This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Exact Analysis of the Age of Information in the Multi-Source M/GI/1 Queueing System

Yoshiaki Inoue and Tetsuya Takine This work was supported in part by JSPS KAKENHI Grant Number 21H03399. Y. Inoue and T. Takine are with Department of Information and Communications Technology, Graduate School of Engineering, Osaka University, Suita 565-0871, Japan. E-mail:{yoshiaki, takine}@comm.eng.osaka-u.ac.jp.
Abstract

We consider a situation that multiple monitoring applications (each with a different sensor-monitor pair) compete for a common service resource such as a communication link. Each sensor reports the latest state of its own time-varying information source to its corresponding monitor, incurring queueing and processing delays at the shared resource. The primary performance metric of interest is the age of information (AoI) of each sensor-monitor pair, which is defined as the elapsed time from the generation of the information currently displayed on the monitor. Although the multi-source first-come first-served (FCFS) M/GI/1 queue is one of the most fundamental model to describe such competing sensors, its exact analysis has been an open problem for years. In this paper, we show that the Laplace-Stieltjes transform (LST) of the stationary distribution of the AoI in this model, as well as the mean AoI, is given by a simple explicit formula, utilizing the double Laplace transform of the transient workload in the M/GI/1 queue.

I Introduction

We consider a situation that multiple time-varying information sources are monitored remotely. Specifically, each information source is attached with a sensor node, which reports observed data to a remote monitor that displays the latest state information received. Each observed data needs to be “served” by an intermediate service resource (such as a communication link and a processor) before it is received by the monitor, so that it experiences queueing and processing delay there. The service resource is shared by the source-monitor pairs in the system and the degree of congestion is highly affected by the sampling rate of the sensors.

Because the state of the information source changes over time, the value of displayed information degenerates with time. It is thus important for such a system that the information displayed on each monitor is kept sufficiently fresh. The age of information (AoI) [1, 2] is a widely used metric of the freshness of information, which is defined as the elapsed time from the generation time of currently displayed information. Specifically, letting KK denote the number of sensor-monitor pairs, the AoI AtkA_{t}^{\langle k\rangle} of the kkth (k=0,1,,K1k=0,1,\ldots,K-1) sensor-monitor pair at time tt is defined as

Atk=tηtk,t,A_{t}^{\langle k\rangle}=t-\eta_{t}^{\langle k\rangle},\quad t\in\mathbb{R}, (1)

where ηtk\eta_{t}^{\langle k\rangle} denotes the generation time of the displayed information at time tt.

The mean AoI E[Ak]\mathrm{E}[A^{\langle k\rangle}] is the primary performance measure of interest, which is defined as

E[Ak]=limT1TT/2T/2Atkdt.\mathrm{E}[A^{\langle k\rangle}]=\lim_{T\to\infty}\frac{1}{T}\int_{-T/2}^{T/2}A_{t}^{\langle k\rangle}\mathrm{d}t.

Under a fairly general setting, the mean AoI E[Ak]\mathrm{E}[A^{\langle k\rangle}] is given in terms of the mean inter-sampling time, the mean delay, and a cross-term of the inter-sampling time and the delay [1]. A more detailed characterization of the AoI process is provided by the probability distribution FAk(x)F_{A^{\langle k\rangle}}(x) (x0x\geq 0) of the AoI, which is defined as a long-run fraction of time that the AoI process does not exceed xx:

FAk(x)=limT1TT/2T/2𝟙{Atkx}dt,F_{A^{\langle k\rangle}}(x)=\lim_{T\to\infty}\frac{1}{T}\int_{-T/2}^{T/2}\mathbbm{1}\{A_{t}^{\langle k\rangle}\leq x\}\mathrm{d}t,

which is known to be given in terms of the probability distributions of the system delay and the peak AoI value just before updates [3].

To characterize the AoI process formally, the system is usually modeled as a queueing system, where generation of information packets by sensors corresponds to “arrivals”, and their reception by monitors corresponds to “departures”. Analytical results for the AoI in various queueing systems have been reported in the literature, e.g., first-come first-served (FCFS) queues [1, 3], last-come first-served (LCFS) queues [4, 5, 6], and multi-server queues [7, 8]. The FCFS queueing model represents the situation where the application layer does not have any control over the data transmission performed by lower layer protocols, so that the AoI performance in this framework is interpreted as that of monitoring applications implemented on top of the conventional communication mechanisms. The LCFS queueing model, on the other hand, gives a priority to newer information in transmission, which in general leads to better performance at the expense of implementation/operation costs, especially due to the necessity of the cross-layer design and intelligent functionalities of network components.

The models mentioned above [1, 3, 4, 5, 6, 7] assume that there is only a single sensor node in the system (i.e., K=1K=1), so that the “congestion” of the system refers to the accumulation of buffered packets generated by the sensor itself. In order to incorporate the existence of competing sensors that share the communication resources, it is necessary to consider multi-source queueing models, where each source corresponds to a sensor node. In fact, significant research efforts have recently been put into generalizing the AoI analysis to multi-source queueing models. Such generalization has been particularly successful for the multi-source buffer-less M/G/1/1 queues [11, 12, 13, 14, 15, 16], whose analysis is less complicated due to the small state space.

