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

Occupancy distributions of homogeneous queueing systems under opportunistic schedulingthanks: An earlier version of this manuscript appeared at the Information Theory and Applications Workshop, UCSD, 2008.

Murat Alanyali and Maxim Dashouk
Department of Electrical and Computer Engineering
Boston University
Abstract

We analyze opportunistic schemes for transmission scheduling from one of nn homogeneous queues whose channel states fluctuate independently. Considered schemes consist of the LCQ policy, which transmits from a longest connected queue in the entire system, and its low-complexity variants that transmit from a longest queue within a randomly chosen subset of connected queues. A Markovian model is studied where mean packet transmission time is n1n^{-1} and packet arrival rate is λ<1\lambda<1 per queue. Transient and equilibrium distributions of queue occupancies are obtained in the limit as the system size nn tends to infinity.

1 Introduction

We analyze a queueing system that arises under opportunistic scheduling of packet transmissions from a collection of queues with time-varying service rates. The system of interest is motivated by cellular data communications in which a single transceiver serves multiple mobile stations through distinct channels. Transmission scheduling has been well-studied in this context under the guiding principle of opportunism, which broadly refers to exploiting channel variations to maximize transmission capacity in the long term. In this we paper consider two generic opportunistic scheduling policies and obtain asymptotically exact descriptions of the resulting queue length distributions in symmetric systems of statistically identical queues.

Explicit analysis of queue lengths under opportunistic scheduling is generally difficult due to model complexities and lack of closed-form expressions. In related work Tassiulas and Ephremides [10] considered a queueing system under an on/off channel model in which each queue is independently either connected, and in turn it is eligible for service at a standard rate, or disconnected and it cannot be serviced. It is shown that transmitting from a longest connected queue stabilizes queue lengths if that is at all feasible, and that it minimizes occupancy of symmetric systems in which queues have identical load and channel statistics. This policy is coined LCQ. Explicit description of queue length distributions under LCQ is not available, but several bounds for mean packet delay are obtained in [3, 6] for LCQ and some of its variants. In more general models that admit multiple transmission rates and simultaneous transmissions, max-weight scheduling policies and their variations are shown in [9, 7] to asymptotically minimize a range of occupancy measures along a certain heavy-traffic limit. In the special case when one queue can transmit at a time, max-weight transmits from a queue that maximizes the product of instantaneous queue length and transmission rate. Tails of queue length distributions under such policies are studied in [8, 12] via large deviations analysis.

Here we consider a system of nn queues under an on/off channel model in which each queue is connected independently with probability q(0,1]q\in(0,1]. A continuous-time Markovian model is adopted where packet transmission rate is nn and packet arrival rate is λ<1\lambda<1 per queue. It can be seen that λ\lambda is also the load factor of the system; hence the condition λ<1\lambda<1 is necessary to have positive-recurrent queue lengths. We analyze this system for large values of the system size nn, under the LCQ scheduling policy and under its low-complexity variant, namely LCQ(d)LCQ(d), that transmits from a longest queue within d1d\geq 1 randomly selected connected queues. It is apparent that LCQ(dd) is not particularly suitable for non-symmetric systems, yet our goal here is to obtain a generic evaluation of its underlying principle, which may be tailored to specific circumstances.

We establish that as the system size nn increases equilibrium distribution of queue occupancies under the LCQ policy converges to the deterministic distribution centered at 0. Hence asymptotically almost all queues are empty in equilibrium. The number of queues with one packet is Θ(1)\Theta(1) and the number of queues with more than one packet is o(1)o(1) as nn\rightarrow\infty. In particular maximum queue size tends to one. The total number of packets in the system is therefore given by the number of nonempty queues, and this number is shown to have the same equilibrium distribution as the positive-recurrent birth-death process with birth rate λ\lambda and death rate 1(1q)j1-(1-q)^{j} at state j+j\in\mathbb{Z}_{+}. Note that the latter rate is equal to the probability of having at least one connected queue within a given set of jj queues, and that the nature of the total system occupancy may be anticipated once maximum queue size is determined to be one. The obtained description leads to asymptotic mean packet delay via Little’s law as the rate of packet arrivals to the system is readily seen to be nλn\lambda.

The analysis technique applied to LCQ can be extended, although with excessive tediousness, to symmetric max-weight policies in cases when each queue can be serviced independently at rate nRnR for some random variable RR. Above conclusions about LCQ offer substantial insight about queue occupancies in that more general setting. Namely if RR exceeds λ\lambda with positive probability (note that this condition is necessary for positive recurrence of queue lengths), then stochastic coupling with a related LCQ system yields that the maximum queue length in equilibrium tends to 1 as nn\rightarrow\infty. In turn, equilibrium distribution of total system occupancy should be expected to resemble that of a birth-death process with birth rate λ\lambda and death rate E[max{R1,R2,,Rj}]E[\max\{R_{1},R_{2},\cdots,R_{j}\}] at state jj, where R1,R2,,RjR_{1},R_{2},\cdots,R_{j} are independent copies of RR.

We obtain the equilibrium distribution {pk}k=0\{p_{k}\}_{k=0}^{\infty} of individual queue occupancy under the LCQ(dd) policy in the limit as nn\rightarrow\infty. Specifically pk=vkvk+1p_{k}=v^{*}_{k}-v^{*}_{k+1} where v0=1v^{*}_{0}=1 and

vk=11λvk1d,k=1,2,.v^{*}_{k}=1-\sqrt[d]{1-\lambda v^{*}_{k-1}},~~~~~~k=1,2,\cdots.

This distribution is shown to have tails that decay as Θ((λ/d)k)\Theta((\lambda/d)^{k}) as queue size kk\rightarrow\infty. Hence, in terms of tail occupancy probabilities, the choice parameter dd has the equivalent effect of reducing the system load by the same factor. Yet, for any fixed dd, system occupancy under LCQ(dd) is larger than that of LCQ by a factor of order nn. Numerical values of asymptotic mean packet delay under LCQ and LCQ(dd) are illustrated in Figure 1. We also conclude that if dd is allowed to depend on nn so that dd\rightarrow\infty but d/n0d/n\rightarrow 0 then order of the alluded disparity reduces to n/dn/d. This suggests that for moderate values of dd and nn LCQ(dd) and LCQ may be expected to have comparable packet delay.

The present analysis is based on approximating system dynamics via asymptotically exact differential representations that amount to functional laws of large numbers. Hence besides the mentioned equilibrium properties the paper also describes transient behavior of queue occupancies. The present analysis of LCQ(dd) is inspired by the work of Vvedenskaya et al. [11] which concerns an analogue of this policy that arises in routing and load balancing. It should perhaps be noted that the choice parameter dd appears to have a substantially more pronounced effect in the routing context. Our conclusions about the LCQ policy require a more elaborate technical approach. Here we apply a technique due to Kurtz [5] to obtain a suitable asymptotic description of the system. Related applications of this technique can be found in [1, 4, 13].

Refer to caption
(a) LCQ
Refer to caption
(b) LCQ(dd)
Figure 1: Mean packet delays as nn\rightarrow\infty. Note the normalization in (a).

The rest of this paper is organized as follows. We continue in Section 2 with formal description of the model and the notation adopted in the paper. The policies LCQ(d)(d) and LCQ are analyzed respectively in Sections 3 and 4. The paper concludes with final remarks in Section 5.

2 Queueing Model

Consider nn queues each serving a dedicated stream of packet arrivals as illustrated in Figure 2. Arrivals of each stream occur according to an independent Poisson process with rate λ<1\lambda<1 packets per unit time and transmission time of each packet is exponentially distributed with mean n1n^{-1}, chosen independently of the prior history of the system. Each queue is serviced by a designated channel but at most one channel can transmit at a time. Channel states fluctuate randomly and each channel is eligible for transmission with probability q(0,1]q\in(0,1] independently of other channels. Queues with eligible channels are called connected. We assume that channel states remain constant during packet transmission and that they are determined anew, independently of each other and of the current queue lengths, just before the next transmission decision.

Refer to caption
Figure 2: Sketch of the considered queueing system. At most one queue is serviced at a time, and each queue ii can be serviced only when channel ii is eligible for transmission.

Let mk(t)m_{k}(t) denote the number of queues with kk or more packets at time tt, and let

uk(t)=mk(t)n,k=0,1,2,.u_{k}(t)=\frac{m_{k}(t)}{n},~~~~~k=0,1,2,\cdots.

be the fraction of such queues in the system. In particular

1=u0(t)u1(t)0,1=u_{0}(t)\geq u_{1}(t)\geq\cdots\geq 0, (1)

the sequence {1uk(t)}k=0\{1-u_{k}(t)\}_{k=0}^{\infty} is the empirical cumulative distribution function of queue occupancies, and k1uk(t)\sum_{k\geq 1}u_{k}(t) is the empirical average queue occupancy in the system at time tt. We denote by PnP_{n} and EnE_{n} respectively probabilities and expectations associated with system size nn. In particular if qi(t)q_{i}(t) denotes the occupancy of the ii-th queue at time tt then by the symmetry of the model

En[uk(t)]=Pn(qi(t)k).E_{n}[u_{k}(t)]=P_{n}(q_{i}(t)\geq k).

Let UU denote the collection of sequences 𝐮={uk}k=0\mathbf{u}=\{u_{k}\}_{k=0}^{\infty} that satisfy relation (1), and endow UU with metric ρ\rho that is defined by

ρ(𝐮,𝐮)=supk>0|ukuk|k,𝐮,𝐮U.\rho(\mathbf{u},\mathbf{u}^{\prime})=\sup_{k>0}\frac{|u_{k}-u_{k}^{\prime}|}{k},~~~~~\mathbf{u},\mathbf{u}^{\prime}\in U.

