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

Stochastic Geometry Analysis of Sojourn Time in Multi-Tier Cellular Networks

Mohammad Salehi and Ekram Hossain The authors are with the Department of Electrical and Computer Engineering at the University of Manitoba, Canada (emails: [email protected], [email protected]). E. Hossain is the corresponding author.
Abstract

Impact of mobility will be increasingly important in future generation wireless services and the related challenges will need to be addressed. Sojourn time, the time duration that a mobile user stays within a cell, is a mobility-aware parameter that can significantly impact the performance of mobile users and it can also be exploited to improve resource allocation and mobility management methods in the network. In this paper, we derive the distribution and mean of the sojourn time in multi-tier cellular networks, where spatial distribution of base stations (BSs) in each tier follows an independent homogeneous Poisson point process (PPP). To obtain the sojourn time distribution in multi-tier cellular networks with maximum biased averaged received power association, we derive the linear contact distribution function and chord length distribution of each tier. We also study the relation between mean sojourn time and other mobility-related performance metrics. We show that the mean sojourn time is inversely proportional to the handoff rate, and the complementary cumulative distribution function (CCDF) of sojourn time is bounded from above by the complement of the handoff probability. Moreover, we study the impact of user velocity and network parameters on the sojourn time.

Index Terms:
Multi-tier cellular network, user mobility, sojourn time, handoff probability, handoff rate, Poisson point process (PPP).

I Introduction

I-A Background and Related Work

The next generations of cellular wireless networks are expected to support communications for highly mobile users and devices [1] with applications in new vertical sectors such as railway, unmanned aerial vehicle (UAV), and autonomous car. Therefore, addressing the mobility related challenges is necessary for the development of the next generation cellular networks. Impact of user/device mobility on its performance in cellular networks can be measured through mobility-aware performance metrics such as handoff rate, handoff probability, and sojourn time [1]. Sojourn time (or dwell time), time duration that a mobile user stays within a cell, is a key network parameter which allows studying other important network parameters such as channel occupancy time, new call and handoff call dropping probabilities [2]. Therefore, it is imperative to incorporate the sojourn time distribution in resource allocation and mobility management for improving the network performance. In general, modeling and analysis of mobility-related parameters and performances is however challenging in multi-tier (or heterogeneous) cellular networks (e.g. a two-tier macrocell-small cell network) since it needs to consider different aspects such as how to model the distributions of base stations (BSs) at the different tiers, how to model the user mobility and traffic at the different tiers, and how to model the radio access network performance at the different tiers [3].

In this above context, [4] derived the sojourn time distribution for the hexagonal (deterministic) cellular networks and Poisson (random) cellular networks, where the BSs are distributed according to a homogeneous Poisson point process (PPP). In [5], mean sojourn time of two-tier cellular networks was approximately derived, where the coverage areas of macro cells and small cells have regular shapes (circles) and within each macro cell multiple small cells are irregularly deployed. [6, 7] derived the mean sojourn time in small cells of two-tier cellular networks. The BSs of each tier are distributed following an independent homogeneous PPP. [8] also derived the mean sojourn time in two-tier cellular networks. However, it was assumed that a handoff occurs only when the mobile user crosses the boundary of a macro cell. Therefore, the mean sojourn time in [8] is similar to that in a single-tier network as in [4].

Moreover, the handoff rate, i.e. the expected number of handoffs in unit time, was derived in [4] for single-tier Poisson cellular networks and in [9] for multi-tier Poisson cellular networks. The handoff probability, i.e. the probability that the mobile user handoffs to a new BS at the end of a movement period, was also studied in [10] and [11] for single-tier and multi-tier Poisson networks, respectively. To derive the mean sojourn time (or distribution of the sojourn time), [4, 7, 8] used the chord length distribution (or linear contact distribution function) of Poisson Voronoi cells. However, in multi-tier networks with different transmission power and bias factor for each tier, we need the chord length distribution (or linear contact distribution function) of weighted Poisson Voronoi cells which is not available in the literature. For single-tier networks, [1] used the handoff probability to derive the distribution of sojourn time in the cell where connection is initiated. In single-tier networks, since the Voronoi cells are convex, we can directly use the handoff probability to derive the distribution of the sojourn time. However, in multi-tier networks, cells may not be convex. Therefore, the analytical method in [1] cannot be used for multi-tier networks.

A handoff is considered to be unnecessary when the dwell time of the mobile user in the new cell after the handoff is less than a predefined threshold. In [12, 13], handoff skipping schemes are employed to avoid unnecessary handoffs. Moreover, important system parameters such as channel occupancy time, new call and handoff call dropping probabilities depend on the sojourn time [2]. Therefore, sojourn time is fundamental for analysis and design of the mobile cellular networks. In [12, 2, 14, 15, 16], different distributions such as exponential, Erlang, gamma, Pareto, and Weibull were used for modeling the sojourn time distribution. Due to the principal role of the sojourn time in mobility management and resource allocation, in this paper, we derive the sojourn time distribution in multi-tier cellular networks.

I-B Contributions

To analyze the sojourn time distribution in multi-tier scenarios with PPP distributed BSs, the existing works either assume that the mobile user is always associated to only one of the tiers, or only focus on the small tier (in two-tier scenarios). For both the cases, the results are no different from the single-tier scenarios. In single-tier networks with maximum averaged received power association (nearest BS association), (Voronoi) cells are convex; however, in multi-tier networks with maximum biased averaged received power association, cells may not be convex depending on the transmission power and bias factor of each tier. Therefore, analysis of sojourn time of multi-tier cellular networks is more complicated compared to the single-tier networks. In this regard, the contributions of this paper can be summarized as follows:

  • We derive the distribution and mean of the sojourn time for multi-tier cellular networks. We show that the mean sojourn time is inversely proportional to velocity. We also study the impact of network parameters on the sojourn time.

  • To obtain the analytical results, we derive the linear contact distribution and chord length distribution of each tier.

  • We show that the mean sojourn time is inversely proportional to handoff rate. Also, using handoff rate and sojourn time, we calculate the ping-pong rate (i.e. rate of unnecessary handoffs) for each tier.

  • We show that the complement of the handoff probability provides an upper bound for the complementary cumulative distribution function (CCDF) of the sojourn time. We also discuss the scenarios where the CCDF of the sojourn time is equal to the complement of the handoff probability.

The rest of this paper is organized as follows: In Section II, the system model is presented. In Section III, we state the methodology for deriving the analytical results. In Sections IV and V, we obtain the main results related to the distribution and mean of the sojourn time and also discuss the effects of network parameters. Numerical and simulation results are provided in Section VI. Finally, in Section VII, we conclude the paper.

II System Model and Notations

Consider a KK-tier heterogeneous cellular network with KK classes of BSs and let 𝒦={1,2,,K}\mathcal{K}=\{1,2,...,K\}. The spatial distribution of BSs of kk-th tier, k{1,2,,K}k\in\{1,2,...,K\}, follows an independent homogeneous PPP Φk\Phi_{k} of intensity λk\lambda_{k}. Different tiers of BSs transmit at different power levels. PkP_{k} denotes the transmission power of the kk-th tier BSs.

Consider a typical mobile user which moves in a straight line with a constant velocity vv. Due to the stationarity of the homogeneous PPP, i.e. its distribution is invariant under translation [17], we can assume that the typical mobile user is located at the origin oo at time 0. Since homogeneous PPP is isotropic, i.e. its distribution is invariant under rotation with respect to the origin [17], we can also assume that the typical mobile user moves along the positive xx-axis. Therefore, at time tt, the typical mobile user is located at x(t)=(vt,0)\text{x}(t)=\left(vt,0\right).

The mobile user is always associated to the BS which provides the maximum biased averaged received power. Let us denote the serving BS at time tt by BS(t)BS(t). Therefore,

BS(t)=argmaxxΦk,k𝒦BkPkx(t)xα,\displaystyle BS(t)=\text{arg}\max\limits_{x\in\Phi_{k},\forall k\in\mathcal{K}}B_{k}P_{k}\|\text{x}(t)-x\|^{-\alpha}, (1)

where BkB_{k} is the cell range expansion bias factor for tier-kk, and α\alpha is the path-loss exponent. Let us denote the distance between BS(0)BS(0) and the mobile user at x(t)\text{x}(t) by r0(t)r_{0}(t), i.e. r0(t)=BS(0)x(t)r_{0}(t)=\|BS(0)-\text{x}(t)\|. Given that at time tt the mobile user is associated to a tier-kk BS at distance r(t)r(t), from (1), we have Φj((x(t),r(t)βkj))=0\Phi_{j}\left(\mathcal{B}\left(\text{x}(t),\frac{r(t)}{\beta_{kj}}\right)\right)=0, j𝒦\forall j\in\mathcal{K}, where βkj=(BkPkBjPj)1/α\beta_{kj}=\left(\frac{B_{k}P_{k}}{B_{j}P_{j}}\right)^{1/\alpha}, (x,r)\mathcal{B}(\text{x},r) denotes a ball with radius rr centered at x, and Φj(A)\Phi_{j}(A) is the number of tier-jj BSs in set A2A\subset\mathbb{R}^{2}. For simplicity, we define r0=r0(0)r_{0}=r_{0}(0).

A summary of the major notations is provided in Table I.

TABLE I: Summary of Notations
Notation Description
Φk\Phi_{k}, λk\lambda_{k} PPP of tier-kk BSs, intensity of Φk\Phi_{k}
PkP_{k} Transmit power of the kk-th tier BSs
x(t)\text{x}(t) Location of the mobile user at time tt
vv Velocity of the mobile user
BS(t)BS(t) Serving BS of the mobile user at time tt
BkB_{k} Cell range expansion (bias) factor for tier kk
α\alpha Path-loss exponent
r0(t)r_{0}(t), r0r_{0} Distance between the initially serving BS and the mobile user at time tt, r0(0)r_{0}(0)
βkj\beta_{kj} (BkPkBjPj)1/α\left(\frac{B_{k}P_{k}}{B_{j}P_{j}}\right)^{1/\alpha}
(x,r)\mathcal{B}(\text{x},r) Ball with radius rr centred at x
S~\tilde{S} Sojourn time in the cell where connection is initiated
SS Sojourn time
HkH_{k} Handoff rate from (to) a tier-kk cell to (from) any other cell in the network
HH Handoff rate

III Methodology of Analysis of Sojourn Time in Multi-Tier Cellular Networks


Refer to caption
(a) Single-tier.


Refer to caption
(b) Two-tier.
Figure 1: Voronoi cells in a single-tier and a two-tier cellular networks: (a) λ=0.15\lambda=0.15, (b) λ1=0.05\lambda_{1}=0.05, λ2=0.1\lambda_{2}=0.1, and β12=2\beta_{12}=2.

The sojourn time SS is the duration that the mobile user stays within a particular serving cell before it is handed over to another cell [18]. Analysis of sojourn time in multi-tier cellular networks consists of the following four steps:

  • Step 1: Deriving the conditional distribution of the sojourn time in the cell where connection is initiated, given that the mobile user is initially associated to a tier-kk BS.

  • Step 2: Deriving the linear contact distribution function, given that the mobile user is in a tier-kk cell at time 0.

  • Step 3: Obtaining the chord length distribution for tier kk using linear contact distribution function.

  • Step 4: Deriving the distribution of the sojourn time SS for tier kk.

III-A Step 1 of Analysis

First we focus on the distribution of the sojourn time in the cell where connection is initiated S~\tilde{S}. Specifically, we derive the CCDF of S~\tilde{S}, i.e.

F¯S~(T)=(S~>T)\displaystyle\bar{F}_{\tilde{S}}(T)=\mathbb{P}(\tilde{S}>T) =\displaystyle= (no handoff occurs in the interval [0,T])\displaystyle\mathbb{P}\left(\text{no handoff occurs in the interval }[0,T]\right) (2)
=\displaystyle= (BS(t)=BS(0),t(0,T]).\displaystyle\mathbb{P}\left(BS(t)=BS(0),\forall t\in(0,T]\right).

In single-tier cellular networks, Voronoi cells are convex [19] (as shown in Fig. 1(a)). A set CC is convex if the line segment between any two points in CC lies in CC [20]. Thus, in single-tier cellular networks, when the mobile user is connected to the same BS at time 0 and TT, i.e. when BS(T)=BS(0)BS(T)=BS(0), the serving BS at any time between 0 and TT is also BS(0)BS(0), i.e. BS(t)=BS(0)BS(t)=BS(0), t(0,T)\forall t\in(0,T). Hence, for single-tier cellular networks, (2) can be simplified as

F¯S~(T)=(BS(T)=BS(0)).\displaystyle\bar{F}_{\tilde{S}}(T)=\mathbb{P}\left(BS(T)=BS(0)\right). (3)

However, for multi-tier cellular networks, Voronoi cells may not be convex depending on the values of βkj\beta_{kj}, k,j𝒦k,j\in\mathcal{K} (as shown in Fig. 1(b)). Therefore, even when BS(T)=BS(0)BS(T)=BS(0), there may exist a time tt between 0 and TT for which BS(t)BS(0)BS(t)\neq BS(0). To derive the CCDF of S~\tilde{S} for multi-tier cellular networks, we must use (2), which makes the analysis of sojourn time in multi-tier cellular networks more complicated compared to the single-tier networks. Actually, single-tier scenario can be considered as a special case of multi-tier scenarios. Moreover, note that, (3) provides an upper bound for (2).

