Hostname: page-component-78c5997874-lj6df Total loading time: 0 Render date: 2024-11-11T00:56:25.766Z Has data issue: false hasContentIssue false

Extended generator and associated martingales for M/G/1 retrial queue with classical retrial policy and general retrial times

Published online by Cambridge University Press:  11 January 2022

S. Meziani
Affiliation:
Department of Probability and Statistics, Faculty of Mathematics, University of Sciences and Technology USTHB, Algiers, Algeria. E-mail: smeziani@usthb.dz
T. Kernane
Affiliation:
Department of Probability and Statistics, Faculty of Mathematics, University of Sciences and Technology USTHB, Algiers, Algeria. E-mail: smeziani@usthb.dz Laboratory of Research in Intelligent Informatics and Applied Mathematics (RIIMA), University of Sciences and Technology USTHB, Algiers, Algeria. E-mail: tkernane@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

A retrial queue with classical retrial policy, where each blocked customer in the orbit retries for service, and general retrial times is modeled by a piecewise deterministic Markov process (PDMP). From the extended generator of the PDMP of the retrial queue, we derive the associated martingales. These results are used to derive the conditional expected number of customers in the orbit in the transient regime.

Type
Research Article
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press

1. Introduction

Retrial queueing models are specified by the following feature: an external arriving customer which finds all the servers busy joins a virtual waiting area, called orbit, instead of leaving the system and becomes a blocked customer. The blocked customers in the orbit are allowed to retry their luck to capture the service area after a random amount of time. Time intervals between these attempts are called retrial times or repeated times. Queueing systems with repeated times are regarded as a special part of queueing theory, which has a wide field of possible applications such as computer and communication networks. Telephone exchanges in a call center are a well-known application of retrial queues in the literature. This problem was tackled by Fayolle [Reference Fayolle14] in the case of a constant type of retrial policy where the recall rate is independent of the number of blocked customers. Fayolle [Reference Fayolle14] has analyzed the stationary distribution of the system states and the sojourn time distribution for an M/M/1 queue with a constant retrial policy. This analysis was extended to the case of an M/G/1 queue in Farahmand [Reference Farahmand13], as well as the case of a G/M/1 system in Lillo [Reference Lillo19]. In the work of Kim et al. [Reference Kim, Klimenok and Dudin18], the analysis of the G/M/1 system became simpler compared to Lillo [Reference Lillo19] by using the matrix analytic technique. Alternatively, blocked customers may also attempt to get served at the same time. This refers to the classical retrial policy which states that each customer in the orbit tries to capture the server independently of the other customers. Thus, the retrial rate depends on the number of blocked customers. Retrial queues with classical retrial policy were extensively studied in the literature [Reference Falin11,Reference Falin and Templeton12] .

All of the above-quoted works have supposed the exponential distribution for repeated time. Meanwhile, in our study, we consider general retrial time distribution. Research in this direction is so limited especially when dealing with the case of a blocked customer who acts independently of the other customers in the orbit. The first attempt to generalize the retrial time distribution was done by Kapyrin [Reference Kapyrin16] for a classical retrial policy, but Falin [Reference Falin10] has shown later that his approach was incorrect. In Kernane [Reference Kernane17], the stability condition was determined for an M/G/1 retrial queue with general retrial times and classical retrial policy by assuming also that the process of service times is stationary and ergodic. In the references [Reference Chakravarthy2,Reference Shin and Moon20,Reference Shin and Moon21] , only some approximations and simulations are presented for models with phase-type retrial time. Therefore, dealing with general retrial times in the case of classical retrial policy is a challenging research problem in retrial queueing theory.

In this paper, we analyze the M/G/1 retrial queue as a piecewise deterministic Markov process (PDMP) instead of studying it using the traditional methods, such as the embedded Markov chain and the supplementary variable methods. This new approach makes it possible to study the dynamics of such a complicated model in greater depth as well as to derive the performance quantities, by means of martingales, in a transient regime which is difficult to investigate even for simple Markovian queues. In fact, PDMP modelization has been introduced to analyze several models in the literature including queueing and epidemic models [Reference Breuer1,Reference Clancy3,Reference Gómez-Corral and López-García15] . However, the similarity between the dynamics of the SIR model with general infectious period distribution and those of the M/G/1 queue with general retrial times, which is clearly outstanding in Gómez-Corral and López-García [Reference Gómez-Corral and López-García15], has most motivated us to carry out this work.

In Section 2, we recall the PDMP framework. In Section 3, we model the dynamics of the retrial queue with classical retrial policy and general retrial times by a PDMP. From the extended generator of the PDMP modeling the retrial queue, we derive the associated martingales in Section 4. In Section 5, we utilize these results to derive the conditional expected number of blocked customers in a transient regime. We end the paper with a conclusion in Section 6 giving some areas that can be studied for future works on the subject.

