16
1 Coverage and Rate Analysis of Downlink Cellular Vehicle-to-Everything (C-V2X) Communication Vishnu Vardhan Chetlur, Student Member, IEEE, and Harpreet S. Dhillon, Senior Member, IEEE Abstract—In this paper, we present the downlink coverage and rate analysis of a cellular vehicle-to-everything (C-V2X) com- munication network where the locations of vehicular nodes and road side units (RSUs) are modeled as Cox processes driven by a Poisson line process (PLP) and the locations of cellular macro base stations (MBSs) are modeled as a 2D Poisson point process (PPP). Assuming a fixed selection bias and maximum average received power based association, we compute the probability with which a typical receiver, an arbitrarily chosen receiving node, connects to a vehicular node or an RSU and a cellular MBS. For this setup, we derive the signal-to-interference ratio (SIR)- based coverage probability of the typical receiver. One of the key challenges in the computation of coverage probability stems from the inclusion of shadowing effects. As the standard procedure of interpreting the shadowing effects as random displacement of the location of nodes is not directly applicable to the Cox process, we propose an approximation of the spatial model inspired by the asymptotic behavior of the Cox process. Using this asymptotic characterization, we derive the coverage probability in terms of the Laplace transform of interference power distribution. Further, we compute the downlink rate coverage of the typical receiver by characterizing the load on the serving vehicular nodes or RSUs and serving MBSs. We also provide several key design insights by studying the trends in the coverage probability and rate coverage as a function of network parameters. We observe that the improvement in rate coverage obtained by increasing the density of MBSs can be equivalently achieved by tuning the selection bias appropriately without the need to deploy additional MBSs. Index Terms—Stochastic geometry, Cox process, Poisson line process, coverage probability, C-V2X, rate coverage. I. I NTRODUCTION Vehicular communication networks are essential to the development of intelligent transportation systems (ITS) and improving road safety [2]–[4]. As the in-vehicle sensors can assess only their immediate environment, some information still needs to be communicated from external sources to the vehicle to assist the driver in making critical decisions in advance [5]. Vehicle-to-vehicle (V2V) communication enables the vehicular nodes to share information with each other with- out network assistance. Vehicle-to-infrastructure (V2I) and vehicle-to-network (V2N) communications can support a wide range of applications from basic safety messages and info- tainment to autonomous driving. The underlying technology that enables all these services through long term evolution The authors are with Wireless@VT, Department of ECE, Virginia Tech, Blacksburg, VA (email: {vishnucr, hdhillon}@vt.edu). The support of the US NSF (Grant IIS-1633363) is gratefully acknowledged. This work will be presented in part at IEEE Globecom 2019 [1]. Manuscript last updated: November 29, 2019. (LTE) communication is referred to as cellular vehicle-to- everything (C-V2X) and has been standardized by the third generation partnership project (3GPP) as part of Release 14 [6]. The key technical characteristics of C-V2X include low latency, network independence, and support for high speed vehicular use. Some preliminary system-level simulations are considered in [6] to study the performance of this network under different scenarios. Since simulation-based design ap- proaches are often time consuming and may not be scalable for large number of simulation parameters, it is important to develop complementary analytical approaches to gain design insights and benchmark simulators. For this purpose, tools from stochastic geometry are of particular interest, where the idea is to endow appropriate distributions to the locations of different network entities and then use properties of these distributions to characterize network performance. From the perspective of vehicular networks, the spatial model that is gaining popularity is the so-called Cox process, where the spatial layout of roads is modeled by a Poisson line process (PLP) and the locations of vehicular nodes and road side units (RSUs) on each line (road) are modeled by a 1D Poisson point process (PPP) [7], [8]. While this spatial model has been employed in a few works in the literature, the effect of shadowing has not been considered. In particular, as the signals are often blocked by buildings and other structures in urban areas, shadowing effects have a significant impact on the performance of vehicular networks. Also, the characterization of key performance metrics such as rate coverage, which requires the computation of load on the cellular macro base stations (MBSs), vehicular nodes, and RSUs, is still an open problem. In this paper, we close this knowledge gap and present the coverage and rate analysis for a C-V2X network in the presence of shadowing. A. Related Work Most of the earlier works in the stochastic geometry liter- ature pertaining to vehicular communications were motivated by vehicular ad hoc networks (VANETs) based on dedicated short range communication (DSRC) standard [9]–[15]. Also, the spatial models considered in these works were limited to a single road or an intersection of two roads. That said, there have been some works that consider more sophisticated models, primarily the Cox process, where the spatial layout of roads were modeled by a PLP and the locations of nodes on each road were modeled as a 1D PPP [7], [8], [16]–[22]. The idea of using PLP to model road systems was first proposed in [8] to study the handover rate of vehicles moving across

Coverage and Rate Analysis of Downlink Cellular Vehicle-to

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

1

Coverage and Rate Analysis of Downlink CellularVehicle-to-Everything (C-V2X) Communication

Vishnu Vardhan Chetlur, Student Member, IEEE, and Harpreet S. Dhillon, Senior Member, IEEE

Abstract—In this paper, we present the downlink coverage andrate analysis of a cellular vehicle-to-everything (C-V2X) com-munication network where the locations of vehicular nodes androad side units (RSUs) are modeled as Cox processes driven bya Poisson line process (PLP) and the locations of cellular macrobase stations (MBSs) are modeled as a 2D Poisson point process(PPP). Assuming a fixed selection bias and maximum averagereceived power based association, we compute the probabilitywith which a typical receiver, an arbitrarily chosen receiving node,connects to a vehicular node or an RSU and a cellular MBS.For this setup, we derive the signal-to-interference ratio (SIR)-based coverage probability of the typical receiver. One of the keychallenges in the computation of coverage probability stems fromthe inclusion of shadowing effects. As the standard procedure ofinterpreting the shadowing effects as random displacement of thelocation of nodes is not directly applicable to the Cox process, wepropose an approximation of the spatial model inspired by theasymptotic behavior of the Cox process. Using this asymptoticcharacterization, we derive the coverage probability in termsof the Laplace transform of interference power distribution.Further, we compute the downlink rate coverage of the typicalreceiver by characterizing the load on the serving vehicular nodesor RSUs and serving MBSs. We also provide several key designinsights by studying the trends in the coverage probability andrate coverage as a function of network parameters. We observethat the improvement in rate coverage obtained by increasingthe density of MBSs can be equivalently achieved by tuning theselection bias appropriately without the need to deploy additionalMBSs.

Index Terms—Stochastic geometry, Cox process, Poisson lineprocess, coverage probability, C-V2X, rate coverage.

I. INTRODUCTION

Vehicular communication networks are essential to thedevelopment of intelligent transportation systems (ITS) andimproving road safety [2]–[4]. As the in-vehicle sensors canassess only their immediate environment, some informationstill needs to be communicated from external sources to thevehicle to assist the driver in making critical decisions inadvance [5]. Vehicle-to-vehicle (V2V) communication enablesthe vehicular nodes to share information with each other with-out network assistance. Vehicle-to-infrastructure (V2I) andvehicle-to-network (V2N) communications can support a widerange of applications from basic safety messages and info-tainment to autonomous driving. The underlying technologythat enables all these services through long term evolution

The authors are with Wireless@VT, Department of ECE, Virginia Tech,Blacksburg, VA (email: {vishnucr, hdhillon}@vt.edu). The support of theUS NSF (Grant IIS-1633363) is gratefully acknowledged. This work will bepresented in part at IEEE Globecom 2019 [1]. Manuscript last updated:November 29, 2019.

(LTE) communication is referred to as cellular vehicle-to-everything (C-V2X) and has been standardized by the thirdgeneration partnership project (3GPP) as part of Release 14[6]. The key technical characteristics of C-V2X include lowlatency, network independence, and support for high speedvehicular use. Some preliminary system-level simulations areconsidered in [6] to study the performance of this networkunder different scenarios. Since simulation-based design ap-proaches are often time consuming and may not be scalablefor large number of simulation parameters, it is important todevelop complementary analytical approaches to gain designinsights and benchmark simulators. For this purpose, toolsfrom stochastic geometry are of particular interest, where theidea is to endow appropriate distributions to the locations ofdifferent network entities and then use properties of thesedistributions to characterize network performance. From theperspective of vehicular networks, the spatial model that isgaining popularity is the so-called Cox process, where thespatial layout of roads is modeled by a Poisson line process(PLP) and the locations of vehicular nodes and road side units(RSUs) on each line (road) are modeled by a 1D Poissonpoint process (PPP) [7], [8]. While this spatial model hasbeen employed in a few works in the literature, the effectof shadowing has not been considered. In particular, as thesignals are often blocked by buildings and other structures inurban areas, shadowing effects have a significant impact on theperformance of vehicular networks. Also, the characterizationof key performance metrics such as rate coverage, whichrequires the computation of load on the cellular macro basestations (MBSs), vehicular nodes, and RSUs, is still an openproblem. In this paper, we close this knowledge gap andpresent the coverage and rate analysis for a C-V2X networkin the presence of shadowing.

A. Related Work

Most of the earlier works in the stochastic geometry liter-ature pertaining to vehicular communications were motivatedby vehicular ad hoc networks (VANETs) based on dedicatedshort range communication (DSRC) standard [9]–[15]. Also,the spatial models considered in these works were limitedto a single road or an intersection of two roads. That said,there have been some works that consider more sophisticatedmodels, primarily the Cox process, where the spatial layout ofroads were modeled by a PLP and the locations of nodes oneach road were modeled as a 1D PPP [7], [8], [16]–[22]. Theidea of using PLP to model road systems was first proposedin [8] to study the handover rate of vehicles moving across

Page 2: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

2

cells. This model was further explored in [23], where thedistribution of the distances between the nodes located ona PLP was studied. In [16], the authors have presented thesuccess probability of a typical link in a VANET modeled asa Cox process driven by a PLP. The routing performance ofa VANET for a linear multi-hop relay was studied in [17].In [18], the authors have presented the downlink coverageprobability of a typical receiver in an urban setup where thelocations of base stations operating at millimeter wave fre-quencies were modeled as Cox process driven by a ManhattanPoisson line process (MPLP). The uplink coverage probabilityfor a setup where the typical receiver is chosen from a 2D PPPand the locations of transmitting nodes form a Cox processwas presented in [19]. In [7], [24], the authors have derivedthe downlink coverage probability of a typical receiver fornearest neighbor connectivity in a vehicular network where thetransmitter and receiver nodes are modeled by Cox processesdriven by the same PLP. While these are key steps towardsunderstanding the behavior of vehicular networks, the analysispresented in these works is often limited to V2V and V2Icommunications as they do not include cellular base stationsin their setup.