Refer to caption
Figure 2: System model.

Given that the mobile user is initially connected to a tier-kk BS, the CCDF of S~\tilde{S} can be obtained by

\IEEEeqnarraymulticol3lF¯S~(Ttier=k)=(BS(t)=BS(0),t(0,T]tier=k)\displaystyle\IEEEeqnarraymulticol{3}{l}{\bar{F}_{\tilde{S}}(T\mid\text{tier}=k)=\mathbb{P}\left(BS(t)=BS(0),\forall t\in(0,T]\mid\text{tier}=k\right)} (4)
=\displaystyle= 1π00π(BS(t)=BS(0),t(0,T]r0,θ,tier=k)fR(r0tier=k)dθdr0,\displaystyle\frac{1}{\pi}\int_{0}^{\infty}\int_{0}^{\pi}\mathbb{P}\left(BS(t)=BS(0),\forall t\in(0,T]\mid r_{0},\theta,\text{tier}=k\right)f_{R}(r_{0}\mid\text{tier}=k){\rm d}\theta{\rm d}r_{0},

where θ\theta is the angle between the serving BS at time 0 and direction of the movement (as shown in Fig. 2). θ\theta is uniformly distributed in [0,π][0,\pi]. fR(r0tier=k)f_{R}(r_{0}\mid\text{tier}=k) is the probability density function (PDF) of the serving link distance at time 0, given that BS(0)BS(0) belongs to tier-kk. According to [21],

fR(r0tier=k)=1(tier=k)2λkπr0exp{j𝒦λjπβjk2r02},\displaystyle f_{R}(r_{0}\mid\text{tier}=k)=\frac{1}{\mathbb{P}(\text{tier}=k)}2\lambda_{k}\pi r_{0}\exp\left\{-\sum_{j\in\mathcal{K}}\lambda_{j}\pi\beta_{jk}^{2}r_{0}^{2}\right\}, (5)

where (tier=k)\mathbb{P}(\text{tier}=k) is the probability that BS(0)BS(0) belongs to tier-kk which is given by [21]:

(tier=k)=λkj𝒦λjβjk2.\displaystyle\mathbb{P}(\text{tier}=k)=\frac{\lambda_{k}}{\sum_{j\in\mathcal{K}}\lambda_{j}\beta_{jk}^{2}}. (6)

Using the association strategy (1), we get

\IEEEeqnarraymulticol3l(BS(t)=BS(0),t(0,T]r0,θ,tier=k)\displaystyle\IEEEeqnarraymulticol{3}{l}{\mathbb{P}\left(BS(t)=BS(0),\forall t\in(0,T]\mid r_{0},\theta,\text{tier}=k\right)} (7)
=\displaystyle= (j𝒦Φj((x(t),r0(t)βkj)(0,r0βkj))=0,t(0,T]r0,θ,tier=k),\displaystyle\mathbb{P}\left(\bigcap_{j\in\mathcal{K}}\Phi_{j}\left(\mathcal{B}\left(\text{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right)\setminus\mathcal{B}\left(0,\frac{r_{0}}{\beta_{kj}}\right)\right)=0,\forall t\in(0,T]\mid r_{0},\theta,\text{tier}=k\right),

where (0,r0βkj)\mathcal{B}\left(0,\frac{r_{0}}{\beta_{kj}}\right) is excluded since we know there is no tier jj BS closer than r0βkj\frac{r_{0}}{\beta_{kj}} to the typical mobile user at time 0. Let us define

𝒜kj(r0,θ,v,T,βkj)={t(x(t),r0(t)βkj)t[0,T],r0(t)=r02+v2t22r0vtcosθ}.\displaystyle\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})=\left\{\bigcup_{t}\mathcal{B}\left(\text{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right)\mid t\in[0,T],r_{0}(t)=\sqrt{r_{0}^{2}+v^{2}t^{2}-2r_{0}vt\cos\theta}\right\}. (8)

Using 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}), we can write

\IEEEeqnarraymulticol3l(BS(t)=BS(0),t(0,T]r0,θ,tier=k)\displaystyle\IEEEeqnarraymulticol{3}{l}{\mathbb{P}\left(BS(t)=BS(0),\forall t\in(0,T]\mid r_{0},\theta,\text{tier}=k\right)} (9)
=\displaystyle= (j𝒦Φj(𝒜kj(r0,θ,v,T,βkj)(0,r0βkj))=0r0,θ,tier=k)\displaystyle\mathbb{P}\left(\bigcap_{j\in\mathcal{K}}\Phi_{j}\left(\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})\setminus\mathcal{B}\left(0,\frac{r_{0}}{\beta_{kj}}\right)\right)=0\mid r_{0},\theta,\text{tier}=k\right)
=(a)\displaystyle\stackrel{{\scriptstyle({\text{a}})}}{{=}} j𝒦(Φj(𝒜kj(r0,θ,v,T,βkj)(0,r0βkj))=0r0,θ,tier=k)\displaystyle\prod_{j\in\mathcal{K}}\mathbb{P}\left(\Phi_{j}\left(\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})\setminus\mathcal{B}\left(0,\frac{r_{0}}{\beta_{kj}}\right)\right)=0\mid r_{0},\theta,\text{tier}=k\right)
=(b)\displaystyle\stackrel{{\scriptstyle({\text{b}})}}{{=}} j𝒦exp{λj|𝒜kj(r0,θ,v,T,βkj)(0,r0βkj)|},\displaystyle\prod_{j\in\mathcal{K}}\exp\left\{-\lambda_{j}\left|\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})\setminus\mathcal{B}\left(0,\frac{r_{0}}{\beta_{kj}}\right)\right|\right\},

where |A||A| denotes the area of AA, (a) follows from the independence of different tiers’ point processes, and (b) is obtained by using the void probability of PPP. In Fig. 3, 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}) is illustrated for three different cases: a) βkj<1\beta_{kj}<1, b) βkj=1\beta_{kj}=1, and c) βkj>1\beta_{kj}>1. To derive the distribution of S~\tilde{S}, we need to calculate the area of 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}) for all three cases. Further discussion about 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}) is provided in the next section.


Refer to caption
(a) βkj=0.8\beta_{kj}=0.8.


Refer to caption
(b) βkj=1\beta_{kj}=1.


Refer to caption
(c) βkj=1.2\beta_{kj}=1.2.
Figure 3: 𝒜kj(20,π/3,5,20,βkj)\mathcal{A}_{kj}(20,\pi/3,5,20,\beta_{kj}). (a) βkj<1\beta_{kj}<1, (b) βkj=1\beta_{kj}=1, and (c) βkj>1\beta_{kj}>1. Red circles correspond to (0,r0βkj)\mathcal{B}\left(0,\frac{r_{0}}{\beta_{kj}}\right) and (x(T),r0(T)βkj)\mathcal{B}\left(\text{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right).

III-B Step 2 of Analysis

Given that, at time 0, the mobile user is associated to a tier-kk BS, the origin is almost surely contained in the interior of a tier-kk Voronoi cell. In this paper, we define linear contact distribution function as the probability that a line segment \ell containing the origin with length rr and random orientation crosses the cell boundaries. Therefore, given origin oo is inside a tier-kk cell, linear contact distribution function H(ztier=k)H_{\ell}(z\mid\text{tier}=k) is equal to the probability that intersection of user’s trajectory with length zz and the cell boundaries is nonempty. Using the conditional CCDF of S~\tilde{S}, we can derive the linear contact distribution function as

H(ztier=k)\displaystyle H_{\ell}(z\mid\text{tier}=k) =\displaystyle= 1(S~>zvtier=k)=1F¯S~(zvtier=k).\displaystyle 1-\mathbb{P}(\tilde{S}>\frac{z}{v}\mid\text{tier}=k)=1-\bar{F}_{\tilde{S}}(\frac{z}{v}\mid\text{tier}=k). (10)

III-C Step 3 of Analysis

So far, we have considered the sojourn time in the cell where connection is initiated (S~\tilde{S}). Distribution of the sojourn time (SS), for tier-kk, can be obtained using the chord length distribution. Due to the stationarity of our model, chord length distribution for tier-kk, denoted by FL(ztier=k)F_{L}(z\mid\text{tier}=k), can be computed as follows [22]:

FL(ztier=k)=1𝔼[Ltier=k]ddzH(ztier=k),\displaystyle F_{L}(z\mid\text{tier}=k)=1-\mathbb{E}[L\mid\text{tier}=k]\frac{{\rm d}}{{\rm d}z}H_{\ell}(z\mid\text{tier}=k), (11)

where 𝔼[Ltier=k]\mathbb{E}[L\mid\text{tier}=k] is the mean length of the chords lying in tier-kk cells, and is obtained by [23]

𝔼[Ltier=k]=limz0zH(ztier=k)\displaystyle\mathbb{E}[L\mid\text{tier}=k]=\lim_{z\to 0}\frac{z}{H_{\ell}(z\mid\text{tier}=k)} (12)

III-D Step 4 of Analysis

Finally, we can characterize the sojourn time distribution for tier-kk using the results from previous step. In particular, the mean and CCDF of the sojourn time in tier-kk are

𝔼[Stier=k]\displaystyle\mathbb{E}[S\mid\text{tier}=k] =\displaystyle= 1v𝔼[Ltier=k],\displaystyle\frac{1}{v}\mathbb{E}[L\mid\text{tier}=k], (13)
F¯S(Ttier=k)\displaystyle\bar{F}_{S}(T\mid\text{tier}=k) =\displaystyle= 1FL(vTtier=k).\displaystyle 1-F_{L}(vT\mid\text{tier}=k). (14)

IV First Step of Sojourn Time Analysis: Derivation of |𝒜kj(r0,θ,v,T,βkj)||\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})|

As mentioned in the previous subsection, the first step of sojourn time analysis requires calculation of |𝒜kj(r0,θ,v,T,βkj)||\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})| (area of 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})). In this regard, we consider three cases: I) βkj<1\beta_{kj}<1, II) βkj>1\beta_{kj}>1, and III) βkj=1\beta_{kj}=1.

IV-A Case I: βkj<1\beta_{kj}<1

The following proposition helps us to derive the area of 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}) for this case.

Proposition 1.

When βkj<1\beta_{kj}<1, 𝒜kj(r0,θ,v,T,βkj)=(x(0),r0βkj)(x(T),r0(T)βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})=\mathcal{B}\left(\textup{x}(0),\frac{r_{0}}{\beta_{kj}}\right)\cup\mathcal{B}\left(\textup{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right).

Proof:

See Appendix A. ∎

Note that, depending on radii of the two circles, r0βkj\frac{r_{0}}{\beta_{kj}} and r0(T)βkj\frac{r_{0}(T)}{\beta_{kj}}, and the distance between their centres, i.e. vTvT, three different situations can happen when βkj<1\beta_{kj}<1:

Situation 1: When r0(T)βkjr0βkj+vT\frac{r_{0}(T)}{\beta_{kj}}\geq\frac{r_{0}}{\beta_{kj}}+vT, we have (x(0),r0βkj)(x(T),r0(T)βkj)\mathcal{B}\left(\textup{x}(0),\frac{r_{0}}{\beta_{kj}}\right)\subset\mathcal{B}\left(\textup{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right), which yields 𝒜kj(r0,θ,v,T,βkj)=(x(T),r0(T)βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})=\mathcal{B}\left(\textup{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right).

Situation 2: When r0βkjr0(T)βkj+vT\frac{r_{0}}{\beta_{kj}}\geq\frac{r_{0}(T)}{\beta_{kj}}+vT, we have (x(T),r0(T)βkj)(x(0),r0βkj)\mathcal{B}\left(\textup{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right)\subset\mathcal{B}\left(\textup{x}(0),\frac{r_{0}}{\beta_{kj}}\right), which yields 𝒜kj(r0,θ,v,T,βkj)=(x(0),r0βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})=\mathcal{B}\left(\textup{x}(0),\frac{r_{0}}{\beta_{kj}}\right).

Situation 3: When r0(T)βkj<r0βkj+vT\frac{r_{0}(T)}{\beta_{kj}}<\frac{r_{0}}{\beta_{kj}}+vT and r0βkj<r0(T)βkj+vT\frac{r_{0}}{\beta_{kj}}<\frac{r_{0}(T)}{\beta_{kj}}+vT, 𝒜kj(r0,θ,v,T,βkj)=(x(0),r0βkj)(x(T),r0(T)βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})=\mathcal{B}\left(\textup{x}(0),\frac{r_{0}}{\beta_{kj}}\right)\cup\mathcal{B}\left(\textup{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right). An example of which is illustrated in Fig. 3(a).