For the standard (infinite-buffer) multi-source FCFS systems, however, the exact analysis of the AoI has been reported for the simplest M/M/1 queueing model only [9, 10]. The difficulty in the analysis of the multi-source FCFS M/GI/1 queue is explained in [10], where approximate formulas for the mean AoI are proposed. Roughly speaking, in multi-class FCFS queues, an unbounded number of information packets from other sources can arrive between two successive packets of a single source, which makes it necessary to keep track of the whole dynamics of the system driven by other sources. Because of such difficulty, exact analysis of the AoI in the multi-source FCFS M/GI/1 queue has been an open problem for years.

The purpose of this paper is to provide a solution to this problem, by showing that the Laplace-Stieltjes transform (LST) of the stationary distribution of the AoI in the multi-source FCFS M/GI/1 queue, as well as the mean AoI, is given by a simple explicit formula. The key observation is that the system dynamics between two successive packets from a single source is characterized in terms of the double transform [17, P. 83] of the transient workload process in the M/GI/1 queue. As will be shown later, the complexity of the exact mean AoI formula is more or less the same as that of the single-source M/GI/1 queue.

The rest of this paper is organized as follows. We first provide a formal description of the model in Section II. We then present the exact analysis of the AoI in Section III, where explicit formulas of the LST and the mean of the AoI are derived. Finally, we conclude this paper in Section IV.

II Model

Throughout this paper, we comply with the following rules of notation: for any non-negative random variable YY, its cumulative distribution function (CDF) is denoted by FY(x)F_{Y}(x) (x0x\geq 0) and its LST is denoted by fY(s)f_{Y}^{*}(s) (s>0s>0):

FY(x):=Pr(Yx),fY(s):=E[esY]=0esxdFY(x).F_{Y}(x):=\Pr(Y\leq x),\qquad f_{Y}^{*}(s):=\mathrm{E}[e^{-sY}]=\int_{0}^{\infty}e^{-sx}\mathrm{d}F_{Y}(x).

We also apply this rule in the opposite direction: given a CDF denoted by FY(x)F_{Y}(x) (x0x\geq 0), the corresponding random variable YY is defined accordingly.

We consider an FCFS single-server queue with KK different source-monitor pairs. The kkth (k=0,1,,K1k=0,1,\ldots,K-1) source generates information packets according to a Poisson process with rate λk\lambda^{\langle k\rangle} (λk>0\lambda^{\langle k\rangle}>0), which arrive at the server immediately after their generations. Packets are served by the server in order of arrival and they update the corresponding monitor on completion of the service. Hereafter we shall refer to the kkth source and monitor as source kk and monitor kk for short. We assume that service times of packets from source kk are i.i.d. according to a general non-negative probability distribution with CDF FHk(x)F_{H^{\langle k\rangle}}(x) (x0x\geq 0), i.e., the service time distribution is allowed to differ depending on the source-monitor pair. Let ρk\rho^{\langle k\rangle} denote the traffic intensity of source kk:

ρk=λkE[Hk],k=0,1,,K1.\rho^{\langle k\rangle}=\lambda^{\langle k\rangle}\mathrm{E}[H^{\langle k\rangle}],\quad k=0,1,\ldots,K-1.

We assume k=1Kρk<1\sum_{k=1}^{K}\rho^{\langle k\rangle}<1, so that the system is stable. Also, we assume that the system is stationary unless otherwise mentioned.

To define the AoI process for each monitor, let αnk\alpha_{n}^{\langle k\rangle} (nn\in\mathbb{Z}) denote the generation time of the nnth packet of source kk. Without loss of generality, we set packet indices so that α1k<0α0k\alpha_{-1}^{\langle k\rangle}<0\leq\alpha_{0}^{\langle k\rangle} holds for each kk. Let βnk\beta_{n}^{\langle k\rangle} (nn\in\mathbb{Z}) denote the reception time of the nnth packet by monitor kk. Its system delay DnkD_{n}^{\langle k\rangle} (nn\in\mathbb{Z}) is then given by Dnk=βnkαnkD_{n}^{\langle k\rangle}=\beta_{n}^{\langle k\rangle}-\alpha_{n}^{\langle k\rangle}.

The AoI AtkA_{t}^{\langle k\rangle} (t0t\geq 0, k=1,2,,Kk=1,2,\ldots,K) of monitor kk at time tt is then defined as

Atk=tαnk,βnkt<βn+1k.A_{t}^{\langle k\rangle}=t-\alpha_{n}^{\langle k\rangle},\quad\beta_{n}^{\langle k\rangle}\leq t<\beta_{n+1}^{\langle k\rangle}.

Let Apeak,nkA_{\mathrm{peak},n}^{\langle k\rangle} denote the nnth peak AoI of monitor kk, i.e., the value of AoI just before nnth update of the monitor:

Apeak,nk=limtβnkAtk\displaystyle A_{\mathrm{peak},n}^{\langle k\rangle}=\lim_{t\to\beta_{n}^{\langle k\rangle}-}A_{t}^{\langle k\rangle} =βnkαn1k\displaystyle=\beta_{n}^{\langle k\rangle}-\alpha_{n-1}^{\langle k\rangle}
=βnkαnk+αnkαn1k\displaystyle=\beta_{n}^{\langle k\rangle}-\alpha_{n}^{\langle k\rangle}+\alpha_{n}^{\langle k\rangle}-\alpha_{n-1}^{\langle k\rangle}
=Dnk+Gnk,\displaystyle=D_{n}^{\langle k\rangle}+G_{n}^{\langle k\rangle}, (2)

where Gnk:=αnkαn1kG_{n}^{\langle k\rangle}:=\alpha_{n}^{\langle k\rangle}-\alpha_{n-1}^{\langle k\rangle} denotes the inter-generation time between nnth and (n1)(n-1)st packets of source kk.