Recently, there have been a few works in which cellular-assisted vehicular networks were considered [25]–[27]. Acomprehensive coverage analysis for two different types ofusers (planar and vehicular users) was provided in [25], wherethe locations of transmitting vehicular nodes were modeled bythe Cox process and the locations of cellular base stations weremodeled by 2D PPP. In [27], the authors have attempted toderive the uplink coverage probability for C-V2X communica-tion. However, the effects of shadowing were ignored in theseworks because of the technical complications arising fromthe doubly stochastic nature of this Cox process. Moreover,the rate coverage analysis of a C-V2X network has not beeninvestigated in the literature due to the challenges involvedin the characterization of the load on RSUs and MBSs usingthis spatial model. So, this paper makes two key contributions.First, we present the coverage analysis of a C-V2X networkincluding the shadowing effects in our model. Second, wepresent the rate coverage analysis for a C-V2X network. Moredetails of our contribution are provided next.

B. Contributions

In order to distinguish our contributions from prior art,we will first provide a few details about the system modelconsidered in this paper. We model the locations of vehicularnodes and RSUs by Cox processes driven by a PLP andthe locations of cellular MBSs by a homogeneous 2D PPP.We consider Nakagami-m fading to model the effects ofsmall-scale fading. We consider standard power-law path-lossmodel and log-normal shadowing to model large-scale fadingeffects. We further introduce selection bias to balance the loadbetween MBSs, RSUs and vehicular nodes across the network.Assuming that a typical receiver, which is an arbitrarily chosenreceiver node, connects to the node that yields the highestaverage biased received power among the contending nodes,we derive the signal-to-interference ratio (SIR)-based coverage

probability and rate coverage. We also offer useful designinsights by studying the trends in the performance as a functionof key network parameters. More technical details about theanalysis are provided next.

Asymptotic behavior of the Cox process. Based on theresults provided in [19], we first present the expression forthe void probability of a Cox process driven by a PLP for anyplanar Borel set. Using Choquet’s theorem, we then rigorouslyprove that the Cox process asymptotically converges to ahomogeneous 2D PPP with the same average density of pointsas the line density approaches infinity and the density of pointson each line tends to zero [28, Chapter 6].

Coverage probability. The key technical challenge in thecharacterization of SIR-based coverage probability stems fromthe inclusion of shadowing effects. Traditionally, in wirelessnetworks where the locations of nodes were modeled as aPPP, the effect of shadowing could be conveniently interpretedas a random displacement of the location of the nodes inthe computation of the desired signal power and interferencepower at the receiver [29]. However, this method is notapplicable to the Cox process as the random displacementof each point on a line would disrupt the collinearity ofthe points, thereby making it difficult to exactly characterizethe SIR at the typical receiver. Therefore, inspired by theasymptotic behavior of the Cox process, we propose a tractableand accurate approximation of the spatial model. Using thisasymptotic characterization, we obtain the expression for cov-erage probability in terms of the Laplace transform of theinterference power distribution. Given its nature, the proposedmodel could enable similar analyses that may not otherwise bepossible due to the doubly stochastic nature of a Cox processdriven by a PLP.

Rate Coverage. In order to determine the rate coverageof the typical receiver, one of the important components isthe distribution of load on the serving node of the typicalreceiver. As the load on the serving node is proportional tothe size of its coverage region, we characterize the length ofthe coverage region of the serving vehicular node or RSU.We also determine the mean load on the serving MBS. Usingthese results, along with the distribution of SIR, we completelycharacterize the rate coverage of the typical receiver.

Design Insights. Using the analytical results, we study theimpact of node densities and selection biases on the coverageprobability and rate coverage of the typical receiver. Asexpected, we observe that the rate coverage can be improvedby increasing the node density of MBSs. However, sincethe deployment of additional MBSs is costly (and may notalways be possible because of challenges in site acquisitions),our results concretely demonstrate that we can alternatelyachieve similar performance gains by adjusting bias factorsappropriately. We also observe that an increase in the densityof RSUs has a contrasting effect on coverage probability andrate coverage. While a denser deployment of RSUs decreasescoverage probability due to an increase in interference power,it also reduces the load on the RSUs thereby improving the ratecoverage. Hence, one must consider such trade-offs betweenthese two metrics in the design of the network.

Page 3: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

3

II. MATHEMATICAL PRELIMINARY: ASYMPTOTICBEHAVIOR OF COX PROCESS

In this section, we will explore a key property of the Coxprocess driven by a PLP which will be useful in developinga tractable way to characterize the interference distribution inour analysis later. Specifically, we will analyze the asymptoticbehavior of a Cox process driven by a PLP for extreme valuesof line and point densities. As basic knowledge of PLP andits properties would be useful in understanding this analysis,we will first give a brief introduction to the topic. A detailedaccount on the theory of line processes can be found in [28].

Definition 1. (Poisson line process). An undirected line L

in R2 can be uniquely parameterized by its perpendiculardistance from the origin ⇢ and the angle ✓ subtended bythe perpendicular dropped onto the line w.r.t. the positivex-axis. So, each line in R2 can be uniquely mapped toa point with coordinates (⇢, ✓) on the cylindrical surfaceC ⌘ R+⇥[0, 2⇡), which is termed as the representation space.A random collection of lines that is generated by a PPP in Cis called a PLP.

Definition 2. (Line density). The line density of a line processis defined as the mean line length per unit area.

Definition 3. (Motion-invariance). A line process is said tobe motion-invariant if the distribution of lines is invariant totranslation and rotation.

We will now discuss the construction of the Cox process.First, we consider a motion-invariant PLP ` with line densityµ`. We denote the corresponding 2D PPP in C by C withdensity �`. As shown in [28], the relation between �` and µ`

is given by �` = µ`

⇡. We populate points on the lines of `

such that they form a 1D PPP with density �p on each line.Thus, we obtain a Cox process a driven by the PLP `.Owing to the stationarity of the underlying PLP `, it followsthat the Cox process a is also stationary for any arbitraryvalues of µ` and �p. We begin our analysis by deriving thevoid probability of the Cox process in the following lemma.

Lemma 1. The void probability of the Cox process a is givenby

P (Np(A) = 0)

= exp

��`

Z 2⇡

0

Z

R+

⇥1� exp

���p⌫1

�L(⇢,✓) \A

��⇤d⇢d✓

�,

(1)

where Np(·) denotes the number of points, A ⇢ R2 is a planarBorel set, and ⌫1(·) is the 1D Lebesgue measure.

Proof: See Appendix A.Using Choquet’s theorem [28], we will now show that the

Cox process a asymptotically converges to a homogeneous2D PPP in the following theorem.

Theorem 1. As the line density approaches infinity (�` ! 1)and the density of points on each line tends to zero (�p ! 0)while the overall density of points (average number of pointsper unit area) remains unchanged, the Cox process a con-

verges to that of a 2D PPP with the same average node density⇡�`�p.

Proof: See Appendix B.

III. SYSTEM MODEL

A. Spatial Modeling of Wireless NodesWe consider a vehicular network consisting of vehicular

nodes, RSUs, and cellular MBSs, as illustrated in Fig. 1. Sincethe locations of vehicles and RSUs are restricted to roadways,we first model the irregular spatial layout of roads by a PLP�l with line density µl. We denote the corresponding 2D PPPin the representation space C by �C with density �l. We modelthe locations of transmitting and receiving vehicular nodes oneach line Lj 2 �l by independent and homogeneous 1D PPPs⌅(V )Lj

and ⌦(R)Lj

with densities �v and �r, respectively. Further,we model the locations of RSUs on each line Lj by a 1DPPP ⌅(U)

Ljwith density �u. Thus, the locations of receiving

vehicular nodes, transmitting vehicular nodes and RSUs formCox processes �r ⌘ {⌦(R)

Lj}Lj2�l , �v ⌘ {⌅(V )

Lj}Lj2�l and

�u ⌘ {⌅(U)Lj

}Lj2�l , respectively. Note that �r, �v , and �u arestationary [19]. The locations of cellular MBSs are modeledby a 2D PPP �c with density �c.

Our goal is to characterize the SIR-based coverage prob-ability and rate coverage of a typical receiver, which is anarbitrarily chosen receiving vehicular node from the pointprocess �r. For analytical simplicity, we translate the origino ⌘ (0, 0) to the location of the typical receiver. Since thetypical receiver must be located on a line, we now havea line L0 passing through the origin. Thus, under Palmprobability of the receiver point process, the resulting lineprocess is �l0 ⌘ �l [ L0. This result simply follows fromthe application of Slivnyak’s theorem to the line process�l or equivalently to the corresponding 2D PPP �C in therepresentation space C. Thus, the point process of receivingvehicular nodes �r0 ⌘ �r[⌦(R)

L0[{o} is the superposition of

�r, an independent 1D PPP with density �r on line L0 andan atom at the origin. Consequently, under Palm probabilityof �r0 , the point process of transmitting vehicular nodes�v0 ⌘ �v [ ⌅(V )

L0is the superposition of the point process

�v and an independent 1D PPP with density �v on line L0

[7], [19]. Similarly, under Palm probability of �r0 , the pointprocess of RSUs �u0 ⌘ �u[ L0 is also the superposition of�u and an independent 1D PPP L0 on L0 with density �u.The line L0 will henceforth be referred to as the typical line.

B. Propagation ModelWe assume that all the MBSs have the same transmit powers

Pc. Further, we assume that all the vehicular nodes and RSUshave the same transmit power Pu. Since the vehicular nodesand RSUs predominantly communicate with the nodes on theroad on which they are located, we assume that vehicularnodes and RSUs employ transmit beamforming to maximizethe signal power at the receiver nodes. Based on the directionalantennas considered in [30], we assume a beam pattern inwhich the main lobes are aligned along the road on which thesenodes are located, as shown in Fig. 2. We assume a sectorized

Page 4: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

4

!0

Typical Receiver

Serving node

Transmitting Vehicular nodes

RSUs

Cellular MBS

!0

Coverage region of aVehicular node

Coverage region of aRSU

Fig. 1. Illustration of the system model.

Main lobes

Side lobes

0.2

0.4

0.6

0.8

1.0

!

"!