Using this information, now we can compute |𝒜kj(r0,θ,v,T,βkj)||\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})| when βkj<1\beta_{kj}<1.

|𝒜kj(r0,θ,v,T,βkj)|={πr0(T)2βkj2,if 2r0cosθ+βkj1βkj2vTπr02βkj2,if vT2r0cosθβkj1βkj2πr02βkj2+πr0(T)2βkj2V(r0βkj,r0(T)βkj,vT),if 2r0cosθβkj1βkj2<vT<2r0cosθ+βkj1βkj2\displaystyle|\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})|=\begin{cases}\pi\frac{r_{0}(T)^{2}}{\beta_{kj}^{2}},&\text{if }2r_{0}\frac{cos\theta+\beta_{kj}}{1-\beta_{kj}^{2}}\leq vT\\ \pi\frac{r_{0}^{2}}{\beta_{kj}^{2}},&\text{if }vT\leq 2r_{0}\frac{cos\theta-\beta_{kj}}{1-\beta_{kj}^{2}}\\ \pi\frac{r_{0}^{2}}{\beta_{kj}^{2}}+\pi\frac{r_{0}(T)^{2}}{\beta_{kj}^{2}}-V\left(\frac{r_{0}}{\beta_{kj}},\frac{r_{0}(T)}{\beta_{kj}},vT\right),&\text{if }2r_{0}\frac{cos\theta-\beta_{kj}}{1-\beta_{kj}^{2}}<vT<2r_{0}\frac{cos\theta+\beta_{kj}}{1-\beta_{kj}^{2}}\end{cases} (15)

where V(r0βkj,r0(T)βkj,vT)V\left(\frac{r_{0}}{\beta_{kj}},\frac{r_{0}(T)}{\beta_{kj}},vT\right) is the area of intersection of two circles with radii r0βkj\frac{r_{0}}{\beta_{kj}} and r0(T)βkj\frac{r_{0}(T)}{\beta_{kj}} whose centers are separated by vTvT, i.e., V(r0βkj,r0(T)βkj,vT)=V\left(\frac{r_{0}}{\beta_{kj}},\frac{r_{0}(T)}{\beta_{kj}},vT\right)=

r02βkj2arccos(r02+βkj2v2T2r0(T)22βkjr0vT)+r0(T)2βkj2arccos(r0(T)2+βkj2v2T2r022βkjr0(T)vT)\displaystyle\frac{r_{0}^{2}}{\beta_{kj}^{2}}\arccos\left(\frac{r_{0}^{2}+\beta_{kj}^{2}v^{2}T^{2}-r_{0}(T)^{2}}{2\beta_{kj}r_{0}vT}\right)+\frac{r_{0}(T)^{2}}{\beta_{kj}^{2}}\arccos\left(\frac{r_{0}(T)^{2}+\beta_{kj}^{2}v^{2}T^{2}-r_{0}^{2}}{2\beta_{kj}r_{0}(T)vT}\right)
12(r0βkj+r0(T)βkj+vT)(r0βkj+r0(T)βkjvT)(r0βkjr0(T)βkj+vT)(r0βkj+r0(T)βkj+vT).\displaystyle-\frac{1}{2}\sqrt{\left(\frac{r_{0}}{\beta_{kj}}+\frac{r_{0}(T)}{\beta_{kj}}+vT\right)\left(\frac{r_{0}}{\beta_{kj}}+\frac{r_{0}(T)}{\beta_{kj}}-vT\right)\left(\frac{r_{0}}{\beta_{kj}}-\frac{r_{0}(T)}{\beta_{kj}}+vT\right)\left(-\frac{r_{0}}{\beta_{kj}}+\frac{r_{0}(T)}{\beta_{kj}}+vT\right)}.
(16)


Refer to caption
(a) t=2,2.2t=2,2.2.


Refer to caption
(b) t=0,1,2,3,4,5,6,7,8t=0,1,2,3,4,5,6,7,8.
Figure 4: (x(t),r0(t)βkj)\mathcal{B}\left(\textup{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right) at different time instants. Union of these circles from t=0t=0 till t=8t=8 forms 𝒜kj(20,π/3,5,8,1.2)\mathcal{A}_{kj}(20,\pi/3,5,8,1.2).

IV-B Case II: Bkj>1B_{kj}>1

For this case, to derive the area of 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}), first we study the intersection of (x(t),r0(t)βkj)\mathcal{B}\left(\textup{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right) and (x(t+dt),r0(t+dt)βkj)\mathcal{B}\left(\textup{x}(t+{\rm d}t),\frac{r_{0}(t+{\rm d}t)}{\beta_{kj}}\right) as dt0{\rm d}t\to 0 (Fig. 4(a)). From triangle equations, we have

r0(t+dt)2=r02+v2(t+dt)22r0v(t+dt)cosθ=r0(t)2+v2dt2+2vdt(vtr0cosθ).\displaystyle r_{0}(t+{\rm d}t)^{2}=r_{0}^{2}+v^{2}(t+{\rm d}t)^{2}-2r_{0}v(t+{\rm d}t)\cos\theta=r_{0}(t)^{2}+v^{2}{\rm d}t^{2}+2v{\rm d}t(vt-r_{0}\cos\theta). (17)

Since |vtr0cosθ|r0(t)|vt-r_{0}\cos\theta|\leq r_{0}(t),

r0(t)2+v2dt22r0(t)vdtr0(t+dt)2r0(t)2+v2dt2+2r0(t)vdt.\displaystyle r_{0}(t)^{2}+v^{2}{\rm d}t^{2}-2r_{0}(t)v{\rm d}t\leq r_{0}(t+{\rm d}t)^{2}\leq r_{0}(t)^{2}+v^{2}{\rm d}t^{2}+2r_{0}(t)v{\rm d}t. (18)

Dividing r0(t)vdtr0(t+dt)r0(t)+vdtr_{0}(t)-v{\rm d}t\leq r_{0}(t+{\rm d}t)\leq r_{0}(t)+v{\rm d}t by βkj\beta_{kj} yields

r0(t)βkjvdt(a)r0(t)βkjvdtβkjr0(t+dt)βkjr0(t)βkj+vdtβkj(b)r0(t)βkj+vdt,\displaystyle\frac{r_{0}(t)}{\beta_{kj}}-v{\rm d}t\stackrel{{\scriptstyle(\text{a})}}{{\leq}}\frac{r_{0}(t)}{\beta_{kj}}-\frac{v{\rm d}t}{\beta_{kj}}\leq\frac{r_{0}(t+{\rm d}t)}{\beta_{kj}}\leq\frac{r_{0}(t)}{\beta_{kj}}+\frac{v{\rm d}t}{\beta_{kj}}\stackrel{{\scriptstyle(\text{b})}}{{\leq}}\frac{r_{0}(t)}{\beta_{kj}}+v{\rm d}t,

where (a) and (b) are obtained using βkj>1\beta_{kj}>1. Therefore, as dt0{\rm d}t\to 0, (x(t),r0(t)βkj)\mathcal{B}\left(\textup{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right) and (x(t+dt),r0(t+dt)βkj)\mathcal{B}\left(\textup{x}(t+{\rm d}t),\frac{r_{0}(t+{\rm d}t)}{\beta_{kj}}\right) partially overlap (boundaries of (x(t),r0(t)βkj)\mathcal{B}\left(\textup{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right) and (x(t+dt),r0(t+dt)βkj)\mathcal{B}\left(\textup{x}(t+{\rm d}t),\frac{r_{0}(t+{\rm d}t)}{\beta_{kj}}\right) intersect at two points), and we have

\IEEEeqnarraymulticol3l|(x(t),r0(t)βkj)(x(t+dt),r0(t+dt)βkj)|=πr0(t)2βkj2V(r0(t)βkj,r0(t+dt)βkj,vdt)\displaystyle\IEEEeqnarraymulticol{3}{l}{\left|\mathcal{B}\left(\textup{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right)\setminus\mathcal{B}\left(\textup{x}(t+{\rm d}t),\frac{r_{0}(t+{\rm d}t)}{\beta_{kj}}\right)\right|=\pi\frac{r_{0}(t)^{2}}{\beta_{kj}^{2}}-V\left(\frac{r_{0}(t)}{\beta_{kj}},\frac{r_{0}(t+{\rm d}t)}{\beta_{kj}},v{\rm d}t\right)}
=\displaystyle= πr0(t)2βkj2arccos(r0cosθvtβkjr0(t)+βkj212βkjvdtr0(t))r0(t)2βkj2\displaystyle\pi\frac{r_{0}(t)^{2}}{\beta_{kj}^{2}}-\arccos\left(\frac{r_{0}\cos\theta-vt}{\beta_{kj}r_{0}(t)}+\frac{\beta_{kj}^{2}-1}{2\beta_{kj}}\frac{v{\rm d}t}{r_{0}(t)}\right)\frac{r_{0}(t)^{2}}{\beta_{kj}^{2}}
arccos(vtr0cosθβkjr0(t+dt)+βkj2+12βkjvdtr0(t+dt))r0(t+dt)2βkj2\displaystyle-\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t+{\rm d}t)}+\frac{\beta_{kj}^{2}+1}{2\beta_{kj}}\frac{v{\rm d}t}{r_{0}(t+{\rm d}t)}\right)\frac{r_{0}(t+{\rm d}t)^{2}}{\beta_{kj}^{2}}
+122vdtβkj(r0(t)vtr0cosθβkj)+v2dt2(11βkj2)\displaystyle+\frac{1}{2}\sqrt{\frac{2v{\rm d}t}{\beta_{kj}}\left(r_{0}(t)-\frac{vt-r_{0}\cos\theta}{\beta_{kj}}\right)+v^{2}{\rm d}t^{2}\left(1-\frac{1}{\beta_{kj}^{2}}\right)}
×2vdtβkj(r0(t)+vtr0cosθβkj)v2dt2(11βkj2)\displaystyle\times\sqrt{\frac{2v{\rm d}t}{\beta_{kj}}\left(r_{0}(t)+\frac{vt-r_{0}\cos\theta}{\beta_{kj}}\right)-v^{2}{\rm d}t^{2}\left(1-\frac{1}{\beta_{kj}^{2}}\right)}
=\displaystyle= 2vβkj2[βkj2r0(t)2(vtr0cosθ)2arccos(vtr0cosθβkjr0(t))(vtr0cosθ)]dt+O(dt2),\displaystyle\frac{2v}{\beta_{kj}^{2}}\left[\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-(vt-r_{0}\cos\theta)^{2}}-\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t)}\right)(vt-r_{0}\cos\theta)\right]{\rm d}t+O({\rm d}t^{2}),

where the last result is proved in Appendix B.

Moreover, we can derive the intersection points of these two circles from their equations in Cartesian coordinate system, i.e.,

(x(t),r0(t)βkj)\displaystyle\mathcal{B}\left(\textup{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right) :\displaystyle: [xvt]2+y2=r0(t)2βkj2,\displaystyle\left[x-vt\right]^{2}+y^{2}=\frac{r_{0}(t)^{2}}{\beta_{kj}^{2}},
(x(t+dt),r0(t+dt)βkj)\displaystyle\mathcal{B}\left(\textup{x}(t+{\rm d}t),\frac{r_{0}(t+{\rm d}t)}{\beta_{kj}}\right) :\displaystyle: [xv(t+dt)]2+y2=r0(t+dt)2βkj2.\displaystyle\left[x-v(t+{\rm d}t)\right]^{2}+y^{2}=\frac{r_{0}(t+{\rm d}t)^{2}}{\beta_{kj}^{2}}.

Combining these equations and solving for xx results in

x=v(t+dt2)(11βkj2)+r0cosθβkj2,\displaystyle x=v\left(t+\frac{{\rm d}t}{2}\right)\left(1-\frac{1}{\beta_{kj}^{2}}\right)+\frac{r_{0}\cos\theta}{\beta_{kj}^{2}},

which indicates that, for βkj>1\beta_{kj}>1, the boundaries’ intersection points move along the positive xx-axis as tt increases (Fig. 4(b)). Using this result, we can write

\IEEEeqnarraymulticol3l|i=0n(x(t+idt),r0(t+idt)βkj)|=πr0(t+ndt)2βkj2\displaystyle\IEEEeqnarraymulticol{3}{l}{\left|\bigcup_{i=0}^{n}\mathcal{B}\left(\textup{x}(t+i{\rm d}t),\frac{r_{0}(t+i{\rm d}t)}{\beta_{kj}}\right)\right|=\pi\frac{r_{0}(t+n{\rm d}t)^{2}}{\beta_{kj}^{2}}} (20)
+i=0n1|(x(t+idt),r0(t+idt)βkj)(x(t+(i+1)dt),r0(t+(i+1)dt)βkj)|\displaystyle+\sum_{i=0}^{n-1}\left|\mathcal{B}\left(\textup{x}(t+i{\rm d}t),\frac{r_{0}(t+i{\rm d}t)}{\beta_{kj}}\right)\setminus\mathcal{B}\left(\textup{x}(t+(i+1){\rm d}t),\frac{r_{0}(t+(i+1){\rm d}t)}{\beta_{kj}}\right)\right|

Let the time interval [0,T][0,T] be partitioned by points ti=idtt_{i}=i{\rm d}t, i=0,,Tdti=0,...,\frac{T}{{\rm d}t}. We can calculate the area of 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}) by setting t=0t=0 in (20), i.e.,