Because of the stationarity of the system, the probability distribution of AtkA_{t}^{\langle k\rangle} does not depend on tt. Let AkA^{\langle k\rangle} denote a generic random variable following the stationary distribution of AtkA_{t}^{\langle k\rangle}. Similarly, let GkG^{\langle k\rangle}, DkD^{\langle k\rangle}, and ApeakkA_{\mathrm{peak}}^{\langle k\rangle} denote generic random variables following the stationary distributions of GnkG_{n}^{\langle k\rangle}, DnkD_{n}^{\langle k\rangle} and Apeak,nkA_{\mathrm{peak},n}^{\langle k\rangle}.

III Exact Analysis of the AoI distribution

For better readability, we mainly focus on the AoI A0A^{\langle 0\rangle} of source 0 in the analysis below, which can be readily done without loss of generality.

Because the sequences of amounts of work brought by sources 1,2,,K11,2,\ldots,K-1 (excluding source 0) are all represented as compound Poisson processes, their superposition also forms a compound Poisson process with rate λ+\lambda^{+} and the random jump size H+H^{+}, where

λ+=k=1K1λk,fH+(s)=k=1K1λkλ+fHk(s).\displaystyle\lambda^{+}=\sum_{k=1}^{K-1}\lambda^{\langle k\rangle},\quad f_{H^{+}}^{*}(s)=\sum_{k=1}^{K-1}\frac{\lambda^{\langle k\rangle}}{\lambda^{+}}\cdot f_{H^{\langle k\rangle}}^{*}(s).

Similarly, we define λall\lambda^{\mathrm{all}} and HallH^{\mathrm{all}} as the rate and random jump size of the compound Poisson process composed of all sources including source 0:

λall=λ0+λ+,fHall(s)=λ0fH0(s)+λ+fH+(s)λall.\displaystyle\lambda^{\mathrm{all}}=\lambda^{\langle 0\rangle}+\lambda^{+},\quad f_{H^{\mathrm{all}}}^{*}(s)=\frac{\lambda^{\langle 0\rangle}f_{H^{\langle 0\rangle}}^{*}(s)+\lambda^{+}f_{H^{+}}^{*}(s)}{\lambda^{\mathrm{all}}}. (3)

We further define ρall\rho^{\mathrm{all}} as the total traffic intensity:

ρall=ρ0+ρ+,\rho^{\mathrm{all}}=\rho^{\langle 0\rangle}+\rho^{+}, (4)

where

ρ+=ρ1+ρ2++ρK1.\rho^{+}=\rho_{1}+\rho_{2}+\cdots+\rho_{K-1}.

To make the notation even simpler, we hereafter drop the superscript 0\langle 0\rangle for quantities of source 0. For example, the inter-generation time Gn0G_{n}^{\langle 0\rangle} and system delay Dn0D_{n}^{\langle 0\rangle} of source 0 are written simply as

Gn:=Gn0,Dn:=Dn0.\displaystyle G_{n}:=G_{n}^{\langle 0\rangle},\quad D_{n}:=D_{n}^{\langle 0\rangle}.

Similarly, the AoI and peak AoI of source 0 are written as

At:=At0,Apeak,n=Apeak,n0.A_{t}:=A_{t}^{\langle 0\rangle},\quad A_{\mathrm{peak},n}=A_{\mathrm{peak},n}^{\langle 0\rangle}.

The same rules apply to their stationary versions:

G:=G0,D:=D0,A:=A0.G:=G^{\langle 0\rangle},\quad D:=D^{\langle 0\rangle},\quad A:=A^{\langle 0\rangle}.

Notice that (3) and (4) are rewritten as

λall=λ+λ+,fHall(s)=λfH(s)+λ+fH+(s)λall,ρall=ρ+ρ+.\lambda^{\mathrm{all}}=\lambda+\lambda^{+},\qquad f_{H^{\mathrm{all}}}^{*}(s)=\frac{\lambda f_{H}^{*}(s)+\lambda^{+}f_{H^{+}}^{*}(s)}{\lambda^{\mathrm{all}}},\qquad\rho^{\mathrm{all}}=\rho+\rho^{+}.

The starting point of our analysis is the following relation known in the literature [3]:

Lemma 1.

The LST of the stationary AoI fA(s)f_{A}^{*}(s) of source 0 is represented in terms of the LSTs of its stationary system delay fD(s)f_{D}^{*}(s) and stationary peak AoI fApeak(s)f_{A_{\mathrm{peak}}}^{*}(s) as

fA(s)=λfD(s)fApeak(s)s,s>0.f_{A}^{*}(s)=\lambda\cdot\frac{f_{D}^{*}(s)-f_{A_{\mathrm{peak}}}^{*}(s)}{s},\quad s>0.

