viewed as such a background cluster. For details on the random probability measure associated to the restaurant process with cocktail bar we refer the reader to the remark at the end of Section 5.

After $n\in\mathbb{N}_0$ customers have been seated, typical quantities of interest are the number $N_{n,0}$ of customers seated at the bar, the number $N_{n,i}$ of customers seated at table $i\in\mathbb{N}$ , and the number $K_n$ of occupied (nonempty) tables. Note that $N_{n,i}=0$ for $i>K_n$ and $\sum_{i=0}^{K_n} N_{n,i}=n$ . The process $(N_n)_{n\in\mathbb{N}_0}$ , defined via $N_n\,:\!=\,(N_{n,0},\ldots,N_{n,K_n})$ for all $n\in\mathbb{N}_0$ , is a Markov chain with $N_0=(0)$ and transition probabilities $\mathbb{P}(N_{n+1}=\textbf{n}+\textbf{e}_0 \mid N_n=\textbf{n})=(n_0+\theta_2-\theta_1)/(n+\theta_2)$ , $\mathbb{P}(N_{n+1}=\textbf{n}+\textbf{e}_i \mid N_n=\textbf{n}) = (n_i-\alpha)/(n+\theta_2)$ , $i\in\{1,\ldots,k\}$ , and $\mathbb{P}(N_{n+1}=(n_0,\ldots,n_k,1) \mid N_n=\textbf{n}) = (\alpha k+\theta_1)/(n+\theta_2)$ for all $\textbf{n}=(n_0,\ldots,n_k)$ with $k,n_0\in\mathbb{N}_0$ , $n_1,\ldots,n_k\in\mathbb{N}$ , and $n_0+\cdots+n_k=n$ , where $\textbf{e}_i$ denotes the ith unit vector in $\mathbb{R}^{\{0,\ldots,k\}}$ , $i\in\{0,\ldots,k\}$ .

The article is organized as follows. In Section 2 the model is reinterpreted in terms of a mixture of a two-type Pólya urn and a two-parameter CRP. Section 3 focuses on the number $N_{n,0}$ of customers at the bar at stage n. In Section 4, in analogy to the construction provided by Pitman [Reference Pitman30], a general restaurant process with cocktail bar and random seating plan is introduced; this is useful to derive results on the joint distribution and asymptotics of $N_{n,i}$ , $i\in\mathbb{N}_0$ , which is done in Section 5. The number $K_n$ of occupied tables is studied in Section 6. A formula in terms of generalized Stirling numbers for the distribution of $K_n$ is provided in (14). The simple case $\alpha=0$ is studied briefly in Section 6.1. The main part (Section 6.2) concentrates on the case $\alpha\ne 0$ . For $\alpha\ne 0$ , Corollary 1 provides an exact formula for the moments of $K_n$ . Theorem 3 states that, for $\alpha>0$ , $K_n/n^\alpha$ is asymptotically three-parameter Mittag–Leffler distributed. The restaurant process with cocktail bar can hence be viewed as a basic model where the three-parameter Mittag–Leffler distribution arises naturally in the limit. The work finishes with an independent part (Section 7) on the three-parameter Mittag–Leffler distribution and a more applied part (Section 8) containing new distributional representations for the two- (see Proposition 4) and three-parameter Mittag–Leffler distribution (see Corollary 3) leading to accurate and efficient random number generators for these distributions.

The proofs draw heavily from methods known for random partitions [Reference Pitman27,Reference Pitman28,Reference Pitman30], from martingale methods and results known for Pólya urns and related urn models (see, for example, [Reference Janson18,Reference Janson19]), and from (asymptotic) results known for the two-parameter CRP [Reference Pitman30].

The restaurant process with cocktail bar turns out to be a mathematically well tractable extension of the two-parameter CRP having strong relations to martingales, Pólya urns, and random partitions. Moreover, in the analytic community there is interest (see, for example, [Reference Tomovski, Pogány and Srivastava40]) on (extensions of) Prabhakar’s [Reference Prabhakar32] three-parameter Mittag–Leffler function, which is (up to a factor) the Laplace transform of the three-parameter Mittag–Leffler distribution being the limiting distribution of the number of occupied tables in the restaurant model with cocktail bar provided that $\alpha>0$ (see Theorem 3). The restaurant process with cocktail bar therefore sheds some new light on the three-parameter Mittag–Leffler distribution and therefore bridges analytic and probabilistic aspects around Mittag–Leffler functions. The model may also find its way into applications, for example in the context of extensions of sampling formulas in mathematical population genetics. We refer the reader to the remarks at the end of Section 5 for some more details pointing in this direction.

Throughout the article, $\beta(a,b)$ denotes the beta distribution with parameters $a,b>0$ having density $x\mapsto ({\rm B}(a,b))^{-1}x^{a-1}(1-x)^{b-1}$ , $x\in (0,1)$ , with respect to Lebesgue measure on (0,1), where ${\rm B}(.,.)$ denotes the beta function. We furthermore use for $a,b>0$ the convention $\beta(0,b)\,:\!=\,\delta_0$ (Dirac measure at 0) and $\beta(a,0)\,:\!=\,\delta_1$ (Dirac measure at 1).

2. Reinterpretation as an urn model

The restaurant with cocktail bar can be reinterpreted as a mixture of a two-type Pólya urn with initial weights $\theta_2-\theta_1$ and $\theta_1$ and a standard CRP with parameters $\alpha$ and $\theta_1$ as follows. Consider an urn with balls of two main colors, white and red. Assume furthermore that there is an entire (countable infinite) graduation of red colors, the red balls being split into balls of red subcolors. Let $(W_n,R_{n,1},\ldots,R_{n,K_n})$ be the number of balls of the different colors present after n steps, the total number of red balls being $R_n=\sum_{i=1}^{K_n}R_{n,i}$ . Let $W_0\,:\!=\,K_0\,:\!=\,0$ . Then, recursively, given $(W_n,R_{n,1},\ldots,R_{n,K_n})$ at step n, one adds a white ball with probability $(W_n+\theta_2-\theta_1)/(n+\theta_2)$ or a ball of red subcolor i with probability $(R_{n,i}-\alpha)/(n+\theta_2)$ , $i\in\{1,\ldots,K_n\}$ , or a ball with new red subcolor $K_n+1$ with probability $(\alpha K_n+\theta_1)/(n+\theta_2)$ . This setting corresponds to the restaurant model with cocktail bar with $N_{n,0}=W_n$ and $N_{n,i}=R_{n,i}$ . Now observe that $(W_n+\theta_2-\theta_1,R_n+\theta_1)_{n\in\mathbb{N}_0}$ is a two-type Pólya urn with initial weights $\theta_2-\theta_1$ and $\theta_1$ . Indeed, conditional that there are $W_n=k$ white balls (and hence $R_n=n-k$ red balls) in the urn, a white ball is drawn from the urn with probability

\begin{equation*} \frac{W_n+\theta_2-\theta_1}{(W_n+\theta_2-\theta_1)+(R_n+\theta_1)} = \frac{k+\theta_2-\theta_1}{n+\theta_2}\end{equation*}

and returned to the urn together with an additional white ball. Moreover, the jump evolution of the process $(R_{n,1},\ldots,R_{n,K_n})$ is a standard CRP (see [Reference Pitman30]) with parameters $\alpha$ and $\theta_1$ independent of the Pólya urn. For $${\theta _1} = {\theta _2} = :\theta $$ and $\alpha=0$ the model coincides with the Pólya-like urns studied in [Reference Donnelly7,Reference Hoppe12,Reference Hoppe13,Reference Trieb41], being related to the Ewens sampling formula [Reference Ewens8].

3. Number of customers at the cocktail bar

One of the simplest functionals is the number $N_{n,0}$ of customers seated at the cocktail bar after $n\in\mathbb{N}_0$ customers have entered the restaurant. Clearly, $N_{n,0}$ coincides with the number $W_n$ of white balls of the Pólya urn $(W_n+\theta_2-\theta_1,R_n+\theta_1)_{n\in\mathbb{N}_0}$ introduced in Section 2. The following results on $N_{n,0}=W_n$ are therefore known from standard theory on Pólya urns and we hence keep this section relatively short. From the model it follows that $(W_n)_{n\in\mathbb{N}_0}$ is a Markov chain with state space $\mathbb{N}_0$ , $W_0=0$ , and transition probabilities

(1) \begin{equation} \mathbb{P}(W_{n+1}=k+1 \mid W_n=k) = \frac{k+\theta_2-\theta_1}{n+\theta_2} = 1 - \mathbb{P}(W_{n+1}= k \mid W_n=k), \quad n\in\mathbb{N}_0.\end{equation}

The distribution of $W_n$ depends only on $\theta_1$ and $\theta_2$ but not on $\alpha$ . It is easily checked by induction on $n\in\mathbb{N}_0$ that $W_n$ has distribution

(2) \begin{equation} \mathbb{P}(W_n=k) = \binom{n}{k} \frac{[\theta_2-\theta_1]_k[\theta_1]_{n-k}}{[\theta_2]_n} = \binom{n}{k}\mathbb{E}(P_0^k(1-P_0)^{n-k}), \quad k\in\{0,\ldots,n\},\end{equation}

where $[x]_0\,:\!=\,1$ , $[x]_k\,:\!=\,x(x+1)\cdots(x+k-1)$ for $x\in\mathbb{R}$ and $k\in\mathbb{N}$ , and $P_0\stackrel{\textrm{d}}{=}\beta(\theta_2-\theta_1,\theta_1)$ . The occurrence of $P_0$ is well known from Pólya urns and will also become clearer in Sections 4 and 5. From (2) it follows that, given $P_0$ , the distribution of $W_n$ is binomial with parameters n and $P_0$ . In particular, $W_n$ has descending factorial moments

(3) \begin{equation} \mathbb{E}((W_n)_j) = (n)_j\mathbb{E}(P_0^j),\qquad j\in\mathbb{N}_0,\end{equation}

where $(x)_0\,:\!=\,1$ and $(x)_j\,:\!=\,x(x-1)\cdots(x-j+1)$ for $x\in\mathbb{R}$ and $j\in\mathbb{N}$ . For $\theta_1=\theta_2$ the bar stays empty ( $W_n=0$ ). We therefore focus without loss of generality on the case $0<\theta_1<\theta_2$ . Conditional on $P_0$ , each ball is independently white with probability $P_0$ . Thus, if $0<\theta_1<\theta_2$ and, hence, $P_0>0$ almost surely, it follows that $W_n\to\infty$ almost surely as $n\to\infty$ .

Theorem 1. (Convergence of scaled $W_n$ .) Assume that $0<{\theta}_{1}\le\theta_{2}$ . Then, as $n\to\infty$ , $W_n/n\to P_0$ almost surely and in $L^p$ for any $p>0$ , where $P_0 \stackrel{\textrm{d}}{=}\beta(\theta_2-\theta_1,\theta_1)$ . In particular, the convergence $\lim_{n\to\infty}\mathbb{E}((W_n/n)^p)=\mathbb{E}(P_0^p)$ , $p\ge 0$ , of all moments holds.

Remark 1. For $\theta_1=\theta_2$ the bar stays empty and the result holds with $P_0\stackrel{\textrm{d}}{=}\delta_0=\beta(0,\theta_1)$ .

Proof. For $\theta_1=\theta_2$ the bar stays empty and the result holds with $P_0\stackrel{\textrm{d}}{=}\delta_0=\beta(0,\theta_1)$ . Assume now that $\theta_1<\theta_2$ . The proof is based on a martingale method known from Pólya urns. For $n\in\mathbb{N}_0$ define $M_n\,:\!=\,(W_n+\theta_2-\theta_1)/(n+\theta_2)$ . Using (1) it is easily checked that $\mathbb{E}(M_{n+1} \mid W_n)=M_n$ almost surely, $n\in\mathbb{N}_0$ . Thus, $(M_n)_{n\in\mathbb{N}_0}$ is a nonnegative martingale with respect to the filtration $(\mathcal{F}_n)_{n\in\mathbb{N}_0}$ , where $\mathcal{ F}_n$ denotes the $\sigma$ -algebra generated by $W_0,\ldots,W_n$ . By martingale convergence there exists $M_\infty$ such that $M_n\to M_\infty$ almost surely as $n\to\infty$ . Since $W_n\to\infty$ almost surely as $n\to\infty$ , it follows that $M_n\sim W_n/n$ and, hence, $W_n/n\to M_\infty$ almost surely. The distribution of $M_\infty$ is obtained as follows. Let $P_0\stackrel{\textrm{d}}{=}\beta(\theta_2-\theta_1,\theta_1)$ . For all $j\in\mathbb{N}_0$ , by (3), $\mathbb{E}((W_n)_j)=(n)_j\mathbb{E}(P_0^j)\sim n^j\mathbb{E}(P_0^j)$ , $n\to\infty$ . Thus, the convergence $\lim_{n\to\infty}\mathbb{E}((W_n/n)^j)=\mathbb{E}(P_0^j)$ , $j\in\mathbb{N}_0$ , of all moments holds. Since $\beta(\theta_2-\theta_1,\theta_1)$ is uniquely determined by its moments, this convergence of moments implies (see, for example, [Reference Billingsley3, Theorem 30.2]) the convergence $W_n/n\to P_0$ in distribution as $n\to\infty$ . In particular, the distribution of the almost sure limit $M_\infty$ must coincide with the distribution $\beta(\theta_2-\theta_1,\theta_1)$ of the distributional limit $P_0$ . The almost sure convergence implies the convergence in probability. Convergence in probability together with the convergence of all moments implies (see, for example, [Reference Kallenberg20, p. 68, Proposition 4.12]) the convergence in $L^p$ and, hence, the convergence of the pth moments for any $p>0$ .

Remark 2. Under the assumption $0<\theta_1<\theta_2<\infty$ the first passage times $\tau_k\,:\!=\,\inf\{n\in\mathbb{N}_0:W_n=k\}$ , $k\in\mathbb{N}_0$ , satisfy $\tau_k/k\to\tau\,:\!=\,1/P_0$ almost surely as $k\to\infty$ , where $\tau$ has an inverted beta distribution with density $x\mapsto ({\rm B}(\theta_2-\theta_1,\theta_1))^{-1}x^{-\theta_2}(x-1)^{\theta_1-1}$ , $x>1$ .

Remark 3. The number $\sum_{i=1}^{K_n}N_{n,i}$ of customers seated at all tables coincides with the number $R_n$ of red balls of the Pólya urn introduced in Section 2. Under the assumption $0<\theta_1\le\theta_2$ it follows from $W_n+R_n=n$ and Theorem 1 that, as $n\to\infty$ , $R_n/n\to R$ almost surely and in $L^p$ for any $p>0$ , where $R\,:\!=\,1-P_0\stackrel{\textrm{d}}{=}\beta(\theta_1,\theta_2-\theta_1)$ . In particular, the convergence $\lim_{n\to\infty}\mathbb{E}((R_n/n)^p)=\mathbb{E}(R^p)$ , $p\ge 0$ , of all moments holds.

In order to study the number of customers $N_{n,i}$ at table $i\in\mathbb{N}$ after n customers have entered the restaurant it will be helpful to introduce the following more general restaurant process.

4. Random seating plan

We slightly extend Pitman’s [Reference Pitman30, p. 59] random seating plan model as follows. Let $P_0, P_1,\ldots$ be random variables with $P_i\ge 0$ and $\sum_{i=0}^\infty P_i\le 1$ . Given $P_0,P_1,\ldots$ , if at stage $n\in\mathbb{N}_0$ there are k occupied tables the next customer is seated

  • at the bar with probability $P_0$ ,

  • at occupied table $i\in\{1,\ldots,k\}$ with probability $P_i$ , and

  • at the new table $k+1$ with probability $1-\sum_{i=0}^k P_i$ .