30°

60°

90°

120°

150°

180°

210°

270°300°

330°

240°

Fig. 2. Illustration of the antenna radiation pattern for avehicular node or an RSU.

Main lobe

Side lobe

0.2

0.4

0.6

0.8

1.0 !

"!0°

30°

60°

90°

120°

150°

180°

210°

270°300°

330°

240°

Fig. 3. Illustration of the antenna radiation pattern for acellular MBS.

antenna pattern with main lobe and side lobe gains denotedby Gu and gu, respectively. We assume that the cellularMBSs also use transmit beamforming with main lobe and sidelobe gains denoted by Gc and gc, respectively, as shown inFig. 3. Unlike vehicular nodes and RSUs with fixed beampatterns, MBSs scan the space for the best beam alignmentduring the association process. Therefore, the main lobe ofthe serving MBS is directed towards the typical receiver, i.e.,the beamforming gain from the serving MBS is always Gc.However, the main lobes of other MBSs may not necessarilybe in the direction of the typical receiver. So, we assumethat the main lobe of an interfering MBS is directed towardsthe typical receiver with a probability qc, which depends onthe main lobe beamwidth. Thus, the beamforming gain froman interfering MBS is Gc with a probability qc and gc withprobability (1�qc). As the typical receiver may not have priorknowledge of the direction of arrival of the desired signal, weassume an omni-directional antenna at the typical receiver.

As the vehicular nodes and RSUs have identical transmitpowers and beamforming patterns, we consider them as wire-less nodes of a single tier in the coverage and rate analysispresented in this paper. Therefore, the vehicular nodes andRSUs will henceforth be collectively referred to as tier 2 nodesand the cellular MBSs will be referred to as tier 1 nodes.Thus, under Palm probability of the receiver point process,the locations of the tier 2 nodes form a Cox process �2

driven by the line process �l0 , where the locations of nodes oneach line Lj form a homogeneous 1D PPP ⌅Lj with density�2 = �u + �v . In order to be consistent with the notation, wedenote the set of locations of tier 1 nodes by �1 with density�1 = �c. We also update the notation of transmit powers,main lobe and side lobe gains of tier 1 and tier 2 nodes to{P1, G1, g1} and {P2, G2, g2}, respectively.

As the channel fading characteristics could vary signifi-cantly from rural to urban areas, we choose Nakagami-mfading to mimic a wide range of fading environments. Wedenote the fading parameter for the link between the typicalreceiver and tier 1 nodes by m1 and the channel fading gainsby H1. The tier 2 nodes on the typical line have a higherlikelihood of a line-of-sight (LOS) link to the typical receiverthan those nodes located on the other lines. So, in orderto distinguish the severity of fading, we denote the fadingparameters for the links to tier 2 nodes on the typical lineand the tier 2 nodes on other lines by m20 and m21, andthe corresponding fading gains by H20 and H21, respectively.The fading gains Hi, where i 2 {1, 20, 21} follow a Gammadistribution with probability density function (PDF)

fHi(h) =m

mii

hmi�1

�(mi)exp(�mih), m 2 {m1,m20,m21}.

(2)

In the interest of analytical tractability, we restrict the valuesof m1, m20 and m21 to integers.

We will now discuss the modeling of large scale fadingeffects. We assume a standard power-law path loss model withexponent ↵ > 2 for all the wireless links. We model the shad-owing effects for the links from the typical receiver to the tier 1nodes, tier 2 nodes on the typical line, and tier 2 nodes on otherlines by the random variables X1, X20, and X21, respectively.We assume that Xi, where i 2 {1, 20, 21}, follows a log-normal distribution such that 10 log10 Xi ⇠ N (!i,�

2i) with

mean !i and standard deviation �i in dB. Thus, the receivedsignal power at the typical receiver from a node located at xis

Pr(x) =

8><

>:

P1GxH1X1kxk�↵, x 2 �1,

P2G2H20X20kxk�↵, x 2 ⌅L0 ,

P2g2H21X21kxk�↵, x 2 �2 \ ⌅L0 ,

(3)

where Gx 2 {G1, g1} denotes the beamforming gain fromthe tier 1 node located at x, and k · k denotes the Euclideannorm. Note that we have assumed that the beamforming gainfrom all the tier 2 nodes that are not located on the typicalline would be g2. However, if the typical receiver is locatedexactly at the intersection of two roads, then there would betwo lines which contain nodes with a beamforming gain of G2.

Page 5: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

5

This case can be handled by slightly modifying the equationfor received power given in (3) and then following the sameanalytical procedure. However, since it is a zero measure eventin our model, we do not have to consider it explicitly in thispaper.

C. Association Policy

We assume that the typical receiver connects to the nodewhich yields the highest average biased received power. Asmentioned earlier, tier 1 nodes scan the space for the bestbeam alignment during the association process. Therefore, thecandidate serving node from tier 1 is the node located at x⇤1 =arg maxx2�1

P1G1X1kxk�↵. A well-known procedure in theliterature to handle the effect of shadowing is to express it asa random displacement of the location of the node w.r.t. thetypical receiver [29]. Thus, the average received power at thetypical receiver from a tier 1 node can be alternately writtenas Pr(y) = P1G1kyk�↵, where y = X� 1

↵1 x. By displacement

theorem [31], for a homogeneous PPP ⇢ R2 with density�, if each point xk 2 is transformed to y = X� 1

↵k

xk, thenthe transformed points also form a homogeneous PPP withdensity E[X� 2

↵ ]�. Thus, the candidate serving node from tier1 is the node located at y⇤1 = arg max

y2�(e)1

P1G1kyk�↵,

where �(e)1 is the equivalent 2D PPP with density �

(e)1 =

E[X� 2↵

1 ]�1 = exp⇣

!1 ln 105↵ + 1

2

��1 ln 10

5↵

�2⌘�1. Therefore, the

candidate serving node from tier 1 is the closest node to thetypical receiver from the point process �(e)

1 .Among the tier 2 nodes, the information obtained at the

typical receiver from the nodes on the typical line is morerelevant than the information from the nodes on the otherlines. So, we consider that the typical receiver would connectto only those tier 2 nodes that lie on the typical line. Therefore,the candidate tier 2 serving node is the node located atx⇤2 = arg maxx2⌅L0

P2G2X20kxk�↵. As shown in case oftier 1 nodes, the impact of shadowing can be expressed as arandom displacement of the location of the node at x 2 ⌅L0 .Therefore, the candidate serving node from tier 2 is the nodelocated at y⇤2 = arg max

y2⌅(e)L0

P2G2kyk�↵, where ⌅(e)L0

is the

equivalent 1D PPP with density �(e)2 = E[X� 1

↵20 ]�2. Thus, the

candidate serving node from tier 2 is the closest node to thetypical receiver from the point process ⌅(e)

L0.

Therefore, the typical receiver either connects to its closesttier 1 node from �(e)

1 or the closest tier 2 located on the typicalline ⌅(e)

L0. Further, we introduce selection bias factors B1 and

B2 for the tier 1 and tier 2 nodes, respectively, in order to bal-ance the load between them across the network. So, among thetwo candidate nodes from tier 1 and tier 2, the typical receiverconnects to the node at y⇤ = arg maxy2{y⇤

j } BjPjGjkyk�↵,where j 2 {1, 2}.

For expositional simplicity, we assume that the system isinterference limited and hence, the thermal noise is neglectedin our analysis. The aggregate interference at the typicalreceiver is composed of the interference from the tier 1 nodesI1, interference from the tier 2 nodes on the typical line I20,

and the interference from the tier 2 nodes on the other linesI21. Thus, the SIR measured at the typical receiver is

SIR =Pr(y⇤)

I1 + I20 + I21, (4)

where

I1 =X

x2�1

P1GxX1kxk�↵<Pr(y

⇤)

P1GxH1X1kxk�↵

=X

y2�(e)1

P1Gykyk�↵<Pr(y

⇤)

P1GyH1kyk�↵;Gy =

(G1 with prob. qc,g1 with prob. 1� qc.

(5)

I20 =X

x2⌅L0

P2G2X20kxk�↵<Pr(y

⇤)

P2G2H20X20kxk�↵ =X

y2⌅(e)L0

P2G2kyk�↵<Pr(y

⇤)

P2G2H20kyk�↵,

(6)and

I21 =X

x2�2\⌅L0

P2g2X21kxk�↵

=X

Lj2{�l0\L0}

X

x2⌅Lj

P2g2X21H21kxk�↵. (7)

While the terms I1 and I20 were simplified by the applica-tion of displacement theorem, it is not possible for I21. Thecharacterization of I21 is the key challenge that needs to behandled carefully in the computation of coverage probability.

IV. COVERAGE ANALYSIS

In this section, we will determine the SIR-based coverageprobability of the typical receiver in the network. Recallthat the typical receiver would connect to its closest tier1 node from the 2D PPP �(e)

1 or the closest tier 2 nodefrom the 1D PPP ⌅(e)

L0and we denote these events by E1

and E2, respectively. As a result, the interference measuredat the typical receiver is different in these two cases andhence we handle them separately in our analysis. We firstdetermine the probability of occurrence of these events andthen characterize the desired signal power by computing thecumulative distribution function (CDF) of the serving distanceR conditioned on these events.

A. Association ProbabilityIn this subsection, we derive the probability of the occur-

rence of the events E1 and E2. We denote the distances fromthe typical receiver to its closest node in �(e)

1 and ⌅(e)L0

by R1

and R2, respectively. As is well known in the literature, theCDF and PDF of R1 and R2 are given by

CDF: FR1(r1) = 1� exp⇣��

(e)1 ⇡r

21

⌘, (8)

FR2(r2) = 1� exp⇣��

(e)2 2r2

⌘, (9)

PDF: fR1(r1) = 2⇡�(e)1 r1 exp

⇣��

(e)1 ⇡r

21

⌘, (10)

fR2(r2) = 2�(e)2 exp

⇣��

(e)2 2r2

⌘. (11)

Page 6: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

6

Using these distributions, we will now derive the associationprobabilities of the typical receiver with tier 1 and tier 2 nodesin the following Lemma. As the shadowing terms are alreadyhandled, this derivation follows from the standard procedurebut is given for completeness.