Because the stationary waiting time of packets of source 0 is identical to that in the single-input M/GI/1 queue with the arrival rate λall\lambda^{\mathrm{all}} and service time LST fHall(s)f_{H^{\mathrm{all}}}^{*}(s), we have

fW(s)=(1ρall)ssλall+λallfHall(s)=(1ρρ+)ssλλ++λfH(s)+λ+fH+(s).f_{W}^{*}(s)=\frac{(1-\rho^{\mathrm{all}})s}{s-\lambda^{\mathrm{all}}+\lambda^{\mathrm{all}}f_{H^{\mathrm{all}}}^{*}(s)}=\frac{(1-\rho-\rho^{+})s}{s-\lambda-\lambda^{+}+\lambda f_{H}^{*}(s)+\lambda^{+}f_{H^{+}}^{*}(s)}. (5)

It then follows from fD(s)=fW(s)fH(s)f_{D}^{*}(s)=f_{W}^{*}(s)f_{H}^{*}(s) that

fD(s)\displaystyle f_{D}^{*}(s) =(1ρρ+)ssλλ++λfH(s)+λ+fH+(s)fH(s).\displaystyle=\frac{(1-\rho-\rho^{+})s}{s-\lambda-\lambda^{+}+\lambda f_{H}^{*}(s)+\lambda^{+}f_{H^{+}}^{*}(s)}\cdot f_{H}^{*}(s). (6)

We thus consider fApeak(s)f_{A_{\mathrm{peak}}}^{*}(s) below.

Lemma 2.

The LST fApeak(s)f_{A_{\mathrm{peak}}}^{*}(s) of the stationary peak AoI of source 0 is given by

fApeak(s)=λfH(s)s+λϕ(s){fD(s)sfD(ψ(s+λ))ψ(s+λ)},f_{A_{\mathrm{peak}}}^{*}(s)=\frac{\lambda f_{H}^{*}(s)}{s+\lambda-\phi(s)}\left\{f_{D}^{*}(s)-\frac{sf_{D}^{*}\left(\psi(s+\lambda)\right)}{\psi(s+\lambda)}\right\}, (7)

where ϕ(s)\phi(s) (s>0s>0) is defined as

ϕ(s)=sλ++λ+fH+(s),\phi(s)=s-\lambda^{+}+\lambda^{+}f_{H^{+}}^{*}(s), (8)

and ψ(ω)\psi(\omega) (ω>0\omega>0) denotes the inverse function of ϕ(s)\phi(s), i.e.,

ϕ(ψ(ω))=ω.\phi(\psi(\omega))=\omega. (9)
Remark 1.

It is readily verified that ϕ(s)\phi(s) is an increasing continuous function for s>0s>0:

dϕ(s)ds=1+λ+ddsE[esH+]=1ρ+E[esH+]>0.\frac{\mathrm{d}\phi(s)}{\mathrm{d}s}=1+\lambda^{+}\cdot\frac{\mathrm{d}}{\mathrm{d}s}\mathrm{E}[e^{-sH^{+}}]=1-\rho^{+}\mathrm{E}[e^{-sH^{+}}]>0.

Also, we have lims0+ϕ(s)=0\lim_{s\to 0+}\phi(s)=0 and limsϕ(s)=\lim_{s\to\infty}\phi(s)=\infty, which ensures the existence of the inverse function ψ(ω)\psi(\omega) (ω>0\omega>0).

Proof.

From (2), we have

Apeak,n+1=Gn+1+Dn+1=Gn+1+Wn+1+Hn+1,A_{\mathrm{peak},n+1}=G_{n+1}+D_{n+1}=G_{n+1}+W_{n+1}+H_{n+1},

where WnW_{n} denotes the waiting time of the nnth packet of source 0. We then have

fApeak(s)\displaystyle f_{A_{\mathrm{peak}}}^{*}(s) =E[esApeak,n+1]\displaystyle=\mathrm{E}\left[e^{-sA_{\mathrm{peak},n+1}}\right]
=E[es(Gn+1+Wn+1+Hn+1)]\displaystyle=\mathrm{E}\left[e^{-s(G_{n+1}+W_{n+1}+H_{n+1})}\right]
=fH(s)E[es(Gn+1+Wn+1)]\displaystyle=f_{H}^{*}(s)\cdot\mathrm{E}\left[e^{-s(G_{n+1}+W_{n+1})}\right]
=fH(s)x=0dFDn(x)y=0E[es(y+Wn+1)Dn=x,Gn=y]λeλydy\displaystyle=f_{H}^{*}(s)\int_{x=0}^{\infty}\mathrm{d}F_{D_{n}}(x)\int_{y=0}^{\infty}\mathrm{E}\left[e^{-s(y+W_{n+1})}\mid D_{n}=x,G_{n}=y\right]\lambda e^{-\lambda y}\mathrm{d}y
=fH(s)x=0dFD(x)y=0E[esWn+1Dn=x,Gn=y]λe(s+λ)ydy.\displaystyle=f_{H}^{*}(s)\int_{x=0}^{\infty}\mathrm{d}F_{D}(x)\int_{y=0}^{\infty}\mathrm{E}\left[e^{-sW_{n+1}}\mid D_{n}=x,G_{n}=y\right]\lambda e^{-(s+\lambda)y}\mathrm{d}y. (10)

Note here that the waiting time Wn+1W_{n+1} of the (n+1n+1)st packet of source 0 is given in terms of the transient workload VtV_{t} of the M/GI/1 queue with arrival rate λ+\lambda^{+} and the service time distribution FH+(x)F_{H^{+}}(x). More specifically, its conditional LST given DnD_{n} and GnG_{n} is given by

E[esWn+1Dn=x,Gn=y]=E[esVyV0=x].\displaystyle\mathrm{E}\left[e^{-sW_{n+1}}\mid D_{n}=x,G_{n}=y\right]=\mathrm{E}[e^{-sV_{y}}\mid V_{0}=x].