Note that convergence in UU is equivalent to coordinate-wise convergence, and that UU is compact as each coordinate lies in a compact interval.

For each time tt let 𝐮(t)\mathbf{u}(t) denote the sequence {uk(t)}k=0\{u_{k}(t)\}_{k=0}^{\infty}. We represent the trajectory (𝐮(t):t0)(\mathbf{u}(t):t\geq 0) by the symbol 𝐮()\mathbf{u}(\cdot), and say that 𝐮()\mathbf{u}(\cdot) converges to a given trajectory 𝐯()\mathbf{v}(\cdot) uniformly on compact time-sets (uoc) if for all t>0t>0

sup0stρ(𝐮(s),𝐯(s))0 a.s. as n.\sup_{0\leq s\leq t}\rho(\mathbf{u}(s),\mathbf{v}(s))\rightarrow 0~~\mbox{ a.s. as }n\rightarrow\infty.

3 LCQ(dd)

We focus on LCQ(dd) which randomly and independently selects dd connected queues and transmits from a longest queue within this collection. For convenience of analysis we assume that repetitions are allowed in the selection procedure, and that if all selected queues are empty or no connected queue exists at a scheduling instant then the scheduler makes a new selection after idling for the transmission time of a hypothetical packet. This latter assumption can be seen to imply that scheduling instances form a Poisson process of rate nn.

For k=1,2,3,k=1,2,3,\cdots let 𝐞k={ek(i)}i=0\mathbf{e}_{k}=\{e_{k}(i)\}_{i=0}^{\infty} where

ek(i)=1{i=k}.e_{k}(i)=1\{i=k\}.

Here and in the rest of the paper 1{}1\{\cdot\} denotes 1 if its argument is true and 0 otherwise. Jumps of the process 𝐮()\mathbf{u}(\cdot) are of the form ±n1𝐞k\pm n^{-1}\mathbf{e}_{k} for some kk. Namely 𝐮()\mathbf{u}(\cdot) changes by +n1𝐞k+n^{-1}\mathbf{e}_{k} whenever some queue with exactly k1k-1 packets has a new arrival, and by n1𝐞k-n^{-1}\mathbf{e}_{k} whenever a packet transmission is scheduled from a queue with exactly kk packets. The number of queues with k1k-1 packets at time tt is given by n(uk1(t)uk(t))n(u_{k-1}(t)-u_{k}(t)); hence the former event occurs at instantaneous rate nλ(uk1(t)uk(t))n\lambda(u_{k-1}(t)-u_{k}(t)). The latter event occurs if and only if, upon completion of a packet transmission, (i)(i) there exists a connected queue and (ii)(ii) the scheduler inspects at least one connected queue with kk packets but none with more than kk packets. To determine the instantaneous rate of this event let τ\tau be a scheduling instant and let

αn=1(1q)n.\alpha_{n}=1-(1-q)^{n}.

Namely αn\alpha_{n} is the probability that there exists a connected queue at time τ\tau. Since channel states are assigned independently of queue lengths at time τ\tau, a connected queue at this time has strictly less than kk packets with probability 1uk(τ)1-u_{k}(\tau). Therefore, given that a connected queue exists, the maximum queue length inspected by the scheduler is equal to kk with (conditional) probability (1uk+1(τ))d(1uk(τ))d(1-u_{k+1}(\tau))^{d}-(1-u_{k}(\tau))^{d}. Since scheduling instants occur at constant rate nn, instantaneous rate of transmissions from a queue of size kk is nαn((1uk+1(t))d(1uk(t))d)n\alpha_{n}((1-u_{k+1}(t))^{d}-(1-u_{k}(t))^{d}). In particular 𝐮()\mathbf{u}(\cdot) is a time-homogeneous Markov process whose generator can be sketched as

𝐮{𝐮+n1𝐞k at rate nλ(uk1uk)𝐮n1𝐞k at rate nαn((1uk+1)d(1uk)d),k=1,2.\mathbf{u}\leftarrow\left\{\begin{array}[]{ll}\mathbf{u}+n^{-1}\mathbf{e}_{k}&\mbox{ at rate }~~n\lambda(u_{k-1}-u_{k})\\ \mathbf{u}-n^{-1}\mathbf{e}_{k}&\mbox{ at rate }~~n\alpha_{n}((1-u_{k+1})^{d}-(1-u_{k})^{d}),\end{array}\right.~~~~~~k=1,2\cdots. (2)

It offers some convenience in the subsequent discussion to represent the process 𝐮()\mathbf{u}(\cdot) via the “random time change” construction of [2, Chapter 6]. Namely

𝐮(t)\displaystyle\mathbf{u}(t) =\displaystyle= 𝐮(0)+k=1n1𝐞kAk1(nλ0tuk1(s)uk(s)ds)\displaystyle\mathbf{u}(0)+\sum_{k=1}^{\infty}n^{-1}\mathbf{e}_{k}A_{k-1}\left(n\lambda\int_{0}^{t}u_{k-1}(s)-u_{k}(s)ds\right) (3)
k=1n1𝐞kDk(nαn0t(1uk+1(s))d(1uk(s))dds)\displaystyle~~~~~~~~~~~~-\sum_{k=1}^{\infty}n^{-1}\mathbf{e}_{k}D_{k}\left(n\alpha_{n}\int_{0}^{t}(1-u_{k+1}(s))^{d}-(1-u_{k}(s))^{d}ds\right)

where Ak1(),Dk()A_{k-1}(\cdot),~D_{k}(\cdot), k=1,2,k=1,2,\cdots, are mutually independent Poisson processes each with unit rate. In informal terms, the processes Ak()A_{k}(\cdot) and Dk()D_{k}(\cdot) clock respectively arrivals to and departures from some queue with length kk, and the construction (3) is based on suitably expediting these processes to match the instantaneous transition rates given in (2). Martingale decomposition of the Poisson processes used in (3) yields

𝐮(t)=𝐮(0)+k=1𝐞k0t(λ(uk1(s)uk(s))dsαn((1uk+1(s))d(1uk(s))d))𝑑s+𝜺(t),\mathbf{u}(t)=\mathbf{u}(0)+\sum_{k=1}^{\infty}\mathbf{e}_{k}\int_{0}^{t}\left(\lambda(u_{k-1}(s)-u_{k}(s))ds-\alpha_{n}((1-u_{k+1}(s))^{d}-(1-u_{k}(s))^{d})\right)ds+\mbox{\boldmath$\varepsilon$}(t), (4)

where 𝜺(t)={εk(t)}k=0\mbox{\boldmath$\varepsilon$}(t)=\{\varepsilon_{k}(t)\}_{k=0}^{\infty} is such that each coordinate process εk()\varepsilon_{k}(\cdot) is a real-valued martingale adapted to the filtration generated by 𝐮()\mathbf{u}(\cdot).

Theorem 3.1

Every subsequence of {n}\{n\} has a further subsequence along which 𝐮()\mathbf{u}(\cdot) converges in distribution to a differentiable process 𝐯()\mathbf{v}(\cdot) such that v0(t)1v_{0}(t)\equiv 1 and

ddtvk(t)=λ(vk1(t)vk(t))(1vk+1(t))d+(1vk(t))dk=1,2.\frac{d}{dt}v_{k}(t)=\lambda(v_{k-1}(t)-v_{k}(t))-(1-v_{k+1}(t))^{d}+(1-v_{k}(t))^{d}~~~~~~k=1,2\cdots. (5)

Proof. The sequence of processes 𝐮():n=1,2,\mathbf{u}(\cdot):n=1,2,\cdots is tight in the Skorokhod space DU[0,)D_{U}[0,\infty) of right continuous functions with left limits in UU [2, Chapter 3.5]. Therefore every subsequence has a further subsequence that converges in distribution. By Skorokhod’s Embedding Theorem [2, Theorem 3.1.8] the processes can be reconstructed in an appropriate probability space if necessary so that the convergence occurs almost surely. Since jumps of 𝐮()\mathbf{u}(\cdot) have magnitudes that scale with n1n^{-1}, the limit process is continuous and convergence of 𝐮()\mathbf{u}(\cdot) can be taken uoc [2, Theorem 3.10.1]. To describe a limit process 𝐯()\mathbf{v}(\cdot) note that u0(t)1u_{0}(t)\equiv 1, and so ε0(t)0\varepsilon_{0}(t)\equiv 0. For k=1,2,k=1,2,\cdots the martingale εk()\varepsilon_{k}(\cdot) is square integrable. This process has O(n)O(n) jumps per unit time and each jump is of size n1n^{-1}, hence its quadratic variation vanishes as nn\rightarrow\infty. In turn, Doob’s L2L^{2} inequality [2, Proposition 2.2.16] implies that εk()0\varepsilon_{k}(\cdot)\rightarrow 0 uoc as nn\rightarrow\infty. Since αn1\alpha_{n}\rightarrow 1 and convergence of 𝐮()\mathbf{u}(\cdot) is uoc, the kkth integral in equality (4) converges to

0t(λ(vk1(s)vk(s))ds((1vk+1(s))d(1vk(s))d))𝑑s.\int_{0}^{t}\left(\lambda(v_{k-1}(s)-v_{k}(s))ds-((1-v_{k+1}(s))^{d}-(1-v_{k}(s))^{d})\right)ds.

Therefore 𝐯()\mathbf{v}(\cdot) satisfies equality (4) with αn=1\alpha_{n}=1 and 𝜺(t)𝟎\mbox{\boldmath$\varepsilon$}(t)\equiv\mbox{\boldmath$0$}. Differential representation of that equality is (5). \bf\Box

Let UoU_{o} denote the set of system states in which average queue occupancy is finite. That is,

Uo={𝐮U:k=1uk<}.U_{o}=\{\mathbf{u}\in U:\sum_{k=1}^{\infty}u_{k}<\infty\}.