Lemma 2. The probability of occurrence of the events E1 andE2 are

P(E1)

= 1�

vuut (�(e)2 )2

�(e)1 ⇣

� 2↵

21

exp

"(�(e)

2 )2

�(e)1 ⇡⇣

� 2↵

21

#erfc

0

@ �(e)2q

�(e)1 ⇡⇣

� 2↵

21

1

A ,

(12)P(E2)

=

vuut (�(e)2 )2

�(e)1 ⇣

� 2↵

21

exp

"(�(e)

2 )2

�(e)1 ⇡⇣

� 2↵

21

#erfc

0

@ �(e)2q

�(e)1 ⇡⇣

� 2↵

21

1

A ,

(13)

where ⇣21 = P2B2G2P1B1G1

.Proof: See Appendix C.

B. Serving Distance DistributionThe next key step in computing the coverage probability is

to characterize the desired signal power at the typical receiverwhich depends on the serving distance. Therefore, we willnow derive the serving distance distribution conditioned onthe events E1 and E2 in the following Lemmas.

Lemma 3. Conditioned on the event E1, the PDF of theserving distance is given by

fR(r|E1) =2⇡�(e)

1 r exp⇣�⇡�

(e)1 r

2 � 2⇣1↵21�

(e)2 r

P(E1).

Proof: The conditional CDF of R can be obtained as

FR(r|E1) = 1� P(R > r, E1)P(E1)

(a)= 1� 1

P(E1)P R1 > r,R1 <

✓P2B2G2

P1B1G1

◆� 1↵

R2

!

(b)= 1� 1

P(E1)

Z 1

1↵21 r

P⇣r < R1 < ⇣

� 1↵

21 r2|R2

⌘fR2(r2)dr2

= 1� 1

P(E1)

Z 1

1↵21 r

hFR1

⇣⇣� 1

↵21 r2

⌘� FR1(r)

ifR2(r2)dr2,

where (a) follows form the condition for the occurrence ofthe event E1, (b) follows from substituting ⇣21 = P2B2G2

P1B1G1and

conditioning on R2, and the final expression for CDF can beobtained by substituting the expressions for FR1(·) and fR2(·)in the last step and simplifying the resulting integral. The PDFof the serving distance conditioned on E1 can then be obtainedby taking the derivative of FR(r|E1) w.r.t. r.Lemma 4. Conditioned on the event E2, the PDF of theserving distance is given by

fR(r|E2) =2�(e)

2 exp⇣�2�(e)

2 r � ⇡�(e)1 ⇣

� 2↵

21 r2⌘

P (E2),

where ⇣21 = P2B2G2P1B1G1

.

Proof: The proof follows along the same lines as that ofLemma 3 and hence skipped.

C. Approximation of the Cox Process ModelIn order to determine the coverage probability, we need to

compute the conditional Laplace transform of the interferencepower distribution. Recall that the aggregate interference atthe typical receiver can be decomposed into three independentcomponents I1, I20, and I21, caused by tier 1 nodes, tier 2nodes on the typical line and the tier 2 nodes on other lines,respectively. The key technical challenge in the computation ofLaplace transform of interference stems from the inclusion ofshadowing effects. As explained in section III, a well-knownapproach in the literature is to interpret the effect of shadowingas a random displacement of the location of the nodes andthen compute the received power at the typical receiver fromthe nodes whose locations are given by the equivalent pointprocess. So, we can apply this technique to the 2D PPP of tier1 nodes �1 and 1D PPP of tier 2 nodes on the typical line⌅L0 and rewrite the interference powers I1 and I20 in termsof the equivalent points processes �(e)

1 and ⌅(e)L0

, respectively,as given in (5) and (6). However, upon applying this techniqueto the Cox process of tier 2 nodes excluding the nodes on thetypical line �0

2 = �2 \⌅L0 , the collinearity of the locations ofthe nodes is disrupted in the resulting point process �0(e)

2 =

{y : y = X� 1↵

21 x, x 2 �02} due to the random displacement of

each point on a line. Thus, in order to exactly characterize theinterference, it is necessary to find the exact distribution ofpoints in �0(e)

2 , which is quite hard. This is the key motivationbehind proposing a tractable yet accurate approximation forthe interference analysis of this component.

From Theorem 1, we already know that the Cox processasymptotically converges to a 2D PPP. Further, as mentionedearlier, the effect of shadowing interpreted as random displace-ment of points also disturbs the collinear structure of pointson all the lines except the typical line. Motivated by thesetwo facts, we make the following assumption which enablesa tractable interference analysis.Assumption 1. We assume that the spatial distribution ofnodes in �0(e)

2 follows a 2D PPP with density E[X� 2↵

2 ]⇡�l�2.This density follows from the result given in Theorem 1combined with the interpretation of shadowing as randomdisplacement of nodes.

Under Assumption 1, the approximate spatial model for thetier 2 nodes with the inclusion of shadowing effects is �̃(e)

2 =⌅(e)L0

[ �(a)2 , where �(a)

2 is a 2D PPP with density �(a)2 =

E[X� 2↵

2 ]⇡�l�2. In other words, the proposed approximationfor the tier 2 nodes is the superposition of the 1D PPP ⌅(e)

L0

with density �(e)2 on the typical line and the 2D PPP �(a)

2 withdensity �

(a)2 .

Remark 1. The Ripley’s K-function of a Cox process drivenby PLP with line density µl is given by K(r) = 2r

µl+ ⇡r

2.The first term in K(r) corresponds to the points located on thetypical line. The second term, which corresponds to the rest of

Page 7: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

7

the points, is identical to the K-function of a homogeneous 2DPPP. This is further evidence that the proposed approximationis quite reasonable. More numerical evidence will be providedin Section VI.

Now that we have clearly established the approximate spa-tial model for the tier 2 nodes, we proceed with the coverageanalysis by computing the Laplace transform of interferencein the next subsection.

D. Laplace Transform of Interference Power DistributionIn this subsection, we will compute the Laplace transform of

the distribution of interference power measured at the typicalreceiver conditioned on the serving distance R and the eventsE1 and E2. First, let us consider E1. In case of E1, as the typicalreceiver connects to its closest tier 1 node at a distance R, therecan not be any tier 1 node closer than R. Also, the averagereceived power from any tier 2 nodes on the typical line cannot exceed the received power from the serving node. Thismeans that there can not be any tier 2 node on the typical linewhose distance to the typical receiver is lesser than ⇣

1↵21R. We

have similar restrictions on the spatial distribution of nodes incase of E2. Incorporating these conditions, we determine theLaplace transform of interference conditioned on E1 and E2 inthe following Lemmas.Lemma 5. Conditioned on the event E1 and the servingdistance R, the Laplace transform of the interference at thetypical receiver is

LI(s|R, E1)

= exp

"-2⇡qc�

(e)1

Z 1

r

1-✓1 +

sP1G1y�↵

m1

◆�m1!ydy

-2⇡(1� qc)�(e)1

Z 1

r

1-

1 +

sP1g1y�↵

m1

!�m1!ydy

-2�(e)2

Z 1

1↵21 r

1-✓1 +

sP2G2y�↵

m20

◆�m20!dy

-2⇡�(a)2

Z 1

0

1-✓1 +

sP2g2y�↵

m21

◆�m21!ydy

#. (14)

Proof: See Appendix D.

Lemma 6. Conditioned on the event E2 and the servingdistance R, the Laplace transform of the interference at thetypical receiver is

LI(s|R, E2)

= exp

"-2⇡qc�

(e)1

Z 1

⇣� 1

↵21 r

1-✓1 +

sP1G1y�↵

m1

◆�m1!ydy

-2⇡(1� qc)�(e)1

Z 1

⇣� 1

↵21 r

1-

1 +

sP1g1y�↵

m1

!�m1!ydy

-2�(e)2

Z 1

r

1-✓1 +

sP2G2y�↵

m20

◆�m20!dy

-2⇡�(a)2

Z 1

0

1-✓1 +

sP2g2y�↵

m21

◆�m21!ydy

#. (15)

Proof: The proof follows along the same lines as that ofLemma 5 and hence skipped.

E. Coverage Probability

The coverage probability is formally defined as the prob-ability with which the SIR measured at the typical receiverexceeds a predetermined threshold �. Using the conditionaldistributions of serving distance and the conditional Laplacetransform of interference power distribution, we derive thecoverage probability in the following theorem.

Theorem 2. The coverage probability of the typical receiveris

Pc = P(E1)m1�1X

k=0

Z 1

0

(-m1�r↵)k

(P1G1)kk!

@k

@skLI(s|R, E1)

s=m1�r↵

P1G1

⇥ fR(r|E1)dr + P(E2)m20�1X

k=0

Z 1

0

(-m20�r↵)k

(P2G2)kk!

⇥@k

@skLI(s|R, E2)

s=m20�r↵

P2G2

fR(r|E2)dr, (16)

where P(E1) and P(E2) are given in Lemma 2, LI(s|R, E1)and LI(s|R, E2) are given in Lemmas 5 and 6, and fR(r|E1)and fR(r|E2) are given in Lemmas 3 and 4.

Proof: The proof follows along the same lines as that ofTheorem 1 in [7] and is hence skipped.

Note that this is the exact coverage probability for theasymptotic case which corresponds to the dense layout ofroads and sparsely distributed vehicular nodes and RSUs.However, as will be demonstrated by the numerical resultsprovided in Section VI, this result is quite accurate even fornominal values of line and node densities.

V. DOWNLINK RATE COVERAGE

In this section, we will compute the downlink rate coverageof the typical receiver by characterizing the load on the servingtier 1 and tier 2 nodes. This characterization of load is one ofthe key contributions of this paper and is useful in studyingseveral other metrics such as latency and packet receptionrate [6]. Assuming that the bandwidth resource is equallyshared by all the users that are connected to the serving node,the downlink rate achievable at the typical receiver can becomputed using Shannon’s theorem as

R =W

Jlog2 (1 + SIR) bits/sec, (17)

where W is the available bandwidth and J is the loadon the serving node of the typical receiver. Note that SIRand J are in general negatively correlated. However, in theinterest of analytical tractability, we consider these two randomvariables to be independent. This assumption does not affectthe accuracy of our final results. We further assume a full-buffer traffic model for the downlink communications. We willrefer to the serving node of the typical receiver as the taggednode and the corresponding cell as the tagged cell [32], [33].We denote the load on the tagged tier 1 and tier 2 nodes by

Page 8: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

8

J1 and J2, respectively. As we have already characterized theSIR at the typical receiver in the previous section, we willnow focus on the distribution of J1 and J2 in the followingsubsections.