\IEEEeqnarraymulticol3l|𝒜kj(r0,θ,v,T,βkj)|=limdt0|i=0Tdt(x(idt),r0(idt)βkj)|\displaystyle\IEEEeqnarraymulticol{3}{l}{|\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})|=\lim_{{\rm d}t\to 0}\left|\bigcup_{i=0}^{\frac{T}{{\rm d}t}}\mathcal{B}\left(\textup{x}(i{\rm d}t),\frac{r_{0}(i{\rm d}t)}{\beta_{kj}}\right)\right|}
=\displaystyle= πr0(T)2βkj2+limdt0i=0Tdt1|(x(idt),r0(idt)βkj)(x((i+1)dt),r0((i+1)dt)βkj)|\displaystyle\pi\frac{r_{0}(T)^{2}}{\beta_{kj}^{2}}+\lim_{{\rm d}t\to 0}\sum_{i=0}^{\frac{T}{{\rm d}t}-1}\left|\mathcal{B}\left(\textup{x}(i{\rm d}t),\frac{r_{0}(i{\rm d}t)}{\beta_{kj}}\right)\setminus\mathcal{B}\left(\textup{x}((i+1){\rm d}t),\frac{r_{0}((i+1){\rm d}t)}{\beta_{kj}}\right)\right|
=(a)\displaystyle\stackrel{{\scriptstyle\text{(a)}}}{{=}} πr0(T)2βkj2+limdt0i=0Tdt1\displaystyle\pi\frac{r_{0}(T)^{2}}{\beta_{kj}^{2}}+\lim_{{\rm d}t\to 0}\sum_{i=0}^{\frac{T}{{\rm d}t}-1}
2vβkj2[βkj2r0(ti)2(vtir0cosθ)2arccos(vtir0cosθβkjr0(ti))(vtir0cosθ)]dt+O(dt2)\displaystyle\frac{2v}{\beta_{kj}^{2}}\left[\sqrt{\beta_{kj}^{2}r_{0}(t_{i})^{2}-(vt_{i}-r_{0}\cos\theta)^{2}}-\arccos\left(\frac{vt_{i}-r_{0}\cos\theta}{\beta_{kj}r_{0}(t_{i})}\right)(vt_{i}-r_{0}\cos\theta)\right]{\rm d}t+O({\rm d}t^{2})
=(b)\displaystyle\stackrel{{\scriptstyle\text{(b)}}}{{=}} πr0(T)2βkj2+2vβkj20Tβkj2r0(t)2(vtr0cosθ)2arccos(vtr0cosθβkjr0(t))(vtr0cosθ)dt,\displaystyle\pi\frac{r_{0}(T)^{2}}{\beta_{kj}^{2}}+\frac{2v}{\beta_{kj}^{2}}\int_{0}^{T}\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-(vt-r_{0}\cos\theta)^{2}}-\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t)}\right)(vt-r_{0}\cos\theta){\rm d}t,

where (a) is obtained by using (LABEL:eq:infinitesimal) and (b) follows from the Riemann integral.

IV-C Case III: βkj=1\beta_{kj}=1

For this case, similar to Appendix A, we can prove 𝒜kj(r0,θ,v,T,1)=(x(0),r0)(x(T),r0(T))\mathcal{A}_{kj}(r_{0},\theta,v,T,1)=\mathcal{B}\left(\textup{x}(0),r_{0}\right)\cup\mathcal{B}\left(\textup{x}(T),r_{0}(T)\right). Since |r0vT|r0(T)r0+vT|r_{0}-vT|\leq r_{0}(T)\leq r_{0}+vT, (x(0),r0)\mathcal{B}\left(\textup{x}(0),r_{0}\right) and (x(T),r0(T))\mathcal{B}\left(\textup{x}(T),r_{0}(T)\right) partially overlap. Therefore,

|𝒜kj(r0,θ,v,T,βkj)|=πr02βkj2+πr0(T)2βkj2V(r0βkj,r0(T)βkj,vT),\displaystyle|\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})|=\pi\frac{r_{0}^{2}}{\beta_{kj}^{2}}+\pi\frac{r_{0}(T)^{2}}{\beta_{kj}^{2}}-V\left(\frac{r_{0}}{\beta_{kj}},\frac{r_{0}(T)}{\beta_{kj}},vT\right), (22)

where βkj=1\beta_{kj}=1 and V(r0βkj,r0(T)βkj,vT)V\left(\frac{r_{0}}{\beta_{kj}},\frac{r_{0}(T)}{\beta_{kj}},vT\right) is given in (16).

IV-D Closed-form Expression for |𝒜kj(r0,θ,v,T,βkj)||\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})|

Theorem 1.

Area of 𝒜kj(r0,θ,v,T,βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}) can be obtained by

\IEEEeqnarraymulticol3l|𝒜kj(r0,θ,v,T,βkj)|=|𝒜kj(r0,θ,vT,1,βkj)|=\displaystyle\IEEEeqnarraymulticol{3}{l}{|\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})|=|\mathcal{A}_{kj}(r_{0},\theta,vT,1,\beta_{kj})|=}
{πg(vT,1)2βkj2+2vTβkj201βkj2g(vT,u)2(vTur0cosθ)2arccos(vTur0cosθβkjg(vT,u))(vTur0cosθ)du,if (βkj>1)πr02βkj2+πg(vT,1)2βkj2V(r0βkj,g(vT,1)βkj,vT),if (βkj=1) or (βkj<1 and 2r0cosθβkj1βkj2<vT<2r0cosθ+βkj1βkj2)πr02βkj2,if (βkj<1 and vT2r0cosθβkj1βkj2)πg(vT,1)2βkj2,if (βkj<1 and 2r0cosθ+βkj1βkj2vT),\displaystyle\begin{cases}\pi\frac{g(vT,1)^{2}}{\beta_{kj}^{2}}+\frac{2vT}{\beta_{kj}^{2}}\int\limits_{0}^{1}\sqrt{\beta_{kj}^{2}g(vT,u)^{2}-(vTu-r_{0}\cos\theta)^{2}}-\arccos\left(\frac{vTu-r_{0}\cos\theta}{\beta_{kj}g(vT,u)}\right)(vTu-r_{0}\cos\theta){\rm d}u,\\ \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\!\!\text{if }\left(\beta_{kj}>1\right)\\ \pi\frac{r_{0}^{2}}{\beta_{kj}^{2}}+\pi\frac{g(vT,1)^{2}}{\beta_{kj}^{2}}-V\left(\frac{r_{0}}{\beta_{kj}},\frac{g(vT,1)}{\beta_{kj}},vT\right),\\ \qquad\qquad\qquad\qquad\qquad\qquad\!\text{if }\left(\beta_{kj}=1\right)\text{ or }\left(\beta_{kj}<1\text{ and }2r_{0}\frac{cos\theta-\beta_{kj}}{1-\beta_{kj}^{2}}<vT<2r_{0}\frac{cos\theta+\beta_{kj}}{1-\beta_{kj}^{2}}\right)\\ \pi\frac{r_{0}^{2}}{\beta_{kj}^{2}},\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\text{if }\left(\beta_{kj}<1\text{ and }vT\leq 2r_{0}\frac{cos\theta-\beta_{kj}}{1-\beta_{kj}^{2}}\right)\\ \pi\frac{g(vT,1)^{2}}{\beta_{kj}^{2}},\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\!\!\!\text{if }\left(\beta_{kj}<1\text{ and }2r_{0}\frac{cos\theta+\beta_{kj}}{1-\beta_{kj}^{2}}\leq vT\right)\end{cases},
(23)

where g(vT,u)=r0(Tu)=r02+v2T2u22r0vTucosθg(vT,u)=r_{0}(Tu)=\sqrt{r_{0}^{2}+v^{2}T^{2}u^{2}-2r_{0}vTu\cos\theta} and V(r0βkj,g(vT,1)βkj,vT)V\left(\frac{r_{0}}{\beta_{kj}},\frac{g(vT,1)}{\beta_{kj}},vT\right) is given in (16).

Proof:

The proof follows from combining (15), (LABEL:eq:area_beta_gt_1), and (22). For, βkj>1\beta_{kj}>1, we have used change of variable tT=u\frac{t}{T}=u. ∎

It is worth mentioning that, according to Theorem 1, in multi-tier networks, tier-kk cells are convex, if BkPkBjPjB_{k}P_{k}\leq B_{j}P_{j} (or equivalently, βkj1\beta_{kj}\leq 1), j𝒦\forall j\in\mathcal{K}. Therefore, in this case, we can derive the sojourn time distribution of tier-kk similar to the single-tier scenario using (3) (instead of (2)). This is the reason why some works in the literature only focus on the sojourn time in small cells of two-tier networks.

V Main Results on Sojourn Time and Handoff Rate and Effects of Network Parameters

V-A Conditional CCDF and Mean of Sojourn Time

Since (0,r0βkj)𝒜kj(r0,θ,v,T,βkj)\mathcal{B}\left(0,\frac{r_{0}}{\beta_{kj}}\right)\subset\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}), we can further simplify (9) as

(BS(t)=BS(0),t(0,T]r0,θ,tier=k)=exp{j𝒦λj(|𝒜kj(r0,θ,vT,1,βkj)|πr02βkj2)},\mathbb{P}\left(BS(t)=BS(0),\forall t\in(0,T]\mid r_{0},\theta,\text{tier}=k\right)=\\ \exp\left\{-\sum_{j\in\mathcal{K}}\lambda_{j}\left(\left|\mathcal{A}_{kj}(r_{0},\theta,vT,1,\beta_{kj})\right|-\pi\frac{r_{0}^{2}}{\beta_{kj}^{2}}\right)\right\}, (24)

The CCDF of the sojourn time of a connection in a cell where it is initiated, S~\tilde{S} can be obtained by substituting (5) and (24) in (4).

\IEEEeqnarraymulticol3lF¯S~(Ttier=k)=\displaystyle\IEEEeqnarraymulticol{3}{l}{\bar{F}_{\tilde{S}}(T\mid\text{tier}=k)=} (25)
1(tier=k)00π2λkr0exp{j𝒦λj|𝒜kj(r0,θ,vT,1,βkj)|}dθdr0,\displaystyle\frac{1}{\mathbb{P}(\text{tier}=k)}\int_{0}^{\infty}\int_{0}^{\pi}2\lambda_{k}r_{0}\exp\left\{-\sum_{j\in\mathcal{K}}\lambda_{j}\left|\mathcal{A}_{kj}(r_{0},\theta,vT,1,\beta_{kj})\right|\right\}{\rm d}\theta{\rm d}r_{0},

where (tier=k)\mathbb{P}(\text{tier}=k) is given in (6), and |𝒜kj(r0,θ,vT,1,βkj)|\left|\mathcal{A}_{kj}(r_{0},\theta,vT,1,\beta_{kj})\right| is given in Theorem 1.

As discussed before, in Step 2, we use (25) to derive the linear contact distribution function given that the mobile user is in a tier-kk cell at time 0, i.e.

\IEEEeqnarraymulticol3lH(ztier=k)=1F¯S~(zvtier=k)=\displaystyle\IEEEeqnarraymulticol{3}{l}{H_{\ell}(z\mid\text{tier}=k)=1-\bar{F}_{\tilde{S}}(\frac{z}{v}\mid\text{tier}=k)=} (26)
11(tier=k)00π2λkr0exp{j𝒦λj|𝒜kj(r0,θ,z,1,βkj)|}dθdr0.\displaystyle 1-\frac{1}{\mathbb{P}(\text{tier}=k)}\int_{0}^{\infty}\int_{0}^{\pi}2\lambda_{k}r_{0}\exp\left\{-\sum_{j\in\mathcal{K}}\lambda_{j}\left|\mathcal{A}_{kj}(r_{0},\theta,z,1,\beta_{kj})\right|\right\}{\rm d}\theta{\rm d}r_{0}.

To derive the chord length distribution in tier-kk cells, according to (11), we need ddzH(ztier=k)\frac{{\rm d}}{{\rm d}z}H_{\ell}(z\mid\text{tier}=k) and 𝔼[Ltier=k]\mathbb{E}[L\mid\text{tier}=k]. From (26), we have

ddzH(ztier=k)\displaystyle\frac{{\rm d}}{{\rm d}z}H_{\ell}(z\mid\text{tier}=k) =\displaystyle= 1(tier=k)00π2λkr0(j𝒦λjddz|𝒜kj(r0,θ,z,1,βkj)|)\displaystyle\frac{1}{\mathbb{P}(\text{tier}=k)}\int_{0}^{\infty}\int_{0}^{\pi}2\lambda_{k}r_{0}\left(\sum_{j\in\mathcal{K}}\lambda_{j}\frac{{\rm d}}{{\rm d}z}\left|\mathcal{A}_{kj}(r_{0},\theta,z,1,\beta_{kj})\right|\right) (27)
×exp{j𝒦λj|𝒜kj(r0,θ,z,1,βkj)|}dθdr0,\displaystyle\times\exp\left\{-\sum_{j\in\mathcal{K}}\lambda_{j}\left|\mathcal{A}_{kj}(r_{0},\theta,z,1,\beta_{kj})\right|\right\}{\rm d}\theta{\rm d}r_{0},