Let 𝐯={vk}k=0U\mathbf{v}^{*}=\{v^{*}_{k}\}_{k=0}^{\infty}\in U be defined by setting v0=1v^{*}_{0}=1 and

vk=11λvk1d,k=1,2,.v^{*}_{k}=1-\sqrt[d]{1-\lambda v^{*}_{k-1}},~~~~~~k=1,2,\cdots. (6)

Since 11λvk1dλvk11-\sqrt[d]{1-\lambda v^{*}_{k-1}}\leq\lambda v^{*}_{k-1} it follows that vkλkv^{*}_{k}\leq\lambda^{k}; in particular 𝐯Uo\mathbf{v}^{*}\in U_{o}. It can be readily verified by substitution that 𝐯\mathbf{v}^{*} is an equilibrium point for the differential system (5). The following lemma establishes that 𝐯\mathbf{v}^{*} is the unique stable equilibrium for trajectories that start in UoU_{o}.

Lemma 3.1

Let 𝐯()\mathbf{v}(\cdot) solve the differential system (5) with initial state 𝐯(0)Uo\mathbf{v}(0)\in U_{o}. Then

limtvk(t)=vk,k=1,2,.\lim\limits_{t\rightarrow\infty}v_{k}(t)=v^{*}_{k},~~~~~~~k=1,2,\cdots. (7)

We provide a proof based on the following auxiliary result:

Lemma 3.2

Let 𝐯+()\mathbf{v}^{+}(\cdot) and 𝐯()\mathbf{v}^{-}(\cdot) solve the differential system (5) with respective initial conditions 𝐯+(0),𝐯(0)U\mathbf{v}^{+}(0),\mathbf{v}^{-}(0)\in U such that vk+(0)vk(0)v_{k}^{+}(0)\geq v^{-}_{k}(0) for all kk. Then vk+(t)vk(t)v^{+}_{k}(t)\geq v^{-}_{k}(t) for all kk and all t>0t>0.

Proof. Suppose that the lemma is incorrect and let t>0t>0 be the first instant such that

vk+(t)=vk(t) and ddtvk+(t)<ddtvk(t) for some k.v^{+}_{k}(t)=v^{-}_{k}(t)~~\mbox{ and }~~\frac{d}{dt}v_{k}^{+}(t)<\frac{d}{dt}v^{-}_{k}(t)~~~\mbox{ for some }k.

Let ii be the largest index kk that satisfies this condition at time tt. Then by (5)

ddtvi+(t)ddtvi(t)=λ(vi1+(t)vi1(t))+(1vi+1(t))d(1vi+1+(t))d.\frac{d}{dt}v_{i}^{+}(t)-\frac{d}{dt}v^{-}_{i}(t)=\lambda(v^{+}_{i-1}(t)-v^{-}_{i-1}(t))+(1-v^{-}_{i+1}(t))^{d}-(1-v^{+}_{i+1}(t))^{d}.

The right hand side of this equality is nonnegative due to the choice of tt (since otherwise either the condition vi1+(t)vi1(t)v^{+}_{i-1}(t)\geq v^{-}_{i-1}(t) and or the condition vi+1+(t)vi+1(t)v^{+}_{i+1}(t)\geq v^{-}_{i+1}(t) must be violated before time tt). This contradicts with the definition of tt; therefore no such tt exists and the lemma holds. \bf\Box

Proof of Lemma 7 Let 𝐯+()\mathbf{v}^{+}(\cdot) and 𝐯()\mathbf{v}^{-}(\cdot) be solutions to (5) with respective initial states 𝐯+(0)\mathbf{v}^{+}(0) and 𝐯(0)\mathbf{v}^{-}(0) that are defined by setting vk+(0)=max{vk(0),vk}v^{+}_{k}(0)=\max\{v_{k}(0),v^{*}_{k}\} and vk(0)=min{vk(0),vk}v^{-}_{k}(0)=\min\{v_{k}(0),v^{*}_{k}\} for k=0,1,2,k=0,1,2,\cdots. By Lemma 3.2

vk(t)vk(t),vkvk+(t), for all k,t.v^{-}_{k}(t)~\leq~v_{k}(t),v^{*}_{k}~\leq~v^{+}_{k}(t),~~~\mbox{ for all }k,t. (8)

Equality (5) and definition (6) of 𝐯\mathbf{v}^{*} give

ddti=kvi±(t)\displaystyle\frac{d}{dt}\sum_{i=k}^{\infty}v^{\pm}_{i}(t) =\displaystyle= λvk1±(t)+(1vk±(t))d1\displaystyle\lambda v^{\pm}_{k-1}(t)+(1-v^{\pm}_{k}(t))^{d}-1
=\displaystyle= λ(vk1±(t)vk1)+(1vk±(t))d(1vk)d,\displaystyle\lambda(v^{\pm}_{k-1}(t)-v^{*}_{k-1})+(1-v^{\pm}_{k}(t))^{d}-(1-v^{*}_{k})^{d},

or, in integral form,

i=kvi±(t)i=kvi±(0)=0tλ(vk1±(s)vk1)𝑑s+0t((1vk±(s))d(1vk)d)𝑑s.\sum_{i=k}^{\infty}v^{\pm}_{i}(t)-\sum_{i=k}^{\infty}v^{\pm}_{i}(0)=\int_{0}^{t}\lambda(v^{\pm}_{k-1}(s)-v^{*}_{k-1})ds+\int_{0}^{t}((1-v^{\pm}_{k}(s))^{d}-(1-v^{*}_{k})^{d})ds. (9)

Note that since v1+(t)v1=11λdv^{+}_{1}(t)\geq v^{*}_{1}=1-\sqrt[d]{1-\lambda} it follows that

ddti=1vk+(t)=λ+(1v1+(t))d10.\frac{d}{dt}\sum_{i=1}^{\infty}v^{+}_{k}(t)~=~\lambda+(1-v^{+}_{1}(t))^{d}-1~\leq~0.

Hence i=kvi+(t)\sum_{i=k}^{\infty}v^{+}_{i}(t), and therefore i=kvi(t)\sum_{i=k}^{\infty}v^{-}_{i}(t), is bounded by k=1vk+(0)\sum_{k=1}^{\infty}v_{k}^{+}(0) uniformly for all tt. In turn equality (9) yields

|0tλ(vk1±(s)vk1)𝑑s+0t((1vk±(s))d(1vk)d)𝑑s|i=kvi+(0).\left|~\int_{0}^{t}\lambda(v^{\pm}_{k-1}(s)-v^{*}_{k-1})ds+\int_{0}^{t}((1-v^{\pm}_{k}(s))^{d}-(1-v^{*}_{k})^{d})ds~\right|~\leq~\sum_{i=k}^{\infty}v^{+}_{i}(0).

The bound on the right hand side is finite since 𝐯+(0)Uo\mathbf{v}^{+}(0)\in U_{o} due to the hypothesis 𝐯(0)Uo\mathbf{v}(0)\in U_{o}. Note that, owing to the inequality (8), neither one of the two integrands above changes sign. Hence if the first integral converges as tt\rightarrow\infty then so does the second one, implying further that

limtvk±(t)=vk.\lim_{t\rightarrow\infty}v^{\pm}_{k}(t)=v^{*}_{k}. (10)

Since v0±(t)v0=1v^{\pm}_{0}(t)\equiv v^{*}_{0}=1, this is clearly the case for k=1k=1. Induction on kk confirms that equality (10) holds for all kk. The desired conclusion (7) now follows from the property (8). \bf\Box

Theorem 5, which establishes convergence over finite time intervals, is complemented next by showing that equilibrium distribution of 𝐮()\mathbf{u}(\cdot) converges as nn\rightarrow\infty to the deterministic measure concentrated at 𝐯\mathbf{v}^{*}.

Theorem 3.2

The process 𝐮()\mathbf{u}(\cdot) is ergodic. Let 𝐮={uk}k=0\mathbf{u}^{*}=\{u^{*}_{k}\}_{k=0}^{\infty} denote the equilibrium random variable. For ε>0\varepsilon>0

limnPn(ρ(𝐮,𝐯)>ε)=0.\lim_{n\rightarrow\infty}P_{n}(\rho(\mathbf{u}^{*},\mathbf{v}^{*})>\varepsilon)=0. (11)

In particular limnEn[uk]=vk\lim_{n\rightarrow\infty}E_{n}[u^{*}_{k}]=v^{*}_{k} for k=0,1,2,k=0,1,2,\cdots.

Proof. Let Un={𝐮U:nuk+ for k0}U_{n}=\{\mathbf{u}\in U:nu_{k}\in\mathbb{Z}_{+}\mbox{ for }k\geq 0\}. Note that (𝐮(t):t0)(\mathbf{u}(t):t\geq 0) is irreducible in UnU_{n} and UnU_{n} is compact; therefore the process is ergodic and has a unique equilibrium distribution concentrated on UnU_{n}. In that equilibrium the rate of arrivals to queues with occupancy kk or higher should be equal to the rate of departures from such queues. That is,

En[λuk1]=1En[(1uk)d]En[uk], for k1,E_{n}[\lambda u^{*}_{k-1}]~=~1-E_{n}[(1-u^{*}_{k})^{d}]~\geq~E_{n}[u^{*}_{k}],~~~\mbox{ for }k\geq 1,

where the inequality follows since (1uk)d(1uk)(1-u^{*}_{k})^{d}\leq(1-u^{*}_{k}). Therefore En[uk]λkE_{n}[u^{*}_{k}]\leq\lambda^{k} and in turn En[k=1uk]λ/(1λ)E_{n}[\sum_{k=1}^{\infty}u^{*}_{k}]\leq\lambda/(1-\lambda). Let Uo,λ{𝐮U:k=1ukλ/(1λ)}U_{o,\lambda}\triangleq\{\mathbf{u}\in U:\sum_{k=1}^{\infty}u_{k}\leq\lambda/(1-\lambda)\} so that Pn(𝐮Uo,λ)=1P_{n}(\mathbf{u}^{*}\in U_{o,\lambda})=1.