In particular, the first customer is seated at the cocktail bar with probability $\mathbb{E}(P_0)$ and at table 1 with probability $1-\mathbb{E}(P_0)$ . For $P_0=0$ the bar stays empty, hence $K_1=1$ and the model coincides with that of [Reference Pitman30, p. 59]. Let $i\in\mathbb{N}_0$ . The law of large numbers ensures that $N_{n,i}/n\to P_i$ almost surely as $n\to\infty$ . The latter convergence also holds in $L^p$ for any $p>0$ , since $0\le N_{n,i}/n\le 1$ for all $n\in\mathbb{N}_0$ . The dominated convergence theorem yields

\begin{equation*} \lim_{n\to\infty}\mathbb{E}((N_{n,0}/n)^{q_0}\cdots(N_{n,j}/n)^{q_j}) = \mathbb{E}(P_0^{q_0}\cdots P_j^{q_j}),\qquad j\in\mathbb{N}_0,q_0,\ldots,q_j\in [0,\infty). \end{equation*}

By conditioning on $(P_0,P_1,\ldots)$ it follows that $N_n\,:\!=\,(N_{n,0},\ldots,N_{n,K_n})$ has distribution

(4) \begin{equation} \mathbb{P}(N_n=\textbf{n}) = C(\textbf{n})\,p(\textbf{n}), \end{equation}

$\textbf{n}=(n_0,\ldots,n_k)$ with $k,n_0\in\mathbb{N}_0$ , $n_1,\ldots,n_k\in\mathbb{N}$ , and $n_0+\cdots+n_k=n$ , where

(5) \begin{equation} p(\textbf{n}) \,:\!=\, \mathbb{E}\bigg( P_0^{n_0}\prod_{i=1}^k P_i^{n_i-1} \prod_{i=0}^{k-1}\bigg(1-\sum_{j=0}^i P_j\bigg)\bigg) \end{equation}

and the combinatorial factor

(6) \begin{equation} C(\textbf{n}) \,:\!=\, \frac{n!}{n_0!\prod_{i=1}^k((n_i+\cdots+n_k)(n_i-1)!)} \end{equation}

counts the number of permutations of the n-tuple

\begin{equation*} (\underbrace{0,\ldots,0}_{n_0\rm\;times},\underbrace{1,\ldots,1}_{n_1\rm\;times},\ldots, \underbrace{k,\ldots,k}_{n_k\rm\;times}) \end{equation*}

such that left to the first 1 there are only 0s, left to the first 2 there are only 0s and 1s, $\ldots$ , left to the first k there are only 0s, 1s, $\ldots$ , $(k-1)$ s. Let $S_k$ denote the set of all permutations of $\{1,\ldots,k\}$ . From the identity $\sum_{\sigma\in S_k}\prod_{i=1}^k 1/(n_{\sigma i}+\cdots+n_{\sigma k})=1/(n_1\cdots n_k)$ it follows that

(7) \begin{equation} \sum_{\sigma\in S_k} C(n_0,n_{\sigma 1},\ldots,n_{\sigma k}) = \frac{n!}{n_0!\cdots n_k!}. \end{equation}

For $k=0$ , (4) reduces to $\mathbb{P}((N_{n,0},\ldots,N_{n,K_n})=(n))=\mathbb{P}(N_{n,0}=n,K_n=0)=\mathbb{E}(P_0^n)$ , which is the probability that the first n customers all sit at the bar. For $P_0=0$ (and hence $n_0=0$ ) the three formulas (4), (5), and (6) are in agreement with the corresponding formulas (see [Reference Pitman27, Eqs. (21), (7), and (20)]) for partially exchangeable partitions. An adaptation of [Reference Pitman27, Corollary 17] yields the following. Given $P_0$ , the number $N_{n,0}$ of customers at the bar has a binomial distribution with parameters n and $P_0$ . Thus, $\mathbb{P}(N_{n,0}=k)=\binom{n}{k}\mathbb{E}(P_0^k(1-P_0)^{n-k})$ , $k\in\{0,\ldots,n\}$ . For $i\in\mathbb{N}_0$ , given $P_0,\ldots,P_{i+1}$ and $N_{n,0},\ldots,N_{n,i}$ with $N_{n,0}+\cdots+N_{n,i}<n$ , the random variable $N_{n,i+1}-1$ has a binomial distribution with parameters $n-\sum_{j=0}^i{N_{n,j}}-1$ and $W_{i+1}\,:\!=\,P_{i+1}/(1-\sum_{j=0}^i P_j)$ .

The model generates a random partition $\Pi$ of $\mathbb{N}$ by defining $i,j\in\mathbb{N}$ to belong to the same block of $\Pi$ if and only if customers i and j are both seated at the bar or at the same table. Let $C_0$ be the set of customers at the bar and $C_i$ the set of customers at table $i\in\mathbb{N}$ . In general, $\Pi=\{C_i:i\in\mathbb{N}_0,C_i\ne\emptyset\}$ is not exchangeable, but given $C_0$ , the partition corresponding to the blocks $C_i$ , $i\in\mathbb{N}$ with $C_i\ne\emptyset$ , is a partially exchangeable partition of $\mathbb{N}\setminus C_0$ in the sense of [Reference Pitman27].

5. Joint convergence

We now come back to the model introduced in Section 1 and study the asymptotics of $N_{n,i}$ , $i\in\mathbb{N}_0$ , as $n\to\infty$ . Our main result is the following.

Theorem 2. Assume that $0\le\alpha\le 1$ and $0<\theta_1\le\theta_2<\infty$ . Then, as $n\to\infty$ , $(N_{n,i}/n)_{i\in\mathbb{N}_0}\to (P_i)_{i\in\mathbb{N}_0}$ almost surely, where $(P_i)_{i\in\mathbb{N}_0}$ is a stick-breaking sequence satisfying $(P_i)_{i\in\mathbb{N}_0}\stackrel{\textrm{d}}{=}((B_i)\prod_{j=0}^{i-1}(1-B_j))_{i\in\mathbb{N}_0}$ with independent random variables $B_0,B_1,\ldots$ having beta distributions $B_0\stackrel{\textrm{d}}{=}\beta(\theta_2-\theta_1,\theta_1)$ and $B_i\stackrel{\textrm{d}}{=}\beta(1-\alpha,\alpha i+\theta_1)$ , $i\in\mathbb{N}$ . Moreover, the convergence

(8) \begin{equation} \lim_{n\to\infty} \mathbb{E}((N_{n,0}/n)^{q_0}\cdots(N_{n,j}/n)^{q_j}) = \mathbb{E}(P_0^{q_0}\cdots P_j^{q_j}), \qquad j\in\mathbb{N}_0, \quad q_0,\ldots,q_j\in [0,\infty),\end{equation}

of all finite joint moments holds.

Remark 4. For $\alpha=1$ each customer is seated either at the bar or at a new table. Thus, $N_{n,i}=1$ , $i\in\{1,\ldots,K_n\}$ , and the theorem holds with $P_0\stackrel{\textrm{d}}{=}\beta(\theta_2-\theta_1,\theta_1)$ and $P_i=0$ for all $i\in\mathbb{N}$ .

Two proofs of Theorem 2 are provided. The first is self-contained and based on the random seating plan model introduced in Section 4. The second exploits the reinterpretation in terms of a mixture of a Pólya urn and a two-parameter CRP described in Section 2.

Proof 1 of Theorem 2. Let $B_0,B_1,\ldots$ be independent random variables with $B_0\stackrel{\textrm{d}}{=}\beta(\theta_2-\theta_1,\theta_1)$ and $B_i\stackrel{\textrm{d}}{=}\beta(1-\alpha,\alpha i+\theta_1)$ , $i\in\mathbb{N}$ . Note that $B_0=0$ if $\theta_1=\theta_2$ and that $B_i=0$ for all $i\in\mathbb{N}$ if $\alpha=1$ . For $i\in\mathbb{N}_0$ define $\overline{B}_i\,:\!=\,1-B_i$ and $P_i\,:\!=\,\overline{B}_0\cdots\overline{B}_{i-1}B_i$ . Clearly, $1-\sum_{j=0}^iP_j=\overline{B}_0\cdots\overline{B}_i$ , $i\in\mathbb{N}_0$ . For $\textbf{n}\,:\!=\,(n_0,\ldots,n_k)$ with $k,n_0\in\mathbb{N}_0$ and $n_1,\ldots,n_k\in\mathbb{N}$ ,

\begin{align*} p(\textbf{n}) & \,:\!=\, \mathbb{E}\bigg(P_0^{n_0}\prod_{i=1}^k P_i^{n_i-1}\prod_{i=0}^{k-1}\bigg(1-\sum_{j=0}^i P_j\bigg)\bigg)\\[3pt] & = \mathbb{E}\big(B_0^{n_0}\overline{B}_0^{n_1+\cdots+n_k}\big)\prod_{i=1}^k \mathbb{E}\big(B_i^{n_i-1}\overline{B_i}^{n_{i+1}+\cdots+n_k}\big)\\[3pt] & = \frac{[\theta_2-\theta_1]_{n_0}[\theta_1]_{n_1+\cdots+n_k}}{[\theta_2]_n} \prod_{i=1}^k \frac{[1-\alpha]_{n_i-1}[\alpha i+\theta_1]_{n_{i+1}+\cdots+n_k}} {[1+\alpha(i-1)+\theta_1]_{n_i+\cdots+n_k-1}}, \end{align*}

since $X\stackrel{\textrm{d}}{=}\beta(a,b)$ satisfies $\mathbb{E}(X^n(1-X)^m)=[a]_n[b]_m/[a+b]_{n+m}$ for all $n,m\in\mathbb{N}_0$ . Define $[x|y]_0\,:\!=\,1$ and $[x|y]_k\,:\!=\,\prod_{j=0}^{k-1}(x+jy)$ for $x,y\in\mathbb{R}$ and $k\in\mathbb{N}$ . Since $[1+\alpha(i-1)+\theta_1]_{n_i+\cdots+n_k-1}=[\alpha(i-1)+\theta_1]_{n_i+\cdots+n_k}/(\alpha(i-1)+\theta_1)$ , the formula above reduces to

(9) \begin{align} p(\textbf{n}) & = \frac{[\theta_2-\theta_1]_{n_0}[\theta_1]_{n_1+\cdots+n_k}} {[\theta_2]_n} \prod_{i=1}^k (\alpha(i-1)+\theta_1) \frac{[1-\alpha]_{n_i-1}[\alpha i+\theta_1]_{n_{i+1}+\cdots+n_k}} {[\alpha(i-1)+\theta_1]_{n_i+\cdots+n_k}}\nonumber\\[3pt] & = \frac{[\theta_2-\theta_1]_{n_0}}{[\theta_2]_n} [\theta_1 \mid \alpha]_k \prod_{i=1}^k [1-\alpha]_{n_i-1}. \end{align}

Note that (9) is symmetric in $n_1,\ldots,n_k$ (excluding $n_0$ ). From (9) it follows (by distinguishing the cases $i=0$ and $i\in\{1,\ldots,k\}$ ) that $N_n\,:\!=\,(N_{n,0},\ldots,N_{n,K_n})$ satisfies $\mathbb{P}(N_{n+1}=\textbf{n}+\textbf{e}_i \mid N_n=\textbf{n}) = p(\textbf{n}+\textbf{e}_i)/p(\textbf{n})$ for $i\in\{0,\ldots,k\}$ and $\textbf{n}=(n_0,\ldots,n_k)$ with $k,n_0\in\mathbb{N}_0$ , $n_1,\ldots,n_k\in\mathbb{N}$ , and $n\,:\!=\,n_0+\cdots+n_k$ . We are therefore in the situation of the model with random seating plan studied in Section 4 corresponding to the sequence $(P_0,P_1,\ldots)$ . From the results in Section 4 it follows that, for every $i\in\mathbb{N}_0$ , as $n\to\infty$ , $N_{n,i}/n\to P_i$ almost surely and in $L^p$ for any $p>0$ . Moreover, the convergence (8) of all finite joint moments holds.

Proof 2 of Theorem 2. Recall the reinterpretation of the model as a mixture of a Pólya urn and a two-parameter CRP (see Section 2). The number $N_{n,0}$ of customers at the bar coincides with the number $W_n$ of white balls in the urn. From standard theory on Pólya urns (see also Section 3) it therefore follows that, as $n\to\infty$ , $N_{n,0}/n=W_n/n\to B_0$ almost surely, where $B_0\stackrel{\textrm{d}}{=}\beta(\theta_2-\theta_1,\theta_1)$ . Let $i\in\mathbb{N}$ . The number $N_{n,i}$ of customers at table i coincides with the number $R_{n,i}$ of balls of red subcolor i. Therefore,

\begin{equation*} \frac{N_{n,i}}{n} = \frac{R_{n,i}}{n} = \frac{R_n}{n}\frac{R_{n,i}}{R_n}.\end{equation*}

As $n\to\infty$ , $R_n/n=1-W_n/n\to 1-B_0$ almost surely. Since the jump evolution of the process $(R_{n,1},\ldots,R_{n,K_n})$ is a two-parameter CRP with parameters $\alpha$ and $\theta_1$ independent of the Pólya urn, it follows from [Reference Pitman30, Theorem 3.2] (see also [Reference Pitman27]) that, as $n\to\infty$ , $R_{n,i}/R_n\to B_i\prod_{j=1}^{i-1}(1-B_j)$ almost surely, where $B_1,B_2,\ldots$ are independent (and independent of $B_0$ ) with $B_j\stackrel{\textrm{d}}{=}\beta(1-\alpha,\alpha j+\theta_1)$ . Thus, $N_{n,i}/n\to P_i$ almost surely as $n\to\infty$ , $i\in\mathbb{N}_0$ , with $P_i$ as given in Theorem 2. By dominated convergence the convergence (8) holds.

Remark 5. Let $i\in\mathbb{N}$ . The almost sure convergence of $N_{n,i}/n$ as $n\to\infty$ can alternatively be derived via a martingale method as follows. Define

(10) \begin{equation} M_{n,i} \,:\!=\, \frac{\max(N_{n,i},1)-\alpha}{n+\theta_2}, \qquad n\in\mathbb{N}_0. \end{equation}

For $k\in\mathbb{N}$ we have $N_{n+1,i}\ge 1$ on $\{N_{n,i}=k\}$ and, hence,

\begin{eqnarray*} & & \mathbb{E}(M_{n+1,i} \mid N_{n,i}=k) = \mathbb{E}\bigg(\frac{N_{n+1,i}-\alpha}{n+1+\theta_2} \,\Big|\, N_{n,i}=k\bigg)\\[3pt] & = & \frac{k-\alpha}{n+1+\theta_2} \bigg(1-\frac{k-\alpha}{n+\theta_2}\bigg) + \frac{k+1-\alpha}{n+1+\theta_2}\frac{k-\alpha}{n+\theta_2}\\[3pt] & = & \frac{k-\alpha}{n+1+\theta_2} \bigg(1-\frac{k-\alpha}{n+\theta_2}+\frac{k+1-\alpha}{n+\theta_2}\bigg) = \frac{k-\alpha}{n+\theta_2},\qquad n\in\mathbb{N}_0,\ k\in\mathbb{N}. \end{eqnarray*}

On $\{N_{n,i}=0\}$ the random variable $N_{n+1,i}$ takes the two values 0 and 1. From (10) we conclude that, on $\{N_{n,i}=0\}$ , the random variable $M_{n+1,i}$ is constant, equal to $(1-\alpha)/(n+1+\theta_2)$ . Thus, $\mathbb{E}(M_{n+1,i} \mid N_{n,i}=0) =(1-\alpha)/(n+1+\theta_2)\le(1-\alpha)/(n+\theta_2)$ . Clumping the two cases $k\in\mathbb{N}$ and $k=0$ together leads to $\mathbb{E}(M_{n+1,i} \mid N_{n,i}) \le (\max(N_{n,i},1)-\alpha)/(n+\theta_2)=M_{n,i}$ almost surely. Thus, $(M_{n,i})_{n\in\mathbb{N}_0}$ is a nonnegative supermartingale with respect to the filtration $(\mathcal{ F}_n)_{n\in\mathbb{N}_0}$ , where $\mathcal{ F}_n$ denotes the $\sigma$ -algebra generated by $N_{0,i},\ldots,N_{n,i}$ . By martingale convergence there exists $M_{\infty,i}$ such that $M_{n,i}\to M_{\infty,i}$ almost surely as $n\to\infty$ . For $\alpha<1$ it can be checked that $N_{n,i}\to\infty$ as $n\to\infty$ . Thus, $M_{n,i}\sim N_{n,i}/n$ and, hence, $N_{n,i}/n\to M_{\infty,i}$ almost surely as $n\to\infty$ .

However, this martingale approach alone does not seem to be useful to derive the distribution of $M_{\infty,i}$ . Proof 1 based on the random seating plan model has the advantage that the limiting beta distributions are obvious by construction. In the second proof the limiting beta distributions follow from the asymptotic results known for Pólya urns and for the two-parameter CRP.

Remark 6. (Sampling formula) Define

\begin{equation*} \tilde{p}(\textbf{n}) \,:\!=\, \frac{1}{k!}\sum_{\sigma\in S_k} \mathbb{P}(N_n=(n_0,n_{\sigma 1},\ldots,n_{\sigma k})). \end{equation*}

Since $p(\textbf{n})$ is in our case (see (9)) symmetric in $n_1,\ldots,n_k$ , it follows from (4) and (7) that

(11) \begin{equation} \tilde{p}(\textbf{n}) = \frac{n!}{k!\,n_0!\cdots n_k!}\,p(\textbf{n}) = \frac{n!}{k!n_0!\cdots n_k!}\frac{[\theta_2-\theta_1]_{n_0}}{[\theta_2]_n} [\theta_1 \mid \alpha]_k \prod_{i=1}^k [1-\alpha]_{n_i-1}, \end{equation}

where the last equality holds by (9). The sampling formula (11) for the restaurant process with cocktail bar is symmetric in $n_1,\ldots,n_k$ but not in $n_0,\ldots,n_k$ . Formulas of this form might be of interest for applications in mathematical population genetics, where a distinguished type 0 occurs in the population. Clearly, for $\theta_1=\theta_2 \!=: \theta>0$ the cocktail bar stays empty ( $n_0=0$ ) and (11) turns into the symmetric formula for the two-parameter CRP,

\begin{equation*} \tilde{p}(n_1,\ldots,n_k)\ =\ \frac{n!}{k!} \frac{[\theta|\alpha]_k}{[\theta]_n}\prod_{i=1}^k \frac{[1-\alpha]_{n_i-1}}{n_i!}, \end{equation*}

$1\le k\le n$ , $n_1,\ldots,n_k\in\mathbb{N}$ , $n_1+\cdots+n_k=n$ , which for $\alpha=0$ is in agreement with Ewens’s sampling formula (see, for example, [Reference Möhle25, Eq. (1)]), $\tilde{p}(n_1,\ldots,n_k)=n!\theta^k/(k![\theta]_n n_1\cdots n_k)$ , $1\le k\le n$ , $n_1,\ldots,n_k\in\mathbb{N}$ , $n_1+\cdots+n_k=n$ .