where

\IEEEeqnarraymulticol3lddz|𝒜kj(r0,θ,z,1,βkj)|=\displaystyle\IEEEeqnarraymulticol{3}{l}{\frac{{\rm d}}{{\rm d}z}|\mathcal{A}_{kj}(r_{0},\theta,z,1,\beta_{kj})|=}
{π2(zr0cosθ)βkj2+2βkj201βkj2g(z,u)2(zur0cosθ)2arccos(zur0cosθβkjg(z,u))(zur0cosθ)du+2zβkj201βkj2g(z,u)2(zur0cosθ)2u(zur0cosθ)g(z,u)2arccos(zur0cosθβkjg(z,u))udu,if (βkj>1)π2(zr0cosθ)βkj2+βkj21βkj2r024βkj2r02((βkj21)z+2r0cosθ)2+βkj2+12cos2θβkj2r02βkj21βkj2r0zcosθ4βkj2g(z,1)2((βkj2+1)z2r0cosθ)2arccos((βkj2+1)z2r0cosθ2βkjg(z,1))2(zr0cosθ)βkj2+(βkj21)z+2r0(βkjcosθ)(βkj21)z+2r0(βkj+cosθ)(βkj21)z+r0(βkj+cosθ)2βkj2+(βkj21)z+2r0(βkj+cosθ)(βkj21)z+2r0(βkjcosθ)(βkj21)z+r0(βkjcosθ)2βkj2if (βkj=1) or (βkj<1 and 2r0cosθβkj1βkj2<z<2r0cosθ+βkj1βkj2)0,if (βkj<1 and z2r0cosθβkj1βkj2)π2(zr0cosθ)βkj2,if (βkj<1 and 2r0cosθ+βkj1βkj2z).\displaystyle\begin{cases}\pi\frac{2(z-r_{0}\cos\theta)}{\beta_{kj}^{2}}+\frac{2}{\beta_{kj}^{2}}\int\limits_{0}^{1}\sqrt{\beta_{kj}^{2}g(z,u)^{2}-(zu-r_{0}\cos\theta)^{2}}-\arccos\left(\frac{zu-r_{0}\cos\theta}{\beta_{kj}g(z,u)}\right)(zu-r_{0}\cos\theta){\rm d}u\\ +\frac{2z}{\beta_{kj}^{2}}\int\limits_{0}^{1}\sqrt{\beta_{kj}^{2}g(z,u)^{2}-(zu-r_{0}\cos\theta)^{2}}\frac{u(zu-r_{0}\cos\theta)}{g(z,u)^{2}}-\arccos\left(\frac{zu-r_{0}\cos\theta}{\beta_{kj}g(z,u)}\right)u{\rm d}u,\\ \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\!\!\text{if }\left(\beta_{kj}>1\right)\\ \pi\frac{2(z-r_{0}\cos\theta)}{\beta_{kj}^{2}}+\frac{\frac{\beta_{kj}^{2}-1}{\beta_{kj}^{2}}r_{0}^{2}}{\sqrt{4\beta_{kj}^{2}r_{0}^{2}-\left(\left(\beta_{kj}^{2}-1\right)z+2r_{0}\cos\theta\right)^{2}}}+\frac{\frac{\beta_{kj}^{2}+1-2\cos^{2}\theta}{\beta_{kj}^{2}}r_{0}^{2}-\frac{\beta_{kj}^{2}-1}{\beta_{kj}^{2}}r_{0}z\cos\theta}{\sqrt{4\beta_{kj}^{2}g(z,1)^{2}-\left(\left(\beta_{kj}^{2}+1\right)z-2r_{0}\cos\theta\right)^{2}}}\\ -\arccos\left(\frac{\left(\beta_{kj}^{2}+1\right)z-2r_{0}\cos\theta}{2\beta_{kj}g(z,1)}\right)\frac{2(z-r_{0}\cos\theta)}{\beta_{kj}^{2}}+\sqrt{\frac{-\left(\beta_{kj}^{2}-1\right)z+2r_{0}\left(\beta_{kj}-\cos\theta\right)}{\left(\beta_{kj}^{2}-1\right)z+2r_{0}\left(\beta_{kj}+\cos\theta\right)}}\frac{\left(\beta_{kj}^{2}-1\right)z+r_{0}\left(\beta_{kj}+\cos\theta\right)}{2\beta_{kj}^{2}}\\ +\sqrt{\frac{\left(\beta_{kj}^{2}-1\right)z+2r_{0}\left(\beta_{kj}+\cos\theta\right)}{-\left(\beta_{kj}^{2}-1\right)z+2r_{0}\left(\beta_{kj}-\cos\theta\right)}}\frac{-\left(\beta_{kj}^{2}-1\right)z+r_{0}\left(\beta_{kj}-\cos\theta\right)}{2\beta_{kj}^{2}}\\ \qquad\qquad\qquad\qquad\qquad\qquad\;\;\text{if }\left(\beta_{kj}=1\right)\text{ or }\left(\beta_{kj}<1\text{ and }2r_{0}\frac{cos\theta-\beta_{kj}}{1-\beta_{kj}^{2}}<z<2r_{0}\frac{cos\theta+\beta_{kj}}{1-\beta_{kj}^{2}}\right)\\ 0,\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\quad\quad\quad\;\;\;\;\text{if }\left(\beta_{kj}<1\text{ and }z\leq 2r_{0}\frac{cos\theta-\beta_{kj}}{1-\beta_{kj}^{2}}\right)\\ \pi\frac{2(z-r_{0}\cos\theta)}{\beta_{kj}^{2}},\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\!\text{if }\left(\beta_{kj}<1\text{ and }2r_{0}\frac{cos\theta+\beta_{kj}}{1-\beta_{kj}^{2}}\leq z\right)\end{cases}.
(28)

𝔼[Ltier=k]\mathbb{E}[L\mid\text{tier}=k] is also provided in the following theorem.

Theorem 2.

The mean length of the chords lying in tier-kk cells is as

𝔼[Ltier=k]=π(j𝒦λjβjk2)1/2j𝒦λj(βkj),\displaystyle\mathbb{E}[L\mid\emph{tier}=k]=\pi\frac{\left(\sum_{j\in\mathcal{K}}\lambda_{j}\beta_{jk}^{2}\right)^{1/2}}{\sum_{j\in\mathcal{K}}\lambda_{j}\mathcal{I}(\beta_{kj})},

where

(β)={1β20πβ2+12cos2θβ2cos2θdθ,if β11β2arccos(β)πarccos(β)β2+12cos2θβ2cos2θdθ,if β<1.\displaystyle\mathcal{I}(\beta)=\begin{cases}\frac{1}{\beta^{2}}\int_{0}^{\pi}\frac{\beta^{2}+1-2\cos^{2}\theta}{\sqrt{\beta^{2}-\cos^{2}\theta}}{\rm d}\theta,&\text{if }\beta\geq 1\\ \frac{1}{\beta^{2}}\int_{\arccos(\beta)}^{\pi-\arccos(\beta)}\frac{\beta^{2}+1-2\cos^{2}\theta}{\sqrt{\beta^{2}-\cos^{2}\theta}}{\rm d}\theta,&\text{if }\beta<1\end{cases}. (29)
Proof:

See Appendix C. ∎

Using these results, the mean and the CCDF of the sojourn time are

𝔼[Stier=k]\displaystyle\mathbb{E}[S\mid\text{tier}=k] =\displaystyle= πv(j𝒦λjβjk2)1/2j𝒦λj(βkj),\displaystyle\frac{\pi}{v}\frac{\left(\sum_{j\in\mathcal{K}}\lambda_{j}\beta_{jk}^{2}\right)^{1/2}}{\sum_{j\in\mathcal{K}}\lambda_{j}\mathcal{I}(\beta_{kj})}, (30)
F¯S(Ttier=k)\displaystyle\bar{F}_{S}(T\mid\text{tier}=k) =\displaystyle= 𝔼[Ltier=k]ddzH(ztier=k)|z=vT.\displaystyle\mathbb{E}[L\mid\text{tier}=k]\frac{{\rm d}}{{\rm d}z}H_{\ell}(z\mid\text{tier}=k)\Big{|}_{z=vT}. (31)

V-B Handoff Rate

Refer to caption
Figure 5: (β)\mathcal{I}(\beta) and (β)\mathcal{F}(\beta) for β1\beta\geq 1.

In [9], rates of different handoff types in multi-tier cellular networks are provided. For k,j𝒦k,j\in\mathcal{K}, the type kk-jj handoff rate HkjH_{kj}, defined as the mean number of handoffs made from a tier-kk cell to a tier-jj cell in unit time, is222 HkjH_{kj} in (32) is obtained by further simplifying the result in [9]. Specifically, we have used βij=βikβjk\beta_{ij}=\frac{\beta_{ik}}{\beta_{jk}} besides (1β)=β3(β)\mathcal{F}(\frac{1}{\beta})=\beta^{3}\mathcal{F}(\beta).

Hkj=vπλkλj(βkj)(i𝒦λiβik2)3/2,\displaystyle H_{kj}=\frac{v}{\pi}\frac{\lambda_{k}\lambda_{j}\mathcal{F}(\beta_{kj})}{\left(\sum_{i\in\mathcal{K}}\lambda_{i}\beta_{ik}^{2}\right)^{3/2}}, (32)

where

(β)=1β20πβ2+12βcosθdθ.\displaystyle\mathcal{F}(\beta)=\frac{1}{\beta^{2}}\int_{0}^{\pi}\sqrt{\beta^{2}+1-2\beta\cos\theta}{\rm d}\theta. (33)

Therefore,

Hk=vπλkj𝒦λj(βkj)(i𝒦λiβik2)3/2\displaystyle H_{k}=\frac{v}{\pi}\lambda_{k}\frac{\sum_{j\in\mathcal{K}}\lambda_{j}\mathcal{F}(\beta_{kj})}{\left(\sum_{i\in\mathcal{K}}\lambda_{i}\beta_{ik}^{2}\right)^{3/2}} (34)

is the mean number of handoffs from (to) a tier-kk cell to (from) any other cell in the network.

In Fig. 5, we can see (β)=(β)\mathcal{I}(\beta)=\mathcal{F}(\beta) for β1\beta\geq 1. Since (1β)=β3(β)\mathcal{F}(\frac{1}{\beta})=\beta^{3}\mathcal{F}(\beta) and (1β)=β3(β)\mathcal{I}(\frac{1}{\beta})=\beta^{3}\mathcal{I}(\beta), we can conclude (β)=(β)\mathcal{I}(\beta)=\mathcal{F}(\beta) for any β>0\beta>0. Using this result, the relation between mean sojourn time and handoff rate for tier-kk is as follows:

𝔼[Stier=k]=(tier=k)Hk.\displaystyle\mathbb{E}[S\mid\text{tier}=k]=\frac{\mathbb{P}(\text{tier}=k)}{H_{k}}. (35)

An important metric in mobility analysis is the fraction of time the mobile user stays in tier-kk cells during a movement period since it considers both handoff rate and sojourn time.

Corollary 1.

In high velocity scenarios or mobility models with low direction switch rate, the fraction of time the mobile user stays in tier-kk, during the movement period, is (tier=k)\mathbb{P}(\emph{tier}=k).

Proof:

Let us denote the number of times that the mobile user enters a tier-kk cell during the movement period by 𝒩k\mathcal{N}_{k}, k𝒦k\in\mathcal{K}. Also, ti(k)t_{i}^{(k)} denotes the ii-th dwell time in the tier-kk cell, where i=1,,𝒩ki=1,...,\mathcal{N}_{k} and k𝒦k\in\mathcal{K}. The fraction of time that the mobile user stays in tier-kk can be obtained by

i=1𝒩kti(k)j𝒦i=1𝒩jti(j)=i=1𝒩kti(k)𝒩k×𝒩kj𝒦i=1𝒩jti(j)=(a)𝔼[Stier=k]×Hk=(b)(tier=k),\displaystyle\frac{\sum_{i=1}^{\mathcal{N}_{k}}t_{i}^{(k)}}{\sum_{j\in\mathcal{K}}\sum_{i=1}^{\mathcal{N}_{j}}t_{i}^{(j)}}=\frac{\sum_{i=1}^{\mathcal{N}_{k}}t_{i}^{(k)}}{\mathcal{N}_{k}}\times\frac{\mathcal{N}_{k}}{\sum_{j\in\mathcal{K}}\sum_{i=1}^{\mathcal{N}_{j}}t_{i}^{(j)}}\stackrel{{\scriptstyle\text{(a)}}}{{=}}\mathbb{E}[S\mid\text{tier}=k]\times H_{k}\stackrel{{\scriptstyle\text{(b)}}}{{=}}\mathbb{P}(\text{tier}=k),

where (a) is obtained since we have assumed user crosses a large number of cell boundaries during a movement period, i.e. vTvT is large compared to the average cell size. (b) also follows from (35). ∎

V-C Unconditional CCDF and Mean of Sojourn Time

So far we have focused on F¯S(Ttier=k)\bar{F}_{S}(T\mid\text{tier}=k) and 𝔼[Stier=k]\mathbb{E}[S\mid\text{tier}=k], i.e. the CCDF and average of sojourn time in a tier-kk cell. In the following corollaries, we derive the unconditional CCDF and mean.

Corollary 2.

In high velocity scenarios or mobility models with low direction switch rate, the CCDF of sojourn time, during the movement period, can be obtained by

F¯S(T)=k𝒦F¯S(Ttier=k)HkH\displaystyle\bar{F}_{S}(T)=\sum_{k\in\mathcal{K}}\bar{F}_{S}(T\mid\emph{tier}=k)\frac{H_{k}}{H}

where H=k𝒦HkH=\sum_{k\in\mathcal{K}}H_{k} is the mean number of handoffs in unit time.

Proof:

Using the same notation as in the proof of Corollary 1, we can write

F¯S(T)=𝔼[𝟏(S>T)]\displaystyle\bar{F}_{S}(T)=\mathbb{E}\left[\mathbf{1}(S>T)\right] =\displaystyle= k𝒦i=1𝒩k𝟏(ti(k)>T)j𝒦𝒩j\displaystyle\frac{\sum_{k\in\mathcal{K}}\sum_{i=1}^{\mathcal{N}_{k}}\mathbf{1}(t_{i}^{(k)}>T)}{\sum_{j\in\mathcal{K}}\mathcal{N}_{j}}
=\displaystyle= k𝒦i=1𝒩k𝟏(ti(k)>T)𝒩k×𝒩kj𝒦i=1𝒩jti(j)×j𝒦i=1𝒩jti(j)j𝒦𝒩j.\displaystyle\sum_{k\in\mathcal{K}}\frac{\sum_{i=1}^{\mathcal{N}_{k}}\mathbf{1}(t_{i}^{(k)}>T)}{\mathcal{N}_{k}}\times\frac{\mathcal{N}_{k}}{\sum_{j\in\mathcal{K}}\sum_{i=1}^{\mathcal{N}_{j}}t_{i}^{(j)}}\times\frac{\sum_{j\in\mathcal{K}}\sum_{i=1}^{\mathcal{N}_{j}}t_{i}^{(j)}}{\sum_{j\in\mathcal{K}}\mathcal{N}_{j}}.