2. The PDMP framework

Piecewise deterministic Markov processes were introduced by Davis [Reference Davis7] as a general family of non-diffusion stochastic models. These Markov processes consist of a mixture of deterministic motion and random jumps. The class of PDMPs provides a framework for studying optimization problems especially in queueing systems [Reference Breuer1,Reference Davis8] . As shown in Davis [Reference Davis7] and then extensively indicated in Davis [Reference Davis8], a PDMP can be explicitly determined by means of three parameters $(\mathfrak {X},\lambda,Q)$. Let $X(t)$ be a PDMP in a state space $\mathbb {E}$ defined as follows. Let $K$ be a countable set and $d:K\longrightarrow \mathbb {N}$ be a given function. For each $\upsilon \in K$, $M_{\upsilon }$ is an open set of $\mathbb {R}^{d(\upsilon )}$. Then, $\mathbb {E}=\bigcup _{\upsilon \in K}M_{\upsilon }=\{(\upsilon,\xi ): \upsilon \in K, \xi \in M_{\upsilon }\}$. Denote by $\mathcal {E}$ the $\sigma$-algebra generated by the Borel subsets of $\mathbb {E}$. The state of the process will be denoted by $X(t)=(\upsilon (t),\xi (t))$ and the characteristics $\mathfrak {X}$, $\lambda$ and $Q$ are defined as follows:

  • $\mathfrak {X}=\{\mathfrak {X}_{\upsilon }, \upsilon \in K\}$ is a locally Lipschitz continuous vector fields in $\mathbb {E}$ with flow functions $\phi _{\upsilon }(t,\xi )$ for each $\xi \in M_{\upsilon }$.

  • $\lambda : \mathbb {E}\rightarrow \mathbb {R}_{+}$ is the jump rate. It is assumed that this function is measurable and for all $x=(\upsilon,\xi ) \in \mathbb {E}$, there exists $\varepsilon > 0$ such that $\int _{0}^{\varepsilon } \lambda (\phi _{\upsilon }(t,\xi )) \,dt$ exists.

  • $Q:(\mathbb {E} \cup \partial ^{*} \mathbb {E} )\times \mathcal {E} \rightarrow [0,1]$, with $\partial ^{*} \mathbb {E} =\bigcup _{\upsilon \in K}\partial ^{*}M_{\upsilon }$ where $\partial ^{*}M_{\upsilon }=\{\xi ^{\prime }\in \partial M_{\upsilon }: \phi _{\upsilon }(t,\xi )=\xi ^{\prime },\text { for all } (t,\xi )\in \mathbb {R}_{+}\times M_{\upsilon }\}$, a transition measure specifying the post-jump locations with $Q(x; \{x\})=0$. Note that $\partial M_{\upsilon }$ is the boundary of $M_{\upsilon }$ and $\partial ^{*} M_{\upsilon }$ represents those boundary points at which the flow exits from $M_{\upsilon }$.

With a convenient choice of the state space $\mathbb {E}$ and parameters $\mathfrak {X}$, $\lambda$ and $Q$ it is possible to model almost all non-diffusion processes found in the literature. Several important applications were presented in Davis [Reference Davis7,Reference Davis8] . The motion of the PDMP depends on the characteristics $\mathfrak {X}$, $\lambda$ and $Q$ in the following way. Starting from a point $x=(\upsilon,\xi )\in \mathbb {E}$, we select a jump time $T_{1}$ with distribution function $P_{x}(T_{1}>t)=\mathbf {I}_{t< t_{*}(x)}\exp \{-\int _{0}^{t}\lambda (\phi _{\upsilon }(s,\xi )) \,ds\}$, where $t_{*}(x)= \inf \{t \in \mathbb {R}_{+}: \phi _{\upsilon }(t,\xi )\in \partial ^{*} M_{\upsilon } \}$. The deterministic variable $t_{*}(x)$ denotes the time until the set $\partial ^{*} \mathbb {E}$ is reached from a state $x \in \mathbb {E}$. Now select a random variable $Z_{1}$ having distribution $Q(\phi _{\upsilon }(T_{1},\xi );\cdot )$. Hence, the trajectory of $X(t)$ for $t\leq T_{1}$ is given by

$$X(t)=\begin{cases} (\upsilon,\phi_{\upsilon}(t,\xi)) & \text{for}\ t< T_{1};\\ Z_{1} & \text{for}\ t= T_{1}. \end{cases}$$