Remark 7. (Random probability measure, RPM) Section 5 of [Reference Green, Bingham and Goldie11] studies Dirichlet mixtures of Dirichlet processes of the form $\xi\,:\!=\,\sum_j W_j \xi_j$ , where the $\xi_j$ are independent Dirichlet processes and $(W_j)_j$ is Dirichlet distributed and independent of $(\xi_j)_j$ . Note that each $\xi_j$ is an RPM on some appropriate measure space, say $(S,\mathcal{S})$ . The construction is easily generalized by allowing $\xi_j$ to be, for example, a two-parameter Poisson–Dirichlet process or a general RPM $\xi_j:\Omega\times\mathcal{ S}\to [0,1]$ .

Our model has two colors $j\in\{0,1\}$ (white and red), and $(W_0,W_1)$ is Dirichlet distributed with parameters $\theta_2-\theta_1$ and $\theta_1$ , the initial weights of the two-type Pólya urn from Section 2. Comparing the stick-breaking construction from [Reference Green, Bingham and Goldie11, p. 16] with Theorem 2, we see that $\xi_1$ is a two-parameter Poisson–Dirichlet process $\xi_1={\rm PDP}(\alpha,\theta_1,Q_1)$ with parameters $\alpha$ and $\theta_1$ . Here, $Q_1$ denotes some given (usually diffuse) base probability measure on $(S,\mathcal{ S})$ . Moreover, $\xi_0$ has the form $\xi_0(\omega,A)=\delta_{X_0(\omega)}(A)$ , $\omega\in\Omega$ , $A\in\mathcal{ S}$ , with $X_0\stackrel{\textrm{d}}{=}Q_0$ , where again $Q_0$ is a base probability measure on $(S,\mathcal{ S})$ . Note that $\xi_0$ can be viewed as a degenerate Dirichlet process ${\rm DP}(a,Q_0)$ in the limit $a\to 0$ , corresponding to the fact that the white-colored segment of the stick does not undergo any stick-breaking procedure. As noted in [Reference James16, Eq. (6)], $\xi_1$ can be represented as $\xi_1(\omega,A)=\sum_{i\ge 1}J_i(\omega) \delta_{X_i(\omega)}(A)$ , $\omega\in\Omega$ , $A\in\mathcal{ S}$ , where $X_1,X_2,\ldots$ are independent and identically distributed with distribution $Q_1$ and $J=(J_1,J_2,\ldots)$ has the two-parameter Poisson–Dirichlet distribution ${\rm PD}(\alpha,\theta_1)$ and is independent of $(X_i)_{i\in\mathbb{N}}$ . Note that $J_1>J_2>\cdots$ and $\sum_{i\ge 1} J_i=1$ . The RPM $\xi$ associated to the restaurant process with cocktail bar thus has the form

\begin{equation*} \xi(\omega,A) = W_0(\omega)\delta_{X_0(\omega)}(A) + W_1(\omega) \sum_{i\ge 1} J_i(\omega)\delta_{X_i(\omega)}(A), \qquad \omega\in\Omega,\ A\in\mathcal{ S}. \end{equation*}

6. Number of occupied tables

In order to analyze the number of occupied tables it will turn out to be convenient to introduce, for $a,b,r\in\mathbb{R}$ , the generalized Stirling numbers $S(n,k;a,b,r)$ , $n,k\in\mathbb{N}_0$ , having vertical generating function (see [Reference Hsu and Shiue14, Theorem 2])

(12) \begin{equation} k!\sum_{n\ge 0}S(n,k;a,b,r)\frac{t^n}{n!} = \left\{ \begin{array}{cl} \displaystyle(1+at)^{r/a}\bigg(\frac{(1+at)^{b/a}-1}{b}\bigg)^k & \mbox{if $ab\ne 0$,}\\[3pt] \displaystyle \textrm{e}^{rt}\bigg(\frac{\textrm{e}^{bt}-1}{b}\bigg)^k & \mbox{if $a=0$ and $b\ne 0$,}\\[3pt] \displaystyle(1+at)^{r/a}\bigg(\frac{\log(1+at)}{a}\bigg)^k & \mbox{if $a\ne 0$ and $b=0$,}\\[3pt] \textrm{e}^{rt}t^k & \mbox{if $a=b=0$.} \end{array} \right.\end{equation}

We now turn to the number $K_n$ of occupied tables at stage $n\in\mathbb{N}_0$ . Clearly, $(K_n)_{n\in\mathbb{N}_0}$ is a Markov chain with state space $\mathbb{N}_0$ , $K_0=0$ , and transition probabilities

(13) \begin{equation} \mathbb{P}(K_{n+1}=k+1 \mid K_n=k) = \frac{\alpha k+\theta_1}{n+\theta_2} = 1-\mathbb{P}(K_{n+1}=k \mid K_n=k),\qquad n\in\mathbb{N}_0.\end{equation}

Exploiting the recursion [Reference Hsu and Shiue14, Theorem 1] for the generalized Stirling numbers it is easily checked by induction on $n\in\mathbb{N}_0$ that $K_n$ has distribution

(14) \begin{equation} \mathbb{P}(K_n=k) = \frac{[\theta_1 \mid \alpha]_k}{[\theta_2]_n}S(n,k;-1,-\alpha,\theta_2-\theta_1), \qquad k\in\mathbb{N}_0.\end{equation}

For $\alpha=-\kappa<0$ and $\theta_1=m\kappa$ it follows that $[\theta_1 \mid \alpha]_k=[m\kappa \mid-\kappa ]_k=0 $ for $k>m$ , confirming that $\mathbb{P}(K_n\le m)=1$ in this case. For $\theta_1=\theta_2 \!=: \theta$ , the random variable $K_n$ counts the number of occupied tables in Pitman’s [Reference Pitman30] two-parameter $(\alpha,\theta)$ -CRP. Although the distribution (14) of $K_n$ is known, it will turn out to be helpful to derive recursions for some functionals of $K_n$ such as the moments and the Laplace transform.

Lemma 1. (Recursion for the moments and the Laplace transform of $K_n.$ )

The moments $\mu_{n,m}\,:\!=\,\mathbb{E}(K_n^m)$ , $n,m\in\mathbb{N}_0$ , of $K_n$ satisfy the linear recursion

(15) \begin{equation} \mu_{n+1,m} = \frac{n+\theta_2+\alpha m}{n+\theta_2}\mu_{n,m} + \frac{1}{n+\theta_2}\sum_{j=0}^{m-1} \bigg(\alpha\binom{m}{j-1}+\theta_1\binom{m}{j}\bigg)\mu_{n,j}, \end{equation}

$n,m\in\mathbb{N}_0$ , with initial conditions $\mu_{0,m}\,:\!=\,\delta_{0,m}$ (Kronecker symbol), $m\in\mathbb{N}_0$ , and the usual convention $\binom{m}{j-1}\,:\!=\,0$ for $j=0$ . The Laplace transform $\psi_n$ of $K_n$ satisfies $\psi_0(\lambda)=1$ and

(16) \begin{equation} \psi_{n+1}(\lambda) = \psi_n(\lambda)-\frac{1-e^{-\lambda}}{n+\theta_2} \big(\theta_1\psi_n(\lambda)-\alpha\psi_n{\!\!'}\kern2pt(\lambda)\big), \qquad n\in\mathbb{N}_0, \lambda\ge 0. \end{equation}

Proof. Clearly, $\mu_{0,m}=\delta_{0,m}$ and $\psi_0(\lambda)=1$ , since $K_0=0$ . For $n\in\mathbb{N}_0$ and $k\in\{1,\ldots,n\}$ ,

\begin{align*} \mathbb{E}(K_{n+1}^m \mid K_n=k) & = k^m\bigg(1-\frac{\alpha k+\theta_1}{n+\theta_2}\bigg) + (k+1)^m\frac{\alpha k+\theta_1}{n+\theta_2}\\[3pt] & = k^m + \frac{\alpha k+\theta_1}{n+\theta_2}\big((k+1)^m-k^m\big)\\[3pt] & = k^m+\frac{\alpha k+\theta_1}{n+\theta_2} \sum_{j=0}^{m-1}\binom{m}{j}k^j\\[3pt] & = \bigg(1+\frac{\alpha m}{n+\theta_2}\bigg)k^m + \frac{1}{n+\theta_2}\sum_{j=0}^{m-1} \bigg(\alpha\binom{m}{j-1}+\theta_1\binom{m}{j}\bigg)k^j, \end{align*}

where $\binom{m}{j-1}\,:\!=\,0$ for $j=0$ . Multiplying by $\mathbb{P}(K_n=k)$ and summing over all k yields (15). Similarly, $\mathbb{E}(\textrm{e}^{-\lambda K_{n+1}} \mid K_n=k) = \big(1-\frac{\alpha k+\theta_1}{n+\theta_2}\big)\textrm{e}^{-\lambda k}+\frac{\alpha k+\theta_1}{n+\theta_2}\textrm{e}^{-\lambda(k+1)} = \big(1-(1-\textrm{e}^{-\lambda})\frac{\alpha k+\theta_1}{n+\theta_2}\big)\textrm{e}^{-\lambda k}$ , $\lambda\ge 0$ . Multiplication by $\mathbb{P}(K_n=k)$ and summation over all k yields (16).

Remark 8. (First passage time) For $k\in\mathbb{N}_0$ let $T_k\,:\!=\,\inf\{n\in\mathbb{N}_0:K_n=k\}$ denote the time the chain K reaches state k. In other words, $T_k$ is the first hitting time or first passage time of the state k. Clearly, $T_0=0$ , since $K_0=0$ . For $k\in\mathbb{N}$ the distribution of $T_k$ is given by

(17) \begin{align} \mathbb{P}(T_k=n) & = \mathbb{P}(K_n=k,K_{n-1}=k-1)\nonumber\\[3pt] & = \mathbb{P}(K_n=k \mid K_{n-1}=k-1)\mathbb{P}(K_{n-1}=k-1)\nonumber\\[3pt] & = \frac{\alpha(k-1)+\theta_1}{n-1+\theta_2}\mathbb{P}(K_{n-1}=k-1), \qquad n\in\{k,k+1,\ldots\}. \end{align}

From (14) we conclude that

(18) \begin{align} \mathbb{P}(T_k=n) & = \frac{\alpha(k-1)+\theta_1}{n-1+\theta_2} \frac{[\theta_1 \mid \alpha]_{k-1}}{[\theta_2]_{n-1}}S(n-1,k-1;-1,-\alpha,\theta_2-\theta_1)\nonumber\\[3pt] & = \frac{[\theta_1 \mid \alpha]_k}{[\theta_2]_n} S(n-1,k-1;-1,-\alpha,\theta_2-\theta_1),\qquad n\in\{k,k+1,\ldots\} . \end{align}

The asymptotics of $K_n$ as $n\to\infty$ and of $T_k$ as $k\to\infty$ are provided in Sections 6.1 and 6.2 for $\alpha=0$ and $\alpha\ne 0$ respectively.

6.1. The case ${\alpha=0}$

We briefly discuss the behavior of $K_n$ for $\alpha=0$ . In this case it follows, for example from (16), that $K_n\stackrel{\textrm{d}}{=}\sum_{j=0}^{n-1}B_j$ , $n\in\mathbb{N}$ , where $B_0,B_1,\ldots$ are independent Bernoulli random variables with mean $\mathbb{E}(B_j)=\theta_1/(j+\theta_2)$ , $j\in\mathbb{N}_0$ . Thus, $K_n$ has probability-generating function

\begin{equation*} f_n(s) = \prod_{j=0}^{n-1} \frac{j+\theta_2-\theta_1+\theta_1 s}{j+\theta_2} = \frac{\Gamma(\theta_2)\Gamma(n+\theta_2-\theta_1+\theta_1 s)} {\Gamma(n+\theta_2)\Gamma(\theta_2-\theta_1+\theta_1 s)}, \qquad n\in\mathbb{N}_0,s\in\mathbb{C},\end{equation*}

and integer moments

\begin{align*} \mathbb{E}(K_n^m) & = \mathbb{E}((B_0+\cdots+B_{n-1})^m)\\[3pt] & = \sum_{\genfrac{}{}{0pt}{2}{m_0,\ldots,m_{n-1}\ge 0}{m_0+\cdots+m_{n-1}=m}} \frac{m!}{m_0!\cdots m_{n-1}!} \mathbb{E}\big(B_0^{m_0}\big)\cdots\mathbb{E}\big(B_{n-1}^{m_{n-1}}\big)\\[3pt] & = m!\sum_{I\subseteq\{0,\ldots,n-1\}} \sum_{\genfrac{}{}{0pt}{2}{m_i,i\in I}{\sum_{i\in I} m_i=m}} \prod_{i\in I}\frac{\mathbb{E}(B_i)}{m_i!}\\[3pt] & = m!\sum_{I\subseteq\{0,\ldots,n-1\}} \bigg(\prod_{i\in I}\mathbb{E}(B_i)\bigg) \sum_{\genfrac{}{}{0pt}{2}{m_i,i\in I}{\sum_{i\in I}m_i=m}} \prod_{i\in I}\frac{1}{m_i!}\\[3pt] & = \sum_{I\subseteq\{0,\ldots,n-1\}}|I|!S(m,|I|) \bigg(\prod_{i\in I}\mathbb{E}(B_i)\bigg)\\[3pt] & = \sum_{j=1}^m S(m,j) \sum_{\genfrac{}{}{0pt}{2}{i_1,\ldots,i_j=0}{\rm all\;distinct}}^{n-1} \mathbb{E}(B_{i_1})\cdots\mathbb{E}(B_{i_j})\\[3pt] & = \sum_{j=1}^m S(m,j) \sum_{\genfrac{}{}{0pt}{2}{i_1,\ldots,i_j=0}{\rm all\;distinct}}^{n-1} \frac{\theta_1^j}{(i_1+\theta_2)\cdots(i_j+\theta_2)}, \qquad n,m\in\mathbb{N},\end{align*}

where $S(\!\cdot\,,\cdot\!)$ denote Stirling numbers of the second kind. Clearly,

\begin{equation*} \mathbb{E}(K_n) = \sum_{i=0}^{n-1}\frac{\theta_1}{i+\theta_2} \sim \theta_1\log n,\qquad n\to\infty,\end{equation*}

and

\begin{equation*} {\rm Var}(K_n) = \sum_{i=0}^{n-1}\frac{\theta_1(i+\theta_2-\theta_1)}{(i+\theta_2)^2} \sim \theta_1\log n,\qquad n\to\infty.\end{equation*}

By the central limit theorem, $(K_n-\theta_1\log n)/\sqrt{\theta_1\log n}$ is asymptotically standard normal distributed. This convergence holds also locally, i.e.

\begin{equation*} \lim_{n\to\infty}\sup_{k\in\mathbb{Z}}|\sqrt{\theta_1\log n}\mathbb{P}(K_n=k)-(2\pi)^{-1/2}\exp\{-(k-\theta_1\log n)^2/(2\theta_1\log n)\}| = 0,\end{equation*}

which follows from [Reference Petrov26, p.193, Theorem 3].

Let us now turn to the first passage time $T_k$ .

Proposition 1. If $\alpha=0$ then, as $k\to\infty$ , $(\theta_1\log T_k-k)/\sqrt{k}$ is asymptotically standard normal distributed.

Proof. We known that $(K_n-\theta_1\log n)/\sqrt{\theta_1\log n}$ converges in distribution to the standard normal law. Fix $x\in\mathbb{R}$ . For $k\in\mathbb{N}$ define $n\,:\!=\,n(k)\,:\!=\,\lfloor\exp((x\sqrt{k}+k)/\theta_1)\rfloor\in\mathbb{N}_0$ . For all $k\in\mathbb{N}$ we have

\begin{align*} \mathbb{P}\bigg(\frac{\theta_1\log T_k-k}{\sqrt{k}}\le x\bigg) & = \mathbb{P}(T_k\le n) = \mathbb{P}(K_n\ge k) = 1-\mathbb{P}(K_n\le k-1)\\[3pt] & = 1-\mathbb{P}\bigg( \frac{K_n-\theta_1\log n}{\sqrt{\theta_1\log n}}\le\frac{k-1-\theta_1\log n}{\sqrt{\theta_1\log n}} \bigg)\\[3pt] & \to 1-\Phi(-x)\ =\ \Phi(x) \end{align*}

as $k\to\infty$ , since $(k-1-\theta_1\log n)/\sqrt{\theta_1\log n}\to-x$ as $k\to\infty$ .