Suppose that (11) is false so that for some infinite subsequence {n}\{n^{\prime}\} of {n}\{n\} and some δ>0\delta>0

Pn(ρ(𝐮,𝐯)>ε)>δ.P_{n^{\prime}}(\rho(\mathbf{u}^{*},\mathbf{v}^{*})>\varepsilon)>\delta. (12)

Due to Lemma 7 and compactness of Uo,λU_{o,\lambda} one can choose t(ε)t(\varepsilon) such that if 𝐯(0)Uo,λ\mathbf{v}(0)\in U_{o,\lambda} then ρ(𝐯(t),𝐯)<ε/2\rho(\mathbf{v}(t),\mathbf{v}^{*})<\varepsilon/2 for tt(ε)t\geq t(\varepsilon). Let 𝐮(0)\mathbf{u}(0) have the same distribution as 𝐮\mathbf{u}^{*} and let 𝐯(0)=𝐮(0)\mathbf{v}(0)=\mathbf{u}(0). By Theorem 5 there is a further subsequence {n′′}\{n^{\prime\prime}\} of {n}\{n^{\prime}\} such that Pn′′(ρ(𝐮(t(ε)),𝐯(t(ε)))>ε/2)<δP_{n^{\prime\prime}}(\rho(\mathbf{u}(t(\varepsilon)),\mathbf{v}(t(\varepsilon)))>\varepsilon/2)<\delta whenever n′′n^{\prime\prime} is large enough. Since Pn′′(𝐮Uo,λ)=1P_{n^{\prime\prime}}(\mathbf{u}^{*}\in U_{o,\lambda})=1, the choice of t(ε)t(\varepsilon) implies that Pn′′(ρ(𝐮(t(ε)),𝐯)>ε)<δP_{n^{\prime\prime}}(\rho(\mathbf{u}(t(\varepsilon)),\mathbf{v}^{*})>\varepsilon)<\delta for those values of n′′n^{\prime\prime}. However 𝐮(t(ε))\mathbf{u}(t(\varepsilon)) and 𝐮\mathbf{u}^{*} have identical distributions as the latter is in equilibrium; leading to a contradiction with (12). Hence no sequence {n}\{n^{\prime}\} and constant δ>0\delta>0 satisfy (12); so (11) holds. By definition of ρ\rho equality (11) implies that each entry uku^{*}_{k} of 𝐮\mathbf{u}^{*} converges in probability to the constant vkv^{*}_{k}; since 0uk10\leq u^{*}_{k}\leq 1, so does En[uk]E_{n}[u_{k}^{*}]. \bf\Box

We conclude the discussion of LCQ(dd) with a relationship between dd and the tail probabilities of equilibrium queue occupancy:

Theorem 3.3

vk=Θ((λ/d)k)v^{*}_{k}=\Theta((\lambda/d)^{k}) as kk\rightarrow\infty.

Proof. The assertion is immediate for d=1d=1 so we consider the case d>1d>1. Equality (6), together with Taylor expansion of 1xd\sqrt[d]{1-x} around x=0x=0 yields

vk=λdvk1+1di=2(λvk1)ii!j=1i1(j1d),k=1,2,.v^{*}_{k}=\frac{\lambda}{d}v^{*}_{k-1}+\frac{1}{d}\sum_{i=2}^{\infty}\frac{(\lambda v^{*}_{k-1})^{i}}{i!}\prod_{j=1}^{i-1}(j-\frac{1}{d}),~~~~~~~~~k=1,2,\cdots. (13)

The second term on the right hand side is nonnegative; therefore vk(λ/d)vkv^{*}_{k}\geq(\lambda/d)v^{*}_{k} and

lim infkvk(λ/d)k1.\liminf\limits_{k\rightarrow\infty}\frac{v^{*}_{k}}{(\lambda/d)^{k}}\geq 1. (14)

We define ckvk/(λ/d)kc_{k}\triangleq v^{*}_{k}/(\lambda/d)^{k} and complete the proof by showing that {ck}k=0\{c_{k}\}_{k=0}^{\infty} is uniformly bounded. Let βk\beta_{k} be defined as

βk1+i=1(λvk1)i2(i+2)!j=2i+1(j1d)<1+i=1(λvk1)i\beta_{k}~\triangleq~1+\sum_{i=1}^{\infty}(\lambda v^{*}_{k-1})^{i}\frac{2}{(i+2)!}\prod_{j=2}^{i+1}(j-\frac{1}{d})~<~1+\sum_{i=1}^{\infty}(\lambda v^{*}_{k-1})^{i} (15)

so that equality (13) can be rearranged as

vk=λdvk1+(λvk1)22d(11d)βk.v^{*}_{k}=\frac{\lambda}{d}v^{*}_{k-1}+\frac{(\lambda v^{*}_{k-1})^{2}}{2d}(1-\frac{1}{d})\beta_{k}. (16)

Since vkλkv^{*}_{k}\leq\lambda^{k} there exists a finite kok_{o} such that vk1<1/dv_{k-1}^{*}<1/d for kkok\geq k_{o}. The bound in (15) implies that βk<1/(1λ/d)<d/(d1)\beta_{k}<1/(1-\lambda/d)<d/(d-1) for such kk; in turn by (16)

ck<ck1+ck12d2(λ/d)k,k>ko.c_{k}<c_{k-1}+c_{k-1}^{2}\frac{d}{2}(\lambda/d)^{k},~~~~~~~~~k>k_{o}.

It can be verified by induction on k>kok>k_{o} that

ck<cko(1+λ+λ2++λkko):c_{k}<c_{k_{o}}(1+\lambda+\lambda^{2}+\cdots+\lambda^{k-k_{o}}): (17)

Namely, if (17) holds for kk then it holds also for k+1k+1 if cko(1+λ+λ2++λkko)2λko/(2dk)<1c_{k_{o}}(1+\lambda+\lambda^{2}+\cdots+\lambda^{k-k_{o}})^{2}\lambda^{k_{o}}/(2d^{k})<1. This latter condition can be verified based on the bound cko(λ/d)ko=vko<1/dc_{k_{o}}(\lambda/d)^{k_{o}}=v_{k_{o}}<1/d, which follows from the definition of kok_{o}. Inequality (17) implies the uniform bound ck<cko/(1λ)c_{k}<c_{k_{o}}/(1-\lambda); therefore

lim supkvk(λ/d)k<cko1λ.\limsup\limits_{k\rightarrow\infty}\frac{v^{*}_{k}}{(\lambda/d)^{k}}<\frac{c_{k_{o}}}{1-\lambda}. (18)

The theorem follows due to (14) and (18). \bf\Box

4 LCQ

Given 𝐦(τ)\mathbf{m}(\tau) at a scheduling instant τ\tau, the maximum occupancy over all connected queues at time τ\tau is equal to k=1,2,k=1,2,\cdots with probability (1q)mk+1(τ)(1q)mk(τ)(1-q)^{m_{k+1}(\tau)}-(1-q)^{m_{k}(\tau)}; hence under the LCQ policy 𝐮()\mathbf{u}(\cdot) is a time-homogenous Markov process with jump rates

𝐮{𝐮+n1𝐞k at rate nλ(uk1uk)𝐮n1𝐞k at rate n((1q)nuk+1(1q)nuk).\mathbf{u}\leftarrow\left\{\begin{array}[]{ll}\mathbf{u}+n^{-1}\mathbf{e}_{k}&\mbox{ at rate }~~n\lambda(u_{k-1}-u_{k})\\ \mathbf{u}-n^{-1}\mathbf{e}_{k}&\mbox{ at rate }~~n\left((1-q)^{nu_{k+1}}-(1-q)^{nu_{k}}\right).\end{array}\right. (19)

This process can be constructed as in Section 3, so that

uk(t)=uk(0)+0t(λ(uk1(s)uk(s))(1q)mk+1(s)+(1q)mk(s))𝑑s+εk(t)u_{k}(t)=u_{k}(0)+\int_{0}^{t}\left(\lambda(u_{k-1}(s)-u_{k}(s))-(1-q)^{m_{k+1}(s)}+(1-q)^{m_{k}(s)}\right)ds+\varepsilon_{k}(t) (20)

where εk()\varepsilon_{k}(\cdot) is a martingale that vanishes as nn\rightarrow\infty. The sequence of processes 𝐮():n=1,2,\mathbf{u}(\cdot):n=1,2,\cdots converges in distribution along subsequences of {n}\{n\}, but identifying a limit is relatively more involved than for the LCQ(d)(d) policy since the process 𝐦()\mathbf{m}(\cdot) fluctuates persistently for all values of nn and the integrand in (20) does not converge. Rather than this integrand, here we study the behavior of the integral in (20) via an averaging technique due to Kurtz [5]. In reading this section the reader may find it helpful to consult related applications of this technique in [1, 4, 13].