Starting from $X(T_{1})$ we now select the next inter-jump time $T_{2}-T_{1}$ and the post-jump location $X(T_{2})$ in a similar way. This gives a piecewise deterministic trajectory $(X(t))$ with jump times $T_{1},T_{2},\ldots$ and post-jump locations $Z_{1},Z_{2},\ldots$. The sequence of random variables $\{Z_{n}\}$ is the Markov chain associated with the original process $X(t)$, so that previous results on the well-known discrete-time Markov chains were used to investigate the stability and ergodicity of PDMPs, see [Reference Costa4,Reference Costa and Dufour5,Reference Dufour and Costa9] . Furthermore, it is assumed that there are only finite many jumps of $X(t)$ in any finite time interval (see assumption 3.1 in Davis [Reference Davis7] or assumption 24.4 in Davis [Reference Davis8]). Thus, if $Z_{0}$ is the initial point then the associated Markov chain $\{Z_{n}, n\geq 0\}$ has the transition measure $\mathcal {P}: \mathbb {E} \cup \partial ^{*} \mathbb {E} \times \mathcal {E} \rightarrow [0,1]$ given by:

\begin{align*} \mathcal{P}(x; A)& =\int_{0}^{t_{*}(x)}Q(\phi_{\upsilon}(s,\xi);A)\lambda(\phi_{\upsilon}(s,\xi))\exp(-\Lambda(s,x))\,ds\\ & \quad +Q(\phi_{\upsilon}(t_{*}(x),\xi);A)\exp(-\Lambda(t_{*}(x),x)). \end{align*}

where $\Lambda (t,x)=\int _{0}^{t}\lambda (\phi _{\upsilon }(s,\xi ))\,ds$ for all $x\in \mathbb {E}$ and $0\leq t\leq t_{*}(x)$.

Note that $\Lambda (t_{*}(x),x)=\infty$ whenever $t_{*}(x)=\infty$.

3. M/G/1 queue with general retrial times as a PDMP

The model considered is an M/G/1 retrial queue. Customers arrive to the system according to a Poisson process with rate $\lambda$. Each incoming customer that finds the server busy joins the orbit. Service times are i.i.d. with the same general distribution function $F$. After a random time in the orbit, each blocked customer repeats his attempt to enter service. Retrial times are i.i.d. with the same general distribution function $G$. Inter-arrivals, service periods and retrial times are assumed to be mutually independent. The model studied can be represented as a PDMP in the following way. Let $(X(t)=(\upsilon (t),\xi (t));t\in \mathbb {R}_{+})$, where $\upsilon (t)=(C(t),N(t))$ and $\xi (t)=(Y(t),\mathbf {R}(t))$, be the PDMP describing the system state at time $t$. The component $C(t)$ represents the state of the server ($C(t)$ equals 1 or 0 according as the server is busy or free), $N(t)$ is the number of the blocked customers, $Y(t)$ is the residual service time and $\mathbf {R}(t)=(R_{1}(t),R_{2}(t),\ldots,R_{N(t)}(t))$ where $R_{k}(t)$ denotes the remaining retrial time of the $k$th blocked customer for $1\leq k \leq N(t)$. The considered process is defined on the state space $\mathbb {E}=(\mathbb {E}_{0}\cup \partial ^{*}\mathbb {E}_{0}) \cup (\mathbb {E}_{1}\cup \partial ^{*}\mathbb {E}_{1})$, where $\mathbb {E}_{0}=\{(0,n,0,r_{1},r_{2},\ldots,r_{n}),\ n\in \mathbb {N},\ 0< r_{1}< r_{2}<\cdots < r_{n}\}$, $\partial ^{*}\mathbb {E}_{0}=\{(0,n,0,0,r_{2},\ldots,r_{n}),\ n\in \mathbb {N}^{*},\ 0< r_{2}<\cdots < r_{n}\}$, $\mathbb {E}_{1}=\{(1,n,y,r_{1},r_{2},\ldots,r_{n}),\ n\in \mathbb {N},\ y>0,\ 0< r_{1}< r_{2}<\cdots < r_{n}\}$ and $\partial ^{*} \mathbb {E}_{1}=\{(1,n,0,r_{1},r_{2},\ldots,r_{n}),\ n\in \mathbb {N},\ 0< r_{1}< r_{2}<\cdots < r_{n}\}$. This particular formulation of the state space $\mathbb {E}$ allows us to specify the transition measure $Q(x;\cdot )$ of $X(t)$ effortlessly and one can see that it remains suitable to our case where all blocked customers retry to capture the service simultaneously.