Remark 9. Useful formulas for some functionals of $T_k$ are available for $\alpha=0$ and $\theta_1=\theta_2=1$ . In this case it follows from (17) and (18) that $T_k$ has distribution

\begin{equation*} \mathbb{P}(T_k=n) = \frac{1}{n}\mathbb{P}(K_{n-1}=k-1) = \frac{|s(n-1,k-1)|}{n!},\qquad n\in\{k,k+1,\ldots\}, \end{equation*}

where $s(\cdot\,,\cdot)$ are the Stirling numbers of the first kind. It follows that $T_k$ has probability-generating function

\begin{align*} \mathbb{E}(s^{T_k}) & = \sum_{n=k}^\infty |s(n-1,k-1)|\frac{s^n}{n!}\\[3pt] & = \int_0^s \sum_{n=k}^\infty |s(n-1,k-1)|\frac{u^{n-1}}{(n-1)!}\,{\rm d}u\\[3pt] & = \int_0^s \frac{(-\log(1-u))^{k-1}}{(k-1)!}\,{\rm d}u\\[3pt] & = 1-(1-s)\sum_{j=0}^{k-1}\frac{(-\log(1-s))^j}{j!},\qquad 0\le s\le 1. \end{align*}

6.2. The case ${\alpha\ne 0}$

For $\alpha\ne 0$ the right-hand side of the recursion (16) for the Laplace transform $\psi_n$ of $K_n$ depends not only on $\psi_n$ but also on the derivative of $\psi_n$ and is hence much more involved than for $\alpha=0$ . Therefore, at first glance it seems to be hopeless to solve the recursion (16). The following proposition, however, provides a rather useful solution to this recursion on the interval $[0,\log 2)$ .

Proposition 2. (Laplace transform of $K_n$ .) If $\alpha\ne 0$ then the Laplace transform $\psi_n$ of $K_n$ satisfies

(19) \begin{equation} \psi_n(\lambda) = \textrm{e}^{\gamma\lambda}\sum_{j=0}^\infty \frac{[\theta_2+\alpha j]_n}{[\theta_2]_n} \frac{[\gamma]_j}{j!}(1-\textrm{e}^\lambda)^j, \qquad n\in\mathbb{N}_0,\ 0\le\lambda<\log 2, \end{equation}

where $\gamma\,:\!=\,\theta_1/\alpha$ .

Remark 10. For $\alpha>0$ the series on the right-hand side of (19) is absolutely convergent since $|1-\textrm{e}^\lambda|<1$ for $0\le\lambda<\log 2$ . For $\alpha=-\kappa<0$ and $\theta_1=m\kappa$ for some $m\in\mathbb{N}$ we have $\gamma\,:\!=\,\theta_1/\alpha=-m\in-\mathbb{N}$ and the series on the right-hand side in (19) reduces to a finite sum, since in this case $[\gamma]_j=0$ for $j>m$ . Clearly, knowledge of the Laplace transform $\psi_n$ on the interval $[0,\log 2)$ fully determines the distribution of $K_n$ .

Proof. It suffices to verify that $\psi_n$ , defined via (19), solves the recursion (16) with initial condition $\psi_0\equiv 1$ . Equivalently, it suffices to verify that $(f_n)_{n\in\mathbb{N}_0}$ , defined via

(20) \begin{equation} f_n(\lambda) \,:\!=\, \sum_{j=0}^\infty \frac{[\theta_2+\alpha j]_n}{[\theta_2]_n} \frac{[\gamma]_j}{j!}\big(1-\textrm{e}^\lambda\big)^j, \qquad n\in\mathbb{N}_0,\ 0\le\lambda<\log 2, \end{equation}

solves the recursion

(21) \begin{equation} f_{n+1}(\lambda) = f_n(\lambda) + \frac{\alpha}{n+\theta_2}(1-\textrm{e}^{-\lambda})\,{f\kern0.5pt}_n{\!\!'}\kern2pt(\lambda),\qquad n\in\mathbb{N}_0,\ 0\le\lambda<\log 2 \end{equation}

with initial condition $f_0(\lambda)=\textrm{e}^{-\gamma\lambda}$ , $0\le\lambda<\log 2$ . Obviously,

\begin{equation*} f_0(\lambda)=\sum_{j=0}^\infty \binom{j+\gamma-1}{j}(1-\textrm{e}^\lambda)^j=\sum_{j=0}^\infty \binom{-\gamma}{j}(\textrm{e}^\lambda-1)^j=\big(1+\big(\textrm{e}^\lambda-1\big)\big)^{-\gamma}=\textrm{e}^{-\gamma\lambda} \end{equation*}

for all $0\le\lambda<\log 2$ by binomial series expansion. Taking the derivative with respect to $\lambda$ in (20) yields

\begin{equation*} {f\kern0.5pt}_n{\!\!'}\kern2pt(\lambda) = \sum_{j=0}^\infty \frac{[\theta_2+\alpha j]_n}{[\theta_2]_n} \frac{[\gamma]_j}{j!} j\big(1-\textrm{e}^\lambda\big)^{j-1}\big(-\textrm{e}^\lambda\big). \end{equation*}

Multiplication by $1-\textrm{e}^{-\lambda}$ leads to

(22) \begin{equation} (1-\textrm{e}^{-\lambda}){f\kern0.5pt}_n{\!\!'}\kern1pt(\lambda) = \sum_{j=0}^\infty \frac{[\theta_2+\alpha j]_n}{[\theta_2]_n} \frac{[\gamma]_j}{j!}j\big(1-\textrm{e}^\lambda\big)^j. \end{equation}

Now, by (20),

\begin{align*} f_{n+1}(\lambda) & = \sum_{j=0}^\infty \frac{[\theta_2+\alpha j]_{n+1}}{[\theta_2]_{n+1}} \frac{[\gamma]_j}{j!}\big(1-\textrm{e}^\lambda\big)^j\\[3pt] & = \sum_{j=0}^\infty \frac{[\theta_2+\alpha j]_n}{[\theta_2]_n} \bigg(1+\frac{\alpha j}{\theta_2+n}\bigg) \frac{[\gamma]_j}{j!} \big(1-\textrm{e}^\lambda\big)^j\\[3pt] & = f_n(\lambda) + \frac{\alpha}{\theta_2+n} \sum_{j=0}^\infty \frac{[\theta_2+\alpha j]_n}{[\theta_2]_n} \frac{[\gamma]_j}{j!} j\big(1-\textrm{e}^\lambda\big)^j\\[3pt] & = f_n(\lambda) + \frac{\alpha}{\theta_2+n}\big(1-\textrm{e}^{-\lambda}\big)\kern1pt{f\kern0.5pt}_n{\!\!'}\kern1pt (\lambda), \end{align*}

where the last equality holds by (22). Hence, $(f_n)_{n\in\mathbb{N}_0}$ solves the recursion (21).

Corollary 1 provides a formula for $\mathbb{E}(K_n^i)$ , $i\in\mathbb{N}_0$ , showing that $\mathbb{E}(K_n^i)$ is a linear combination of terms of the form $[\theta_2+\alpha j]_n/[\theta_2]_n$ , $j\in\{0,\ldots,i\}$ , where each term occurs according to a particular weight $\kappa_{i,j}$ . Similar formulas have been obtained in [Reference Zhang, Chen and Mahmoud42, Theorem 3.1] for the moments of the number of balls of a given type for balanced triangular Pólya urns. The proof in [Reference Zhang, Chen and Mahmoud42] is elementary but based on a double induction on n and i. Slightly less precise statements (the weights $\kappa_{i,j}$ of the linear combinations are not provided explicitly) for certain urn models can be also found in [Reference Flajolet, Dumas and Puyhaubert10, Proposition 15]. Our result follows directly from our Proposition 2.

Corollary 1. (Moments of $K_n$ .) If $\alpha\ne 0$ then

(23) \begin{equation} \mathbb{E}(K_n^i) = \frac{1}{[\theta_2]_n} \sum_{j=0}^i \kappa_{i,j} [\theta_2+\alpha j]_n,\qquad n,i\in\mathbb{N}_0, \end{equation}

where the coefficients $\kappa_{i,j}$ only depend on $\gamma\,:\!=\,\theta_1/\alpha$ and are given by

(24) \begin{equation} \kappa_{i,j} \,:\!=\, [\gamma]_j(-1)^{i-j}S_\gamma(i,j), \qquad i\in\mathbb{N}_0,\ j\in\{0,\ldots,i\}, \end{equation}

with $S_\gamma(i,j)\,:\!=\,S(i,j;0,1,\gamma)$ ; see (12). In particular, if $\alpha>0$ then, for all $i\in\mathbb{N}_0$ ,

(25) \begin{equation} \mathbb{E}(K_n^i) \sim \kappa_{i,i}\frac{[\theta_2+\alpha i]_n}{[\theta_2]_n} \sim [\gamma]_i \frac{\Gamma(\theta_2)}{\Gamma(\alpha i+\theta_2)}n^{\alpha i}, \qquad n\to\infty, \end{equation}

and, if $\alpha<0$ , then, for all $i\in\mathbb{N}_0$ ,

(26) \begin{equation} \lim_{n\to\infty}\mathbb{E}(K_n^i) = \kappa_{i,0} = (-\gamma)^i\ =\ m^i. \end{equation}

Remark 11. Roughly speaking, for $\alpha>0$ the ith moment $\mathbb{E}(K_n^i)$ is of order $n^{\alpha i}$ , whereas for $\alpha<0$ the ith moment $\mathbb{E}(K_n^i)$ is bounded and converges to $m^i$ as $n\to\infty$ , in agreement with the fact that, for $\alpha<0$ , the number of occupied tables never exceeds m.

Proof. Plugging the expansion

\begin{equation*} \textrm{e}^{\gamma\lambda}\big(1-\textrm{e}^\lambda\big)^j = (-1)^j \textrm{e}^{\gamma\lambda}(\textrm{e}^\lambda-1)^j = (-1)^jj!\sum_{i=j}^\infty S_\gamma(i,j)\frac{\lambda^i}{i!} \end{equation*}

into (19) leads to

\begin{align*} \mathbb{E}(\textrm{e}^{-\lambda K_n}) & = \sum_{j=0}^\infty \frac{[\theta_2+\alpha j]_n}{[\theta_2]_n} [\gamma]_j(-1)^j \sum_{i=j}^\infty S_\gamma(i,j)\frac{\lambda^i}{i!}\\[3pt] & = \sum_{i=0}^\infty \frac{(-\lambda)^i}{i!} \sum_{j=0}^i \frac{[\theta_2+\alpha j]_n}{[\theta_2]_n} [\gamma]_j (-1)^{i-j} S_\gamma(i,j), \quad n\in\mathbb{N}_0,\ 0\le\lambda<\log 2. \end{align*}

Comparing the coefficients in this expansion with those of

\begin{equation*} \mathbb{E}(\textrm{e}^{-\lambda K_n}) = \sum_{i=0}^\infty \mathbb{E}(K_n^i)\frac{(-\lambda)^i}{i!} \end{equation*}

shows that $K_n$ has moments (23) with weights $\kappa_{i,j}$ defined via (24). The asymptotic formulas (25) and (26) follow immediately.

We now study the asymptotics of $K_n$ as $n\to\infty$ and start with the rather simple case $\alpha=-\kappa<0$ and $\theta_1=m\kappa$ for some $m\in\mathbb{N}$ . In this case we conclude from (14) and $[\theta_1|\alpha]_k=[m\kappa|-\kappa]_k=\kappa^km!/(m-k)!$ that, for $k\in\{0,\ldots,m\wedge n\}$ ,

\begin{align*} \mathbb{P}(K_n=k) & = \frac{[\theta_1|\alpha]_k}{[\theta_2]_n} S(n,k;-1,\kappa,\theta_2-\theta_1)\nonumber\\[3pt] & = \kappa^k\frac{m!}{(m-k)!} \frac{1}{[\theta_2]_n}S(n,k;-1,\kappa,\theta_2-\theta_1).\end{align*}

Proposition 3. If $\alpha=-\kappa<0$ and $\theta_1=m\kappa$ for some $m\in\mathbb{N}$ , then, as $n\to\infty$ , $K_n\to m$ almost surely and in $L^p$ for any $p>0$ .

Proof. For $\alpha<0$ the chain $(K_n)_{n\in\mathbb{N}_0}$ has finite state space $\{0,\ldots,m\}$ . Thus, as $n\to\infty$ , the chain converges almost surely to its absorbing state m. For all $i\in\mathbb{N}_0$ , by (26), $\lim_{n\to\infty}\mathbb{E}(K_n^i)=m^i$ . This convergence of all moments together with the almost sure convergence implies (see [Reference Kallenberg20, p. 68, Proposition 4.12]) the convergence $K_n\to m$ in $L^p$ for any $p>0$ .

Example 1. For $\kappa=1$ the Stirling numbers $S(n,k;-1,\kappa,0)$ coincide with the Lah numbers $S(n,k;-1,1,0)=\frac{n!}{k!}\binom{n-1}{k-1}$ . Choosing in addition $\theta_2\,:\!=\,\theta_1$ ( $=m$ ) yields

\begin{equation*} \mathbb{P}(K_n=k) = \frac{m!}{(m-k)!}\frac{(m-1)!}{(m+n-1)!} \frac{n!}{k!}\binom{n-1}{k-1} = \frac{\binom{m-1}{k-1}\binom{n}{k}} {\binom{m+n-1}{n-1}}, \quad k\in\{0,\ldots,m\wedge n\}, \end{equation*}

so $K_n-1$ has a hypergeometric distribution with parameters $m+n-1$ , $m-1$ , and $n-1$ .

More exciting is the case $\alpha>0$ . Before we state the main result (Theorem 3), let us turn to Mittag–Leffler distributions. Let $0\le\alpha\le 1$ and $\beta,\gamma>0$ with $\beta\ge\alpha\gamma$ . By Lemma 2 provided in Section 7 there exists a positive random variable $\eta$ uniquely determined by its integer moments,

(27) \begin{equation} \mathbb{E}(\eta^j) = \frac{\Gamma(\beta)}{\Gamma(\gamma)}\frac{\Gamma(j+\gamma)}{\Gamma(\alpha j+\beta)},\qquad j\in\mathbb{N}_0.\end{equation}

Since the Laplace transform $\psi$ of $\eta$ is related to Prabhakar’s [Reference Prabhakar32] three-parameter Mittag–Leffler function $E_{\alpha,\beta}^\gamma$ via $\psi(\lambda)=\Gamma(\beta)E_{\alpha,\beta}^\gamma(-\lambda)$ , $\lambda\ge 0$ , we call the distribution of $\eta$ the three-parameter Mittag–Leffler distribution and denote it by ${\rm ML}(\alpha,\beta,\gamma)$ . The distribution ${\rm ML}(\alpha,\beta)\,:\!=\,{\rm ML}(\alpha,\beta,\beta/\alpha)$ is called the two-parameter Mittag–Leffler distribution and for $\beta\,:\!=\,\gamma\,:\!=\,1$ we recover the one-parameter Mittag–Leffler distribution ${\rm ML}(\alpha)\,:\!=\,{\rm ML}(\alpha,1,1)$ having moments $\Gamma(j+1)/\Gamma(\alpha j+1)$ , $j\in\mathbb{N}_0$ . For more information on Mittag–Leffler distributions we refer the reader to Sections 7 and 8. Pitman [Reference Pitman30, Theorem 3.8] showed for the two-parameter CRP with parameters $0<\alpha<1$ and $\theta>-\alpha$ that, as $n\to\infty$ , $K_n/n^\alpha\to K$ almost surely and in $L^p$ for any $p>0$ , where $K\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\theta)$ . The following convergence result is the natural extension for the restaurant process with cocktail bar.

Theorem 3. (Convergence of scaled $K_n$ .) If $\alpha>0$ then, as $n\to\infty$ , $K_n/n^\alpha\to K$ almost surely and in $L^p$ for any $p>0$ , where $K\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ is three-parameter Mittag–Leffler distributed with parameters $0<\alpha<1$ , $\beta\,:\!=\,\theta_2$ , and $\gamma\,:\!=\,\theta_1/\alpha$ . In particular, the convergence

\begin{equation*} \lim_{n\to\infty}\mathbb{E}\bigg(\bigg(\frac{K_n}{n^\alpha}\bigg)^p\bigg) = \mathbb{E}(K^p) = \frac{\Gamma(\beta)\Gamma(p+\gamma)} {\Gamma(\gamma)\Gamma(\alpha p+\beta)}, \qquad p\ge 0, \end{equation*}

of all moments holds.