Let Ω\Omega denote the set of sequences 𝝎={ωk}k=0\mbox{\boldmath$\omega$}=\{\omega_{k}\}_{k=0}^{\infty} such that ωk+{}\omega_{k}\in\mathbb{Z}_{+}\cup\{\infty\}, ω0=\omega_{0}=\infty, and ωkωk+1\omega_{k}\geq\omega_{k+1}. Define the mapping h:Ω[0,1]h:\Omega\mapsto[0,1]^{\infty} by setting

h(𝝎)={(1+ωk)1}k=0,𝝎Ω,h(\mbox{\boldmath$\omega$})=\{(1+\omega_{k})^{-1}\}_{k=0}^{\infty},~~~~~\mbox{\boldmath$\omega$}\in\Omega,

with the understanding that 1+=1+\infty=\infty and 1/=01/\infty=0. Let Ω\Omega be endowed with metric ρo\rho_{o} defined by

ρo(𝝎,𝝎)=ρ(h(𝝎),h(𝝎)),𝝎,𝝎Ω.\rho_{o}(\mbox{\boldmath$\omega$},\mbox{\boldmath$\omega$}^{\prime})=\rho(h(\mbox{\boldmath$\omega$}),h(\mbox{\boldmath$\omega$}^{\prime})),~~~~~\mbox{\boldmath$\omega$},\mbox{\boldmath$\omega$}^{\prime}\in\Omega.

In particular Ω\Omega is compact with respect to the induced topology. We denote by \mathcal{L} the collection of measures μ\mu on the product space [0,)×Ω[0,\infty)\times\Omega such that μ([0,t)×Ω)=t\mu([0,t)\times\Omega)=t for each t>0t>0. Let \mathcal{L} be endowed with the topology corresponding to weak convergence of measures restricted to [0,t)×Ω[0,t)\times\Omega for each tt. Since Ω\Omega is compact, so is \mathcal{L} due to Prohorov’s Theorem.

Let ξ\xi be a random member of \mathcal{L} defined by

ξ([0,t)×A)=0t1{𝐦(s)A}𝑑s,t>0,A(Ω).\xi([0,t)\times A)=\int_{0}^{t}1\{\mathbf{m}(s)\in A\}ds,~~~t>0,~A\in\mathcal{B}(\Omega).

Here (Ω)\mathcal{B}(\Omega) denotes Borel sets of Ω\Omega. Note that equality (20) can be expressed in terms of ξ\xi as

uk(t)=uk(0)+0tλ(uk1(s)uk(s))𝑑sϕk+1(t)+ϕk(t)+εk(t)u_{k}(t)~=~u_{k}(0)+\int_{0}^{t}\lambda(u_{k-1}(s)-u_{k}(s))ds-\phi_{k+1}(t)+\phi_{k}(t)+\varepsilon_{k}(t)

where

ϕk(t)0t(1q)mk(s)𝑑s=[0,t)×Ω(1q)ωkξ(ds×d𝝎).\phi_{k}(t)~\triangleq~\int_{0}^{t}(1-q)^{m_{k}(s)}ds~=~\int_{[0,t)\times\Omega}(1-q)^{\omega_{k}}\xi(ds\times d\mbox{\boldmath$\omega$}).

Compactness of \mathcal{L} implies that each subsequence of {n}\{n\} has a further subsequence along which ξ\xi converges in distribution. This property is also possessed by 𝐮()\mathbf{u}(\cdot), and therefore by the pair (𝐮(),ξ)(\mathbf{u}(\cdot),\xi).

The following definition is useful in characterizing possible limits of (𝐮(),ξ)(\mathbf{u}(\cdot),\xi): For fixed 𝐮U\mathbf{u}\in U let ω𝐮()\omega^{\mathbf{u}}(\cdot) denote the Markov process with states in Ω\Omega and with the following transition rates:

ω𝐮{ω𝐮+𝐞k at rate λ(uk1uk)ω𝐮𝐞k at rate (1q)ωk+1𝐮(1q)ωk𝐮.\omega^{\mathbf{u}}\leftarrow\left\{\begin{array}[]{ll}\omega^{\mathbf{u}}+\mathbf{e}_{k}&\mbox{ at rate }\lambda(u_{k-1}-u_{k})\\ \omega^{\mathbf{u}}-\mathbf{e}_{k}&\mbox{ at rate }(1-q)^{\omega^{\mathbf{u}}_{k+1}}-(1-q)^{\omega^{\mathbf{u}}_{k}}.\end{array}\right. (21)

See Figure 3 for a partial illustration of this process. The process ω𝐮()\omega^{\mathbf{u}}(\cdot) bears a certain resemblance to 𝐦()=n𝐮()\mathbf{m}(\cdot)=n\mathbf{u}(\cdot), which can be observed by inspecting the generators (19) and (21), though it should be noted that in (21) 𝐮={uk}k=0\mathbf{u}=\{u_{k}\}_{k=0}^{\infty} is a constant and has no binding to instantaneous values of ω𝐮()\omega^{\mathbf{u}}(\cdot). We also point out that ω𝐮()\omega^{\mathbf{u}}(\cdot) evolves on a compactified state space and it is reducible due to the states that involve \infty; hence it has multiple equilibrium distributions in general.

Refer to caption
Figure 3: Transition rates of ωk𝐮()\omega^{\mathbf{u}}_{k}(\cdot), that is, the kkth coordinate of ω𝐮()\omega^{\mathbf{u}}(\cdot). The process has also an isolated state \infty which is not shown. The coordinate process ωk𝐮()\omega^{\mathbf{u}}_{k}(\cdot) is generally not Markovian due to its dependence on ωk+1𝐮()\omega^{\mathbf{u}}_{k+1}(\cdot).
Theorem 4.1

Let (𝐯(),χ)(\mathbf{v}(\cdot),\chi) be the limit of (𝐮(),ξ)(\mathbf{u}(\cdot),\xi) along a convergent subsequence of {n}\{n\}.

a) The limit measure χ\chi satisfies

χ([0,t)×A)=0tπ𝐯(s)(A)𝑑s,t>0,A(Ω),\chi([0,t)\times A)=\int_{0}^{t}\pi_{\mathbf{v}(s)}(A)ds,~~~~~~~~~t>0,~A\in\mathcal{B}(\Omega),

where, for each s>0s>0, π𝐯(s)\pi_{\mathbf{v}(s)} is an equilibrium distribution for the process ω𝐯(s)()\omega^{\mathbf{v}(s)}(\cdot) such that

π𝐯(s)(ωk=)=1 if vk(s)>0.\pi_{\mathbf{v}(s)}\left(\omega_{k}=\infty\right)=1~~\mbox{ if }~~v_{k}(s)>0.

b) The limit trajectory 𝐯()\mathbf{v}(\cdot) satisfies

ddtvk(t)=λ(vk1(t)vk(t))Eπ𝐯(t)[(1q)ωk+1(1q)ωk],\hskip 0.0pt\frac{d}{dt}v_{k}(t)=\lambda(v_{k-1}(t)-v_{k}(t))-E_{\pi_{\mathbf{v}(t)}}[(1-q)^{\omega_{k+1}}-(1-q)^{\omega_{k}}], (22)

where k=1,2,k=1,2,\cdots and Eπ𝐯(t)E_{\pi_{\mathbf{v}(t)}} denotes expectation with respect to distribution π𝐯(t)\pi_{\mathbf{v}(t)}.

Proof. Let all processes be constructed on a common probability space so that convergence of (𝐮(),ξ)(\mathbf{u}(\cdot),\xi) is almost sure. Convergence of 𝐮()\mathbf{u}(\cdot) is then uoc. We start by consulting [5, Lemma 1.4] to verify that the limit measure χ\chi possesses a density so that

χ([0,t)×A)=0tγs(A)𝑑s,t>0,A(Ω),\chi([0,t)\times A)=\int_{0}^{t}\gamma_{s}(A)ds,~~~~~~~~~t>0,~A\in\mathcal{B}(\Omega), (23)

where, for each ss, γs\gamma_{s} is a probability distribution on Ω\Omega. We proceed by identifying these distributions.

Let \mathcal{F} denote the collection of bounded continuous functions f:Ωf:\Omega\mapsto\mathbb{R} such that f(𝝎)f(\mbox{\boldmath$\omega$}) depends on a finite number of entries in the sequence 𝝎={ωk}k=0Ω\mbox{\boldmath$\omega$}=\{\omega_{k}\}_{k=0}^{\infty}\in\Omega. Given ff\in\mathcal{F} define the function Gf:Ω×UG^{f}:\Omega\times U\mapsto\mathbb{R} by setting

Gf(𝝎,𝐮)k=1(f(𝝎+𝐞k)f(𝝎))λ(uk1uk)+(f(𝝎𝐞k)f(𝝎))((1q)ωk+1(1q)ωk)G^{f}(\mbox{\boldmath$\omega$},\mathbf{u})\triangleq\sum_{k=1}^{\infty}(f(\mbox{\boldmath$\omega$}+\mathbf{e}_{k})-f(\mbox{\boldmath$\omega$}))\lambda(u_{k-1}-u_{k})+(f(\mbox{\boldmath$\omega$}-\mathbf{e}_{k})-f(\mbox{\boldmath$\omega$}))((1-q)^{\omega_{k+1}}-(1-q)^{\omega_{k}})

for each 𝝎Ω\mbox{\boldmath$\omega$}\in\Omega and 𝐮={uk}k=0U\mathbf{u}=\{u_{k}\}_{k=0}^{\infty}\in U. GfG^{f} is continuous due to the continuity of ff, and continuity of GfG^{f} is uniform since the product space Ω×U\Omega\times U is compact.