Corollary 3.

The mean sojourn time during a movement period can be obtained by

𝔼[S]=1H=πv(k𝒦λkj𝒦λj(βkj)(i𝒦λiβik2)3/2)1.\displaystyle\mathbb{E}\left[S\right]=\frac{1}{H}=\frac{\pi}{v}\left(\sum_{k\in\mathcal{K}}\lambda_{k}\frac{\sum_{j\in\mathcal{K}}\lambda_{j}\mathcal{F}(\beta_{kj})}{\left(\sum_{i\in\mathcal{K}}\lambda_{i}\beta_{ik}^{2}\right)^{3/2}}\right)^{-1}.
Proof:

The mean sojourn time can be obtained by following the same approach as in proof of Corollary 2. ∎

V-D Effect of Network Parameters

In this subsection, we study the effect user velocity, transmit power, bias factor, and BS intensity on the distribution and mean of the sojourn time.

Proposition 2.

The CCDF and the mean of the sojourn time decrease as the mobile user’s velocity increases.

Proof:

This can be understood from the definition of the sojourn time. (This can also be proven from the derived analytical results.) ∎

Proposition 3.

In multi-tier networks, the sojourn time of kk-th tier increases as transmit power or bias factor of tier-kk increases, while sojourn time in other tiers decreases. In single-tier networks, sojourn time is independent of transmit power and bias factor.

Proof:

Assume that a user at location y2\text{y}\in\mathbb{R}^{2} is served by a tier-kk BS at xx, i.e, y is in the cell of xx. Therefore, from (1), we have

BkPkyxα\displaystyle B_{k}P_{k}\|\text{y}-x\|^{-\alpha} \displaystyle\geq BkPkyzα,zΦk,\displaystyle B_{k}P_{k}\|\text{y}-z\|^{-\alpha},\qquad z\in\Phi_{k}, (36)
BkPkyxα\displaystyle B_{k}P_{k}\|\text{y}-x\|^{-\alpha} \displaystyle\geq BjPjyzα,j𝒦{k}, and zj𝒦{k}Φj.\displaystyle B_{j}P_{j}\|\text{y}-z\|^{-\alpha},\qquad j\in\mathcal{K}\setminus\{k\},\text{ and }z\in\cup_{j\in\mathcal{K}\setminus\{k\}}\Phi_{j}. (37)

From these equations, when BkPkBkPkB^{\prime}_{k}P^{\prime}_{k}\geq B_{k}P_{k}, we obtain

BkPkyxα\displaystyle B^{\prime}_{k}P^{\prime}_{k}\|\text{y}-x\|^{-\alpha} \displaystyle\geq BkPkyzα,zΦk,\displaystyle B^{\prime}_{k}P^{\prime}_{k}\|\text{y}-z\|^{-\alpha},\qquad z\in\Phi_{k},
BkPkyxα\displaystyle B^{\prime}_{k}P^{\prime}_{k}\|\text{y}-x\|^{-\alpha} \displaystyle\geq BjPjyzα,j𝒦{k}, and zj𝒦{k}Φj,\displaystyle B_{j}P_{j}\|\text{y}-z\|^{-\alpha},\qquad j\in\mathcal{K}\setminus\{k\},\text{ and }z\in\cup_{j\in\mathcal{K}\setminus\{k\}}\Phi_{j},

i.e. y is still in the cell of xx after increasing BkPkB_{k}P_{k}. Therefore, in multi-tier networks, the size of the kk-th tier cells increases as transmit power or bias factor of tier-kk increases. Similarly, we can show that the size of other tiers’ cells decreases with increasing transmit power or bias factor of tier-kk. On the other hand, in single-tier networks, according to (36), cell sizes are independent of transmit power and bias factor. Finally, using these results and the fact that the sojourn time is directly proportional to the size of cells, we can obtain Proposition 3. ∎

Proposition 4.

In multi-tier networks, when the BS intensity of tier-kk increases, the sojourn time for other tiers decreases.

Proof:

According to the superposition property of PPP [17], increasing λk\lambda_{k} to λk\lambda^{\prime}_{k} is similar to adding a new tier of BSs with intensity λkλk\lambda^{\prime}_{k}-\lambda_{k}, transmission power PkP_{k}, and bias factor BkB_{k}. Therefore, the size of cells of other tiers decreases when the BS intensity of kk-th tier increases. ∎

VI Numerical and Simulation Results

VI-A Distribution of Sojourn Time


Refer to caption
(a) CCDF of S~\tilde{S}.


Refer to caption
(b) CCDF of SS.
Figure 6: Distribution of S~\tilde{S} and SS in a two-tier cellular network (for λ1=0.002\lambda_{1}=0.002, λ2=0.005\lambda_{2}=0.005, β12=(12)1/4\beta_{12}=\left(\frac{1}{2}\right)^{1/4}, and v=5v=5).

For a two-tier cellular network, in Fig. 6(a), the distribution of the sojourn time S~\tilde{S} in the cell where the connection is initiated is illustrated for tier-kk, k{1,2}k\in\{1,2\}. In Fig. 6(b), the distribution of (conditional and unconditional) SS, for this network, is provided. As can be seen, the simulation results match the derived analytical results. According to Fig. 6, at high velocities, the sojourn time for tier-kk stochastically dominates the sojourn time of tier-jj when BkPk>BjPjB_{k}P_{k}>B_{j}P_{j}, i.e. βkj>1\beta_{kj}>1.

In practice, when a mobile user crosses a cell boundary, it starts a Time to Trigger (TTT) timer. The mobile user does not make a handoff to the new BS, if it leaves the new BS’s cell before the end of TTT timer [6]. The derived results for handoff rates, in the literature, usually assume TTT is 0, i.e. the handoff rates provided (in [4] and [9] for example) are actually the mean number of intersections between the user trajectory and cell boundaries per unit time. In practice, the handoff rate for tier-kk is Hk(S>TTTtier=k)=HkF¯S(TTTtier=k)H_{k}\mathbb{P}(S>\text{TTT}\mid\text{tier}=k)=H_{k}\bar{F}_{S}(\text{TTT}\mid\text{tier}=k), where HkH_{k} is given in (34). When the network parameters are as in Fig. 6, H1=0.13H_{1}=0.13 and H2=0.41H_{2}=0.41. For this network, with TTT=0.2\text{TTT}=0.2, the handoff rate for tier-one is 0.12 and for tier-two is 0.39. Although the difference between HkH_{k} and the handoff rate for these parameters is negligible, it is noticeable for high velocity scenarios.

Moreover, using the distribution of sojourn time, we can study the ping-pong rate (unnecessary handoff rate). If, after a handoff, the time duration that the mobile user is inside the new cell be less than a threshold TpT_{p}, the handoff is considered unnecessary [6]. Therefore, for tier-kk, the ping-pong rate can be obtained by [6]

Hk((S<Tptier=k)(S<TTTtier=k))=Hk((S>TTTtier=k)(S>Tptier=k)).H_{k}\left(\mathbb{P}(S<T_{p}\mid\text{tier}=k)-\mathbb{P}(S<\text{TTT}\mid\text{tier}=k)\right)=\\ H_{k}\left(\mathbb{P}(S>\text{TTT}\mid\text{tier}=k)-\mathbb{P}(S>T_{p}\mid\text{tier}=k)\right). (38)

For TTT=0.2\text{TTT}=0.2 and Tp=0.5T_{p}=0.5, for tier-one, the ping-pong rate is less than 0.010.01 and for tier-two, the ping-pong rate is 0.030.03.

Refer to caption
Figure 7: Sojourn time and complement of the handoff probability in a three-tier network. For tier-1, λ1=0.01\lambda_{1}=0.01 and B1P1=10B_{1}P_{1}=10, for tier-2, λ2=0.005\lambda_{2}=0.005 and B2P2=50B_{2}P_{2}=50, and for tier-3, λ3=0.001\lambda_{3}=0.001 and B3P3=100B_{3}P_{3}=100. α=4\alpha=4 and v=5v=5.

As discussed earlier, we obtain the distribution of S~\tilde{S} in multi-tier networks from (2). For convex cells, (2) can be further simplified as (3) which is the complement of the handoff probability. Therefore, we can use (3) to derive the CCDF of the sojourn time in single-tier networks and also in multi-tier networks for the tier with the smallest BPBP (multiplication of bias factor and transmission power). However, for other tiers in multi-tier networks, (3) provides an upper bound for the CCDF of the sojourn time. This is also illustrated in Fig. 7 for a three-tier network. It is worth mentioning that the gap between the CCDF of the sojourn time and its upper bound (obtained from (3)) increases as the intensity of tiers with lower BPBP increases.

VI-B Mean Sojourn Time

Refer to caption
Figure 8: Mean sojourn time with respect to velocity. We have compared the results for a two-tier cellular network with λ1=0.002\lambda_{1}=0.002, λ2=0.005\lambda_{2}=0.005, and β12=(12)1/4\beta_{12}=\left(\frac{1}{2}\right)^{1/4} with two single-tier networks.

In Fig. 8, the mean sojourn time for a two-tier cellular network with λ1=0.002\lambda_{1}=0.002 and λ2=0.005\lambda_{2}=0.005 is illustrated as a function of velocity. We compare the results for the two-tier network with two single-tier scenarios where the mobile user is associated to only one of the tiers. When the number of tiers increases, through increased spectral reuse, users can transmit with higher data rates. However, there is more undesired overhead transmission due to the higher handoff rate (lower mean sojourn time). The sojourn time distribution is helpful in mobility management where the mobile user can skip unnecessary handoffs with a negligible spectral efficiency loss [13].


Refer to caption
(a) Effect of transmission power (or bias factor).


Refer to caption
(b) Effect of BS intensity.
Figure 9: Effect of network parameters on the mean sojourn time in a two-tier cellular network with λ1=0.002\lambda_{1}=0.002, B1P1=100B_{1}P_{1}=100, α=4\alpha=4, and v=5v=5. a) Effect of increasing transmission power (or bias factor) when λ2=0.005\lambda_{2}=0.005. b) Effect of increasing BS intensity when B2P2=50B_{2}P_{2}=50.

In Fig. 9(a), the effect of transmit power (or bias factor) on the mean sojourn time in a two-tier cellular network is illustrated. As discussed in Proposition 3, the mean sojourn time of tier-two increases as transmit power (or bias factor) of tier-two increases, while mean sojourn time of other tier decreases. As can be seen, the (unconditional) mean sojourn time in the network does not change with increasing transmit power or bias factor of tier-two. In Fig. 9(b), the effect of BS intensity on the mean sojourn time is shown. As can be seen, the mean sojourn time for all tiers decreases with increasing the BS intensity of tier-two. This is also mentioned in Proposition 4.

VII Conclusion

We have derived the distribution and mean of the sojourn time of multi-tier cellular networks. The existing works assume that a mobile user is always associated to only one of the tiers, or focus on the sojourn time in small cells (for two-tier scenario). Since in both the cases the cells are convex, the sojourn time distribution (or mean) can be easily obtained similar to single-tier scenarios by using the chord length distribution in Poisson Voronoi tessellation. However, in multi-tier networks with maximum biased averaged received power association we need the chord length distribution in weighted Poisson Voronoi tessellation, which is not available in the literature. In this paper, we have derived the linear contact distribution function in weighted Poisson Voronoi tessellation from which we obtained the chord length distribution.

We have studied the relation between mean sojourn time and other mobility-related performance metrics. Specifically, We have shown that mean sojourn time is inversely proportional to the handoff rate. Also, the complementary cumulative distribution function of sojourn time is upper bounded by complement of the handoff probability. In addition, we have studied the impact of user velocity and network parameters on the distribution and mean of the sojourn time. The sojourn time distribution can be used to derive the ping-pong rate which is important in mobility management where the mobile user can skip unnecessary handoffs with a negligible spectral efficiency loss. Moreover, it can be used for studying channel occupancy time which can be exploited for improving resource allocation.

Appendix A: Proof of Proposition 1

Refer to caption
Figure 10: A geometric illustration for proof of Proposition 1.

From (8), we have (x(0),r0βkj)(x(T),r0(T)βkj)𝒜kj(r0,θ,v,T,βkj)\mathcal{B}\left(\textup{x}(0),\frac{r_{0}}{\beta_{kj}}\right)\cup\mathcal{B}\left(\textup{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right)\subset\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj}). To complete the proof we need to show that 𝒜kj(r0,θ,v,T,βkj)(x(0),r0βkj)(x(T),r0(T)βkj)\mathcal{A}_{kj}(r_{0},\theta,v,T,\beta_{kj})\subset\mathcal{B}\left(\textup{x}(0),\frac{r_{0}}{\beta_{kj}}\right)\cup\mathcal{B}\left(\textup{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right).