Two proofs are provided. The first is self-contained, more elementary than that provided in [Reference Pitman30] for the two-parameter CRP, and based on a martingale argument similar to those used in the context of Pólya urns. The distribution of K is determined via Corollary 1. The second proof is based on reinterpretation of the model in terms of a mixture of a Pólya urn and a two-parameter CRP. Both proofs are relatively short.

First proof of Theorem 3. For $n\in\mathbb{N}_0$ define

(28) \begin{equation} M_n \,:\!=\, \frac{\Gamma(n+\beta)}{\Gamma(n+\alpha+\beta)}(K_n+\gamma). \end{equation}

For all $n\in\mathbb{N}_0$ , by (13),

\begin{align*} \mathbb{E}(K_{n+1}+\gamma \mid K_n) & = (K_n+\gamma)\bigg(1-\frac{\alpha(K_n+\gamma)}{n+\beta}\bigg) + (K_n+1+\gamma)\frac{\alpha(K_n+\gamma)}{n+\beta}\\[3pt] & = (K_n+\gamma)\bigg(1-\frac{\alpha(K_n+\gamma)}{n+\beta} +\frac{\alpha(K_n+1+\gamma)}{n+\beta} \bigg)\\[3pt] & = (K_n+\gamma)\bigg(1+\frac{\alpha}{n+\beta}\bigg) = (K_n+\gamma)\frac{n+\alpha+\beta}{n+\beta} \end{align*}

almost surely. Multiplication by $\Gamma(n+1+\beta)/\Gamma(n+1+\alpha+\beta)$ yields $\mathbb{E}(M_{n+1} \mid K_n)=M_n$ almost surely for all $n\in\mathbb{N}_0$ . Thus, $(M_n)_{n\in\mathbb{N}_0}$ is a nonnegative martingale with respect to the filtration $(\mathcal{ F}_n)_{n\in\mathbb{N}_0}$ , where $\mathcal{ F}_n$ denotes the $\sigma$ -algebra generated by $K_0,\ldots,K_n$ . By martingale convergence there exists K such that $M_n\to K$ almost surely as $n\to\infty$ . Note that $K_n\to\infty$ almost surely as $n\to\infty$ , which follows from Theorem 2 and the fact that all of the $P_i$ , $i\in\mathbb{N}$ , are strictly positive almost surely. Thus, it follows from (28) that $M_n\sim (K_n+\gamma)/n^\alpha\sim K_n/n^\alpha$ almost surely as $n\to\infty$ , and hence $K_n/n^\alpha\to K$ almost surely.

Now Corollary 1 is used to determine the distribution of K. Let $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ be three-parameter Mittag–Leffler distributed with parameters $\alpha,\beta\,:\!=\,\theta_2$ and $\gamma\,:\!=\,\theta_1/\alpha$ . For all $j\in\mathbb{N}_0$ , by (25) and (27),

\begin{equation*} \mathbb{E}(K_n^j) \sim \frac{\Gamma(j+\gamma)}{\Gamma(\gamma)} \frac{\Gamma(\beta)}{\Gamma(\alpha j+\beta)}n^{\alpha j} = \mathbb{E}\big(\eta^j\big)n^{\alpha j},\qquad n\to\infty. \end{equation*}

Thus, $\lim_{n\to\infty}\mathbb{E}((K_n/n^\alpha)^j)=\mathbb{E}(\eta^j)$ , $j\in\mathbb{N}_0$ . Since (see Section 7, Lemma 2) the distribution ${\rm ML}(\alpha,\beta,\gamma)$ of $\eta$ is uniquely determined by its moments, this convergence of moments implies (see, for example, [3, Theorem 30.2]) the convergence $K_n/n^\alpha\to\eta$ in distribution as $n\to\infty$ . In particular, the distribution of the almost sure limit K must coincide with the distribution ${\rm ML}(\alpha,\beta,\gamma)$ of the distributional limit $\eta$ . By (41), K has moments $\mathbb{E}(K^p)=(\Gamma(\beta)\Gamma(p+\gamma))/(\Gamma(\gamma)\Gamma(\alpha p+\beta))$ , $p\ge 0$ . The rest of the proof works as in the proof of Theorem 1.

Second proof of Theorem 3. Recall the reinterpretation of the model in terms of a mixture of a two-type Pólya urn and a standard CRP described in Section 2. We have

\begin{equation*} \frac{K_n}{n^\alpha} = \frac{K_n}{R_n^{\alpha}} \bigg(\frac{R_n}{n}\bigg)^\alpha. \end{equation*}

Since the jump evolution of the process $(R_{n,1},\ldots,R_{n,K_n})_{n\in\mathbb{N}}$ is a standard CRP with parameters $\alpha$ and $\theta_1$ , it follows from [Reference Pitman30, Theorem 3.8] that, as $n\to\infty$ , $K_n/R_n^\alpha\to L$ almost surely and in $L^p$ for any $p>0$ , where $L\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\theta_1)$ . By standard results for Pólya urns (see also the remark at the end of Section 3), as $n\to\infty$ , $R_n/n\to R$ almost surely and in $L^p$ for any $p>0$ , where $R\stackrel{\textrm{d}}{=}\beta(\theta_1,\theta_2-\theta_1)$ . Thus, $K_n/n^\alpha\to LR^\alpha \!=: K$ almost surely. Moreover, L and R are independent. By the independence structure we also have convergence of all moments $\mathbb{E}((K_n/n^\alpha)^j)=\mathbb{E}((K_n/R_n^\alpha)^j)\mathbb{E}((R_n/n)^{\alpha j})\to\mathbb{E}(L^j)\mathbb{E}(R^{\alpha j})=\mathbb{E}(K^j)$ , $j\in\mathbb{N}_0$ . The almost sure convergence implies convergence in probability. Convergence in probability together with the convergence of all moments implies (see, for example, [Reference Kallenberg20, p. 68, Proposition 4.12]) convergence in $L^p$ and, hence, the convergence of the pth moments for any $p>0$ .

The distribution of K is determined via moment calculations as follows. Using the density formula [Reference Pitman30, Eq. (3.27)] of L it is easily checked that L has moments (see also [Reference Pitman30, Eq. (3.47)])

(29) \begin{equation} \mathbb{E}(L^j) = \frac{\Gamma(\theta_1)\Gamma\Big(j+\frac{\theta_1}{\alpha}\Big)} {\Gamma\Big(\frac{\theta_1}{\alpha}\Big)\Gamma(\alpha j+\theta_1)}, \qquad j\in\mathbb{N}_0. \end{equation}

Moreover, the beta distributed random variable R satisfies

(30) \begin{equation} \mathbb{E}(R^{\alpha j}) = \frac{\Gamma(\theta_2)\Gamma(\alpha j+\theta_1)} {\Gamma(\theta_1)\Gamma(\alpha j+\theta_2)},\qquad j\in\mathbb{N}_0. \end{equation}

Since L and R are independent it follows from (29) and (30) that $K=LR^\alpha$ has moments

\begin{equation*} \mathbb{E}(K^j) = \mathbb{E}(L^j)\mathbb{E}(R^{\alpha j}) = \frac{\Gamma(\theta_2)\Gamma\Big(j+\frac{\theta_1}{\alpha}\Big)} {\Gamma\Big(\frac{\theta_1}{\alpha}\Big)\Gamma(\alpha j+\theta_2)} = \frac{\Gamma(\beta)\Gamma(j+\gamma)}{\Gamma(\gamma)\Gamma(\alpha j+\beta)},\qquad j\in\mathbb{N}_0. \end{equation*}

These are the moments of ${\rm ML}(\alpha,\beta,\gamma)$ . Since (see Section 7, Lemma 2) the distribution ${\rm ML}(\alpha,\beta,\gamma)$ is uniquely determined by its moments, it follows that $K\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ .

Let us now turn to the first passage times $T_k\,:\!=\,\inf\{n\in\mathbb{N}_0:K_n=k\}$ , $k\in\mathbb{N}_0$ .

Corollary 2. If $\alpha>0$ then, as $k\to\infty$ , $T_k/k^{1/\alpha}\to K^{-1/\alpha}$ almost surely, where $K\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ with $\beta\,:\!=\,\theta_2$ and $\gamma\,:\!=\,\theta_1/a$ .

Proof. From $K_n\to\infty$ almost surely as $n\to\infty$ it follows that $T_k\to\infty$ almost surely as $k\to\infty$ . Therefore, as $k\to\infty$ , $k/T_k^\alpha=K_{T_k}/T_k^\alpha\to K$ almost surely by Theorem 3, or, equivalently, $T_k/k^{1/\alpha}\to K^{-1/\alpha}$ almost surely.

Remark 12. (Siegmund duality) It is readily seen that the chain $K=(K_n)_{n\in\mathbb{N}_0}$ is stochastically monotone, i.e. $\mathbb{P}(K_{n+1}\ge j \mid K_n=i)$ is increasing in $i\in\mathbb{N}_0$ for all $n,j\in\mathbb{N}_0$ . Moreover, $\mathbb{P}(K_{n+1}\ge j \mid K_n=i)\to 1$ as $i\to\infty$ for all $n,j\in\mathbb{N}_0$ . Thus, there exists a Markov chain $X=(X_n)_{n\in\mathbb{N}_0}$ which is Siegmund dual to K, i.e. $\mathbb{P}(X_{n+1}\le j \mid X_n=i)=\mathbb{P}(K_{n+1}\ge i \mid K_n=j)$ for all $n,i,j\in\mathbb{N}_0$ . For general information on Siegmund duality we refer the reader to [Reference Siegmund34]. It is easily seen that the chain X has absorbing state 0 and transition probabilities

\begin{equation*} \mathbb{P}(X_{n+1}=i-1 \mid X_n=i) = \frac{\alpha(i-1)+\theta_1}{n+\theta_2} = 1-\mathbb{P}(X_{n+1}=i \mid X_n=i),\qquad i\in\mathbb{N}.\end{equation*}

It might be interesting to study properties of the chain X. The same could be done for $N_{n,i}$ (i fixed) instead of $K_n$ .

7. The three-parameter Mittag–Leffler distribution

We provide an independent section on a natural extension of the one- and two-parameter Mittag–Leffler distributions. At least two types of Mittag–Leffler distributions are known from the literature. Type 1 distributions are heavy tailed, whereas type 2 distributions have finite moments of all orders. We are dealing here with the three-parameter Mittag–Leffler distribution of type 2. The key lemma is the following.

Lemma 2. Let $0<\alpha<1$ and $\beta,\gamma>0$ with $\beta\ge\alpha\gamma$ . There exists a strictly positive random variable $\eta$ uniquely determined by its integer moments

(31) \begin{equation} \mathbb{E}(\eta^j) = \frac{\Gamma(\beta)}{\Gamma(\gamma)} \frac{\Gamma(j+\gamma)}{\Gamma(\alpha j+\beta)}, \qquad j\in\mathbb{N}_0. \end{equation}

The random variable $\eta$ has strictly positive density

(32) \begin{equation} g(x) \,:\!=\, g_{\alpha,\beta,\gamma}(x) \,:\!=\, \frac{\Gamma(\beta)}{\Gamma(\gamma)}x^{\gamma-1} \sum_{n=0}^\infty \frac{(-x)^n}{n!}\frac{1}{\Gamma(\beta-\alpha\gamma-\alpha n)}, \qquad x>0, \end{equation}

with respect to Lebesgue measure on $(0,\infty)$ , with the usual convention that $1/\Gamma(s)\,:\!=\,0$ for $s\in-\mathbb{N}_0\,:\!=\,\{0,-1,-2,\ldots\}$ .

Remark 13. It will be clarified later (see (41)) that (31) also holds with j replaced by any real number $p>-\gamma$ .

Proof. For $j\in\mathbb{N}_0$ define $a_j\,:\!=\,\Gamma(\beta)\Gamma(j+\gamma)/(j!\Gamma(\gamma)\Gamma(\alpha j+\beta))$ . Clearly,

\begin{equation*} \frac{a_{j+1}}{a_j} = \frac{j+\gamma}{j+1}\frac{\Gamma(\alpha j+\beta)}{\Gamma(\alpha j+\alpha+\beta)} \sim \frac{\Gamma(\alpha j+\beta)}{\Gamma(\alpha j+\alpha+\beta)} \sim \frac{(\alpha j)^\beta\Gamma(\alpha j)} {(\alpha j)^{\alpha+\beta}\Gamma(\alpha j)} = \frac{1}{(\alpha j)^\alpha} \to 0 \end{equation*}

as $j\to\infty$ . The series $\sum_{j=0}^\infty a_jt^j$ is therefore absolutely convergent for every $t\in\mathbb{R}$ . In particular, the function $\psi:[0,\infty)\to\mathbb{R}$ , defined via

(33) \begin{equation} \psi(\lambda) \,:\!=\, \sum_{j=0}^\infty a_j(-\lambda)^j = \frac{\Gamma(\beta)}{\Gamma(\gamma)}\sum_{j=0}^\infty \frac{\Gamma(j+\gamma)}{\Gamma(\alpha j+\beta)} \frac{(-\lambda)^j}{j!}, \qquad\lambda\ge 0, \end{equation}

is well defined. Clearly, $\psi(0)=1$ . Moreover, $\psi(\lambda)=\Gamma(\beta) E_{\alpha,\beta}^\gamma(-\lambda)$ , where $E_{\alpha,\beta}^\gamma$ denotes Prabhakar’s [Reference Prabhakar32] three-parameter Mittag–Leffler function. Equation (22) of [Reference Tomovski, Pogány and Srivastava40], which is valid for $t\,:\!=\,1$ , $0<\alpha<1$ , and $\beta,\gamma>0$ satisfying $\beta\ge\alpha\gamma$ , implies that $\psi(\lambda)=\int_0^\infty \textrm{e}^{-\lambda x}g(x)\,{\rm d}x$ , where $g(x)\,:\!=\, (\Gamma(\beta)/\Gamma(\gamma))x^{(\beta-1)/\alpha-1}\Psi(\alpha,\beta,\gamma;x^{-1/\alpha})$ , $x>0$ , and $\Psi$ is defined as in [Reference Tomovski, Pogány and Srivastava40, Eq. (21)]. By [36, p. 122] (see also [Reference Stanković35]), $\Psi$ , and hence g, is strictly positive. Thus, $\psi$ is completely monotone and, consequently, a Laplace transform of some nonnegative random variable, say $\eta$ . From (33) we conclude that $\eta$ has moments (31). The moment-generating function $t\mapsto\sum_{j=0}^\infty a_jt^j=\sum_{j=0}^\infty t^j\mathbb{E}(\eta^j)/j!$ of $\eta$ has infinite radius of convergence. The distribution of $\eta$ is hence (see [3, Theorem 30.1]) uniquely determined by its moments. Obviously, g is a density of $\eta$ . The series formula (32) for g(x) follows from [Reference Tomovski, Pogány and Srivastava40, Eqs. (19), (21)].

The proof shows that the Laplace transform $\psi$ of $\eta$ is related to Prabhakar’s [Reference Prabhakar32] generalized three-parameter Mittag–Leffler function $E_{\alpha,\beta}^\gamma$ via $\psi(\lambda)=\Gamma(\beta)E_{\alpha,\beta}^\gamma(-\lambda)$ . The following definition is hence natural.

Definition 1. (Three-parameter Mittag–Leffler distribution.) Let $0<\alpha<1$ and $\beta,\gamma>0$ with $\beta\ge\alpha\gamma$ . The distribution of the random variable $\eta$ from Lemma 2 is called the three-parameter Mittag–Leffler distribution (of type 2). We denote this distribution by ${\rm ML}(\alpha,\beta,\gamma)$ .

Remark 14. The density (32) and the Mellin transform

\begin{equation*} s \mapsto \mathbb{E}(\eta^{s-1}) = \frac{\Gamma(\beta)\Gamma(s-1+\gamma)} {\Gamma(\gamma)\Gamma(\alpha(s-1)+\beta)}, \qquad s>1-\gamma, \end{equation*}

of ${\rm ML}(\alpha,\beta,\gamma)$ occur (with other parameter notation and without mentioning the name of the distribution) in [Reference Flajolet, Dumas and Puyhaubert10, Eqs. (80), (81)]. The moments of this distribution also occur (with other parameter notation) in [Reference Janson19, Theorem 8.8].

Remark 15. The distribution ${\rm ML}(\alpha,\beta,\gamma)$ is also defined for $\alpha=0$ and $\alpha=1$ . For $\alpha=0$ the density (32) reduces to $g(x)=x^{\gamma-1}\textrm{e}^{-x}/\Gamma(\gamma)$ , showing that it is appropriate to define ${\rm ML}(0,\beta,\gamma)$ to be the gamma distribution with shape parameter $\gamma$ and scale parameter 1 (independent of $\beta$ ). For $\alpha=1$ the moments (31) coincide with those of the beta distribution with parameters $\gamma$ and $\beta-\gamma$ (with the usual convention to be the Dirac measure in 1 for $\beta=\gamma$ ). It is hence appropriate to define ${\rm ML}(1,\beta,\gamma)$ to be the beta distribution with parameters $\gamma$ and $\beta-\gamma$ .