A. Load on the Tagged Tier 2 Node

As the load on a node is proportional to the size of itscell, we need to characterize the size of the tagged tier 2cell. First, we will focus on the cell of a typical tier 2 node,which is an arbitrarily chosen tier 2 node. We assume thatthe typical tier 2 node is located at xtyp on a line L. Asthe tier 2 nodes serve only the users located on their ownlines, the typical cell is a line segment on the line L andis denoted by Ztyp = {x : x 2 L,P2G2B2kx � xtypk�↵ �PyGyBykx�yk�↵

, 8y 2 �(e)1 [⌅(e)

L}, where Py, Gy, and By

correspond to the transmit power, beamforming gain, and biasof the node located at y, respectively. We begin our analysisby computing the distribution of the length of the typical tier2 cell, which is denoted by Ztyp = ⌫1(Ztyp).

Lemma 7. The PDF of the length of a typical tier 2 cell is

fZtyp(z) ⇡Z

z

k+12k z

fZ1,1(z � z0)fZ0(z0)dz0

+

Z k+12k z

k�12k z

fZ1,2(z � z0)fZ0(z0)dz0

+

Z k�12k z

0fZ1,3(z � z0)fZ0(z0)dz0, (18)

where

fZ1,1(z1) = 2�(e)2 exp(-2�(e)

2 z1), 0 < z1 <k � 1

k + 1z0,

fZ1,2(z1) =

✓2�(e)

2 + �(e)1

@�2,2(z0, z1)

@z1

◆exp

-2�(e)

2 z1

-�(e)1 �2,2(z0, z1)

�,k � 1

k + 1z0 < z1 <

k + 1

k � 1z0,

fZ1,3(z1) = (2�(e)2 + 2�(e)

1 ⇡k2z1) exp

-2�(e)

2 z1-�(e)1 ⇡k

2z21

+ �(e)1 ⇡k

2z20

�,k + 1

k � 1z0 < z1 < 1,

fZ0(z0) =⇣2�(e)

2 + 2�(e)1 ⇡⇣

� 2↵

21 z0

⇥ exph�2�(e)

2 z0 � �(e)1 ⇡⇣

� 2↵

21 z20

i,

�2,2(z0, z1) = ⇡k2z21 � (kz0)

2(✓ � 1

2sin(2✓))

� (kz1)2(�� 1

2sin(2�)),

✓ = arccos

(z0 + z1)2 + (kz0)2 � (kz1)2

2kz0(z0 + z1)

!,

� = arccos

(z0 + z1)2 � (kz0)2 + (kz1)2

2kz1(z0 + z1)

!,

and k = ⇣� 1

↵21 .

Proof: See Appendix E.

!0

Typical ReceiverReceiving Vehicularusers

Tier 2 nodesTagged Tier 1 node

Coverage region ofTier 2 nodes

Fig. 4. Illustration of the coverage region of a tagged tier 1 node.

In order to compute the load on the tagged tier 2 nodelocated at y⇤2 , we now need to determine the length of thetagged tier 2 cell. Recall that the tagged tier 2 cell is thecell that contains the typical receiver located at the origin andis denoted by Ztag = {x : x 2 L0, P2G2B2kx � y⇤2k�↵ �PyGyBykx � yk�↵

, 8y 2 �(e)1 [ ⌅(e)

L0}. Hence, the length of

the tagged tier 2 cell is larger on average than that of a typicaltier 2 cell. Thus, the PDF of the length of the tagged tier 2cell Ztag can be computed from the PDF of the length of atypical tier 2 cell, as given in the following Lemma.

Lemma 8. The PDF of the length of the tagged tier 2 cell is

fZtag(w) =wR1

0 wfZtyp(w)dwfZtyp(w), (19)

where fZtyp(·) is given in Lemma 7.

Proof: As the typical receiver is expected to lie in a longercell than a typical tier 2 cell, the PDF of the length of thetagged cell is directly proportional to its length [34, Lemma3], [35, Chapter 6]. Using appropriate normalization for thePDF of Ztag, we obtain the final expression.

Having determined the PDF of the tagged tier 2 cell length,we now compute the distribution of load on the tagged tier 2node in the following Lemma.

Lemma 9. The probability mass function (PMF) of the loadon the tagged tier 2 node is

P(J2 = j2 + 1|E2) =Z 1

0

exp(��rw)(�rw)j2

j2!fZtag(w)dw,

j2 = 0, 1, 2, . . . . (20)

Proof: The proof follows from the Poisson distribution ofreceiving vehicular user nodes on the typical line.

B. Load on Tagged Tier 1 NodeIn contrast to the tier 2 nodes with 1D coverage regions, the

tier 1 nodes have 2D coverage regions which partition R2 intomultiple cells, as shown in Fig. 4. As all the tier 1 nodes havethe same transmit power and beamforming gains during theassociation process, their coverage regions are simply givenby the Voronoi cells constructed from the 2D PPP �(e)

1 . Thus,the tagged tier 1 cell is the Voronoi cell containing the originand is denoted by Vtag = {x : kx�y⇤1k kx�yk8y 2 �(e)

1 }.Although the vehicular users located on the line segmentsinside Vtag receive the highest average biased power among

Page 9: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

9

the tier 1 nodes from the node at y⇤1 , some of the vehicularusers inside the cell are served by tier 2 nodes located on theseline segments, as illustrated in Fig. 4. Thus, a few segments ofthe lines inside the Voronoi cell are not part of the tagged celland hence, the load on the tagged tier 1 node is the numberof vehicular users located on the line segments inside Vtag

which are not covered by any tier 2 node. As the load isproportional to the length of these line segments, we needto compute the exact distribution of the total length of linesegments that are not covered by tier 2 nodes inside Vtag,which is intractable. Hence, we propose to compute the ratecoverage using the mean load on the tagged node, which willbe formally explained in the next subsection. So, we willnow compute the mean load on the tagged tier 1 node inthis network, which has not been studied in the literature.This result will be derived using the results discussed in theprevious subsection.

Lemma 10. The mean load on a tagged tier 1 node is

J̄1 = 1 + �r

1.28⇡�l

�(e)1

+3.216

q�(e)1

!

⇥✓1� �

(e)2

Z 1

0zfZtyp(z)dz

◆. (21)

Proof: See Appendix F.

C. Rate Coverage

Using the results derived thus far, we will now compute therate coverage of the typical receiver which is formally definedas the probability with which the data rate achievable at thetypical receiver exceeds a predefined target rate T . Before weproceed to the computation of overall rate coverage, we willfirst focus on the case where the typical receiver is connectedto a tier 1 node. As discussed in the previous subsection, it isquite hard to exactly characterize the load on the tagged tier1 node. So, we make the following assumption which enablesthe computation of the rate coverage without much loss in theaccuracy of our final results.

Assumption 2. We assume that all the tier 1 nodes areoperating at the mean load.

Under this assumption, the overall rate coverage of thetypical receiver is given in the following theorem.

Theorem 3. The rate coverage of the typical receiver is

Rc ⇡ P(E1)m1�1X

k=0

Z 1

0

(-m1�1r↵)k

(P1G1)kk!

@k

@skLI(s|R, E1)

s=m1�1r↵

P1G1

⇥ fR(r|E1)dr + P(E2)1X

j2=1

P(J2 = j2)

⇥m20�1X

k=0

Z 1

0

(-m20�2r↵)k

(P2G2)kk!

@k

@skLI(s|R, E2)

s=m20�2r↵

P2G2

⇥ fR(r|E2)dr, (22)

where �1 = 2TJ̄1W � 1 and �2 = 2

Tj2W -1.

Proof: The rate coverage can be computed as

Rc = P(R > T ) =X

i2{1,2}

P(R > T |Ei)P(Ei)

=X

i2{1,2}

EJi

P(R > T |Ei, Ji)

�P(Ei)

=X

i2{1,2}

1X

ji=1

P⇣SIR > 2

TjiW -1|Ei, Ji

⌘P(Ji = ji|Ei)P(Ei).

(23)

Although the SIR measured at the typical receiver and theload on the tagged node are correlated, we assume that theyare independent in the interest of analytical tractability andproceed with our analysis. Thus, the rate coverage is given by

Rc ⇡X

i2{1,2}

1X

ji=1

P⇣SIR > 2

TjiW � 1|Ei

⌘P(Ji = ji|Ei)P(Ei)

(a)= P

⇣SIR > 2

TJ̄1W � 1|E1

⌘P(E1)

+ P(E2)1X

j2=1

P⇣SIR > 2

Tj2W -1|E2

⌘P(J2 = j2|E2), (24)

where (a) follows from Assumption 2. This completes theproof.

VI. NUMERICAL RESULTS AND DISCUSSION

In this section, we will present the numerical results forthe coverage and rate analysis of the vehicular network. Wewill verify the accuracy of our analytical results by comparingthem with the empirical results obtained from Monte-Carlosimulations by running the scripts provided in [36]. We willalso discuss the effect of various parameters such as selectionbias factors, and densities of nodes on the performance ofthe network. We will then provide a few design insights thatcould aid in improving the performance of the network withoutdeploying additional infrastructure.

A. Coverage ProbabilityWe first simulate the network model described in section III

in MATLAB to compute the empirical distribution of coverageprobability and rate coverage. For all the numerical resultspresented in this section, we assume that �r = 15 nodes/km,↵ = 4, W = 10 MHz, m1 = m20 = m21 = 1 and!1 = !20 = !21 = 0 dB. We assume that the ratio of mainlobe to side lobe gains for both tier 1 and tier 2 nodes is 20dB. We also assume that the beamwidth of the main lobe of atier 1 node is 0.1⇡ and its orientation is uniformly distributedin [0, 2⇡). Thus, the probability qc that the main lobe of aninterfering tier 1 node is aligned towards the typical receiver is0.05. We verify our analytical results for coverage probabilityby numerically comparing them with the empirical resultsin Fig. 5. As expected, the coverage probability evaluatedusing the analytical expressions in Theorem 2 matches closelywith the results obtained from simulations. This shows thatthe approximation of the Cox process model presented inSection IV-C is remarkably accurate for the coverage analysis.

Page 10: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

10

-50 0 50

SIR threshold, β (dB)

0

0.2

0.4

0.6

0.8

1

Coverageprobab

ility,Pc

SimulationTheorem 2

Fig. 5. Coverage probability of the typical receiver as a functionof SIR threshold (µl = 10 km�1, �1 = 0.5 nodes/km2, �2 = 4nodes/km, B1 = B2 = 0 dB, P1 = 40 dBm, P2 = 23 dBm and[�1 �20 �21] = [4 2 4] dB).