Consider a point y(x(t),r0(t)βkj)\text{y}\in\mathcal{B}\left(\text{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right), 0tT0\leq t\leq T. Let us represent y in polar coordinates as (ζ,φ)(\zeta,\varphi), where ζ\zeta is the distance between y and the origin (x(0)\text{x}(0)) and φ\varphi is the angle made between the line segment from the origin to y and the positive xx-axis (user’s trajectory) (Fig. 10). Using triangle equations, we have yx(t)=ζ2+v2t22ζvtcosφ\|\text{y}-\text{x}(t)\|=\sqrt{\zeta^{2}+v^{2}t^{2}-2\zeta vt\cos\varphi}. Since y(x(t),r0(t)βkj)\text{y}\in\mathcal{B}\left(\text{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right), 0tT0\leq t\leq T,

ζ2+v2t22ζvtcosφr0(t)2βkj2=r02+v2t22r0vtcosθβkj2.\displaystyle\zeta^{2}+v^{2}t^{2}-2\zeta vt\cos\varphi\leq\frac{r_{0}(t)^{2}}{\beta_{kj}^{2}}=\frac{r_{0}^{2}+v^{2}t^{2}-2r_{0}vt\cos\theta}{\beta_{kj}^{2}}.

Rewriting the above inequality gives

ζ2r02βkj2(1βkj21)v2t2+2vt(ζcosφr0cosθβkj2),0tT.\displaystyle\zeta^{2}-\frac{r_{0}^{2}}{\beta_{kj}^{2}}\leq\left(\frac{1}{\beta_{kj}^{2}}-1\right)v^{2}t^{2}+2vt\left(\zeta\cos\varphi-\frac{r_{0}\cos\theta}{\beta_{kj}^{2}}\right),\qquad 0\leq t\leq T.

For βkj<1\beta_{kj}<1 (0<1βkj210<\frac{1}{\beta_{kj}^{2}}-1), the right hand side of the above inequality is a convex function with respect to tt. When x[a,b]x\in[a,b], for a convex function ff, we have f(x)max{f(a),f(b)}f(x)\leq\max\{f(a),f(b)\}. Using this property of convex functions yields,

ζ2r02βkj2max{0,(1βkj21)v2T2+2vT(ζcosφr0cosθβkj2)}.\displaystyle\zeta^{2}-\frac{r_{0}^{2}}{\beta_{kj}^{2}}\leq\max\left\{0,\left(\frac{1}{\beta_{kj}^{2}}-1\right)v^{2}T^{2}+2vT\left(\zeta\cos\varphi-\frac{r_{0}\cos\theta}{\beta_{kj}^{2}}\right)\right\}.

Therefore, depending on the parameters, we have

ζ2r02βkj20,orζ2r02βkj2(1βkj21)v2T2+2vT(ζcosφr0cosθβkj2).\displaystyle\zeta^{2}-\frac{r_{0}^{2}}{\beta_{kj}^{2}}\leq 0,\quad\text{or}\quad\zeta^{2}-\frac{r_{0}^{2}}{\beta_{kj}^{2}}\leq\left(\frac{1}{\beta_{kj}^{2}}-1\right)v^{2}T^{2}+2vT\left(\zeta\cos\varphi-\frac{r_{0}\cos\theta}{\beta_{kj}^{2}}\right). (A.1)

We can rewrite the inequalities in (A.1) as

ζ2r02βkj2,orζ2+v2T22ζvTcosφr02+v2T22r0vTcosθβkj2.\displaystyle\zeta^{2}\leq\frac{r_{0}^{2}}{\beta_{kj}^{2}},\quad\text{or}\quad\zeta^{2}+v^{2}T^{2}-2\zeta vT\cos\varphi\leq\frac{r_{0}^{2}+v^{2}T^{2}-2r_{0}vT\cos\theta}{\beta_{kj}^{2}}. (A.2)

ζ2r02βkj2\zeta^{2}\leq\frac{r_{0}^{2}}{\beta_{kj}^{2}} is equivalent to yx(0)r0βkj\|\text{y}-\text{x}(0)\|\leq\frac{r_{0}}{\beta_{kj}}, and ζ2+v2T22ζvTcosφr02+v2T22r0vTcosθβkj2\zeta^{2}+v^{2}T^{2}-2\zeta vT\cos\varphi\leq\frac{r_{0}^{2}+v^{2}T^{2}-2r_{0}vT\cos\theta}{\beta_{kj}^{2}} is equivalent to yx(T)r0(T)βkj\|\text{y}-\text{x}(T)\|\leq\frac{r_{0}(T)}{\beta_{kj}}. Thus, y(x(0),r0βkj)(x(T),r0(T)βkj)\text{y}\in\mathcal{B}\left(\textup{x}(0),\frac{r_{0}}{\beta_{kj}}\right)\cup\mathcal{B}\left(\textup{x}(T),\frac{r_{0}(T)}{\beta_{kj}}\right).

Appendix B: Proof of (LABEL:eq:infinitesimal)

Before proving (LABEL:eq:infinitesimal), we provide the Taylor series expansion of arccos\arccos and arcsin\arcsin. They help us to derive the final result in (LABEL:eq:infinitesimal).

The arcsin\arcsin function has a Taylor expansion:

arcsin(x)=n=0(2n)!22n(n!)2x2n+12n+1,\displaystyle\arcsin(x)=\sum_{n=0}^{\infty}\frac{(2n)!}{2^{2n}(n!)^{2}}\frac{x^{2n+1}}{2n+1}, (B.1)

By taking derivative with respect to xx from both sides of (B.1), we get

11x2=n=0(2n)!22n(n!)2x2n.\displaystyle\frac{1}{\sqrt{1-x^{2}}}=\sum_{n=0}^{\infty}\frac{(2n)!}{2^{2n}(n!)^{2}}x^{2n}. (B.2)

Using arccos(x)=π2arcsin(x)\arccos(x)=\frac{\pi}{2}-\arcsin(x), we can write

arccos(x)=π2n=0(2n)!22n(n!)2x2n+12n+1.\displaystyle\arccos(x)=\frac{\pi}{2}-\sum_{n=0}^{\infty}\frac{(2n)!}{2^{2n}(n!)^{2}}\frac{x^{2n+1}}{2n+1}. (B.3)

When dx0{\rm d}x\to 0,

arccos(x+dx)\displaystyle\arccos(x+{\rm d}x) =\displaystyle= π2n=0(2n)!22n(n!)2(x+dx)2n+12n+1\displaystyle\frac{\pi}{2}-\sum_{n=0}^{\infty}\frac{(2n)!}{2^{2n}(n!)^{2}}\frac{(x+{\rm d}x)^{2n+1}}{2n+1} (B.4)
=\displaystyle= π2n=0(2n)!22n(n!)2x2n+1+(2n+1)x2ndx+O(dx2)2n+1\displaystyle\frac{\pi}{2}-\sum_{n=0}^{\infty}\frac{(2n)!}{2^{2n}(n!)^{2}}\frac{x^{2n+1}+(2n+1)x^{2n}{\rm d}x+O({\rm d}x^{2})}{2n+1}
=\displaystyle= π2n=0(2n)!22n(n!)2x2n+12n+1n=0(2n)!22n(n!)2x2ndx+O(dx2)\displaystyle\frac{\pi}{2}-\sum_{n=0}^{\infty}\frac{(2n)!}{2^{2n}(n!)^{2}}\frac{x^{2n+1}}{2n+1}-\sum_{n=0}^{\infty}\frac{(2n)!}{2^{2n}(n!)^{2}}x^{2n}{\rm d}x+O({\rm d}x^{2})
=(a)\displaystyle\stackrel{{\scriptstyle\text{(a)}}}{{=}} arccos(x)dx1x2+O(dx2),\displaystyle\arccos(x)-\frac{{\rm d}x}{\sqrt{1-x^{2}}}+O({\rm d}x^{2}),

where (a) is obtained using (B.2) and (B.3).

From (B.4), we obtain

\IEEEeqnarraymulticol3larccos(r0cosθvtβkjr0(t)+βkj212βkjvdtr0(t))=\displaystyle\IEEEeqnarraymulticol{3}{l}{\arccos\left(\frac{r_{0}\cos\theta-vt}{\beta_{kj}r_{0}(t)}+\frac{\beta_{kj}^{2}-1}{2\beta_{kj}}\frac{v{\rm d}t}{r_{0}(t)}\right)=} (B.5)
arccos(r0cosθvtβkjr0(t))(βkj21)vdt2βkj2r0(t)2(vtr0cosθ)2+O(dt2)=(a)\displaystyle\arccos\left(\frac{r_{0}\cos\theta-vt}{\beta_{kj}r_{0}(t)}\right)-\frac{\left(\beta_{kj}^{2}-1\right)v{\rm d}t}{2\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-\left(vt-r_{0}\cos\theta\right)^{2}}}+O({\rm d}t^{2})\stackrel{{\scriptstyle\text{(a)}}}{{=}}
πarccos(vtr0cosθβkjr0(t))(βkj21)vdt2βkj2r0(t)2(vtr0cosθ)2+O(dt2),\displaystyle\pi-\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t)}\right)-\frac{\left(\beta_{kj}^{2}-1\right)v{\rm d}t}{2\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-\left(vt-r_{0}\cos\theta\right)^{2}}}+O({\rm d}t^{2}),

where (a) follows from arccos(x)=πarccos(x)\arccos(-x)=\pi-\arccos(x).

Using r0(t+dt)=r0(t)(1+vdtr0(t)2(vtr0cosθ)+O(dt2))r_{0}(t+{\rm d}t)=r_{0}(t)\left(1+\frac{v{\rm d}t}{r_{0}(t)^{2}}(vt-r_{0}\cos\theta)+O({\rm d}t^{2})\right), when dt0{\rm d}t\to 0, (B.4), and Taylor expansion of (1+x)1(1+x)^{-1} also yields

\IEEEeqnarraymulticol3larccos(vtr0cosθβkjr0(t+dt)+βkj2+12βkjvdtr0(t+dt))=\displaystyle\IEEEeqnarraymulticol{3}{l}{\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t+{\rm d}t)}+\frac{\beta_{kj}^{2}+1}{2\beta_{kj}}\frac{v{\rm d}t}{r_{0}(t+{\rm d}t)}\right)=} (B.6)
arccos(vtr0cosθβkjr0(t)(vtr0cosθ)2r0(t)2vdtβkjr0(t)+βkj2+12βkjvdtr0(t)+O(dt2))=\displaystyle\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t)}-\frac{(vt-r_{0}\cos\theta)^{2}}{r_{0}(t)^{2}}\frac{v{\rm d}t}{\beta_{kj}r_{0}(t)}+\frac{\beta_{kj}^{2}+1}{2\beta_{kj}}\frac{v{\rm d}t}{r_{0}(t)}+O({\rm d}t^{2})\right)=
arccos(vtr0cosθβkjr0(t))vdtβkj2r0(t)2(vtr0cosθ)2(βkj212+1(vtr0cosθ)2r0(t)2)\displaystyle\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t)}\right)-\frac{v{\rm d}t}{\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-\left(vt-r_{0}\cos\theta\right)^{2}}}\left(\frac{\beta_{kj}^{2}-1}{2}+1-\frac{(vt-r_{0}\cos\theta)^{2}}{r_{0}(t)^{2}}\right)

From binomial series expansion, we also get

\IEEEeqnarraymulticol3l2vdtβkj(r0(t)vtr0cosθβkj)+v2dt2(11βkj2)×\displaystyle\IEEEeqnarraymulticol{3}{l}{\sqrt{\frac{2v{\rm d}t}{\beta_{kj}}\left(r_{0}(t)-\frac{vt-r_{0}\cos\theta}{\beta_{kj}}\right)+v^{2}{\rm d}t^{2}\left(1-\frac{1}{\beta_{kj}^{2}}\right)}\times}
2vdtβkj(r0(t)+vtr0cosθβkj)v2dt2(11βkj2)\displaystyle\sqrt{\frac{2v{\rm d}t}{\beta_{kj}}\left(r_{0}(t)+\frac{vt-r_{0}\cos\theta}{\beta_{kj}}\right)-v^{2}{\rm d}t^{2}\left(1-\frac{1}{\beta_{kj}^{2}}\right)} =\displaystyle= 2vdtβkj2βkj2r0(t)2(vtr0cosθ)2+O(dt2)\displaystyle\frac{2v{\rm d}t}{\beta_{kj}^{2}}\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-(vt-r_{0}\cos\theta)^{2}}+O({\rm d}t^{2})

Therefore, when dt0{\rm d}t\to 0,

\IEEEeqnarraymulticol3l|(x(t),r0(t)βkj)(x(t+dt),r0(t+dt)βkj)|=πr0(t)2βkj2\displaystyle\IEEEeqnarraymulticol{3}{l}{\left|\mathcal{B}\left(\textup{x}(t),\frac{r_{0}(t)}{\beta_{kj}}\right)\setminus\mathcal{B}\left(\textup{x}(t+{\rm d}t),\frac{r_{0}(t+{\rm d}t)}{\beta_{kj}}\right)\right|=\pi\frac{r_{0}(t)^{2}}{\beta_{kj}^{2}}} (B.8)
[πarccos(vtr0cosθβkjr0(t))(βkj21)vdt2βkj2r0(t)2(vtr0cosθ)2]r0(t)2βkj2\displaystyle-\left[\pi-\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t)}\right)-\frac{\left(\beta_{kj}^{2}-1\right)v{\rm d}t}{2\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-\left(vt-r_{0}\cos\theta\right)^{2}}}\right]\frac{r_{0}(t)^{2}}{\beta_{kj}^{2}}
[arccos(vtr0cosθβkjr0(t))vdtβkj2r0(t)2(vtr0cosθ)2(βkj212+1(vtr0cosθ)2r0(t)2)]\displaystyle-\left[\arccos\left(\frac{vt-r_{0}\cos\theta}{\beta_{kj}r_{0}(t)}\right)-\frac{v{\rm d}t}{\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-\left(vt-r_{0}\cos\theta\right)^{2}}}\left(\frac{\beta_{kj}^{2}-1}{2}+1-\frac{(vt-r_{0}\cos\theta)^{2}}{r_{0}(t)^{2}}\right)\right]
×r0(t+dt)2βkj2+vdtβkj2βkj2r0(t)2(vtr0cosθ)2+O(dt2).\displaystyle\times\frac{r_{0}(t+{\rm d}t)^{2}}{\beta_{kj}^{2}}+\frac{v{\rm d}t}{\beta_{kj}^{2}}\sqrt{\beta_{kj}^{2}r_{0}(t)^{2}-(vt-r_{0}\cos\theta)^{2}}+O({\rm d}t^{2}).