Remark 16. Let us clarify the relation to the classical one-parameter Mittag–Leffler distribution ${\rm ML}(\alpha)\,:\!=\,{\rm ML}(\alpha,1,1)$ . For $\beta=\gamma=1$ the density (32) reduces to

\begin{align*} g_\alpha(x) & \,:\!=\, g_{\alpha,1,1}(x) = \sum_{n=0}^\infty \frac{(-x)^n}{n!}\frac{1}{\Gamma(1-\alpha(n+1))} = \sum_{k=1}^\infty \frac{(-x)^{k-1}}{(k-1)!}\frac{1}{\Gamma(1-\alpha k)}\\[3pt] & = \sum_{k=1}^\infty \frac{(-x)^{k-1}}{(k-1)!} \frac{1}{(-\alpha k)\Gamma(-\alpha k)} = \frac{1}{\alpha}\sum_{k=1}^\infty \frac{(-1)^kx^{k-1}}{k!} \frac{1}{\Gamma(-\alpha k)},\qquad x>0. \end{align*}

Applying the formula $\pi/\Gamma(1-z)=\Gamma(z)\sin(\pi z)$ with $z\,:\!=\,\alpha k+1$ we recover the density formula (see, for example, [Reference Pitman30, p. 11, Eq. (0.43)])

\begin{align*} g_\alpha(x) & = \frac{1}{\pi\alpha}\sum_{k=1}^\infty \frac{(-1)^kx^{k-1}}{k!}\Gamma(\alpha k+1)\sin(\pi(\alpha k+1))\\[3pt] & = \frac{1}{\pi\alpha}\sum_{k=1}^\infty \frac{(-x)^{k-1}}{k!}\Gamma(\alpha k+1)\sin(\pi\alpha k), \qquad x>0. \end{align*}

In this case $\eta$ has moments $\mathbb{E}(\eta^j)=\Gamma(j+1)/\Gamma(\alpha j+1)$ , $j\in\mathbb{N}_0$ . Alternatively, the density $g_\alpha$ is obtained from the density $g_{\alpha,\beta,\gamma}$ of ${\rm ML}(\alpha,\beta,\gamma)$ by choosing $\gamma=\beta/\alpha$ and letting $\beta\to 0$ , since, by (32), for all $x>0$ ,

\begin{equation*} g_{\alpha,\beta,\frac{\beta}{\alpha}}(x) = \frac{\Gamma(\beta)}{\Gamma\big(\frac{\beta}{\alpha}\big)} x^{\frac{\beta}{\alpha}-1}\sum_{n=1}^\infty \frac{(-x)^n}{n!} \frac{1}{\Gamma(-\alpha n)} \to \frac{1}{\alpha x}\sum_{n=1}^\infty \frac{(-x)^n}{n!} \frac{1}{\Gamma(-\alpha n)} = g_\alpha(x) \end{equation*}

as $\beta\to 0$ .

Remark 17. We collect here some properties of ${\rm ML}(\alpha)$ . It is known (see, for example, [Reference Feller9, p. 453]) that $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha)$ satisfies $\eta\stackrel{\textrm{d}}{=}\xi^{-\alpha}$ , where $\xi$ is a positive $\alpha$ -stable random variable with Laplace transform $\lambda\mapsto \textrm{e}^{-\lambda^\alpha}$ , $\lambda\ge 0$ . In particular, the density $g_\alpha$ of $\eta$ satisfies

\begin{equation*} g_\alpha(x) = \frac{f_\alpha(x^{-1/\alpha})}{\alpha x^{1+1/\alpha}}, \qquad x>0, \end{equation*}

where $f_\alpha$ denotes the density of $\xi$ having series representation

(34) \begin{equation} f_\alpha(x) = \frac{1}{\pi}\sum_{k=1}^\infty \frac{(-1)^{k-1}}{k!}\frac{\Gamma(\alpha k+1)\sin(\pi\alpha k)}{x^{\alpha k+1}},\qquad x>0. \end{equation}

From the representation $\eta\stackrel{\textrm{d}}{=}\xi^{-\alpha}$ and [Reference Kanter21, Corollary 4.1] (see also [Reference Chambers, Mallows and Stuck5]), it follows that $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha)$ satisfies

(35) \begin{equation} \eta \stackrel{\textrm{d}}{=} \frac{\sin(U)}{(\sin(\alpha U))^\alpha} \bigg(\frac{E}{\sin((1-\alpha)U)}\bigg)^{1-\alpha} = \bigg(\frac{E}{a(U)}\bigg)^{1-\alpha}, \end{equation}

where U and E are independent, U is uniformly distributed on $(0,\pi)$ , E is standard exponentially distributed, and $a(\varphi)\,:\!=\,a_\alpha(\varphi)$ is defined via

(36) \begin{align} a_\alpha(\varphi) & \,:\!=\, \bigg(\frac{\sin(\alpha\varphi)}{\sin\varphi}\bigg)^{1/(1-\alpha)} \frac{\sin((1-\alpha)\varphi)}{\sin(\alpha\varphi)}\nonumber\\[3pt] & = \bigg( \frac{\sin(\alpha\varphi)}{\sin\varphi} \bigg)^{\alpha/(1-\alpha)} \frac{\sin((1-\alpha)\varphi)}{\sin\varphi}, \qquad 0<\varphi<\pi. \end{align}

The functions $a_\alpha(\!\cdot\!)$ are strictly increasing on $(0,\pi)$ and have the symmetry property

(37) \begin{equation} (a_\alpha(\varphi))^{1-\alpha} = (a_{1-\alpha}(\varphi))^\alpha, \qquad 0<\alpha<1,\ 0<\varphi<\pi. \end{equation}

Note that $a_\alpha(\varphi)={\rm Re}(z^\alpha-z)$ is the real part of $z^\alpha-z$ for those complex numbers $z=r\textrm{e}^{i\varphi}\in\mathbb{C}$ for which the imaginary part ${\rm Im}(z^\alpha-z)$ vanishes. The representation (35) yields an efficient and accurate pseudo-random number generator for ${\rm ML}(\alpha)$ . From (35) a straightforward calculation shows that $\mathbb{P}(\eta\le x)=\pi^{-1}\int_0^\pi (1-\exp(-a(\varphi)x^{\frac{1}{1-\alpha}}))\,{\rm d}\varphi$ . Taking the derivative with respect to x yields

(38) \begin{equation} g_\alpha(x) = \frac{x^{\frac{1}{1-\alpha}-1}}{\pi(1-\alpha)} \int_0^\pi a(\varphi)\exp\Big(-a(\varphi)x^{\frac{1}{1-\alpha}}\Big)\,{\rm d}\varphi, \qquad x>0. \end{equation}

This integral formula is useful for computing the density $g_\alpha$ of ${\rm ML}(\alpha)$ numerically. The analogous integral formula for the density $f_\alpha$ of the $\alpha$ -stable random variable $\xi$ can be traced back to [Reference Ibragimov and Chernin15, p. 456] and [Reference Mikusiński24, Eq. (2)].

For particular parameter choices the density of $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ is explicitly known. The following examples show that the half-normal distribution and the Rayleigh distribution belong to the class of three-parameter Mittag–Leffler distributions.

Example 2. For $(\alpha,\beta,\gamma)=(\frac{1}{2},1,1)$ the random variable $\eta$ is half-normal distributed with density

\begin{equation*} x \mapsto \sum_{n=0}^\infty \frac{(-x)^n}{n!\Gamma\big(\frac{1}{2}-\frac{n}{2}\big)} = \sum_{m=0}^\infty \frac{x^{2m}}{(2m)!\Gamma\big(\frac{1}{2}-m\big)} = \frac{1}{\Gamma\big(\frac{1}{2}\big)}\sum_{m=0}^\infty \frac{(-x^2/4)^m}{m!} = \frac{\textrm{e}^{-x^2/4}}{\sqrt{\pi}} , \end{equation*}

$x>0$ , having moments $\mathbb{E}(\eta^j)=\Gamma(j+1)/\Gamma\big(\frac{j}{2}+1\big)=2^j\Gamma\big(\frac{j+1}{2}\big)/\sqrt{\pi}$ , where the last equality holds by Legendre’s duplication formula $\Gamma\big(\frac{x}{2}\big)\Gamma\big(\frac{x+1}{2}\big)=2^{1-x}\sqrt{\pi}\Gamma(x)$ .

Example 3. For $(\alpha,\beta,\gamma)=\big(\frac{1}{2},\beta,2\beta\big)$ the random variable $\eta$ has moments $\mathbb{E}(\eta^j)=\Gamma(\beta)\Gamma(j+2\beta)/\big(\Gamma(2\beta)\Gamma\big(\frac{j}{2}+\beta\big)\big)=2^j\Gamma\big(\frac{j+1}{2}+\beta\big)/\Gamma\big(\beta+\frac{1}{2}\big)$ , $j\in\mathbb{N}_0$ , where the last equality again holds by Legendre’s duplication formula. Moreover, $\eta$ has density

\begin{equation*} x \mapsto \frac{\Gamma(\beta)}{\Gamma(2\beta)}\frac{x^{2\beta}}{2\sqrt{\pi}\,} \textrm{e}^{-x^2/4} = \frac{(x/2)^{2\beta}}{\Gamma\big(\beta+\frac{1}{2}\big)}\textrm{e}^{-x^2/4},\qquad x>0. \end{equation*}

For $\beta=\frac{1}{2}$ it follows that $\eta$ is Rayleigh distributed with density $x\mapsto (x/2)\textrm{e}^{-x^2/4}$ , $x>0$ .

Example 4. A technical but straightforward calculation shows that $\eta\stackrel{\textrm{d}}{=}{\rm ML}\big(\frac{1}{2},2,2\big)$ has density $x\mapsto x\big(1-\pi^{-1/2}\int_0^x \textrm{e}^{-u^2/4}\,{\rm d}u\big)=x(1-\mathbb{P}(|X|\le x))$ , $x>0$ , where $X\stackrel{\textrm{d}}{=}N(0,2)$ is normal distributed with variance 2.

For $0<\alpha<1$ and $\beta>0$ the distribution ${\rm ML}(\alpha,\beta)\,:\!=\,{\rm ML}(\alpha,\beta,\beta/\alpha)$ is usually called the two-parameter Mittag–Leffler distribution. The following corollary provides a characterization of ${\rm ML}(\alpha,\beta,\gamma)$ in terms of a beta distribution and a two-parameter Mittag–Leffler distribution.

Corollary 3. Let $0<\alpha<1$ and $\beta,\gamma>0$ with $\beta\ge\alpha\gamma$ . If $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ is three-parameter Mittag–Leffler distributed with parameters $\alpha$ , $\beta$ , and $\gamma$ then the distributional factorization

(39) \begin{equation} \eta \stackrel{\textrm{d}}{=} BL \end{equation}

holds, where B and L are independent, $B\stackrel{\textrm{d}}{=}{\rm ML}(1,\beta/\alpha,\gamma)$ is beta distributed with parameters $\gamma$ and $\beta/\alpha-\gamma$ , and $L\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta)$ is two-parameter Mittag–Leffler distributed.

Proof. Clearly, B has moments $\mathbb{E}(B^j)=\Gamma\big(\frac{\beta}{\alpha}\big)\Gamma(j+\gamma)/(\Gamma(\gamma)\Gamma\big(j+\frac{\beta}{\alpha}\big))$ , $j\in\mathbb{N}_0$ . By (31), L has moments $\mathbb{E}(L^j)=\Gamma(\beta)\Gamma\big(j+\frac{\beta}{\alpha}\big)/(\Gamma\big(\frac{\beta}{\alpha}\big)\Gamma(\alpha j+\beta))$ , $j\in\mathbb{N}_0$ . Since B and L are independent, it follows that BL has moments

\begin{align*} \mathbb{E}((BL)^j) = \mathbb{E}(B^j)\,\mathbb{E}(L^j) & = \frac{\Gamma\big(\frac{\beta}{\alpha}\big)\Gamma(j+\gamma)} {\Gamma(\gamma)\Gamma\big(j+\frac{\beta}{\alpha}\big)} \frac{\Gamma(\beta)\Gamma\big(j+\frac{\beta}{\alpha}\big)} {\Gamma\big(\frac{\beta}{\alpha}\big)\Gamma(\alpha j+\beta)}\\[3pt] & = \frac{\Gamma(\beta)\Gamma(j+\gamma)} {\Gamma(\gamma)\Gamma(\alpha j+\beta)},\qquad j\in\mathbb{N}_0. \end{align*}

These moments coincide with those (see (31)) of $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ . Thus, $\eta\stackrel{\textrm{d}}{=}BL$ .

Remark 18. The factorization (39) allows us to define the three-parameter Mittag–Leffler distribution from the two-parameter Mittag–Leffler distribution. Equation (39) also yields a random number generator for the three-parameter Mittag–Leffler distribution. Accurate and efficient sampling methods for the two-parameter Mittag–Leffler distribution are provided in Section 8.

Remark 19. The random variable $L\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta)$ in Corollary 3 has density (see, for example, [Reference Pitman30, p. 68, Eq. (3.27)])

(40) \begin{equation} g_{\alpha,\beta}(x) \,:\!=\, \frac{\Gamma(\beta+1)}{\Gamma\big(\frac{\beta}{\alpha}+1\big)}x^{\frac{\beta}{\alpha}} g_\alpha(x), \qquad x>0, \end{equation}

where $g_\alpha$ is the density of the one-parameter Mittag–Leffler distribution ${\rm ML}(\alpha)\,:\!=\,{\rm ML}(\alpha,1,1)$ having moments $\Gamma(p+1)/\Gamma(\alpha p+1)$ , $p>-1$ . In other words, $g_{\alpha,\beta}$ is the $(\beta/\alpha)$ -size-biased density of $g_\alpha$ . From (40) it follows that $L^{-1/\alpha}$ has density

\begin{equation*} x \mapsto \alpha x^{-\alpha-1}g_{\alpha,\beta}(x^{-1/\alpha}) = \frac{\Gamma(\beta+1)}{\Gamma\big(\frac{\beta}{\alpha}+1\big)}x^{-\beta} f_\alpha(x),\qquad x>0, \end{equation*}

where $f_\alpha$ is the density (34) of the positive $\alpha$ -stable random variable $\xi$ with Laplace transform $\lambda\mapsto \textrm{e}^{-\lambda^\alpha}$ , $\lambda\ge 0$ . Thus, $L^{-1/\alpha}$ is the $(-\beta)$ -size-biased random variable of $\xi$ .

Remark 20. The moment formula

\begin{align*} \mathbb{E}(L^p) & = \frac{\Gamma(\beta+1)}{\Gamma\big(\frac{\beta}{\alpha}+1\big)} \int_0^\infty x^{p+\frac{\beta}{\alpha}} g_\alpha(x)\,{\rm d}x\\[3pt] & = \frac{\Gamma(\beta+1)}{\Gamma\big(\frac{\beta}{\alpha}+1\big)} \frac{\Gamma\big(p+\frac{\beta}{\alpha}+1\big)}{\Gamma\big(\alpha\big(p+\frac{\beta}{\alpha}\big)+1\big)} = \frac{\Gamma(\beta)\Gamma\big(p+\frac{\beta}{\alpha}\big)}{\Gamma\big(\frac{\beta}{\alpha}\big)\Gamma(\alpha p+\beta)} \end{align*}

holds for all $p>-\frac{\beta}{\alpha}$ . The calculations in the proof of Corollary 3 thus also hold with j replaced by any real parameter $p>-\gamma$ . In particular, if $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ , then

(41) \begin{equation} \mathbb{E}(\eta^p) = \frac{\Gamma(\beta)\Gamma(p+\gamma)} {\Gamma(\gamma)\Gamma(\alpha p+\beta)}, \qquad p>-\gamma. \end{equation}

For $p>-\gamma$ the p-size-biased random variable $\eta^*$ of $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ satisfies $\eta^*\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta+\alpha p,\gamma+p)$ , since it is seen from (41) and (32) that the density $x\mapsto x^pg_{\alpha,\beta,\gamma}(x)/\mathbb{E}(\eta^p)$ , $x>0$ , of $\eta^*$ coincides with $g_{\alpha,\beta+\alpha p,\gamma+p}$ . In particular, for every $p>0$ the family of three-parameter Mittag–Leffler distributions is closed under p-size-biasing.

Remark 21. Corollary 3 is easily generalized as follows. Let $0<\alpha<1$ and $\beta,\gamma,\gamma'>0$ with $\gamma\le\gamma'$ and $\beta\ge\alpha\gamma'$ . If $\eta\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma)$ and $\eta'\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta,\gamma')$ then the distributional factorization $\eta\stackrel{\textrm{d}}{=}B\eta'$ holds, where $B\stackrel{\textrm{d}}{=}\beta(\gamma,\gamma'-\gamma)$ is independent of $\eta'$ . We have not been able to find a natural explanation of this factorization in terms of an urn model or some combinatorial construction, and leave this question as an open problem for future research.