The process f(𝐦())f(\mathbf{m}(\cdot)) satisfies at each instant tt

f(𝐦(t))f(𝐦(0))\displaystyle\!\!\!\!\!\!\!\!\!\!\!f(\mathbf{m}(t))-f(\mathbf{m}(0))
=0tk=1(f(𝐦(s)+𝐞k)f(𝐦(s)))dAk1(0snλ(uk1(τ)uk(τ))𝑑τ)\displaystyle~~~~~~~~~=~~\int_{0}^{t}\sum_{k=1}^{\infty}(f(\mathbf{m}(s)+\mathbf{e}_{k})-f(\mathbf{m}(s)))dA_{k-1}\left(\int_{0}^{s}n\lambda(u_{k-1}(\tau)-u_{k}(\tau))d\tau\right)
+0tk=1(f(𝐦(s)𝐞k)f(𝐦(s)))dDk(0sn((1q)mk+1(τ)(1q)mk(τ))𝑑τ)\displaystyle~~~~~~~~~~~~~+\int_{0}^{t}\sum_{k=1}^{\infty}(f(\mathbf{m}(s)-\mathbf{e}_{k})-f(\mathbf{m}(s)))dD_{k}\left(\int_{0}^{s}n\left((1-q)^{m_{k+1}(\tau)}-(1-q)^{m_{k}(\tau)}\right)d\tau\right)
=n0tk=1(f(𝐦(s)+𝐞k)f(𝐦(s)))λ(uk1(s)uk(s))ds\displaystyle~~~~~~~~~=~~n\int_{0}^{t}\sum_{k=1}^{\infty}(f(\mathbf{m}(s)+\mathbf{e}_{k})-f(\mathbf{m}(s)))\lambda(u_{k-1}(s)-u_{k}(s))ds
+n0tk=1(f(𝐦(s)𝐞k)f(𝐦(s)))((1q)mk+1(s)(1q)mk(s))ds+μf(t)\displaystyle~~~~~~~~~~~~~+n\int_{0}^{t}\sum_{k=1}^{\infty}(f(\mathbf{m}(s)-\mathbf{e}_{k})-f(\mathbf{m}(s)))\left((1-q)^{m_{k+1}(s)}-(1-q)^{m_{k}(s)}\right)ds+\mu^{f}(t)
=n0tGf(𝐦(s),𝐮(s))𝑑s+μf(t),\displaystyle~~~~~~~~~=~~n\int_{0}^{t}G^{f}(\mathbf{m}(s),\mathbf{u}(s))ds+\mu^{f}(t),

where μf()\mu^{f}(\cdot) is a square-integrable martingale. Rearranging the last equality and expressing the integral there in terms of the random measure ξ\xi yields

[0,t)×ΩGf(𝝎,𝐮(s))ξ(ds×d𝝎)=0tGf(𝐦(s),𝐮(s))ds=(f(𝐦(t))f(𝐦(0))/n+μf(t)/n.\int_{[0,t)\times\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{u}(s))\xi(ds\times d\mbox{\boldmath$\omega$})~~=~~\int_{0}^{t}G^{f}(\mathbf{m}(s),\mathbf{u}(s))ds~~=~~(f(\mathbf{m}(t))-f(\mathbf{m}(0))/n+\mu^{f}(t)/n.

Since ff is bounded, the first term on the right hand side vanishes as nn\rightarrow\infty. The martingale μf()\mu^{f}(\cdot) has bounded jumps; in turn by Doob’s L2L^{2} inequality μf(t)/n\mu^{f}(t)/n also vanishes. Therefore

[0,t)×ΩGf(𝝎,𝐮(s))ξ(ds×d𝝎)0.\int_{[0,t)\times\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{u}(s))\xi(ds\times d\mbox{\boldmath$\omega$})\rightarrow 0. (24)

Since 𝐮()\mathbf{u}(\cdot) converges uoc to 𝐯()\mathbf{v}(\cdot) by hypothesis, uniform continuity of GfG^{f} implies

|[0,t)×ΩGf(𝝎,𝐮(s))ξ(ds×d𝝎)[0,t)×ΩGf(𝝎,𝐯(s))ξ(ds×d𝝎)|0.\left|\int_{[0,t)\times\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{u}(s))\xi(ds\times d\mbox{\boldmath$\omega$})-\int_{[0,t)\times\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{v}(s))\xi(ds\times d\mbox{\boldmath$\omega$})\right|~\rightarrow~0. (25)

Finally by the Continuous Mapping Theorem

[0,t)×ΩGf(𝝎,𝐯(s))ξ(ds×d𝝎)[0,t)×ΩGf(𝝎,𝐯(s))χ(ds×d𝝎).\int_{[0,t)\times\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{v}(s))\xi(ds\times d\mbox{\boldmath$\omega$})\rightarrow\int_{[0,t)\times\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{v}(s))\chi(ds\times d\mbox{\boldmath$\omega$}). (26)

Observations (24)–(26) lead to

[0,t)×ΩGf(𝝎,𝐯(s))χ(ds×d𝝎)=0t𝝎ΩGf(𝝎,𝐯(s))γs(𝝎)ds=0,\int_{[0,t)\times\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{v}(s))\chi(ds\times d\mbox{\boldmath$\omega$})~=~\int_{0}^{t}\sum_{\mbox{\boldmath$\omega$}\in\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{v}(s))\gamma_{s}(\mbox{\boldmath$\omega$})ds~=~0,

where the left equality is due to (23). This equality holds for all t>0t>0; therefore

𝝎ΩGf(𝝎,𝐯(s))γs(𝝎)=0\sum_{\mbox{\boldmath$\omega$}\in\Omega}G^{f}(\mbox{\boldmath$\omega$},\mathbf{v}(s))\gamma_{s}(\mbox{\boldmath$\omega$})=0

for almost all s>0s>0. Since ff\in\mathcal{F} is arbitrary (note that \mathcal{F} is dense in continuous bounded functions on Ω\Omega) [2, Proposition 4.9.2] implies that γs\gamma_{s} is an equilibrium distribution for the process 𝝎𝐯(s)()\mbox{\boldmath$\omega$}^{\mathbf{v}(s)}(\cdot).

Let ε>0\varepsilon>0 and [t0,t1][t_{0},t_{1}] be an interval such that vk(t)εv_{k}(t)\geq\varepsilon for t[t0,t1]t\in[t_{0},t_{1}]. Since vk(t)v_{k}(t) is the limit of uk(t)=n1mk(t)u_{k}(t)=n^{-1}m_{k}(t), for any given integer BB

ξ([t0,t1]×{0,1,2,,B})=t0t11{mk(s)B}𝑑s0 as n.\xi([t_{0},t_{1}]\times\{0,1,2,\cdots,B\})~=~\int_{t_{0}}^{t_{1}}1\{m_{k}(s)\leq B\}ds~\rightarrow~0~~~~~\mbox{ as }n\rightarrow\infty.

Hence χ([t0,t1]×+)=0\chi([t_{0},t_{1}]\times\mathbb{Z}_{+})=0 due to the arbitrariness of BB. Since ε\varepsilon can be chosen arbitrarily small it follows that γt(+)=0\gamma_{t}(\mathbb{Z}_{+})=0 for almost all tt such that vk(t)>0v_{k}(t)>0. This completes the proof of part a). Part b) follows from equality (20) since

0t(uk1(s)uk(s))𝑑s0t(vk1(s)vk(s))𝑑s\int_{0}^{t}(u_{k-1}(s)-u_{k}(s))ds\rightarrow\int_{0}^{t}(v_{k-1}(s)-v_{k}(s))ds

due to uoc convergence of 𝐮()\mathbf{u}(\cdot), and

ϕk(t)[0,t)×Ω(1q)ωkχ(ds×d𝝎)=0tEγs[(1q)wk]𝑑s\phi_{k}(t)~\rightarrow~\int_{[0,t)\times\Omega}(1-q)^{\omega_{k}}\chi(ds\times d\mbox{\boldmath$\omega$})~=~\int_{0}^{t}E_{\gamma_{s}}[(1-q)^{w_{k}}]ds

due to the Continuous Mapping Theorem. \bf\Box

Theorem 4.1 explains the extent of the disparity between time scales of two processes, namely 𝐦()\mathbf{m}(\cdot) and its normalized version 𝐮()\mathbf{u}(\cdot): The process 𝐦()\mathbf{m}(\cdot) displays far larger variation than its normalized version, so that, in the limit of large nn, 𝐦()\mathbf{m}(\cdot) settles to equilibrium before 𝐮()\mathbf{u}(\cdot) changes its value. In particular integral of a binary-valued measurable function of 𝐦()\mathbf{m}(\cdot) is well-approximated by integrating an appropriate equilibrium probability. Provided that 𝐮(t)\mathbf{u}(t) remains close to 𝐯(t)\mathbf{v}(t), the process 𝝎𝐯(t)()\mbox{\boldmath$\omega$}^{\mathbf{v}(t)}(\cdot) mimics a slowed-down version of 𝐦()\mathbf{m}(\cdot) observed around time tt; hence the alluded equilibrium distribution pertains to 𝝎𝐯(t)()\mbox{\boldmath$\omega$}^{\mathbf{v}(t)}(\cdot).

Specification of 𝝎𝐯(t)()\mbox{\boldmath$\omega$}^{\mathbf{v}(t)}(\cdot) requires inclusion of \infty since entries of 𝐦()\mathbf{m}(\cdot) can be as large as nn. Compactifying the augmented state-space Ω\Omega of 𝐦()\mathbf{m}(\cdot) via choice of the metric ρo\rho_{o} leads to the representation (22) of a limit trajectory 𝐯()\mathbf{v}(\cdot), but it also entails ambiguity in that representation. Namely, Theorem 4.1 does not specify which equilibrium distribution for ω𝐯(t)()\omega^{\mathbf{v}(t)}(\cdot) should be adopted in (22). While a full account of equilibrium distributions of ω𝐯(t)()\omega^{\mathbf{v}(t)}(\cdot) appears difficult, an important feature of the right distribution can be identified:

Lemma 4.1

Let 𝐯()\mathbf{v}(\cdot) and π𝐯()\pi_{\mathbf{v}(\cdot)} be as specified by Theorem 4.1. Given k=1,2,k=1,2,\cdots

π𝐯(t)(ωk+ and ωk+1=0)=1\pi_{\mathbf{v}(t)}\left(\omega_{k}\in\mathbb{Z}_{+}\mbox{ and }\omega_{k+1}=0\right)=1

for almost all tt such that vk(t)=0v_{k}(t)=0.