Finally, (LABEL:eq:infinitesimal) can be obtained by substituting (17) in (B.8).

Appendix C: Proof of Theorem 2

We can derive 𝔼[Ltier=k]\mathbb{E}[L\mid\text{tier}=k] using (12), i.e,

𝔼[Ltier=k]=limz0zH(ztier=k)=(a)limz01ddzH(ztier=k),\displaystyle\mathbb{E}[L\mid\text{tier}=k]=\lim_{z\to 0}\frac{z}{H_{\ell}(z\mid\text{tier}=k)}\stackrel{{\scriptstyle\text{(a)}}}{{=}}\lim_{z\to 0}\frac{1}{\frac{{\rm d}}{{\rm d}z}H_{\ell}(z\mid\text{tier}=k)}, (C.1)

where (a) follows from L’Hospital’s rule.

Since |𝒜kj(r0,θ,0,1,βkj)|=πr02βkj2\left|\mathcal{A}_{kj}(r_{0},\theta,0,1,\beta_{kj})\right|=\pi\frac{r_{0}^{2}}{\beta_{kj}^{2}}, we have

\IEEEeqnarraymulticol3lddzH(ztier=k)|z=0\displaystyle\IEEEeqnarraymulticol{3}{l}{\frac{{\rm d}}{{\rm d}z}H_{\ell}(z\mid\text{tier}=k)\Big{|}_{z=0}}
=\displaystyle= 1(tier=k)00π2λkr0(j𝒦λjddz|𝒜kj(r0,θ,z,1,βkj)||z=0)exp{j𝒦λjπβjk2r02}dθdr0\displaystyle\frac{1}{\mathbb{P}(\text{tier}=k)}\int_{0}^{\infty}\int_{0}^{\pi}2\lambda_{k}r_{0}\left(\sum_{j\in\mathcal{K}}\lambda_{j}\frac{{\rm d}}{{\rm d}z}\left|\mathcal{A}_{kj}(r_{0},\theta,z,1,\beta_{kj})\right|\Big{|}_{z=0}\right)\exp\left\{-\sum_{j\in\mathcal{K}}\lambda_{j}\pi\beta_{jk}^{2}r_{0}^{2}\right\}{\rm d}\theta{\rm d}r_{0}
=\displaystyle= 1(tier=k)02λkr0(j𝒦λj0πddz|𝒜kj(r0,θ,z,1,βkj)||z=0dθ)exp{j𝒦λjπβjk2r02}dr0.\displaystyle\frac{1}{\mathbb{P}(\text{tier}=k)}\int_{0}^{\infty}2\lambda_{k}r_{0}\left(\sum_{j\in\mathcal{K}}\lambda_{j}\int_{0}^{\pi}\frac{{\rm d}}{{\rm d}z}\left|\mathcal{A}_{kj}(r_{0},\theta,z,1,\beta_{kj})\right|\Big{|}_{z=0}{\rm d}\theta\right)\exp\left\{-\sum_{j\in\mathcal{K}}\lambda_{j}\pi\beta_{jk}^{2}r_{0}^{2}\right\}{\rm d}r_{0}.

(Note that in the above equations we have used βjk=1βkj\beta_{jk}=\frac{1}{\beta_{kj}}.)

By setting z=0z=0 in (28), we get

\IEEEeqnarraymulticol3lddz|𝒜kj(r0,θ,z,1,βkj)||z=0=\displaystyle\IEEEeqnarraymulticol{3}{l}{\frac{{\rm d}}{{\rm d}z}|\mathcal{A}_{kj}(r_{0},\theta,z,1,\beta_{kj})|\Big{|}_{z=0}=}
{2r0βkj2[βkj2cos2θcosθarccos(cosθβkj)],if (βkj1) or (βkj<1 and arccos(βkj)<θ<πarccos(βkj))0,if (βkj<1 and θarccos(βkj))2πr0cosθβkj2if (βkj<1 and πarccos(βkj)θ).\displaystyle\begin{cases}\frac{2r_{0}}{\beta_{kj}^{2}}\left[\sqrt{\beta_{kj}^{2}-\cos^{2}\theta}-\cos\theta\arccos\left(\frac{\cos\theta}{\beta_{kj}}\right)\right],\\ \qquad\qquad\qquad\qquad\quad\quad\text{if }\left(\beta_{kj}\geq 1\right)\text{ or }\left(\beta_{kj}<1\text{ and }\arccos(\beta_{kj})<\theta<\pi-\arccos(\beta_{kj})\right)\\ 0,\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\quad\qquad\quad\!\text{if }\left(\beta_{kj}<1\text{ and }\theta\leq\arccos(\beta_{kj})\right)\\ -\frac{2\pi r_{0}\cos\theta}{\beta_{kj}^{2}}\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\text{if }\left(\beta_{kj}<1\text{ and }\pi-\arccos(\beta_{kj})\leq\theta\right)\end{cases}.

Therefore,

0πddz|𝒜kj(r0,θ,z,1,βkj)||z=0dθ=2r0(βkj).\displaystyle\int_{0}^{\pi}\frac{{\rm d}}{{\rm d}z}\left|\mathcal{A}_{kj}(r_{0},\theta,z,1,\beta_{kj})\right|\Big{|}_{z=0}{\rm d}\theta=2r_{0}\mathcal{I}\left(\beta_{kj}\right). (C.3)

Substituting (C.3) in (LABEL:eq:C2), we obtain

\IEEEeqnarraymulticol3lddzH(ztier=k)|z=0\displaystyle\IEEEeqnarraymulticol{3}{l}{\frac{{\rm d}}{{\rm d}z}H_{\ell}(z\mid\text{tier}=k)\Big{|}_{z=0}} (C.4)
=\displaystyle= j𝒦λj(βkj)(tier=k)04λkr02exp{j𝒦λjπβjk2r02}dr0\displaystyle\frac{\sum_{j\in\mathcal{K}}\lambda_{j}\mathcal{I}\left(\beta_{kj}\right)}{\mathbb{P}(\text{tier}=k)}\int_{0}^{\infty}4\lambda_{k}r_{0}^{2}\exp\left\{-\sum_{j\in\mathcal{K}}\lambda_{j}\pi\beta_{jk}^{2}r_{0}^{2}\right\}{\rm d}r_{0}
=(a)\displaystyle\stackrel{{\scriptstyle\text{(a)}}}{{=}} j𝒦λj(βkj)π(j𝒦λjβjk2)1/2,\displaystyle\frac{\sum_{j\in\mathcal{K}}\lambda_{j}\mathcal{I}\left(\beta_{kj}\right)}{\pi\left(\sum_{j\in\mathcal{K}}\lambda_{j}\beta_{jk}^{2}\right)^{1/2}},

where (a) follows from change of variable j𝒦λjπβjk2r02=t\sum_{j\in\mathcal{K}}\lambda_{j}\pi\beta_{jk}^{2}r_{0}^{2}=t. Finally, Theorem 2 is derived by substituting (C.4) in (C.1).

References

  • [1] H. Tabassum, M. Salehi, and E. Hossain, “Fundamentals of mobility-aware performance characterization of cellular networks: A tutorial,” IEEE Communications Surveys & Tutorials, vol. 21, no. 3, pp. 2288–2308, thirdquarter 2019.
  • [2] A. L. E. Corral-Ruiz, F. A. Cruz-Pérez, and G. Hernandez-Valdez, “Channel holding time in mobile cellular networks with heavy-tailed distributed cell dwell time,” in 2011 IEEE Wireless Communications and Networking Conference, March 2011, pp. 1242–1247.
  • [3] E. Hossain, L. Le, and D. Niyato, Radio Resource Management in Multi-Tier Cellular Wireless Networks.   Wiley, 2013.
  • [4] X. Lin, R. K. Ganti, P. J. Fleming, and J. G. Andrews, “Towards understanding the fundamentals of mobility in cellular networks,” IEEE Transactions on Wireless Communications, vol. 12, no. 4, pp. 1686–1698, April 2013.
  • [5] S. Shin, U. Lee, F. Dressler, and H. Yoon, “Analysis of cell sojourn time in heterogeneous networks with small cells,” IEEE Communications Letters, vol. 20, no. 4, pp. 788–791, April 2016.
  • [6] X. Xu, Z. Sun, X. Dai, T. Svensson, and X. Tao, “Modeling and analyzing the cross-tier handover in heterogeneous networks,” IEEE Transactions on Wireless Communications, vol. 16, no. 12, pp. 7859–7869, Dec 2017.
  • [7] Y. Hong, X. Xu, M. Tao, J. Li, and T. Svensson, “Cross-tier handover analyses in small cell networks: A stochastic geometry approach,” in 2015 IEEE International Conference on Communications (ICC), June 2015, pp. 3429–3434.
  • [8] C. Lee and Z. Syu, “Handover analysis of macro-assisted small cell networks,” in 2014 IEEE International Conference on Internet of Things (iThings), and IEEE Green Computing and Communications (GreenCom) and IEEE Cyber, Physical and Social Computing (CPSCom), Sep. 2014, pp. 604–609.
  • [9] W. Bao and B. Liang, “Stochastic geometric analysis of user mobility in heterogeneous wireless networks,” IEEE Journal on Selected Areas in Communications, vol. 33, no. 10, pp. 2212–2225, 2015.
  • [10] S. Sadr and R. S. Adve, “Handoff rate and coverage analysis in multi-tier heterogeneous networks,” IEEE Transactions on Wireless Communications, vol. 14, no. 5, pp. 2626–2638, May 2015.
  • [11] S. Hsueh and K. Liu, “An equivalent analysis for handoff probability in heterogeneous cellular networks,” IEEE Communications Letters, vol. 21, no. 6, pp. 1405–1408, June 2017.
  • [12] H. Fu, P. Lin, and Y. Lin, “Reducing signaling overhead for femtocell/macrocell networks,” IEEE Transactions on Mobile Computing, vol. 12, no. 8, pp. 1587–1597, Aug 2013.
  • [13] R. Arshad, H. Elsawy, S. Sorour, T. Y. Al-Naffouri, and M. Alouini, “Handover management in 5g and beyond: A topology aware skipping approach,” IEEE Access, vol. 4, pp. 9073–9081, 2016.
  • [14] Y. Fang, I. Chlamtac, and Yi-Bing Lin, “Call performance for a pcs network,” IEEE Journal on Selected Areas in Communications, vol. 15, no. 8, pp. 1568–1581, Oct 1997.
  • [15] F. Khan and D. Zeghlache, “Effect of cell residence time distribution on the performance of cellular mobile networks,” in 1997 IEEE 47th Vehicular Technology Conference. Technology in Motion, vol. 2, May 1997, pp. 949–953 vol.2.
  • [16] S. Thajchayapong and J. M. Peha, “Mobility patterns in microcellular wireless networks,” IEEE Transactions on Mobile Computing, vol. 5, no. 1, pp. 52–63, Jan 2006.
  • [17] M. Haenggi, Stochastic Geometry for Wireless Networks.   Cambridge University Press, 2012.
  • [18] A. Merwaday, I. Güvenç, W. Saad, A. Mehbodniya, and F. Adachi, “Sojourn time-based velocity estimation in small cell poisson networks,” IEEE Communications Letters, vol. 20, no. 2, pp. 340–343, Feb 2016.
  • [19] J. Moller, Lectures on Random Voronoi Tessellations.   Springer Science & Business Media, 2012, vol. 87.
  • [20] S. Boyd and L. Vandenberghe, Convex Optimization.   Cambridge university press, 2004.
  • [21] H.-S. Jo, Y. J. Sang, P. Xia, and J. G. Andrews, “Heterogeneous cellular networks with flexible cell association: A comprehensive downlink sinr analysis,” IEEE Transactions on Wireless Communications, vol. 11, no. 10, pp. 3484–3495, 2012.
  • [22] S. N. Chiu, D. Stoyan, W. S. Kendall, and J. Mecke, Stochastic Geometry and Its Applications.   John Wiley & Sons, 2013.
  • [23] L. Heinrich, “Contact and chord length distribution of a stationary voronoi tessellation,” Advances in Applied Probability, vol. 30, no. 3, pp. 603–618, 1998.