8. Sampling from the two- and three-parameter Mittag–Leffler distribution

By Corollary 3 and Remark 18 thereafter it suffices to focus on the two-parameter Mittag–Leffler distribution ${\rm ML}(\alpha,\beta)$ . To the best of the author’s knowledge random number generators for ${\rm ML}(\alpha,\beta)$ are not known from the literature. The key to developing random number generators is the following Proposition 4, which provides a useful distributional representation for ${\rm ML}(\alpha,\beta)$ .

Proposition 4. Let $L=L_{\alpha,\beta}\stackrel{\textrm{d}}{=}{\rm ML}(\alpha,\beta)$ be two-parameter Mittag–Leffler distributed with parameters $0<\alpha<1$ and $\beta>0$ . Then the distributional representation

(42) \begin{equation} L \stackrel{\textrm{d}}{=} \bigg(\frac{G}{a(V)}\bigg)^{1-\alpha} \end{equation}

holds, where G is gamma distributed with shape parameter $\frac{\beta}{\alpha}(1-\alpha)+1$ and scale parameter 1 having density $f_G(x)\,:\!=\,x^{\frac{\beta}{\alpha}(1-\alpha)}\textrm{e}^{-x} /\Gamma\big(\frac{\beta}{\alpha}(1-\alpha)+1\big)$ , $x>0$ , the function $a(\!\cdot\!)=a_\alpha(\!\cdot\!)$ is defined via (36), and the $(0,\pi)$ -valued random variable V is independent of G and has density

(43) \begin{equation} h_{\alpha,\beta}(\varphi) \,:\!=\, \frac{c_{\alpha,\beta}}{(a_\alpha(\varphi))^{\frac{\beta}{\alpha}(1-\alpha)}}, \qquad 0<\varphi<\pi, \end{equation}

with normalizing constant

\begin{equation*} c_{\alpha,\beta} \,:\!=\, \frac{\Gamma(\beta+1)\Gamma\big(\frac{\beta}{\alpha}(1-\alpha)+1\big)} {\pi\Gamma\big(\frac{\beta}{\alpha}+1\big)}. \end{equation*}

Proof. For $0<\varphi<\pi$ define $W_\varphi\,:\!=\,(G/a(\varphi))^{1-\alpha}$ . Taking the derivative with respect to x in the equation $\mathbb{P}(W_\varphi\le x)=\mathbb{P}\big(G\le a(\varphi)x^{\frac{1}{1-\alpha}}\big)$ shows that $W_\varphi$ has density $w_\varphi$ given by

\begin{align*} w_\varphi(x) & \,:\!=\, \frac{a(\varphi)}{1-\alpha}x^{\frac{1}{1-\alpha}-1} f_G\big(a(\varphi)x^{\frac{1}{1-\alpha}}\big)\\[3pt] & = \frac{(a(\varphi))^{\frac{\beta}{\alpha}(1-\alpha)+1}} {(1-\alpha)\Gamma\big(\frac{\beta}{\alpha}(1-\alpha)+1\big)} x^{\frac{\beta}{\alpha}+\frac{1}{1-\alpha}-1} \exp\big(-a(\varphi)x^{\frac{1}{1-\alpha}}\big),\qquad x>0. \end{align*}

Note that $w_\varphi$ is the density of a $(\beta/\alpha)$ -size-biased Weibull distribution with distribution function $x\mapsto 1-\exp\big(-a(\varphi)x^{\frac{1}{1-\alpha}}\big)$ , $x>0$ . Now, by (40) and (38), L has density

(44) \begin{align} g_{\alpha,\beta}(x) & = \frac{\Gamma(\beta+1)}{\Gamma\big(\frac{\beta}{\alpha}+1\big)} x^{\frac{\beta}{\alpha}}g_\alpha(x)\nonumber\\[3pt] & = \frac{\Gamma(\beta+1)}{\Gamma\big(\frac{\beta}{\alpha}+1\big)} \frac{x^{\frac{\beta}{\alpha}+\frac{1}{1-\alpha}-1}}{\pi(1-\alpha)} \int_0^\pi a(\varphi) \exp\Big(-a(\varphi)x^{\frac{1}{1-\alpha}}\Big) \,{\rm d}\varphi\nonumber\\[3pt] & = \int_0^\pi w_\varphi(x)\,h_{\alpha,\beta}(\varphi) \,{\rm d}\varphi, \qquad x>0, \end{align}

with $w_\varphi(x)$ as defined above and $h_{\alpha,\beta}(\varphi)$ defined via (43). From $\int_0^\infty w_\varphi(x)\,{\rm d}x=1$ we conclude, by Fubini’s theorem, that $\int_0^\pi h_{\alpha,\beta}(\varphi)\,{\rm d}\varphi =\int_0^\infty \int_0^\pi w_\varphi(x)\,h_{\alpha,\beta}(\varphi) \,{\rm d}\varphi\,{\rm d}x=\int_0^\infty g_{\alpha,\beta}(x)\,{\rm d}x=1$ , so $h_{\alpha,\beta}$ is a density of some $(0,\pi)$ -valued random variable V, which can be chosen such that it is independent of G. Therefore, $(G/a(V))^{1-\alpha}$ has distribution function $x\mapsto\mathbb{P}((G/a(V))^{1-\alpha}\le x) =\int_0^\pi \mathbb{P}((G/a(\varphi))^{1-\alpha}\le x) \,h_{\alpha,\beta}(\varphi)\,{\rm d}\varphi =\int_0^\pi \mathbb{P}(W_\varphi\le x) \,h_{\alpha,\beta}(\varphi)\,{\rm d}\varphi$ and hence density $x\mapsto\int_0^\pi w_\varphi(x)\,h_{\alpha,\beta}(\varphi)\,{\rm d}\varphi$ , $x>0$ , which coincides with the density (44) of L.

Remark 22. Only particular functionals of V are explicitly known. For example, from (42) it follows that the reciprocal of $a_\alpha(V)$ has moments

\begin{align*} \mathbb{E}((a_\alpha(V))^{-q}) & = \frac{\mathbb{E}\big(L^{\frac{q}{1-\alpha}}\big)}{\mathbb{E}(G^q)}\\[3pt] & = \frac{\Gamma(\beta)\Gamma\big(\frac{q}{1-\alpha}+\frac{\beta}{\alpha}\big)} {\Gamma\big(\frac{\beta}{\alpha}\big)\Gamma\big(\alpha\frac{q}{1-\alpha}+\beta\big)} \frac{\Gamma\big(\frac{\beta}{\alpha}(1-\alpha)+1\big)}{\Gamma\big(\frac{\beta}{\alpha}(1-\alpha)+q+1\big)}, \qquad q>-\frac{\beta}{\alpha}(1-\alpha). \end{align*}

For particular parameter values the density (43) of $V=V_{\alpha,\beta}$ further simplifies. For example, for $\alpha=\frac{1}{2}$ , (36) reduces to $a_{\frac{1}{2}}(\varphi)=\big(\sin\big(\frac{\varphi}{2}\big)/\sin(\varphi)\big)^2 =1/\big(2\cos\big(\frac{\varphi}{2}\big)\big)^2$ and Legendre’s duplication formula $\Gamma\big(\frac{x}{2}\big)\Gamma\big(\frac{x+1}{2}\big)=2^{1-x}\sqrt{\pi}\Gamma(x)$ , applied with $x\,:\!=\,2\beta+1$ , leads to

\begin{equation*} h_{\frac{1}{2},\beta}(\varphi) = \frac{\cos^{2\beta}\big(\frac{\varphi}{2}\big)} {{\rm B}\big(\beta+\frac{1}{2},\frac{1}{2}\big)}, \qquad 0<\varphi<\pi. \end{equation*}

It follows that

(45) \begin{equation} V_{\frac{1}{2},\beta} \stackrel{\textrm{d}}{=} 2\arccos\sqrt{B_\beta}, \qquad a_{\frac{1}{2}}\big(V_{\frac{1}{2},\beta}\big) \stackrel{\textrm{d}}{=} \frac{1}{4B_\beta} , \qquad L_{\frac{1}{2},\beta} \stackrel{\textrm{d}}{=} 2\sqrt{G_\beta B_\beta}, \end{equation}

where $B_\beta$ is beta distributed with parameters $\beta+\frac{1}{2}$ and $\frac{1}{2}$ , and $G_\beta$ is independent of $B_\beta$ and gamma distributed with shape parameter $\beta+1$ and scale parameter 1 having density $x\mapsto x^\beta \textrm{e}^{-x}/\Gamma(\beta+1)$ , $x>0$ . The last formula in (45) yields an efficient random number generator for ${\rm ML}(\frac{1}{2},\beta)$ . Proposition 4 also holds for $\beta=0$ . In this case V is uniformly distributed on $(0,\pi)$ , G is standard exponentially distributed, and (42) reduces to (35). From the symmetry relation (37) it follows that $h_{\alpha,\beta}=h_{\tilde\alpha,\tilde\beta}$ with $\tilde\alpha\,:\!=\,1-\alpha$ and $\tilde\beta\,:\!=\,\beta(1-\alpha)/\alpha$ .

Remark 23. (Random number generator) The representation (42) yields a sampling method for ${\rm ML}(\alpha,\beta)$ . Generators for the gamma distribution are well known (see, for example, [Reference Devroye6, Section IX 3.2]). Sampling from V can be implemented by the following rejection technique. Since $a(\!\cdot\!)$ is strictly increasing on $(0,\pi)$ it follows that the density (43) of V is strictly decreasing and hence bounded by $h_{\alpha,\beta}(0+)=c_{\alpha,\beta} (a(0+))^{-\frac{\beta}{\alpha}(1-\alpha)}<\infty$ . The constant $a(0+)=(1-\alpha)\alpha^{\alpha/(1-\alpha)}$ is known. A simple rejection method (see [Reference Devroye6, p. 42]) to sample from V therefore works via Algorithm 1.

The normalizing constant $c_{\alpha,\beta}$ is not involved in this algorithm. Note that the inequality $u\le (a(0+)/a(v))^{\beta(1-\alpha)/\alpha}$ is equivalent to the standard acceptance condition $uh_{\alpha,\beta}(0+)\le h_{\alpha,\beta}(v)$ . The efficiency of Algorithm 1 is discussed in Remark 24. A more efficient rejection method (Algorithm 2) is provided afterwards.

Our next result concerns the asymptotic behavior of $V=V_{\alpha,\beta}$ as $\alpha\to 0$ and as $\beta\to\infty$ respectively. It is remarkable that $V_{\alpha,\beta}$ has a nondegenerate limit as $\alpha\to 0$ .

Proposition 5.

  1. (a) As $\alpha\to 0$ the density $h_{\alpha,\beta}$ of $V_{\alpha,\beta}$ converges uniformly on $(0,\pi)$ to $h_{0,\beta}$ defined via

    (46) \begin{equation} h_{0,\beta}(\varphi) \,:\!=\, \frac{\Gamma(\beta+1)}{\pi} \bigg(\frac{\sin\varphi}{\beta\varphi}\bigg)^\beta \exp(\beta\varphi\cot\varphi), \qquad 0<\varphi<\pi. \end{equation}

    In particular, $V_{\alpha,\beta}\to V_{0,\beta}$ in distribution as $\alpha\to 0$ , where $V_{0,\beta}$ has density (46).

  2. (b) As $\beta\to\infty$ the map $x\mapsto\beta^{-1/2} h_{\alpha,\beta}(\beta^{-1/2}x)$ converges uniformly on $(0,\infty)$ to $h_\alpha$ defined via

    (47) \begin{equation} h_\alpha(x) \,:\!=\, \sqrt{\frac{2(1-\alpha)}{\pi}} \exp\bigg(-\frac{1-\alpha}{2}x^2\bigg),\qquad x>0. \end{equation}
    In particular, $\sqrt{\beta}V_{\alpha,\beta}\to V_\alpha$ in distribution as $\beta\to\infty$ , where $V_\alpha$ is half-normal distributed with density (47).

Proof.

  1. (a) An application of Stirling’s formula $\Gamma(x+1)\sim (x/\textrm{e})^x\sqrt{2\pi x}$ , $x\to\infty$ , shows that the normalizing constants $c_{\alpha,\beta}$ satisfy $c_{\alpha,\beta}\sim\pi^{-1}\Gamma(\beta+1)(\alpha/\beta)^\beta$ as $\alpha\to 0$ . Fix $0<\varphi<\pi$ . From

    \begin{align*} \bigg(\frac{\sin((1-\alpha)\varphi)}{\sin\varphi}\bigg)^{\frac{1}{\alpha}} & = ( \cos(\alpha\varphi)-\sin(\alpha\varphi)\cot\varphi )^{\frac{1}{\alpha}}\\[3pt] & = ( 1 - \alpha\varphi\cot\varphi + O(\alpha^2) )^{\frac{1}{\alpha}} \to \exp(-\varphi\cot\varphi) \end{align*}
    as $\alpha\to 0$ we conclude that
    \begin{equation*} (a_\alpha(\varphi))^{\frac{1}{\alpha}} = \bigg(\frac{\sin(\alpha\varphi)}{\sin\varphi}\bigg)^{1-\alpha} \bigg(\frac{\sin((1-\alpha)\varphi)}{\sin\varphi}\bigg)^{\frac{1}{\alpha}} \sim \frac{\alpha\varphi}{\sin\varphi}\exp(-\varphi\cot\varphi) \end{equation*}
    as $\alpha\to 0$ . Putting these formulas together shows that
    \begin{align*} h_{\alpha,\beta}(\varphi) & = c_{\alpha,\beta} \bigg(\frac{1}{(a_\alpha(\varphi))^{\frac{1}{\alpha}}}\bigg)^{(1-\alpha)\beta}\\[3pt] & \to \frac{\Gamma(\beta+1)}{\pi} \bigg(\frac{\sin\varphi}{\beta\varphi}\bigg)^\beta \exp(\beta\varphi\cot\varphi) = h_{0,\beta}(\varphi),\qquad \alpha\to 0. \end{align*}
    This pointwise convergence holds uniformly on $(0,\pi)$ , since all involved functions $h_{\alpha,\beta}$ and $h_{0,\beta}$ can be continuously extended to $[0,\pi]$ via $h_{\alpha,\beta}(\pi)\,:\!=\,0$ , $h_{0,\beta}(\pi)\,:\!=\,0$ , $h_{\alpha,\beta}(0)\,:\!=\,c_{\alpha,\beta}(a_\alpha(0+))^{\beta(1-\alpha)/\alpha}$ and $h_{0,\beta}(0)\,:\!=\,\pi^{-1}\Gamma(\beta+1)(\textrm{e}/\beta)^\beta$ and all these functions are decreasing, continuous, and bounded on $[0,\pi]$ . In particular $\int_0^\pi h_{0,\beta}(\varphi)\,{\rm d}\varphi=\lim_{\alpha \to 0}\int_0^\pi h_{\alpha,\beta}(\varphi)\,{\rm d}\varphi=1$ , so $h_{0,\beta}$ is a density of some $(0,\pi)$ -valued random variable $V_{0,\beta}$ . Clearly, the uniform convergence of the densities implies the convergence $V_{\alpha,\beta}\to V_{0,\beta}$ in distribution as $\alpha\to 0$ .
  2. (b) An application of Stirling’s formula yields $c_{\alpha,\beta}\sim\sqrt{2(1-\alpha)\beta/\pi}(a_\alpha(0+))^{\frac{\beta}{\alpha}(1-\alpha)}$ as $\beta\to\infty$ . Straightforward calculations show that the function $a_\alpha(\!\cdot\!)$ in (36) has Taylor expansion $a_\alpha(\varphi)=a_\alpha(0+)(1+\frac{\alpha}{2}\varphi^2+O(\varphi^4))$ . Thus, for all $x>0$ and all sufficiently large $\beta$ ,

    \begin{align*} \frac{1}{\sqrt{\beta}}h_{\alpha,\beta}\bigg(\frac{x}{\sqrt{\beta}\,}\bigg) & = \frac{c_{\alpha,\beta}}{\sqrt{\beta}\,} \frac{1}{\big(a_\alpha\big(\frac{x}{\sqrt{\beta}}\big)\big)^{\frac{\beta}{\alpha}(1-\alpha)}} \sim \sqrt{\frac{2(1-\alpha)}{\pi}} \frac{1}{\big(1+\frac{\alpha}{2}\frac{x^2}{\beta}+O(x^4)\big)^{\frac{\beta}{\alpha}(1-\alpha)}}\\[3pt] & \to \sqrt{\frac{2(1-\alpha)}{\pi}}\exp\bigg(-\frac{1-\alpha}{2}x^2\bigg)\ =\ h_\alpha(x), \qquad \beta\to\infty. \end{align*}

This convergence holds uniformly for all $x>0$ since all involved functions are continuous, monotone decreasing, and bounded. In particular, $\sqrt{\beta}V_{\alpha,\beta}\to V_\alpha$ in distribution as $\beta\to\infty$ , where $V_\alpha$ is half-normal distributed with density (47).