It is known that the double-transform of the workload process VtV_{t} in the M/GI/1 queue is given by [17, P. 83]:

0E[esVyV0=x]eωydy=1ωϕ(s){esxseψ(ω)xψ(ω)},\int_{0}^{\infty}\mathrm{E}[e^{-sV_{y}}\mid V_{0}=x]e^{-\omega y}\mathrm{d}y=\frac{1}{\omega-\phi(s)}\left\{e^{-sx}-\frac{se^{-\psi(\omega)x}}{\psi(\omega)}\right\},

where ϕ(s)\phi(s) and ψ(ω)\psi(\omega) are defined as (8) and (9).

Therefore, we have from (10),

fApeak(s)\displaystyle f_{A_{\mathrm{peak}}}^{*}(s) =λfH(s)x=0dFD(x)y=0E[esVyV0=x]e(s+λ)ydy\displaystyle=\lambda f_{H}^{*}(s)\int_{x=0}^{\infty}\mathrm{d}F_{D}(x)\int_{y=0}^{\infty}\mathrm{E}[e^{-sV_{y}}\mid V_{0}=x]e^{-(s+\lambda)y}\mathrm{d}y
=λfH(s)x=01s+λϕ(s){esxseψ(s+λ)xψ(s+λ)}dFD(x),\displaystyle=\lambda f_{H}^{*}(s)\int_{x=0}^{\infty}\frac{1}{s+\lambda-\phi(s)}\left\{e^{-sx}-\frac{se^{-\psi(s+\lambda)x}}{\psi(s+\lambda)}\right\}\mathrm{d}F_{D}(x),

which implies (7). ∎

We then obtain the following expression for the LST of the AoI distribution:

Theorem 1.

The LST fA(s)f_{A}^{*}(s) of the stationary AoI AA is given by

fA(s)=λfH(s)s+λϕ(s){fW(s)(1ρρ+)+λfD(ψ(s+λ))ψ(s+λ)},f_{A}^{*}(s)=\frac{\lambda f_{H}^{*}(s)}{s+\lambda-\phi(s)}\left\{f_{W}^{*}(s)-(1-\rho-\rho^{+})+\frac{\lambda f_{D}^{*}\left(\psi(s+\lambda)\right)}{\psi(s+\lambda)}\right\}, (11)

where fW(s)f_{W}^{*}(s) denotes the LST of the stationary waiting time WW:

fW(s)=(1ρρ+)sϕ(s)λ+λfH(s).f_{W}^{*}(s)=\frac{(1-\rho-\rho^{+})s}{\phi(s)-\lambda+\lambda f_{H}^{*}(s)}.
Proof.

Using ϕ(s)\phi(s) defined in (8), we rewrite the expression for the LST fD(s)f_{D}^{*}(s) of the stationary system delay as

fD(s)=(1ρρ+)sϕ(s)λ+λfH(s)fH(s).f_{D}^{*}(s)=\frac{(1-\rho-\rho^{+})s}{\phi(s)-\lambda+\lambda f_{H}^{*}(s)}\cdot f_{H}^{*}(s). (12)

We then have from Lemma 1 and Lemma 2,

fA(s)\displaystyle f_{A}^{*}(s) =λs{fD(s)fApeak(s)}\displaystyle=\frac{\lambda}{s}\left\{f_{D}^{*}(s)-f_{A_{\mathrm{peak}}}^{*}(s)\right\}
=λs1s+λϕ(s){(s+λϕ(s)λfH(s))fD(s)+sλfH(s)fD(ψ(s+λ))ψ(s+λ)}\displaystyle=\frac{\lambda}{s}\cdot\frac{1}{s+\lambda-\phi(s)}\left\{(s+\lambda-\phi(s)-\lambda f_{H}^{*}(s))f_{D}^{*}(s)+\frac{s\lambda f_{H}^{*}(s)f_{D}^{*}\left(\psi(s+\lambda)\right)}{\psi(s+\lambda)}\right\}
=λs1s+λϕ(s){sfD(s)(1ρρ+)sfH(s)+sλfH(s)fD(ψ(s+λ))ψ(s+λ)},\displaystyle=\frac{\lambda}{s}\cdot\frac{1}{s+\lambda-\phi(s)}\left\{sf_{D}^{*}(s)-(1-\rho-\rho^{+})s\cdot f_{H}^{*}(s)+\frac{s\lambda f_{H}^{*}(s)f_{D}^{*}\left(\psi(s+\lambda)\right)}{\psi(s+\lambda)}\right\},

which implies (11). ∎

Taking the derivative of (6) and letting s0+s\to 0+, we have the mean system delay E[D]\mathrm{E}[D]:

E[D]=λE[H2]+λ+E[(H+)2]2(1ρρ+)+E[H].\mathrm{E}[D]=\frac{\lambda\mathrm{E}[H^{2}]+\lambda^{+}\mathrm{E}[(H^{+})^{2}]}{2(1-\rho-\rho^{+})}+\mathrm{E}[H]. (13)

An explicit formula for the mean AoI is then obtained as follows:

Corollary 1.