Define the flow function $\phi$:

$$\phi(t,x)=\begin{cases} (0,n,0,r_{1}-t,r_{2}-t,\ldots,r_{n}-t) & \text{for}\ x\in \mathbb{E}_{0};\\ (1,n,y-t,r_{1}-t,r_{2}-t,\ldots,r_{n}-t) & \text{for}\ x\in \mathbb{E}_{1}. \end{cases}$$

We also give some notations for further use. Let $\mathcal {B}_{(a,b)}$ be the $\sigma$-algebra of Borel sets on the interval $(a,b)$ and $\beta _{n}$ be the Borel $\sigma$-algebra on the set $F^{(n)}=\{(u_{1},\ldots,u_{n})\in (0,\infty )^{n} : u_{1}<\cdots < u_{n} \}$. We define the following function as well for $A\in \beta _{n}$

$$\mathbf{1}_{A}(u_{1},\ldots,u_{n})=\begin{cases} 1 & \text{if}\ (u_{1},\ldots,u_{n})\in A ,\\ 0 & \text{otherwise}. \end{cases}$$

In the same spirit of Gómez-Corral and López-García [Reference Gómez-Corral and López-García15] and Breuer[Reference Breuer1], in order to describe the jumps that can occur more clearly, we are going to introduce three transition measures $Q_{1},\ Q_{2}$ and $Q_{3}$ which reflect the transitions associated with the arrival process, the service achievement and the removal of a blocked customer from the orbit, respectively.

For states $x\in \mathbb {E}_{0}$, the transition measure $Q_{1}(x;\cdot )$ is given by

$$Q_{1}(x;\{1\}\times \{n\}\times A \times B)= F(A)\mathbf{1}_{B}(r_{1},\ldots,r_{n}),$$

where $A\in \mathcal {B}_{(0,\infty )}$ and $B\in \beta _{n}$.

The transition measure $Q_{1}(x;\cdot )$ captures the transition from state $(0,n,0,r_{1},\ldots,r_{n})$ to state $(1,n,y,r_{1},\ldots,r_{n})$ related to a new external incoming customer to the queueing system that joins the idle server immediately.

For states $x\in \partial ^{*}\mathbb {E}_{0}$, $A\in \mathcal {B}_{(0,\infty )}$ and $B\in \beta _{n-1}$, the transition measure $Q_{3}(x;\cdot )$ is given by

$$Q_{3}(x;\{1\}\times \{n-1\}\times A \times B)= F(A)\mathbf{1}_{B}(r_{2},\ldots,r_{n}).$$

This refers to the transition from state $(0,n,0,0,r_{2},\ldots,r_{n})$ to state $(1,n-1,y,r_{1}^{\prime },\ldots,r_{n-1}^{\prime })$ occurring when a blocked customer joins the idle server.

In a similar manner, we determine the transition measures $Q_{1}(x;\cdot )$ and $Q_{2}(x;\cdot )$ for states $x\in \mathbb {E}_{1}$ and $x\in \partial ^{*} \mathbb {E}_{1}$, respectively. The transition measure $Q_{1}(x;\cdot )$ is specified as follows:

  1. (i) For $A\in \mathcal {B}_{(0,\infty )}$, $B\in \mathcal {B}_{ (0,r_{1})}$ and sets $C\in \beta _{n}$,

    $$Q_{1}(x;\{1\}\times \{n+1\}\times A \times B \times C)=\mathbf{1}_{A}(y)G(B)\mathbf{1}_{C}(r_{1},\ldots,r_{n}).$$
  2. (ii) For $A\in \mathcal {B}_{(0,\infty )}$, sets $B\in \beta _{k}$, $C \in \mathcal {B}_{(r_{k},r_{k+1})}$ and sets $D\in \beta _{n-k}$,

    $$Q_{1}(x;\{1\}\times \{n+1\}\times A \times B \times C \times D)=\mathbf{1}_{A}(y)\mathbf{1}_{B}(r_{1},\ldots,r_{k})G(C)\mathbf{1}_{D}(r_{k+1},\ldots,r_{n}).$$
  3. (iii) $A\in \mathcal {B}_{(0,\infty )}$, sets $B\in \beta _{n}$ and $C \in \mathcal {B}_{(r_{n},\infty )}$,

    $$Q_{1}(x;\{1\}\times \{n+1\}\times A \times B \times C)=\mathbf{1}_{A}(y)\mathbf{1}_{B}(r_{1},\ldots,r_{n})G(C).$$

The transition measure $Q_{1}(x;\cdot )$ captures the transition from state $(1,n,y,r_{1},\ldots,r_{n})$ to state $(1,n+1,y,r_{1}^{\prime },\ldots, r_{n}^{\prime },r_{n+1}^{\prime })$ resulting when an incoming customer finds the server busy, therefore it joins the orbit. Thus, it becomes a blocked customer and a new remaining retrial time generated from $G$ must be added to the vector $(r_{1},\ldots,r_{n})$ in its convenient place yielding to a new vector of remaining retrial times $(r_{1}^{\prime },\ldots, r_{n}^{\prime },r_{n+1}^{\prime })$ with $r_{1}^{\prime }<\cdots < r_{n}^{\prime }< r_{n+1}^{\prime }$.