Remark 24. (Efficiency of Algorithm 1) Let N denote the number of steps the rejection algorithm for V (Algorithm 1) needs to terminate. Clearly, $\mathbb{P}(N=n)=p(1-p)^{n-1}$ , $n\in\mathbb{N}$ , where $p\,:\!=\,1/(\pi h_{\alpha,\beta}(0+))$ denotes the acceptance probability. The rejection algorithm therefore needs, on average,

\begin{align*} \mathbb{E}(N) & = \frac{1}{p} = \pi h_{\alpha,\beta}(0+) = \frac{\pi c_{\alpha,\beta}} {(a_\alpha(0+))^{\frac{\beta}{\alpha}(1-\alpha)}}\\[3pt] & = \frac{\Gamma(\beta+1)\Gamma\big(\frac{\beta}{\alpha}(1-\alpha)+1\big)} {\Gamma\big(\frac{\beta}{\alpha}+1\big)} \big(\alpha(1-\alpha)^{\frac{1-\alpha}{\alpha}}\big)^{-\beta} \end{align*}

steps to terminate. An application of Stirling’s formula yields $\mathbb{E}(N)\sim\sqrt{2\pi(1-\alpha)\beta}$ as $\beta\to\infty$ . Moreover, it can be checked that $\mathbb{E}(N)$ is decreasing in $\alpha\in (0,1)$ . Hence, by Proposition 5(a), we obtain, uniformly for all $\alpha\in (0,1)$ , the upper bound

\begin{equation*} \mathbb{E}(N) \le \pi\lim_{\alpha\to 0}h_{\alpha,\beta}(0+) = \pi h_\beta(0+) = \Gamma(\beta+1)\bigg(\frac{e}{\beta}\bigg)^\beta \sim \sqrt{2\pi\beta},\qquad \beta\to\infty. \end{equation*}

For example, for $\beta=1$ the rejection algorithm needs (independently of $\alpha$ ) on average not more than $\textrm{e} \approx 2.71828$ steps to terminate. For large values of $\beta$ the algorithm becomes less efficient and the average number of steps to terminate grows like $\sqrt{2\pi(1-\alpha)\beta}$ as $\beta\to\infty$ . In summary, Algorithm 1, the rejection algorithm for V, is sufficiently efficient as long as $\beta$ is not too large. An alternative rejection method (Algorithm 2) is provided which is highly efficient for all parameter values $0<\alpha<1$ and $\beta>0$ . For very large values of $\beta$ one may alternatively approximate the distribution of V using the half-normal limit from Proposition 5(b).

Remark 25. As a side effect, Proposition 5(a) implies that the reciprocal gamma function has the integral representation

(48) \begin{equation} \frac{\beta^\beta}{\Gamma(\beta+1)} = \frac{1}{\pi}\int_0^\pi \bigg(\frac{\sin\varphi}{\varphi}\bigg)^\beta \exp(\beta\varphi\cot\varphi) \,{\rm d}\varphi,\qquad \beta>0. \end{equation}

By the identity theorem for holomorphic functions, (48) even holds for all $\beta\in\mathbb{H}\,:\!=\,\{z\in\mathbb{C}:{\rm Re}(z)>0\}$ . Note that both expressions on the left- and right-hand sides in (48) are well-defined holomorphic functions in $\beta\in\mathbb{H}$ . For the left-hand side this is clear since the map $\beta\mapsto\beta^\beta$ and $\Gamma(\!\cdot\!)$ are holomorphic on $\mathbb{H}$ . By the differentiation lemma the right-hand side in (48) is also complex differentiable in any point $\beta\in\mathbb{H}$ , and it is allowed to take the derivative below the integral.

Temme ([Reference Temme38, p. 4, Eq. (2.1.8)], [Reference Temme39, p. 71, Eq. (3.38)]) derives (48), in a slightly different form, via an appropriate contour integration of the reciprocal gamma function and focuses on its usefulness for numerical computations of the gamma function (see also [Reference Schmelzer and Trefethen33]). Despite its elegance and usefulness, (48) does not seem to have achieved much attention in the literature. For one further application of (48) we refer the reader to Example 11.16 of [Reference Steutel and Van Harn37].

We have seen that Algorithm 1 is not very efficient for large values of $\beta$ . The following more sophisticated rejection algorithm for V will turn out to be efficient for all parameters $0<\alpha<1$ and $\beta>0$ . Assume first that $\alpha\le\frac{1}{2}$ . In this case it can be checked that $h_{\alpha,\beta}(\varphi) \le d_{\alpha,\beta}h_{\frac{1}{2},\beta}(\varphi)$ , $ 0<\varphi<\pi$ , $\beta>0$ , with constant

\begin{equation*} d_{\alpha,\beta} \,:\!=\, \frac{h_{\alpha,\beta}(0+)}{h_{\frac{1}{2},\beta}(0+)} = \frac{c_{\alpha,\beta}{\rm B}\big(\beta+\frac{1}{2},\frac{1}{2}\big)} {(a_\alpha(0+))^{\frac{\beta}{\alpha}(1-\alpha)}} = \frac{\Gamma\big(\beta+\frac{1}{2}\big)\Gamma\big(\frac{\beta}{\alpha}(1-\alpha)+1\big)} {\sqrt{\pi}\Gamma\big(\frac{\beta}{\alpha}+1\big)\alpha^\beta(1-\alpha)^{\frac{\beta}{\alpha}(1-\alpha)}}.\end{equation*}

Thus, for $\alpha\le\frac{1}{2}$ and $\beta>0$ , an alternative rejection algorithm for V works via Algorithm 2 (see [Reference Devroye6, p. 42]).

Sampling from $V_{\frac{1}{2},\beta}$ is straightforward by the first formula in (45). The value t simplifies to

\begin{align*} t = \frac{h_{\alpha,\beta}(0+)}{h_{\frac{1}{2},\beta}(0+)} \frac{h_{\frac{1}{2},\beta}(v)}{h_{\alpha,\beta}(v)} & = \bigg( \frac{a_{\frac{1}{2}}(0+)}{a_{\frac{1}{2}}(v)} \bigg)^\beta \bigg( \frac{a_\alpha(v)}{a_\alpha(0+)} \bigg)^{\frac{\beta}{\alpha}(1-\alpha)}\\[3pt] & = \cos^{2\beta}\bigg(\frac{v}{2}\bigg) \bigg(\frac{a_\alpha(v)}{a_\alpha(0+)}\bigg)^{\frac{\beta}{\alpha}(1-\alpha)}\end{align*}

and can hence be computed without much effort. The constants $c_{\alpha,\beta}$ and $d_{\alpha,\beta}$ are not needed during the calculations. Algorithm 2 takes (see [Reference Devroye6, p. 42]) on average $\mathbb{E}(N)=d_{\alpha,\beta}$ steps to terminate. Since $h_{\alpha,\beta}(0+)$ and hence $d_{\alpha,\beta}$ is decreasing in $\alpha$ it follows that

\begin{align*} \mathbb{E}(N) & = \frac{h_{\alpha,\beta}(0+)}{h_{\frac{1}{2},\beta}(0+)} \le \frac{h_{0,\beta}(0+)}{h_{\frac{1}{2},\beta}(0+)}\\[3pt] & = \frac{1}{\pi}\Gamma(\beta+1)\bigg(\frac{\textrm{e}}{\beta}\bigg)^\beta {\rm B}\big(\beta+\mbox{$\frac{1}{2}$},\mbox{$\frac{1}{2}$}\big) = \frac{1}{\sqrt{\pi}\,}\bigg(\frac{\textrm{e}}{\beta}\bigg)^\beta \Gamma\big(\beta+\mbox{$\frac{1}{2}$}\big).\end{align*}

This expression is increasing in $\beta$ and bounded with limit $\sqrt{2}$ as $\beta\to\infty$ , leading to $\mathbb{E}(N)\le \sqrt{2}$ uniformly for all $\alpha\le\frac{1}{2}$ and $\beta>0$ . Algorithm 2 therefore takes on average never more than $\sqrt{2}$ steps to terminate and is hence quite efficient. The case $\alpha>\frac{1}{2}$ can be transformed in advance into $\alpha<\frac{1}{2}$ using the relation $h_{\alpha,\beta}=h_{\tilde\alpha,\tilde\beta}$ with $\tilde\alpha\,:\!=\,1-\alpha$ and $\tilde\beta\,:\!=\,\beta(1-\alpha)/\alpha$ .

Acknowledgements

The author would like to thank two anonymous referees for their constructive reports leading to a significant improvement of the article. He furthermore thanks Helmut Pitters for fruitful discussions and Fernando Cordero for pointing out the integral representation (3.38) on p. 71 of [Reference Temme39].

References

Aldous, D. J. (1985). Exchangeability and Related Topics (Lect. Notes Math. 1117). Springer, New York.10.1007/BFb0099421CrossRefGoogle Scholar
Antoniak, C. E. (1974). Mixtures of Dirichlet processes with applications to Bayesian nonparametric problems. Ann. Statist. 2, 11521174.10.1214/aos/1176342871CrossRefGoogle Scholar
Billingsley, P. (1995). Probability and Measure, 3rd edn. Wiley, New York.Google Scholar
Blackwell, D. and MacQueen, J. B. (1973). Ferguson distributions via Pólya urn schemes. Ann. Statist. 1, 353355.10.1214/aos/1176342372CrossRefGoogle Scholar
Chambers, J. M., Mallows, C. L. and Stuck, B. W. (1976). A method for simulating stable random variables. J. Am. Stat. Assoc. 71, 340344.10.1080/01621459.1976.10480344CrossRefGoogle Scholar
Devroye, L. (1986). Non-Uniform Random Variate Generation. Springer, New York.10.1007/978-1-4613-8643-8CrossRefGoogle Scholar
Donnelly, P. J. (1986). Partition structures, Pólya urns, the Ewens sampling formula and the ages of alleles. Theor. Popul. Biol. 30, 271288.10.1016/0040-5809(86)90037-7CrossRefGoogle Scholar
Ewens, W. J. (1972). The sampling theory of selectively neutral alleles. Theor. Popul. Biol. 3, 87112.10.1016/0040-5809(72)90035-4CrossRefGoogle ScholarPubMed
Feller, W. (1971). An Introduction to Probability and Its Applications, Vol. II. Wiley, New York.Google Scholar
Flajolet, P., Dumas, P. and Puyhaubert, V. (2006). Some exactly solvable models of urn process theory. Discrete Math. Theor. Comp. Sci. AG, 59118.Google Scholar
Green, P. J. (2010). Colouring and breaking sticks: Random distributions and heterogeneous clustering. In Probability and Mathematical Genetics: Papers in Honour of Sir John Kingman, eds. Bingham, N. H. and Goldie, C. M.. Cambridge University Press, pp. 319344.10.1017/CBO9781139107174.015CrossRefGoogle Scholar
Hoppe, F. M. (1984). Pólya-like urns and the Ewens sampling formula. J. Math. Biol. 20, 9194.10.1007/BF00275863CrossRefGoogle Scholar
Hoppe, F. M. (1987). The sampling theory of neutral alleles and an urn model in population genetics. J. Math. Biol. 25, 123159.10.1007/BF00276386CrossRefGoogle Scholar
Hsu, L. C. and Shiue, P. J.-S. (1998). A unified approach to generalized Stirling numbers. Adv. Appl. Math. 20, 366384.10.1006/aama.1998.0586CrossRefGoogle Scholar
Ibragimov, I. A. and Chernin, K. E. (1959). On the unimodality of stable laws. Teor. Veroyatnost. i Primenen 4, 453–456. English translation: Theor. Prob. Appl. 4, 417419 (1959).10.1137/1104043CrossRefGoogle Scholar
James, L. F. (2002). Poisson process partition calculus with applications to exchangeable models and Bayesian nonparametrics. arXiv:math/0205093.Google Scholar
James, L. F. (2005). Bayesian Poisson process partition calculus with an application to Bayesian Lévy moving averages. Ann. Statist. 33, 17711799.10.1214/009053605000000336CrossRefGoogle Scholar
Janson, S. (2004). Functional limit theorems for multitype branching processes and generalized Pólya urns. Stoch. Proc. Appl. 110, 177245.10.1016/j.spa.2003.12.002CrossRefGoogle Scholar
Janson, S. (2006). Limit theorems for triangular urn schemes. Prob. Theory Relat. Fields 134, 417452.10.1007/s00440-005-0442-7CrossRefGoogle Scholar
Kallenberg, O. (2002). Foundations of Modern Probability, 2nd edn. Springer, New York.10.1007/978-1-4757-4015-8CrossRefGoogle Scholar
Kanter, M. (1975). Stable densities under change of scale and total variation inequalities. Ann. Prob. 3, 697707.10.1214/aop/1176996309CrossRefGoogle Scholar
Kingman, J. F. C. (1978). Random partitions in population genetics. Proc. R. Soc. London A, 361, 120.Google Scholar
Lo, A. Y., Brunner, L. J. and Chan, A. T. (1996). Weighted Chinese restaurant processes and Bayesian mixture models. Research report, Hong Kong University of Science and Technology. Available at http://www.utstat.utoronto.ca/~brunner/papers/wcr96.pdf. Google Scholar
Mikusiński, J. (1959). On the function whose Laplace transform is $\textrm{e}^{-s^a}$ . Studia Math. 18, 191198.10.4064/sm-18-2-191-198CrossRefGoogle Scholar
Möhle, M. (2006). On sampling distributions for coalescent processes with simultaneous multiple collisions. Bernoulli 12, 3553.Google Scholar
Petrov, V. V. (1975). Sums of Independent Random Variables. Springer, Berlin.10.1007/978-3-642-65809-9CrossRefGoogle Scholar
Pitman, J. (1995). Exchangeable and partially exchangeable random partitions. Prob. Theory Relat. Fields 102, 145158.10.1007/BF01213386CrossRefGoogle Scholar
Pitman, J. (1996). Random discrete distributions invariant under size-biased permutation. Adv. Appl. Prob. 28, 525539.10.2307/1428070CrossRefGoogle Scholar
Pitman, J. (2003). Poisson–Kingman partitions. In Statistics and Science: A Festschrift for Terry Speed, eds. Speed, T. P. and Goldstein, D. R. (IMS Lect. Notes Monogr. Ser. 40). Institute of Mathematical Statistics, Beechwood, OH, pp. 134.10.1214/lnms/1215091133CrossRefGoogle Scholar
Pitman, J. (2006). Combinatorial Stochastic Processes (Lect. Notes Math. 1875). Springer, Berlin.Google Scholar
Pitman, J. and Yor, M. (1997). The two-parameter Poisson–Dirichlet distribution derived from a stable subordinator. Ann. Prob. 25, 855900.10.1214/aop/1024404422CrossRefGoogle Scholar
Prabhakar, T. R. (1971). A singular integral equation with a generalized Mittag–Leffler function in the kernel. Yokohama Math. J. 19, 715.Google Scholar
Schmelzer, T. and Trefethen, L. N. (2007). Computing the gamma function using contour integrals and related approximations. SIAM J. Numer. Anal. 45, 558571.10.1137/050646342CrossRefGoogle Scholar
Siegmund, D. (1976). The equivalence of absorbing and reflecting barrier problems for stochastically monotone Markov processes. Ann. Prob. 4, 914924.10.1214/aop/1176995936CrossRefGoogle Scholar
Stanković, B. (1954). Sur une fonction du calcul opérationnel. Acad. Serbe Sci. Pub. Inst. Math. 6, 7578.Google Scholar
Stanković, B. (1970). On the function of E. M. Wright. Pub. Inst. Math. Beograd N.S. 10, 113124.Google Scholar
Steutel, F. W. and Van Harn, K. (2004). Infinite Divisibility of Probability Distributions on the Real Line. Marcel Dekker Inc., New York.Google Scholar
Temme, N. M. (1978). The Numerical Computation of Special Functions by Use of Quadrature Rules for Saddle Point Integrals. II. Gamma Functions, Modified Bessel Functions and Parabolic Cylinder Functions. Mathematisch Centrum, Afdeling Toegepaste Wiskunde, Amsterdam.Google Scholar
Temme, N. M. (1996). Special Functions. Wiley, New York.10.1002/9781118032572CrossRefGoogle Scholar
Tomovski, Ž., Pogány, T. K. and Srivastava, H. M. (2014). Laplace type integral expressions for a certain three-parameter family of generalized Mittag-Leffler functions with applications involving complete monotonicity. J. Franklin Inst. 351, 54375454.10.1016/j.jfranklin.2014.09.007CrossRefGoogle Scholar
Trieb, G. (1992). A Pólya urn model and the coalescent. J. Appl. Prob. 29, 110.10.2307/3214786CrossRefGoogle Scholar
Zhang, P., Chen, C. and Mahmoud, H. (2015). Explicit characterization of moments of balanced triangular Pólya urns by an elementary approach. Statist. Prob. Lett. 96, 149153.10.1016/j.spl.2014.09.016CrossRefGoogle Scholar