0 5 10 15 20

Tier 2 Selection Bias, B2 (dB)

0.74

0.76

0.78

0.8

0.82

0.84

0.86

Coverageprobab

ility,Pc

Decreasing λ2 = 10,5, 2 nodes/km2

Fig. 6. Coverage probability of the typical receiver as a functionof selection bias B2 (µl = 5 km�1, �1 = 2 nodes/km2, B1 = 0dB, P1 = 43 dBm, P2 = 23 dBm, [�1 �20 �21] = [4 2 4] dB,and � = 0 dB).

Therefore, the proposed approximation could also enable othersimilar analyses that may not otherwise be possible due to thenature of the Cox process.

Impact of selection bias. We plot the coverage probabilityof the typical receiver as a function of selection bias oftier 2 nodes B2 in Fig. 6. It can be observed that thecoverage probability improves initially as B2 increases andthen it begins to degrade beyond a certain value. This trend issomewhat counterintuitive because biasing forces the typicalreceiver to connect to a tier 2 which yields lower power thanthe other candidate serving node tier 1. Hence, we expect thecoverage probability to decrease as B2 increases. However,when the typical receiver connects to the tier 2 node, theclosest tier 1 node acts an interferer. Recall that the mainlobes of interfering tier 1 nodes are in the direction of thetypical receiver with only a probability qc. Consequently, theinterference caused by the closest tier 1 node is quite low whenits main lobe is not directed towards the typical receiver. Thismay improve the SIR and hence the coverage probability evenwhen the typical receiver connects to a tier 2 node that yields

0 5 10 15 20 25 30

Density of Tier 2 nodes, λ2 (nodes/km)

0.7

0.75

0.8

0.85

0.9

0.95

CoverageProbab

ility

Increasing λ1 = 0.2, 0.5,1, 2 nodes/km2

Fig. 7. Coverage probability of the typical receiver as a functionof the density of tier 2 nodes (µl = 10 km�1, B1 = B2 = 0 dB,P1 = 43 dBm, P2 = 23 dBm, [�1 �20 �21] = [4 2 4] dB, and� = 0 dB).

0.2 0.4 0.6 0.8 1

Density of Tier 1 nodes, λ1 (nodes/km2)

0.66

0.68

0.7

0.72

0.74

0.76

0.78

CoverageProbab

ility,Pc

Increasing B1 = 0, 2, 58, 10 dB

Fig. 8. Coverage probability of the typical receiver as a functionof the density of tier 1 nodes (µl = 5 km�1, �2 = 5 nodes/km,B2 = 0 dB, P1 = 43 dBm, P2 = 23 dBm, [�1 �20 �21] =[4 2 4] dB, and � = 0 dB).

lower power. This is why we observe a slight improvement inthe coverage probability as B2 increases.

Impact of node densities. In Fig. 7, we plot the coverageprobability as a function of the density of tier 2 nodes fordifferent values of the density of tier 1 nodes without anyselection bias (B1 = B2 = 0 dB). We observe that thecoverage probability of the typical receiver initially decreasesas the density of tier 2 nodes increases and then begins toimprove beyond a certain value of �2. While the densificationof tier 2 nodes increases the desired signal power due to thereduced distance from the typical receiver to the serving tier 2node, it also increases the interference power. When operatingat low node density, upon increasing the density of tier 2nodes, the increase in the interference power is more dominantthan the reduction of the serving distance. However, beyond acertain value of �2, the increment in the desired signal powerdue to reduced serving distance exceeds the interference powerand hence the trend is reversed. That said, this trend maychange depending on the selection bias as we observe that thecurves corresponding to different densities cross over beyond

Page 11: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

11

5 10 15 20

Load on the Tagged Tier 2 Node, J2

0

0.2

0.4

0.6

0.8

1

CDFof

J2,FJ2(j

2)

SimulationLemma 9

Fig. 9. CDF of the load on the tagged tier 2 node (µl = 10 km�1,�1 = 0.5 nodes/km2, �2 = 4 nodes/km, B1 = B2 = 0 dB,P1 = 40 dBm, P2 = 23 dBm, and [�1 �20 �21] = [4 2 4] dB).

0 5 10 15 20

Target Rate, T (Mbps)

0

0.2

0.4

0.6

0.8

1

RateCoverage,Rc

SimulationTheorem 3

Fig. 10. Rate coverage of the typical receiver as a function oftarget rate (µl = 10 km�1, �1 = 0.5 nodes/km2, �2 = 4nodes/km, B1 = B2 = 0 dB, P1 = 40 dBm, P2 = 23 dBmand [�1 �20 �21] = [8 4 8] dB).

a certain value of B2, as shown in Fig. 6. Similarly, from Fig.8, we observe that the coverage probability increases with anincrease in the density of tier 1 nodes when there is no bias andthe trend gradually changes as the selection bias B1 increases.

B. Rate CoverageAs the distribution of load is a key component in the com-

putation of the downlink rate, we first compare the CDF of theload on the tagged tier 2 node evaluated using the expressiongiven in Lemma 9 with the results obtained from simulations.We observe that the theoretical results match closely with thesimulations as shown in Fig. 9. This also shows that the PDFof the length of the tagged tier 2 cell given in Lemma 8 is quiteaccurate. We observe that the approximate load distribution,combined with the exact association probabilities and accuratecoverage probability expressions, yields a tight approximationof the rate coverage as depicted in Fig. 10 with only a smallgap between the simulation and analytical results. We will nowstudy the impact of selection bias and node densities on therate coverage of the typical receiver.

0 5 10 15 20

Cellular MBS Selection Bias, B1 (dB)

0.35

0.36

0.37

0.38

0.39

0.4

RateCoverage,Rc

Increasing λ1 = 0.25,0.4, 0.5 nodes/km2

Fig. 11. Rate coverage of the typical receiver as a function ofselection bias B1 (µl = 5 km�1, �2 = 5 nodes/km, B2 = 0 dB,P1 = 43 dBm, P2 = 23 dBm, [�1 �20 �21] = [4 2 4] dB, andT = 10 Mbps).

0 5 10 15 20

Density of RSUs (nodes/km)

0

0.1

0.2

0.3

0.4

0.5

0.6

RateCoverage,Rc

Fig. 12. Rate coverage of the typical receiver as a function ofthe density of tier 2 nodes (µl = 5 km�1, �1 = 0.5 nodes/km2,B1 = B2 = 0 dB, P1 = 43 dBm, P2 = 23 dBm, [�1 �20 �21] =[4 2 4] dB, and T = 10 Mbps).

Impact of selection bias. We plot the rate coverage as afunction of selection bias of tier 1 nodes B1 for different valuesof node densities, as shown in Fig. 11. We observe that therate coverage initially increases with B1 and then degradesbeyond a certain value, thereby yielding an optimal bias factorthat maximizes the rate coverage of the typical receiver fora given set of node densities. Also, the optimal bias factordecreases as the density of tier 1 nodes increases.

Impact of node densities. From Figs. 11 and 12, we observethat the rate coverage improves as the density of tier 1 andtier 2 nodes increases when there is no selection bias. This isbecause of the reduced load on the tagged nodes as the densityof nodes increases while the user density remains unchanged.However, we notice that this trend does not hold for highervalues of selection bias B1, as shown in Fig. 11. This isbecause of the downward trend in coverage probability forhigher values of B1, as shown in Fig. 8.

Page 12: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

12

C. Design Insights

We will first compare the performance of the C-V2Xnetwork considered in the paper with a single-tier setup wherethe vehicular nodes are served only cellular MBSs, whichcorresponds to the case �2 = 0 in the current model. FromFig. 7, we observe that the coverage probability of the typicalreceiver in the C-V2X network is significantly lower thanthat of a single-tier network. Although the addition of tier 2nodes to the network reduces the distance between the typicalreceiver and the serving node, it also significantly increasesthe interference power. Hence, the coverage performance ofthe heterogeneous C-V2X network is worse than that of thesingle-tier setup of cellular MBSs. On the other hand, fromFig. 12, we observe that the rate coverage of the typicalreceiver in the heterogeneous network is remarkably betterthan that of single-tier network due to improved spatial reuseof frequency resources. The deployment of RSUs significantlyreduces the load on the cellular MBSs and hence improves therate coverage of the typical receiver.

We will now provide some key system-level insights into thedeployment of MBSs and RSUs based on the trends observedin coverage probability and rate coverage. From Fig. 11, weobserve that the rate coverage can be improved by increasingthe density of tier 1 nodes. However, deploying more tier 1nodes may not always be an option because of cost and siteacquisition constraints. So, an alternate solution to improvethe performance without additional deployment of nodes is toadjust the selection bias. For instance, we can observe that therate coverage obtained by increasing the node density from�1 = 0.25 nodes/km2 to �1 = 0.5 nodes/km2 without anybias can also be achieved by increasing the selection bias toB1 = 5 dB while keeping the node density unchanged. Unliketier 1 nodes, we observe that an increase in the density of tier2 has different effects on the coverage probability dependingon the regime of operation, as depicted in Fig. 7. However,as mentioned earlier, the rate coverage of the typical receiverimproves with the density of tier 2 nodes as shown in Fig. 12.Hence, it is important to consider the trade-off between thesetwo metrics in the deployment of RSUss.

VII. CONCLUSION

In this paper, we have presented the coverage and rate analy-sis of a C-V2X network in the presence of shadowing. We havemodeled the locations of vehicular nodes and RSUs as Coxprocesses driven by PLP and the locations of cellular MBSsas a 2D PPP. Assuming a fixed selection bias and maximumaverage received power based association, we computed theprobability with which a typical receiver connects to anothervehicular node or an RSU and a cellular MBS. Inspired bythe asymptotic behavior of the Cox process, we approximatedthe Cox process of vehicular nodes and RSUs by a 2D PPPto characterize the interference from those nodes. We thenderived the expression for SIR-based coverage probability interms of the Laplace transform of interference power distri-bution. Further, we computed the rate coverage of the typicalreceiver by characterizing the load on the serving nodes. Wehave also provided several design insights based on the trends

observed in the coverage probability and rate coverage as afunction of network parameters. We have observed that theperformance gain obtained by the densification of MBSs canbe equivalently achieved by choosing an appropriate selectionbias without additional deployment of infrastructure.