Finally, the transition measure $Q_{2}(x;\cdot )$ associated with the transition from state $(1,n,0,r_{1},\ldots,r_{n})$ to state $(0,n,0,r_{1},\ldots,r_{n})$, when the server becomes idle, is given by

$$Q_{2}(x;\{0\}\times \{n\}\times\{0\} \times A)= \mathbf{1}_{A}(r_{1},\ldots,r_{n}),$$

where $A\in \beta _{n}$.

Note that, for our process, the jump rate is exactly the arrival rate $\lambda$. Hence, we have $\Lambda (t,x)=\int _{0}^{t}\lambda (\phi (s,x))\,ds=\lambda t$.

4. The associated martingales

In this section, we will derive the martingales associated with the PDMP that models our retrial queue using its extended generator.

Theorem 1. For $0\leq z_{1} \leq 1$, $0\leq z_{2}\leq 1$, $\gamma \geq 0$ and $\delta \geq 0$ , the function

(4.1)\begin{equation} z_{1}^{C(t)}z_{2}^{N(t)}\,e^{-\gamma Y(t)} \,e^{-\delta \sum_{k=1}^{N(t)}R_{k}(t)}\,e^{\theta_{C(t)}(t)} \end{equation}

with

(4.2)\begin{equation} \theta_{C(t)}(t)=\begin{cases} (\lambda-N(t)\delta)t-\ln\{\lambda z_{2}\varphi_{R}(\delta)+\gamma\}\\ \quad + \ln\{\lambda z_{1}\varphi_{S}(\delta)[e^{-(\lambda z_{2}\varphi_{R}(\delta)+\gamma)t}-1]+\lambda z_{2}\varphi_{R}(\delta)+\gamma\} & \text{for}\ C(t)=0;\\ (\lambda-\lambda z_{2}\varphi_{R}(\delta)-N(t)\delta-\gamma)t & \text{for}\ C(t)=1. \end{cases} \end{equation}

is a martingale for states $x \in \mathbb {E}_{C(t)}$, where

$$\varphi_{S}(\gamma)=\int_{0}^{\infty}e^{-\gamma y}\,dF(y)\quad \mathrm{and}\quad \varphi_{R}(\delta)=\int_{0}^{\infty}e^{-\delta r}\,dG(r).$$

Proof. The infinitesimal generator of the process $X(t)$, acting on a function $f(i,n,y,r_{1},\ldots,r_{n},t)\in \mathfrak {D}(\mathcal {G})$, is given by

(4.3)\begin{align} & \mathcal{G}f(0,n,0,r_{1},r_{2},\ldots,r_{n},t)\nonumber\\ & \quad =\lambda \int_{0}^{\infty}[f(1,n,y,r_{1},r_{2},\ldots,r_{n},t)-f(0,n,0,r_{1},r_{2},\ldots,r_{n},t)]\,dF(y)\nonumber\\ & \qquad - \sum_{k=1}^{n}\dfrac{\partial}{\partial r_{k}} f(0,n,0,r_{1},r_{2},\ldots,r_{n},t)+\dfrac{\partial}{\partial t}f(0,n,0,r_{1},r_{2},\ldots,r_{n},t).\end{align}
(4.4)\begin{align} & \mathcal{G}f(1,n,y,r_{1},r_{2},\ldots,r_{n},t)\nonumber\\ & \quad =\lambda \int_{0}^{\infty}[f(1,n+1,y,r_{1},r_{2},\ldots,r_{n},r_{n+1},t) -f(1,n,y,r_{1},r_{2},\ldots,r_{n},t)]\,dG(r_{n+1})\nonumber\\ & \qquad - \dfrac{\partial}{\partial y}f(1,n,y,r_{1},r_{2},\ldots,r_{n},t) - \sum_{k=1}^{n}\dfrac{\partial}{\partial r_{k}} f(1,n,y,r_{1},r_{2},\ldots,r_{n},t)\nonumber\\ & \qquad +\dfrac{\partial}{\partial t}f(1,n,y,r_{1},r_{2},\ldots,r_{n},t). \end{align}

where $\mathfrak {D}(\mathcal {G})$ is the domain of the generator $\mathcal {G}$ which consists of those functions $f(i,n,y,r_{1},\ldots,r_{n},t)$ that are differentiable with respect to $y,r_{1},\ldots,r_{n}$ and $t$ for all $i,n,y,r_{1},\ldots,r_{n},t$, and satisfy the boundary conditions derived from Eq. (5.4) in Davis [Reference Davis7]