The mean AoI E[A]\mathrm{E}[A] is given by

E[A]=λE[H2]+λ+E[(H+)2]2(1ρρ+)+E[H]+ρ+λ+1ρρ+λfH(γ),\mathrm{E}[A]=\frac{\lambda\mathrm{E}[H^{2}]+\lambda^{+}\mathrm{E}[(H^{+})^{2}]}{2(1-\rho-\rho^{+})}+\mathrm{E}[H]+\frac{\rho^{+}}{\lambda}+\frac{1-\rho-\rho^{+}}{\lambda f_{H}^{*}(\gamma)}, (14)

where γ\gamma is defined as

γ:=ψ(λ),\gamma:=\psi(\lambda),

i.e., it represents the unique solution of the following equation:

xλλ++λ+fH+(x)=0,λxλ+λ+.x-\lambda-\lambda^{+}+\lambda^{+}f_{H^{+}}^{*}(x)=0,\quad\lambda\leq x\leq\lambda+\lambda^{+}. (15)

The proof of Corollary 1 is straightforward; we provide the proof in the appendix.

The expression (14) for the mean AoI E[A]\mathrm{E}[A] largely resembles the known result for the single-source M/GI/1 queue. In fact, it is readily verified that as λ+0+\lambda^{+}\to 0+, we have γλ\gamma\to\lambda and

E[A]λE[H2]2(1ρ)+E[H]+1ρλfH(λ),\mathrm{E}[A]\to\frac{\lambda\mathrm{E}[H^{2}]}{2(1-\rho)}+\mathrm{E}[H]+\frac{1-\rho}{\lambda f_{H}^{*}(\lambda)},

which agrees with the mean AoI in the single-source M/GI/1 queue [3, Eq. (36)].

IV Conclusion

In this paper, we presented the exact analysis of the AoI in the multi-source FCFS M/GI/1 queueing model. We derived the explicit formula for the LST of the stationary distribution of the AoI in Theorem 1, based on the observation that the LST of the stationary peak AoI distribution is tractable using the explicit double-transform of the transient workload in the M/GI/1 queue. We also showed in Corollary 1 that the mean AoI E[A]\mathrm{E}[A] in this model is given by a simple explicit formula. The result shows that the complexity of the mean AoI formula in the multi-source M/GI/1 queue is more or less the same as that of the single-source case, suggesting its usefulness as a simple framework in understanding the AoI performance in multi-source communication systems.

Note first that the existence and uniqueness of the solution of (15) is verified in the same way as Remark 1: the left-hand side of this equation is increasing and continuous for λxλ+λ+\lambda\leq x\leq\lambda+\lambda^{+} and it takes a negative value at x=λx=\lambda and a positive value at x=λ+λ+x=\lambda+\lambda^{+}.

We decompose the LST of the AoI into the following product form (11):

fA(s)=fA,1(s)fA,2(s),f_{A}^{*}(s)=f_{A,1}^{*}(s)\cdot f_{A,2}^{*}(s),

where

fA,1(s)\displaystyle f_{A,1}^{*}(s) :=λfH(s)s+λϕ(s),\displaystyle:=\frac{\lambda f_{H}^{*}(s)}{s+\lambda-\phi(s)}, (16)
fA,2(s)\displaystyle f_{A,2}^{*}(s) :=fW(s)(1ρρ+)+λfD(ψ(s+λ))ψ(s+λ).\displaystyle:=f_{W}^{*}(s)-(1-\rho-\rho^{+})+\frac{\lambda f_{D}^{*}\left(\psi(s+\lambda)\right)}{\psi(s+\lambda)}. (17)

Noting that the definition of γ\gamma and (15) imply ϕ(γ)=λ\phi(\gamma)=\lambda, we have from (12),

fD(γ)=(1ρρ+)γλ.f_{D}^{*}(\gamma)=\frac{(1-\rho-\rho^{+})\gamma}{\lambda}. (18)

It then follows from lims0ϕ(s)=0\lim_{s\to 0}\phi(s)=0 that

lims0+fA,1(s)=1,lims0+fA,2(s)=1.\lim_{s\to 0+}f_{A,1}^{*}(s)=1,\quad\lim_{s\to 0+}f_{A,2}^{*}(s)=1.

Therefore, letting

fA,i(1)(s):=(1)dfA,i(s)ds,i=1,2,f_{A,i}^{(1)}(s):=(-1)\cdot\frac{\mathrm{d}f_{A,i}^{*}(s)}{\mathrm{d}s},\quad i=1,2,

we have

E[A]\displaystyle\mathrm{E}[A] =(1)lims0+dfA(s)ds\displaystyle=(-1)\cdot\lim_{s\to 0+}\frac{\mathrm{d}f_{A}^{*}(s)}{\mathrm{d}s}
=lims0+fA,1(1)(s)+lims0+fA,2(1)(s).\displaystyle=\lim_{s\to 0+}f_{A,1}^{(1)}(s)+\lim_{s\to 0+}f_{A,2}^{(1)}(s). (19)

We thus consider fA,1(1)(s)f_{A,1}^{(1)}(s) and fA,2(1)(s)f_{A,2}^{(1)}(s) below.