Lemma 4.1 will be instrumental in obtaining a sharper description for 𝐯()\mathbf{v}(\cdot), yet an informal explanation may still be useful in putting it in perspective with the queueing system of interest. Note that if vk(t)=0v_{k}(t)=0 and vk1(t)>0v_{k-1}(t)>0 then 𝐯(t)\mathbf{v}(t) reflects a distribution with support {0,1,,k1}\{0,1,\cdots,k-1\}. This property does not immediately translate into a bound on the maximum queue length in the system, since 𝐯(t)\mathbf{v}(t) is the limit of 𝐮(t)=n1𝐦(t)\mathbf{u}(t)=n^{-1}\mathbf{m}(t) and so the number of queues with at least iki\geq k packets, mi(t)m_{i}(t), is o(n)o(n) as nn\rightarrow\infty. By way of interpreting 𝝎𝐯(t)()\mbox{\boldmath$\omega$}^{\mathbf{v}(t)}(\cdot) as a proxy to 𝐦()\mathbf{m}(\cdot) around time tt, Lemma 4.1 indicates that the maximum queue size is at most one larger than what is deduced from 𝐯(t)\mathbf{v}(t) and that the number of maximal queues is O(1)O(1) as nn\rightarrow\infty.

Proof of Lemma 4.1 Let [t0,t1][t_{0},t_{1}] be an interval such that vk(t)=0v_{k}(t)=0 for t[t0,t1]t\in[t_{0},t_{1}]. We prove the lemma by showing that as nn\rightarrow\infty along the convergent subsequence of interest

ξ([t0,t1]×{𝝎:wk+1=0})=t0t11{mk+1(t)1}𝑑t\displaystyle\xi([t_{0},t_{1}]\times\{\mbox{\boldmath$\omega$}:w_{k+1}=0\})~=~\int_{t_{0}}^{t_{1}}1\{m_{k+1}(t)\geq 1\}dt \displaystyle\rightarrow 0,\displaystyle 0, (27)
ξ([t0,t1]×{𝝎:wkZ+})=t0t11{mk(t)Z+}𝑑t\displaystyle\xi([t_{0},t_{1}]\times\{\mbox{\boldmath$\omega$}:w_{k}\in Z_{+}\})~=~\int_{t_{0}}^{t_{1}}1\{m_{k}(t)\in Z_{+}\}dt \displaystyle\rightarrow t1t0.\displaystyle t_{1}-t_{0}. (28)

For each integer ll and time tt let sl(t)i=lmi(t)s_{l}(t)\triangleq\sum_{i=l}^{\infty}m_{i}(t). This quantity increases when some queue with size at least l1l-1 receives a packet, and it decreases when transmission is scheduled from some queue with size at least ll. Given 𝐮(t)\mathbf{u}(t), these events occur at respective instantaneous rates nλul1(t)n\lambda u_{l-1}(t) and n(1(1q)ml(t))n\left(1-(1-q)^{m_{l}(t)}\right). Therefore

En[sl(t1)sl(t0)]=nEn[t0t1λul1(t)(1(1q)ml(t))dt].E_{n}[s_{l}(t_{1})-s_{l}(t_{0})]~=~nE_{n}\left[\int_{t_{0}}^{t_{1}}\lambda u_{l-1}(t)-\left(1-(1-q)^{m_{l}(t)}\right)dt\right]. (29)

Consider this equality for l=k+1l=k+1. By choice of the interval [t0,t1][t_{0},t_{1}]

n1En[sk+1(t)]i=k+1vi(t)=0n^{-1}E_{n}[s_{k+1}(t)]~\rightarrow~\sum_{i=k+1}^{\infty}v_{i}(t)~=~0

and uk(t)vk(t)=0u_{k}(t)\rightarrow v_{k}(t)=0 for all t[t0,t1]t\in[t_{0},t_{1}]. Consequently

En[t0t1(1(1q)mk+1(t))𝑑t]0.E_{n}\left[\int_{t_{0}}^{t_{1}}\left(1-(1-q)^{m_{k+1}(t)}\right)dt\right]\rightarrow 0.

This leads to (27) since

1(1q)mk+1(t)q1{mk+1(t)1}.1-(1-q)^{m_{k+1}(t)}~\geq~q1\{m_{k+1}(t)\geq 1\}.

To complete the proof, note that n1En[sk(t)]0n^{-1}E_{n}[s_{k}(t)]\rightarrow 0 for all t[t0,t1]t\in[t_{0},t_{1}]; therefore (29) evaluated at l=kl=k implies that for any open subset B[t0,t1]B\subset[t_{0},t_{1}]

lim supnEn[B(1(1q)mk(t))𝑑t]=lim supnEn[Bλuk1(t)𝑑t]<B𝑑t.\limsup_{n\rightarrow\infty}E_{n}\left[\int_{B}\left(1-(1-q)^{m_{k}(t)}\right)dt\right]~=~\limsup_{n\rightarrow\infty}E_{n}\left[\int_{B}\lambda u_{k-1}(t)dt\right]~<~\int_{B}dt.

The last inequality is strict since λ<1\lambda<1. Arbitrariness of BB implies (28). \bf\Box

Given positive integer KK let UK={𝐮U:uk=0 for kK}U_{K}=\{\mathbf{u}\in U:u_{k}=0\mbox{ for }k\geq K\}.

Theorem 4.2

Let 𝐯()\mathbf{v}(\cdot) and π𝐯()\pi_{\mathbf{v}(\cdot)} be as specified by Theorem 4.1 with initial state 𝐯(0)UK\mathbf{v}(0)\in U_{K} for some KK. Then for t>0t>0

a) 𝐯(t)UK\mathbf{v}(t)\in U_{K} and

π𝐯(t)(ωK(t)+ and ωK(t)+1=0)=1\pi_{\mathbf{v}(t)}\left(\omega_{K(t)}\in\mathbb{Z}_{+}\mbox{ and }\omega_{K(t)+1}=0\right)=1

where K(t)=min{k:vj(t)=0 for jk}K(t)=\min\{k:v_{j}(t)=0\mbox{ for }j\geq k\}.

b)

ddtvk(t)={λvk1(t)1<0 if k=K(t)10 if kK(t).\frac{d}{dt}v_{k}(t)=\left\{\begin{array}[]{ll}\lambda v_{k-1}(t)-1<0&\mbox{ if }k=K(t)-1\\ 0&\mbox{ if }k\geq K(t).\end{array}\right.

In particular vk(t)=0v_{k}(t)=0 for k>0k>0 and t>K/(1λ)t>K/(1-\lambda).

Proof. Let tt be an instant such that K(t)<K(t)<\infty. Lemma 4.1 implies that

π𝐯(t)(ωK(t)+ and ωK(t)+1=0)=1.\pi_{\mathbf{v}(t)}\left(\omega_{K(t)}\in\mathbb{Z}_{+}\mbox{ and }\omega_{K(t)+1}=0\right)=1. (30)

In particular the coordinate process 𝝎K(t)𝐯(t)()\mbox{\boldmath$\omega$}_{K(t)}^{\mathbf{v}(t)}(\cdot) possesses an equilibrium in +\mathbb{Z}_{+}. The process should have equal rates of up-jumps and down-jumps in that equilibrium, namely

Eπ𝐯(t)[(1q)ωK(t)]=1λvK(t)1(t).E_{\pi_{\mathbf{v}(t)}}[(1-q)^{\omega_{K(t)}}]=1-\lambda v_{K(t)-1}(t). (31)

Since vK(t)1(t)>0v_{K(t)-1}(t)>0 by definition of K(t)K(t), Theorem 4.1.a implies that

Eπ𝐯(t)[(1q)ωK(t)1]=0.E_{\pi_{\mathbf{v}(t)}}[(1-q)^{\omega_{K(t)-1}}]=0. (32)

Substituting (31) and (32) in equality (22) evaluated at k=K(t)1k=K(t)-1 yields

ddtvK(t)1(t)=λvK(t)2(t)1<0.\frac{d}{dt}v_{K(t)-1}(t)~=~\lambda v_{K(t)-2}(t)-1~<~0. (33)

Note also that Eπ𝐯(t)[(1q)ωK(t)+1]=1E_{\pi_{\mathbf{v}(t)}}[(1-q)^{\omega_{K(t)+1}}]=1 due to (30); hence equality (22) for k=K(t)k=K(t) gives

ddtvK(t)(t)=λvK(t)(t)=0.\frac{d}{dt}v_{K(t)}(t)~=~-\lambda v_{K(t)}(t)~=~0. (34)

Since K(0)=K<K(0)=K<\infty by hypothesis, it follows via (33) and (34) that K(t)K(t) is finite and nonincreasing in tt. Part (a) of the theorem now follows by (30). Part (b) is due to (33) and (34). \bf\Box

Corollary 4.1

If 𝐮(0)UK\mathbf{u}(0)\in U_{K} for some KK then

limnPn(m1(t)+,m2(t)=0)=1\lim_{n\rightarrow\infty}P_{n}(m_{1}(t)\in\mathbb{Z}_{+},m_{2}(t)=0)=1 (35)

for tK/(1λ)t\geq K/(1-\lambda). The system occupancy k=1mk(t)\sum_{k=1}^{\infty}m_{k}(t) converges in distribution to the equilibrium value of a birth-death process with constant birth rate λ\lambda and death rate 1(1q)j1-(1-q)^{j} at state jj.