(4.5)\begin{align} f(0,n,0,0,r_{2},\ldots,r_{n},t) & =\int_{0}^{\infty}f(1,n-1,y,r^{\prime}_{1},\ldots,r_{n-1}^{\prime},t)\,dF(y); \end{align}
(4.6)\begin{align} f(1,n,0,r_{1},\ldots, r_{n},t)& =f(0,n,0,r_{1},\ldots,r_{n},t); \end{align}

where $(r^{\prime }_{1},\ldots,r_{n-1}^{\prime })=(r_{2},\ldots,r_{n})$ and verify the integrability conditions

(4.7)\begin{align} & E\left\{\left|\int_{0}^{\infty}f(1,n,y,r_{1},\ldots,r_{n},t)\,dF(y) -f(0,n,0,r_{1},\ldots,r_{n},t)\right|\right\}<\infty; \end{align}
(4.8)\begin{align} & E\left\{\left|\int_{0}^{\infty}f(1,n+1,y,r_{1},\ldots,r_{n},r_{n+1},t)\,dG(r_{n+1}) -f(1,n,y,r_{1},\ldots,r_{n},t)\right|\right\}<\infty. \end{align}

Define now the function

(4.9)\begin{equation} f(i,n,y,r_{1},r_{2},\ldots,r_{n},t)=z_{1}^{i}z_{2}^{n}\,e^{-\gamma y}\,e^{-\delta \sum_{k=1}^{n}r_{k}}\,e^{\theta_{i}(t)}, \end{equation}

where $\theta _{i}(t)$ is given by Eq. (4.2) for $C(t)=i,\ i \in \{0,1\}$.

By substituting Eq. (4.9) in Eq. (4.3) for $i=0$, we get

\begin{align*} \mathcal{G}f(0,n,0,r_{1},r_{2},\ldots,r_{n},t)& =\lambda\int_{0}^{\infty}[z_{1}\,e^{-\gamma y}\,e^{\theta_{1}(t)-\theta_{0}(t)}-1]\,dF(y)+\sum_{k=1}^{n}\delta+\theta_{0}^{\prime}(t)\\ & =\lambda z_{1}\,e^{\theta_{1}(t)-\theta_{0}(t)}\varphi_{S}(\gamma)-\lambda+n\delta+\theta_{0}^{\prime}(t)\\ & =\lambda z_{1}\dfrac{(\lambda z_{2}\varphi_{R}(\delta)+\gamma)\,e^{-(\lambda z_{2}\varphi_{R}(\delta)+\gamma)}}{\lambda z_{1}\varphi_{S}(\gamma)[e^{-(\lambda z_{2}\varphi_{R}(\delta)+\gamma)}-1]+\lambda z_{2}\varphi_{R}(\delta)+\gamma}\varphi_{S}(\gamma)\\ & \quad -\lambda+n\delta+\lambda-n\delta\\ & \quad -\lambda z_{1}\varphi_{S}(\gamma)\dfrac{(\lambda z_{2}\varphi_{R}(\delta)+\gamma)\,e^{-(\lambda z_{2}\varphi_{R}(\delta)+\gamma)}}{\lambda z_{1}\varphi_{S}(\gamma)[e^{-(\lambda z_{2}\varphi_{R}(\delta)+\gamma)}-1]+\lambda z_{2}\varphi_{R}(\delta)+\gamma}\\ & =0. \end{align*}

Similarly, we substitute Eq. (4.9) in Eq. (4.4) for $i=1$.

\begin{align*} \mathcal{G}f(1,n,y,r_{1},r_{2},\ldots,r_{n},t)& =\lambda \int_{0}^{\infty}[z_{2}\,e^{-\delta r_{n+1}-1}]\,dG(r_{n+1})+\gamma+n\delta+\theta_{1}^{\prime}(t)\\ & =\lambda z_{2} \varphi_{R}(\delta)-\lambda+\gamma+n\delta+\theta_{1}^{\prime}(t)\\ & =\lambda z_{2} \varphi_{R}(\delta)-\lambda+\gamma+n\delta+\lambda -\lambda z_{2} \varphi_{R}(\delta)-n\delta-\gamma\\ & =0. \end{align*}

Hence, by the property of the infinitesimal generator, Eq. (4.1) is a martingale for the process $X(t)$.