This work has numerous extensions. First of all, the pro-posed approximation of the spatial model based on the asymp-totic behavior of the Cox process could be applied to lendtractability to similar analyses which may not otherwise bepossible. For instance, the approximate spatial model and someof the intermediate results presented in this paper could beuseful in studying other metrics, such as latency and packetreception rate. Another meaningful extension of this workcould be to study the handover rate of moving vehicularusers from not just one MBS to another but also from anRSU or a vehicular node to a MBS and vice versa. Further,there is scope for improving the system model by consideringcorrelated shadowing where the nodes in close proximityexperience similar loss in signal power due to blockagesfrom the same buildings. Hence, it is worthwhile to developgenerative blockage models that imitate the propagation ofsignals in a physical environment [37], [38].

APPENDIX APROOF OF LEMMA 1

The void probability can be computed as

P�Np(A) = 0

�= E ` [P (Np ([Li2 `{Li \A}) = 0| `)]

(a)= E `

"Y

Li2 `

P (Np(Li \A) = 0| `)

#

(b)= E `

"Y

Li2 `

exp (��p⌫1 (Li \A))

#

(c)= E C

2

4Y

(⇢i,✓i)2 C

exp���p⌫1

�L(⇢i,✓i) \A

��3

5

(d)= exp

-�`

Z 2⇡

0

Z

R+

⇥1- exp

�-�p⌫1

�L(⇢,✓) \A

��⇤d⇢d✓

where (a) follows from the independent distribution of pointson the lines, (b) follows from the void probability of a 1DPPP, (c) follows from rewriting the expression in terms of thepoint process C in the representation space C, and (d) followsfrom the probability generating functional (PGFL) of the 2DPPP C .

APPENDIX BPROOF OF THEOREM 1

In order to prove this result, by Choquet’s theorem [31],it is sufficient to show that the void probability of the Coxprocess converges to that of a 2D PPP. Therefore, we will nowapply the limits �` ! 1 and �p ! 0 on the expression forvoid probability derived in Lemma 1. As the overall densityof nodes in the network �a = ⇡�`�p remains constant, theapplication of the two limits �` ! 1 and �p ! 0 can besimplified to a single limit by substituting �p = �a

⇡�`in the

Page 13: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

13

expression in (1). Thus, the asymptotic void probability canbe evaluated as

lim�`!1

P�Np(A) = 0

= lim�`!1

exp

-�`

Z 2⇡

0

Z

R+

⇥1- exp

�-�p⌫1

�L(⇢,✓) \A

��⇤d⇢d✓

= exp

"lim

�`!1-�`

Z 2⇡

0

Z

R+

1� exp

✓-

�a

⇡�`

⇥ ⌫1

�L(⇢,✓) \A

�◆�d⇢d✓

#

(a)= exp

"lim

�`!1�`

Z 2⇡

0

Z

R+

1X

k=1

�-�a⌫1

�L(⇢,✓) \A

��k

(⇡�`)kk!

d⇢d✓

#

(b)= exp

"lim

�`!1�`

Z 2⇡

0

Z

R+

�-�a⌫1

�L(⇢,✓) \A

��

⇡�`

d⇢d✓

+1X

k=2

Z 2⇡

0

Z

R+

lim�`!1

�-�a⌫1

�L(⇢,✓) \A

��k

⇡k�k�1`

k!d⇢d✓

#

(c)= exp

lim

�`!1� �a

⇡�`

Z 2⇡

0

Z

R+

⌫1

�L(⇢,✓) \A

��`d⇢d✓

(d)= exp

2

4 lim�`!1

� �a

⇡�`

E

2

4X

(⇢,✓)2 C

⌫1

�L(⇢,✓) \A

�3

5

3

5

(e)= exp

lim

�`!1��a

µ`

µ`⌫2(A)

�= e

��a⌫2(A),

where (a) follows from the Taylor series expansion of ex-ponential function, (b) follows from switching the order ofintegral and summation operations and applying DominatedConvergence Theorem (DCT) on the second term, (c) followsfrom the limit of the integrand in the second term whichevaluates to 0 for all k � 2, (d) follows from the Campbell’stheorem for sums over stationary point processes, and (e)follows from the definition of line density of line processes,where ⌫2(A) is the two-dimensional Lebesgue measure (area)of the region A.

APPENDIX CPROOF OF LEMMA 2

The typical receiver connects to a tier 1 node if the averagebiased received power from the closest tier 1 node exceeds theaverage biased received power from the closest tier 2 node onthe typical line. Thus, the probability of occurrence of theevent E1 can be computed as

P(E1) = P�P1B1G1R

�↵

1 > P2B2G2R�↵

2

(a)= ER2

hP⇣R1 < ⇣

� 1↵

21 r2|R2

⌘i

=

Z 1

0FR1

⇣⇣� 1

↵21 r2

⌘fR2(r2)dr2

(b)= 1�

Z 1

0exp

h��

(e)1 ⇡⇣

� 2↵

21 r22

i2�(e)

2 exp⇣�2�(e)

2 r2

⌘dr2

= 1�

vuut (�(e)2 )2

�(e)1 ⇣

� 2↵

21

exp

"(�(e)

2 )2

�(e)1 ⇡⇣

� 2↵

21

#erfc

0

@ �(e)2q

�(e)1 ⇡⇣

� 2↵

21

1

A ,

where (a) follows from substituting ⇣21 = P2B2G2P1B1G1

andconditioning on R2 , and (b) follows from substituting theexpressions for FR1(·) and fR2(·) from (8) and (10) in theprevious step. Since the events E1 and E2 are complementary,P(E2) = 1� P(E1).

APPENDIX DPROOF OF LEMMA 5

As the aggregate interference at the typical receiver canbe decomposed into three independent components I1, I20,and I21, the conditional Laplace transform of interference canbe computed as product of conditional Laplace transforms ofinterference of these individual components, i.e.,

LI(s|R, E1) = LI1(s|R, E1)LI20(s|R, E1)LI21(s|R, E1).(25)

First, we will consider the interference from the tier 1 nodesI1. Recall that the beamforming gain of an interfering tier1 node is G1 with probability qc and g1 with probability(1� qc). So, we can partition the tier 1 interfering nodes intotwo independent 2D PPPs �(e)

11 and �(e)12 with densities qc�

(e)1

and (1 � qc)�(e)1 and with beamforming gains G1 and g1,

respectively. Thus, the Laplace transform of interference fromthe tier 1 nodes can be computed as

LI1(s|R, E1) = E [exp(�sI1)]

= E"exp

� s

X

y2�(e)11 \b(o,R)

P1G1H1kyk�↵

� s

X

y2�(e)12 \b(o,R)

P1g1H1kyk�↵

!#

(a)= E

�(e)11EH1

"Y

y2�(e)11 \b(o,R)

e�sP1G1H1kyk�↵

#

⇥ E�(e)

12EH1

"Y

y2�(e)12 \b(o,R)

e�sP1g1H1kyk�↵

#

(b)= E

�(e)11

"Y

y2�(e)11 \b(o,R)

✓1 +

sP1G1kyk�↵

m1

◆�m1#

⇥ E�(e)

11

"Y

y2�(e)12 \b(o,R)

✓1 +

sP1g1kyk�↵

m1

◆�m1#

(c)= exp

"� 2⇡qc�

(e)1

Z 1

r

1�✓1 +

sP1G1y�↵

m1

◆�m1

ydy

� 2⇡(1� qc)�(e)1

Z 1

r

1�✓1 +

sP1g1y�↵

m1

◆�m1

ydy

#,

(26)

where (a) follows from the independence of �(e)11 and �(e)

12 ,(b) follows from the Nakagami-m fading assumption, and (c)follows from substituting y = kyk and the PGFL of a 2DPPP. The Laplace transform of interference from tier 2 nodes

Page 14: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

14

located on the typical line conditioned on R and E1 can becomputed as LI20(s|R, E1) =

E"exp

� s

X

y2⌅(e)L0

\[�⇣

1↵21 r,⇣

1↵21 r]

P2G2H20kyk�↵

!#

= E⌅(e)

L0

EH20

"Y

y2⌅(e)L0

\[�⇣

1↵21 r,⇣

1↵21 r]

e�sP2G2H20kyk�↵

#

(a)= exp

"� 2�(e)

2

Z 1

1↵21 r

1�✓1 +

sP2G2y�↵

m20

◆�m20

dy

#,

(27)

where (a) simply follows from the Nakagami-m fading as-sumption and the PGFL of 1D PPP. Under Assumption 1, theLaplace transform of I21 conditioned on R and E1 can becomputed as

LI21(s|R, E1) = E"exp

� s

X

y2�(a)2

P2g2H21kyk�↵

!#

= E�(a)

2EH21

"Y

y2�(a)2

e�sP2g2H21kyk�↵

#

(a)= exp

"� 2⇡�(a)

2

Z 1

01�

✓1 +

sP2g2y�↵

m21

◆�m21

ydy

#,

(28)

where (a) follows from the Nakagami-m fading assumptionand PGFL of a 2D PPP.

Substituting (26), (27), and (28) in (25), we obtain thefinal expression for the conditional Laplace transform of theaggregate interference power distribution.

APPENDIX EPROOF OF LEMMA 7

The total length of the typical cell is the sum of the distancesto the farthest points on the line on either side of the typicalnode such that any user closer than that point would connectto the typical node. We denote the farthest points of the typicalcell by z0 and z1 and the distances to these points from thetypical node by Z0 and Z1, respectively. Thus, the length ofthe typical cell is Ztyp = Z0+Z1. Without loss of generality,we assume that z0 is to the right of xtyp and we will firstfocus on the distance Z0.