By definition (cf. (8) and (9)), we have

dϕ(s)ds=1λ+fH+(1)(s),dψ(ω)dω=11λ+fH+(1)(ψ(ω)),\frac{\mathrm{d}\phi(s)}{\mathrm{d}s}=1-\lambda^{+}f_{H^{+}}^{(1)}(s),\quad\frac{\mathrm{d}\psi(\omega)}{\mathrm{d}\omega}=\frac{1}{1-\lambda^{+}f_{H^{+}}^{(1)}(\psi(\omega))},

where

fH+(1)(s):=(1)dfH+(s)ds.f_{H^{+}}^{(1)}(s):=(-1)\cdot\frac{\mathrm{d}f_{H^{+}}^{*}(s)}{\mathrm{d}s}.

It then follows immediately from (16), (17), and (18) that

lims0+fA,1(1)(s)\displaystyle\lim_{s\to 0+}f_{A,1}^{(1)}(s) =E[H]+ρ+λ,\displaystyle=\mathrm{E}[H]+\frac{\rho^{+}}{\lambda}, (20)
lims0+fA,2(1)(s)\displaystyle\lim_{s\to 0+}f_{A,2}^{(1)}(s) =E[W]+λfD(1)(ψ(λ))ψ(λ)limωλdψ(ω)dω+λfD(ψ(λ))(ψ(λ))2limωλdψ(ω)dω\displaystyle=\mathrm{E}[W]+\frac{\lambda f_{D}^{(1)}(\psi(\lambda))}{\psi(\lambda)}\cdot\lim_{\omega\to\lambda}\frac{\mathrm{d}\psi(\omega)}{\mathrm{d}\omega}+\frac{\lambda f_{D}^{*}(\psi(\lambda))}{(\psi(\lambda))^{2}}\cdot\lim_{\omega\to\lambda}\frac{\mathrm{d}\psi(\omega)}{\mathrm{d}\omega}
=E[W]+11λ+fH+(1)(γ)(λγfD(1)(γ)+λfD(γ)γ2)\displaystyle=\mathrm{E}[W]+\frac{1}{1-\lambda^{+}f_{H^{+}}^{(1)}(\gamma)}\left(\frac{\lambda}{\gamma}\cdot f_{D}^{(1)}(\gamma)+\frac{\lambda f_{D}^{*}(\gamma)}{\gamma^{2}}\right)
=E[W]+1{1λ+fH+(1)(γ)}γ(λfD(1)(γ)+1ρρ+),\displaystyle=\mathrm{E}[W]+\frac{1}{\{1-\lambda^{+}f_{H^{+}}^{(1)}(\gamma)\}\gamma}\left(\lambda f_{D}^{(1)}(\gamma)+1-\rho-\rho^{+}\right), (21)

where

fD(1)(s):=(1)dfD(s)ds.f_{D}^{(1)}(s):=(-1)\cdot\frac{\mathrm{d}f_{D}^{*}(s)}{\mathrm{d}s}.

The rest is to determine fD(1)(γ)f_{D}^{(1)}(\gamma), which is straightforward with ϕ(γ)=λ\phi(\gamma)=\lambda:

fD(1)(γ)\displaystyle f_{D}^{(1)}(\gamma) =1ρρ+ϕ(γ)λ+λfH(γ)fH(γ)\displaystyle=-\frac{1-\rho-\rho^{+}}{\phi(\gamma)-\lambda+\lambda f_{H}^{*}(\gamma)}\cdot f_{H}^{*}(\gamma)
+(1ρρ+)γ(ϕ(γ)λ+λfH(γ))2{1λ+fH+(1)(γ)λfH(1)(γ)}fH(γ)\displaystyle\quad{}+\frac{(1-\rho-\rho^{+})\gamma}{(\phi(\gamma)-\lambda+\lambda f_{H}^{*}(\gamma))^{2}}\cdot\{1-\lambda^{+}f_{H^{+}}^{(1)}(\gamma)-\lambda f_{H}^{(1)}(\gamma)\}\cdot f_{H}^{*}(\gamma)
+(1ρρ+)γϕ(γ)λ+λfH(γ)fH(1)(γ)\displaystyle\quad{}+\frac{(1-\rho-\rho^{+})\gamma}{\phi(\gamma)-\lambda+\lambda f_{H}^{*}(\gamma)}\cdot f_{H}^{(1)}(\gamma)
=1ρρ+λfH(γ)fH(γ)\displaystyle=-\frac{1-\rho-\rho^{+}}{\lambda f_{H}^{*}(\gamma)}\cdot f_{H}^{*}(\gamma)
+(1ρρ+)γ(λfH(γ))2{1λ+fH+(1)(γ)λfH(1)(γ)}fH(γ)\displaystyle\quad{}+\frac{(1-\rho-\rho^{+})\gamma}{(\lambda f_{H}^{*}(\gamma))^{2}}\cdot\{1-\lambda^{+}f_{H^{+}}^{(1)}(\gamma)-\lambda f_{H}^{(1)}(\gamma)\}\cdot f_{H}^{*}(\gamma)
+(1ρρ+)γλfH(γ)fH(1)(γ)\displaystyle\quad{}+\frac{(1-\rho-\rho^{+})\gamma}{\lambda f_{H}^{*}(\gamma)}\cdot f_{H}^{(1)}(\gamma)
=1ρρ+λ2fH(γ){λfH(γ)+γγλ+fH+(1)(γ)}.\displaystyle=\frac{1-\rho-\rho^{+}}{\lambda^{2}f_{H}^{*}(\gamma)}\cdot\{-\lambda f_{H}^{*}(\gamma)+\gamma-\gamma\lambda^{+}f_{H^{+}}^{(1)}(\gamma)\}.