Theorem 2. Let $\tau _{0}$ and $\tau _{1}$ be the stopping times that end the server inactivity period (when the process stays in $\mathbb {E}_{0}$) and the server occupation period (when the process stays in $\mathbb {E}_{1}$), respectively. The processes

(4.10)\begin{align} & f(C(t),N(t),Y(t),\mathbf{R}(t),t)-f(0,N(0),0,\mathbf{R}(0),0)\nonumber\\ & \quad -\int_{0}^{t}\mathcal{G}f(0,N(s),0,\mathbf{R}(s),s)\,ds \quad \text{for} \ t\in [0,\tau_{0}] \end{align}

and

(4.11)\begin{align} & f(C(t),N(t),Y(t),\mathbf{R}(t),t)-f(1,N(0),Y(0),\mathbf{R}(0),0)\nonumber\\ & \quad -\int_{0}^{t}\mathcal{G}f(1,N(s),Y(s),\mathbf{R}(s),s)\,ds \quad \text{for} \ t\in [0,\tau_{1}] \end{align}

are martingales for any function $f \in \mathfrak {D}(\mathcal {G})$.

Proof. Define the following process

$$M^{f}(t)=f(X(t))-f(X(0))-\int_{0}^{t}\mathcal{G}f(X(s))\,ds$$

where $\mathcal {G}$ is the infinitesimal generator of the PDMP $X(t)$ and $f$ is a measurable function satisfying (i)–(iii) in Theorem 5.5 of Davis [Reference Davis7]. Hence, according to Proposition 14.13 in Davis [Reference Davis8], $M^{f}(t)$ is a martingale.

Using the generators Eqs. (4.3) and (4.4), the result follows immediately.

Note that according to the treatment of our model as a PDMP, stopping times $\tau _{0}$ and $\tau _{1}$ are given by the random variables $\min (A(t)$, $R_{1}(t))$ and $Y(t)$, respectively. The component $A(t)$ refers to the inter-arrival time which is exponentially distributed with parameter $\lambda$.

5. Mean number of customers in the orbit

In this section, we will derive the expected number of customers blocked in the orbit during the periods of inactivity and occupation of the server separately. To this end, we will mainly use the result of Theorem 2. In fact, authors in Dassios and Zhao [Reference Dassios and Zhao6] have shown that this method based on martingales allows to calculate any moment of $N(t)$ without taking into account the stability condition (see Theorems 3.6 and 3.8).

Theorem 3. The conditional expectation of the number of blocked customers $N(t)$ given $N(0)=n_{0}$ and $Y(0)=y_{0}$ (when $Y(t)\in \mathbb {E}_{1}\cup \partial ^{*}\mathbb {E}_{1}$) is given by:

$$E[N(t)|N(0)=n_{0}]= \begin{cases} n_{0}+ \lambda t & \text{for} \ t \in [0,\tau_{0}];\\ n_{0}+\lambda t+\dfrac{1}{\mu_{1}} (y_{0}-t)-1 & \text{for}\ t \in [0,\tau_{1}]. \end{cases}$$

where $\mu _{1}=\int _{0}^{\infty }y\,dF(y)$.

Proof. By setting $f(i,n,y,r_{1},\ldots,r_{n},t)=y+n\mu _{1}$ and verifying the conditions Eqs. (4.5)–(4.8), we have

$$\mathcal{G}f(0,n,0,r_{1},r_{2},\ldots,r_{n},t)=\lambda \mu_{1} \quad \text{and} \quad \mathcal{G}f(1,n,y,r_{1},r_{2},\ldots,r_{n},t)=\lambda \mu_{1}-1 .$$

According to Theorem 2, $Y(t)+N(t)\mu _{1}-n_{0}\mu _{1}-\int _{0}^{t}\lambda \mu _{1}\,ds$ and $Y(t)+N(t)\mu _{1}-y_{0}-n_{0}\mu _{1}-\int _{0}^{t}(\lambda \mu _{1}-1)\,ds$ are martingales. Hence, for $t \in [0,\tau _{0}]$

(5.1)\begin{equation} E[N(t)|N(0)=n_{0}]= \dfrac{1}{\mu_{1}} E[Y(t)]+\lambda t+n_{0} \end{equation}

and for $t \in [0,\tau _{1}]$

(5.2)\begin{equation} E[N(t)|N(0)=n_{0},Y(0)=y_{0}]=n_{0}+\lambda t+ \dfrac{1}{\mu_{1}} (y_{0}-t-E[Y(t)]). \end{equation}

Besides, we have

$$E[Y(t)]=\begin{cases} 0 & \text{for}\ t\in [0,\tau_{0}];\\ \mu_{1} & \text{for}\ t\in [0,\tau_{1}]. \end{cases}$$

By substituting $E[Y(t)]$ in Eqs. (5.1) and (5.2), the result follows.