Proof. Let {ni}\{n_{i}\} be a subsequence along which (𝐮(),ξ)(\mathbf{u}(\cdot),\xi) converges and let (𝐯(),χ)(\mathbf{v}(\cdot),\chi) denote the limit. Since 𝐮(0)UK\mathbf{u}(0)\in U_{K} it follows that 𝐯(0)UK\mathbf{v}(0)\in U_{K}. Choose t1>t0>K/(1λ)t_{1}>t_{0}>K/(1-\lambda) so that by Theorem 4.2.b 𝐯(t)={1,0,0,0,}\mathbf{v}(t)=\{1,0,0,0,\cdots\} for t[t0,t1]t\in[t_{0},t_{1}]. Let A={𝝎Ω:ω1+,w2=0}A=\{\mbox{\boldmath$\omega$}\in\Omega:\omega_{1}\in\mathbb{Z}_{+},w_{2}=0\}. Then

t0t1Pni(𝐦(t)A)𝑑t=Eni[t0t11{𝐦(t)A}𝑑t]t0t1π𝐯(t)(A)𝑑t=t1t0,\int_{t_{0}}^{t_{1}}P_{n_{i}}(\mathbf{m}(t)\in A)dt~=~E_{n_{i}}\left[\int_{t_{0}}^{t_{1}}1\{\mathbf{m}(t)\in A\}dt\right]~\rightarrow~\int_{t_{0}}^{t_{1}}\pi_{\mathbf{v}(t)}(A)dt~=~t_{1}-t_{0},

where the last equality is due to Theorem 4.2.a. The above limit does not depend on the particular subsequence {ni}\{n_{i}\}; therefore (35) follows. The final claim of the corollary is verified by observing that for t>K/(1λ)t>K/(1-\lambda) the coordinate process ω2𝐯(t)0\omega_{2}^{\mathbf{v}(t)}\equiv 0 in equilibrium; and in turn ω1𝐯(t)\omega_{1}^{\mathbf{v}(t)} is a positive recurrent birth-death process on +\mathbb{Z}_{+} with birth rate λ\lambda and death rate 1(1q)j1-(1-q)^{j} at state jj. \bf\Box

It should be noted that the hypothesis 𝐮(0)UK\mathbf{u}(0)\in U_{K} is necessary for the conclusions of Corollary 4.1: If the initial size of a single queue is allowed to grow without bound with increasing nn then, for large values of nn, that queue receives service whenever it is connected. In effect this reduces the service rate available to the rest of the system by a factor of (1q)(1-q). In such degenerate cases the present analysis applies to the subsystem that is composed of queues with bounded initial occupancies, after appropriate adjustment of the service rate.

5 Final remarks: LCQ(dnd_{n})

Conclusions of Sections 3 and 4 reveal that the system occupancies under LCQ(dd) and LCQ differ by a factor of order nn as nn\rightarrow\infty. More insight on this disparity, especially for moderate values of dd relative to nn, can be gained by considering an asymptotic regime in which dd is allowed to depend on nn. Here we sketch asymptotic analysis of LCQ(dnd_{n}) in the case

limndn= and limndnn=0.\lim_{n\rightarrow\infty}d_{n}=\infty\mbox{ ~~~and~~~ }\lim_{n\rightarrow\infty}\frac{d_{n}}{n}=0.

The present discussion closely follows that of Section 4, hence proofs are omitted.

Under LCQ(dnd_{n}) the representation (4) can be expressed as

uk(t)=uk(0)+0t(λ(uk1(s)uk(s))((1bk+1(s)dn)dn(1bk(s)dn)dn))𝑑s+εk(t)u_{k}(t)=u_{k}(0)+\int_{0}^{t}\left(\lambda(u_{k-1}(s)-u_{k}(s))-\left((1-\frac{b_{k+1}(s)}{d_{n}})^{d_{n}}-(1-\frac{b_{k}(s)}{d_{n}})^{d_{n}}\right)\right)ds+\varepsilon_{k}(t)

where bk(t)dnuk(t)b_{k}(t)\triangleq d_{n}u_{k}(t). Let 𝐛(t)={bk(t)}k=0\mathbf{b}(t)=\{b_{k}(t)\}_{k=0}^{\infty} and let Ωo\Omega_{o} be obtained by augmenting Ω\Omega with nonincreasing sequences that take values in +{}\mathbb{R}_{+}\cup\{\infty\}. Define the random measure ξo\xi_{o} by

ξo([0,t)×A)=0t1{𝐛(s)A}𝑑s,t>0,A(Ωo).\xi_{o}([0,t)\times A)=\int_{0}^{t}1\{\mathbf{b}(s)\in A\}ds,~~~~~~~t>0,~A\in\mathcal{B}(\Omega_{o}).

Consideration of the pair (𝐮(),ξo)(\mathbf{u}(\cdot),\xi_{o}) via an analogue of Theorem 4.1 identifies possible limits 𝐯()\mathbf{v}(\cdot) of 𝐮()\mathbf{u}(\cdot) as solutions to

ddtvk(t)=λ(vk1(t)vk(t))Eπ𝐯(t)[eωk+1eωk],k=1,2,\frac{d}{dt}v_{k}(t)=\lambda(v_{k-1}(t)-v_{k}(t))-E_{\pi_{\mathbf{v}(t)}}[e^{-\omega_{k+1}}-e^{-\omega_{k}}],~~~~~~~~k=1,2,\cdots

where π𝐯(t)\pi_{\mathbf{v}(t)} is a distribution on Ωo\Omega_{o} such that π𝐯(t)(ωk=)=1\pi_{\mathbf{v}(t)}(\omega_{k}=\infty)=1 if vk(t)>0v_{k}(t)>0 and

Eπ𝐯(t)[1{ωk}(λ(vk1(t)vk(t))+eωkeωk+1)]=0.E_{\pi_{\mathbf{v}(t)}}\left[1\{\omega_{k}\neq\infty\}\left(\lambda(v_{k-1}(t)-v_{k}(t))+e^{-\omega_{k}}-e^{-\omega_{k+1}}\right)\right]=0.

The line of reasoning employed in establishing Lemma 4.1 and Theorem 4.2 readily applies to 𝐯()\mathbf{v}(\cdot) and π𝐯()\pi_{\mathbf{v}(\cdot)} here, yielding that

π𝐯(t)(ωk+ and ωk+1=0)=1 if vk(t)=0,\pi_{\mathbf{v}(t)}\left(\omega_{k}\in\mathbb{R}_{+}\mbox{ and }\omega_{k+1}=0\right)=1~~~\mbox{ if }v_{k}(t)=0,

and that v1(t)=0v_{1}(t)=0 for t>K(0)/(1λ)t>K(0)/(1-\lambda). In turn for such tt, b1(t)=O(1)b_{1}(t)=O(1) and b2(t)=o(1)b_{2}(t)=o(1) as nn\rightarrow\infty. The maximum queue size in equilibrium therefore tends to one, but the number of queues at that occupancy is substantially larger than the same number under the LCQ policy. In particular for large enough values of tt the total system occupancy k1mk(t)=(n/dn)k1bk(t)\sum_{k\geq 1}m_{k}(t)=(n/d_{n})\sum_{k\geq 1}b_{k}(t) is O(n/dn)O(n/d_{n}).

References

  • [1] M. Alanyali, “Asymptotically exact analysis of a loss network with channel continuity,” The Annals of Applied Probability, vol. 13, no. 4, pp. 1474–1493, 2003.
  • [2] S. N. Ethier and T. G. Kurtz, Markov processes: Characterization and convergence, Wiley, 1986.
  • [3] A. Ganti, E. Modiano, and J. N. Tsitsiklis, “Optimal transmission scheduling in symmetric communication models with intermittent connectivity,” IEEE Transactions on Information Theory, vol. 53, no. 3, pp. 998–1008, 2007.
  • [4] P. J. Hunt and T. G. Kurtz, “Large loss networks,” Stochastic Processes and Their Applications, vol. 53, pp. 363–378, 1994.
  • [5] T. G. Kurtz, “Averaging for martingale problems and stochastic approximation,” Lecture Notes on Control and Information Sciences, vol. 117, pp. 186–209, Springer-Verlag, 1992.
  • [6] M. J. Neely, “Order optimal delay for opportunistic scheduling in multi-user wireless uplinks and downlinks,” 44th Allerton Conference on Communication, Control, and Computing, 2006.
  • [7] S. Shakkottai, R. Srikant, and A. L. Stolyar, “Pathwise optimality of the exponential scheduling rule for wireless channels,” Advances in Applied Probability, vol. 36, no. 4, pp. 1021-1045, 2004.
  • [8] S. Shakkottai, “Effective capacity and QoS for wireless scheduling,” IEEE Transactions on Automatic Control, vol. 53, no. 3, pp. 749-761, 2008.
  • [9] A. L. Stolyar, “Maxweight scheduling in a generalized switch: State space collapse and workload minimization in heavy traffic,” The Annals of Applied Probability, vol. 14, no. 1, pp. 1-53, 2004.
  • [10] L. Tassiulas and A. Ephremides, “Dynamic server allocation to parallel queues with randomly varying connectivity,” in IEEE Transactions on Information Theory, vol. 39, no. 2, 1993.
  • [11] N. D. Vvedenskaya, R. L. Dobrushin, and F. I. Karpelevich, “Queueing system with selection of the shortest of two queues: an asymptotic approach,” in Problems of Information Transmission, vol. 32, no. 1, pp. 15–27, 1996.
  • [12] L. Ying, R. Srikant, A. Eryilmaz and G. Dullerud, “A large deviations analysis of scheduling in wireless networks,” IEEE Transactions on Information Theory, vol. 52, no. 11, pp. 5088-5098, 2006.
  • [13] S. Zachary and I. Ziedins, “A refinement of the Hunt-Kurtz theory of large loss networks with an application to virtual partitioning,” The Annals of Applied Probability, vol. 12, pp. 1–22, 2002.