Therefore, we have from (19), (20), and (21),

E[A]\displaystyle\mathrm{E}[A] =E[D]+ρ+λ\displaystyle=\mathrm{E}[D]+\frac{\rho^{+}}{\lambda}
+1{1λ+fH+(1)(γ)}γ(1ρρ+λfH(γ){λfH(γ)+γγλ+fH+(1)(γ)}+1ρρ+).\displaystyle\qquad{}+\frac{1}{\{1-\lambda^{+}f_{H^{+}}^{(1)}(\gamma)\}\gamma}\left(\frac{1-\rho-\rho^{+}}{\lambda f_{H}^{*}(\gamma)}\cdot\{-\lambda f_{H}^{*}(\gamma)+\gamma-\gamma\lambda^{+}f_{H^{+}}^{(1)}(\gamma)\}+1-\rho-\rho^{+}\right).

We thus obtain (14) from this equation and (13). ∎

References

  • [1] S. K. Kaul, R. D. Yates, and M. Gruteser, “Real-Time Status: How Often Should One Update?” in Proc. of IEEE INFOCOM 2012, pp. 2731–2735, 2012.
  • [2] R. Yates, Y. Sun, D. R. Brown, S. K. Kaul, E. Modiano, and S. Ulukus, “Age of Information: An Introduction and Survey,” IEEE J. Sel. Areas Commun., vol. 39, no. 5, pp. 1183–1210, 2021.
  • [3] Y. Inoue, H. Masuyama, T. Takine, and T. Tanaka, “A General Formula for the Stationary Distribution of the Age of Information and Its Application to Single-Server Queues,” IEEE Trans. Inf. Theory, vol. 65, no. 12, pp. 8305–8324, 2019.
  • [4] S. K. Kaul, R. D. Yates, and M. Gruteser, “Status Updates through Queues,” in Proc. of CISS 2012, 2012.
  • [5] M. Costa, M. Codreanu, and A. Ephremides, “On the Age of Information in Status Update Systems with Packet Management,” IEEE Trans. Inf. Theory, vol. 62, no. 4, pp. 1897–1910, 2016.
  • [6] J. P. Champati, H. Al-Zubaidy, and J. Gross, “On the Distribution of AoI for the GI/GI/1/1 and GI/GI/1/2* Systems: Exact Expressions and Bounds,” in Proc. of IEEE INFOCOM 2019, pp. 37–45, 2019.
  • [7] C. Kam, S. Kompella, G. D. Nguyen, and A. Ephremides, “Effect of Message Transmission Path Diversity on Status Age,” IEEE Trans. Inf. Theory, vol. 62, no. 3, pp. 1360–1374, 2016.
  • [8] Y. Inoue and T. Kimura, “Age-Effective Information Updating Over Intermittently Connected MANETs,” IEEE J. Sel. Areas Commun., vol. 39, no. 5, pp. 1293–1308, 2021.
  • [9] R. D. Yates S. K. Kaul, “The Age of Information: Real-Time Status Updating by Multiple Sources,” IEEE Trans. Inf. Theory, vol. 65, no. 3, pp. 1807–1827, 2019.
  • [10] M. Moltafet, M. Leinonen, and M. Codreanu, “On the Age of Information in Multi-Source Queueing Models,” IEEE Trans. Commun. vol. 68, no. 8, pp. 5003–5017, 2020.
  • [11] O. Doǧan and N. Akar, “The Multi-Source Probabilistically Preemptive M/PH/1/1 Queue with Packet Errors,” IEEE Trans. Commun, vol. 69, no. 11, pp. 7297–7308, 2021.
  • [12] Y. Jiang and N. Miyoshi, “Joint Performance Analysis of Ages of Information in a Multi-Source Pushout Server,” IEEE Trans. Inf. Theory, vol. 68, no. 2, pp. 965–975, 2022.
  • [13] Z. Chen, D. Deng, C. She, Y. Jia, L. Liang, S. Fang, M. Wang, and Y. Li, “Age of Information: The Multi-Stream M/G/1/1 Non-Preemptive System,” IEEE Trans. Commun., vol. 70, no. 4, pp. 2328–2341, 2022.
  • [14] M. Moltafet, M. Leinonen, and M. Codreanu, “Moment Generating Function of Age of Information in Multisource M/G/1/1 Queueing Systems,” IEEE Trans. Commun., vol. 70, no. 10, pp. 6503–6516, 2022.
  • [15] M. A. Abd-Elmagid and H. S. Dhillon, “Joint Distribution of Ages of Information in Networks,” IEEE Trans. Inf. Theory, vol. 69, no. 9, pp. 5701–5722, 2023.
  • [16] T. Zhang, S. Chen, Z. Chen, Z. Tian, Y. Jia, M. Wang, and D. O. Wu, “AoI and PAoI in the IoT-Based Multisource Status Update System: Violation Probabilities and Optimal Arrival Rate Allocation,” IEEE Internet Things J., vol. 10, no. 23, pp. 20617–20632, 2023.
  • [17] H. Takagi, Queueing Analysis: Vacation and Priority Systems, Part 1, Elsevier Science Publishers, Amsterdam, 1991.