6. Conclusion

The M/G/1 retrial queue with classical retrial policy, where each blocked customer in the orbit retries for service, and general retrial times has been modeled by aPDMP. Using the extended generator of the PDMP, we have derived the associated martingales capturing the dynamics of the retrial queue. These results have been exploited to find the conditional expected number of customers in the orbit in a transient regime. The approach of modeling with PDMPs can be applied to other retrial policies such as the constant retrial policy and the control policy. In further works, we will investigate the stationary regime of the considered model through the PDMP framework.

Acknowledgments

The second author acknowledges the financial support of the Directorate General for Scientific Research and Technological Development (DGRSDT) of Algeria for the Research Laboratory in Intelligent Informatics, Mathematics and Applications (RIIMA).

Competing interests

The authors declare no conflict of interest.

References

Breuer, L. (2008). Continuity of the M/G/c queue. Queueing Systems 58(4): 321331.CrossRefGoogle Scholar
Chakravarthy, S.R. (2013). Analysis of MAP/PH/c retrial queue with phase type retrials–simulation approach. In Belarusian Workshop on Queueing Theory. Springer, pp. 37–49.CrossRefGoogle Scholar
Clancy, D. (2014). SIR epidemic models with general infectious period distribution. Statistics & Probability Letters 85: 15.CrossRefGoogle Scholar
Costa, O.L.V. (1990). Stationary distributions for piecewise-deterministic Markov processes. Journal of Applied Probability 27(1): 6073.CrossRefGoogle Scholar
Costa, O.L.V. & Dufour, F. (2008). Stability and ergodicity of piecewise deterministic Markov processes. SIAM Journal on Control and Optimization 47(2): 10531077.CrossRefGoogle Scholar
Dassios, A. & Zhao, H. (2011). A dynamic contagion process. Advances in Applied Probability 43(3): 814846.CrossRefGoogle Scholar
Davis, M.H. (1984). Piecewise-deterministic Markov processes: A general class of non-diffusion stochastic models. Journal of the Royal Statistical Society: Series B (Methodological) 46(3): 353376.Google Scholar
Davis, M.H. (1993). Markov models & optimization. London, UK: Chapman & Hall.CrossRefGoogle Scholar
Dufour, F. & Costa, O.L. (1999). Stability of piecewise-deterministic Markov processes. SIAM Journal on Control and Optimization 37(5): 14831502.CrossRefGoogle Scholar
Falin, G.I. (1986). Single-line repeated orders queueing systems. Optimization 17(5): 649667.CrossRefGoogle Scholar
Falin, G.I. (1990). A survey of retrial queues. Queueing Systems 7: 127168.CrossRefGoogle Scholar
Falin, G.I. & Templeton, J.G.C. (1997). Retrial queues. London, UK: Chapman & Hall.CrossRefGoogle Scholar
Farahmand, K. (1990). Single line queue with repeated demands. Queueing Systems 6(1): 223228.CrossRefGoogle Scholar
Fayolle, G. (1986). A simple telephone exchange with delayed feedback. In O.J. Boxma, J.W. Cohen, & H.C. Tijms (eds.), Teletraffic analysis, computer performance evaluation, vol. 46. Amsterdam: Elsevier Science, pp. 561–581.Google Scholar
Gómez-Corral, A. & López-García, M. (2017). On SIR epidemic models with generally distributed infectious periods: Number of secondary cases and probability of infection. International Journal of Biomathematics 10(2): 1750024.CrossRefGoogle Scholar
Kapyrin, V. (1977). Stationary characteristics of a queuing system with repeated calls. Cybernetics 13(4): 584590.CrossRefGoogle Scholar
Kernane, T. (2008). Conditions for stability and instability of retrial queueing systems with general retrial times. Statistics & Probability Letters 78(18): 32443248.CrossRefGoogle Scholar
Kim, C., Klimenok, V., & Dudin, A. (2014). A G/M/1 retrial queue with constant retrial rate. Top 22(2): 509529.CrossRefGoogle Scholar
Lillo, R. (1996). A G/M/1-queue with exponential retrial. Top 4(1): 99120.CrossRefGoogle Scholar
Shin, Y.W. & Moon, D.H. (2011). Approximation of M/M/c retrial queue with ph-retrial times. European Journal of Operational Research 213(1): 205209.CrossRefGoogle Scholar
Shin, Y.W. & Moon, D.H. (2014). Approximation of PH/PH/c retrial queue with ph-retrial time. Asia-Pacific Journal of Operational Research 31(2): 1440010.CrossRefGoogle Scholar