The CDF of Z0 can be computed as

FZ0(z0) = 1� P(Z0 > z0)

= 1� P�N2(L \ b (z0, z0)) = 0

⇥ P⇣N1

⇣b

⇣z0, ⇣

� 1↵

21 z0

⌘⌘= 0⌘

= 1� exph�2�(e)

2 z0 � �(e)1 ⇡⇣

� 2↵

21 z20

i,

where N1(·) and N2(·) denote number of tier 1 and tier 2nodes, respectively, and b(c, d) denotes a ball of radius d

centered at c. As the random variables Z0 and Z1 are notindependent, we will now compute the CDF of Z1 conditioned

on Z0. Before we proceed with that derivation, it is importantto note that the boundary of the typical cell z0 is determinedby an adjacent tier 2 node on the same line or a tier 1 node. Inthe former case, conditioning on Z0 implies that there is a tier2 node on the same line at a distance 2z0 from xtyp. On theother hand, if the boundary was determined by a tier 1 node,conditioning on Z0 also means that there exists a tier 1 nodeon the circumference of the disc b(z0, kz0), where k = ⇣

� 1↵

21 .Therefore, in order to exactly characterize the length of thetypical cell, we need to distinguish these two cases in thederivation of the conditional distribution of Z1. However, thiswould result in multiple sub-cases and hence complicate theanalysis. Therefore, in the interest of tractability, we do notdistinguish these two cases in our analysis. As will be shownin the next section, this does not affect the accuracy of ourfinal results. Thus, the CDF of Z1 conditioned on Z0 can becomputed as

FZ1(z1|z0)

⇡ 1� P⇣{N2(L \ b (z1, z1)) = 0}

\ {N1 (b(z1, kz1)) = 0}��N1 (b(z0, kz0)) = 0

= 1� P (N2((L \ b (z1, z1)) = 0)

⇥ P (N1 (b(z1, kz1) \ b(z0, kz0)) = 0)

= 1- exp(�2�(e)2 z1) exp

⇣-�(e)

1 ⌫2(b(z1, kz1) \ b(z0, kz0))⌘.

(29)

The area of the region b(z1, kz1)\ b(z0, kz0) is determined bythe relative ranges of z0 and z1, as shown in Fig. 13. Hence,we obtain a piecewise function for this area which is given by

⌫2

⇣b(z1, kz1) \ b(z0, kz0)

=

8>>>>><

>>>>>:

�2,1(z0, z1), 0 < z1 <k � 1

k + 1z0,

�2,2(z0, z1),k � 1

k + 1z0 < z1 <

k + 1

k � 1z0

�2,3(z0, z1),k + 1

k � 1z0 < z1 < 1

=

8>>>>>>>>><

>>>>>>>>>:

0 0 < z1 <k � 1

k + 1z0,

⇡k2z21-(kz0)

2(✓-1

2sin(2✓))

-(kz1)2(�-1

2sin(2�)),

k � 1

k + 1z0 < z1 <

k + 1

k � 1z0

⇡(kz1)2 � ⇡(kz0)

2,

k + 1

k � 1z0 < z1 < 1,

(30)

where

✓ = arccos

(z0 + z1)2 + (kz0)2 � (kz1)2

2kz0(z0 + z1)

!,

and � = arccos

(z0 + z1)2 � (kz0)2 + (kz1)2

2kz1(z0 + z1)

!.

Substituting (30) in (29), we obtain the conditional CDF of

Page 15: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

15

!" !#

!# + !"

z#z"

(i)

!" !#

z#z"

!# + !"

(ii)

!"

!#

!# + !"z#z"

(iii)

Fig. 13. An illustration of the three cases: (i) 0 < z1 < k�1k+1 z0, (ii) k�1

k+1 z0 < z1 < k+1k�1 z0, and (iii) k+1

k�1 z0 < z1 < 1.

Z1 as follows

FZ1(z1|z0) =8>>>>><

>>>>>:

1� e�2�(e)

2 z1 , 0 < z1 <k � 1

k + 1z0,

1� e�2�(e)

2 z1��(e)1 �2,2(z0,z1),

k � 1

k + 1z0 < z1 <

k + 1

k � 1z0,

1� e�2�(e)

2 z1��(e)1 ⇡k

2z21+�

(e)1 ⇡k

2z20 ,

k + 1

k � 1z0 < z1 < 1.

(31)

Having determined the CDF of Z0 and conditional CDF ofZ1, the CDF of the length of the typical cell can be computedas FZtyp(z)

= P(Z0 + Z1 < z) =

Zz

0P(Z1 < z � z0|Z0)fZ0(z0)dz0

=

Zz

0FZ1(z � z0|z0)fZ0(z0)dz0

=

Zz

k+12k z

FZ1,1(z-z0)fZ0(z0)dz0 +

Z k+12k z

k�12k z

FZ1,2(z-z0)fZ0(z0)dz0

+

Z k�12k z

0FZ1,3(z � z0)fZ0(z0)dz0 (32)

The PDF of Ztyp can be obtained by taking the derivative ofFZtyp(z) w.r.t. z.

APPENDIX FPROOF OF LEMMA 10

The mean load on tagged tier 1 node can be obtainedby determining the difference between the mean number ofvehicular users inside Vtag and the mean number of vehicularusers served by tier 2 nodes inside Vtag. Thus, the mean loadcan be computed as

E[J1] = 1 + E [Nr(Vtag)]� E [N2(Vtag)]E [J 02] , (33)

where Nr(·) represents the number of receiving vehicular usernodes and J

02 denotes the load on a typical tier 2 node. The

first term, which represents the average number of vehicles inVtag, is E [Nr(Vtag)]

= �rE

2

4X

Li2�l0

⌫1(Li \ Vtag)

3

5 =

0

@1.28⇡�l

�(e)1

+3.216

q�(e)1

1

A�r.

(34)

The proof of this result is given in [20, Theorem 2] andis hence skipped. Similarly, the second term in (33) whichrepresents the average number of tier 2 nodes in Vtag is givenby

E [N2(Vtag)] =

0

@1.28⇡�l

�(e)1

+3.216

q�(e)1

1

A�(e)2 . (35)

The third term in (33), which corresponds to the meanload on a typical tier 2 node, can be easily computed usingdistribution of the length of the typical cell as

E [J 02] = �rE[Ztyp] = �r

Z 1

0zfZtyp(z)dz, (36)

where fZtyp(z) is given in Lemma 7. Substituting (34), (35),and (36) in (33), we obtain the final expression for the meanload on the tagged tier 1 node.

REFERENCES

[1] V. V. Chetlur and H. S. Dhillon, “Poisson line Cox process: Asymp-totic characterization and performance analysis of vehicular networks,”accepted to IEEE Global Commun. Conf. (Globecom) 2019.

[2] H. Hartenstein and L. P. Laberteaux, “A tutorial survey on vehicular adhoc networks,” IEEE Commun. Mag., vol. 46, no. 6, pp. 164–171, Jun.2008.

[3] S. Biswas, R. Tatchikou, and F. Dion, “Vehicle-to-vehicle wirelesscommunication protocols for enhancing highway traffic safety,” IEEECommun. Mag., vol. 44, no. 1, pp. 74–82, Jan. 2006.

[4] P. Papadimitratos, A. D. L. Fortelle, K. Evenssen, R. Brignolo, andS. Cosenza, “Vehicular communication systems: Enabling technologies,applications, and future outlook on intelligent transportation,” IEEECommun. Mag., vol. 47, no. 11, pp. 84–95, Nov. 2009.

[5] M. Boban, A. Kousaridas, K. Manolakis, J. Eichinger, and W. Xu,“Connected roads of the future: Use cases, requirements, and designconsiderations for vehicle-to-everything communications,” IEEE Veh.Technol. Mag., vol. 13, no. 3, pp. 110–123, Sep. 2018.

[6] 3GPP TR 36.885, “Study on LTE-based V2X services,” Jul. 2016.[7] V. V. Chetlur and H. S. Dhillon, “Coverage analysis of a vehicular

network modeled as Cox process driven by Poisson line process,” IEEETrans. Wireless Commun., vol. 17, no. 7, pp. 4401–4416, Jul. 2018.

[8] F. Baccelli, M. Klein, M. Lebourges, and S. Zuyev, “Stochastic geom-etry and architecture of communication networks,” TelecommunicationSystems, vol. 7, no. 1, pp. 209–227, Jun. 1997.

[9] B. Blaszczyszyn, P. Muhlethaler, and Y. Toor, “Maximizing through-put of linear vehicular Ad-hoc NETworks (VANETs) – a stochasticapproach,” in European Wireless Conf., May 2009, pp. 32–36.

[10] S. Busanelli, G. Ferrari, and R. Gruppini, “Performance analysis ofbroadcast protocols in VANETs with Poisson vehicle distribution,” inInt. Conf. on ITS Telecommun., Aug 2011, pp. 133–138.

[11] M. Mabiala, A. Busson, and V. Veque, “Inside VANET: Hybrid networkdimensioning and routing protocol comparison,” in Proc., IEEE Veh.Technol. Conf., Apr. 2007, pp. 227–232.

Page 16: Coverage and Rate Analysis of Downlink Cellular Vehicle-to

16

[12] M. Ni, M. Hu, Z. Wang, and Z. Zhong, “Packet reception probabilityof VANETs in urban intersection scenario,” in Int. Conf. on ConnectedVehicles and Expo, Oct. 2015, pp. 124–125.

[13] E. Steinmetz, M. Wildemeersch, T. Q. S. Quek, and H. Wymeersch, “Astochastic geometry model for vehicular communication near intersec-tions,” in Proc., IEEE Globecom Workshops, Dec. 2015, pp. 1–6.

[14] A. Tassi, M. Egan, R. J. Piechocki, and A. Nix, “Modeling and design ofmillimeter-wave networks for highway vehicular communication,” IEEETrans. Veh. Technol., vol. 66, no. 12, pp. 10 676–10 691, 2017.

[15] ——, “Wireless vehicular networks in emergencies: A single frequencynetwork approach,” in Int. Conf. on Recent Advances in Signal Process.,Telecommun. & Computing (SigTelCom), 2017, pp. 27–32.

[16] V. V. Chetlur and H. S. Dhillon, “Success probability and area spectralefficiency of a VANET modeled as a Cox process,” IEEE WirelessCommun. Lett., vol. 7, no. 5, pp. 856–859, Oct. 2018.

[17] B. Blaszczyszyn and P. Muhlethaler, “Random linear multihop relayingin a general field of interferers using spatial Aloha,” IEEE Trans.Wireless Commun., vol. 14, no. 7, pp. 3700–3714, Jul. 2015.

[18] Y. Wang, K. Venugopal, A. F. Molisch, and R. W. Heath, “MmWavevehicle-to-infrastructure communication: Analysis of urban microcellu-lar networks,” IEEE Trans. Veh. Technol., vol. 67, no. 8, pp. 7086–7100,Aug. 2018.

[19] F. Morlot, “A population model based on a Poisson line tessellation,”in Proc., Modeling and Optimization in Mobile, Ad Hoc and WirelessNetworks, May 2012, pp. 337–342.

[20] A. Chattopadhyay, B. Blaszczyszyn, and E. Altman, “Two-tier cellularnetworks for throughput maximization of static and mobile users,” IEEETrans. Wireless Commun., vol. 18, no. 2, pp. 997–1010, Feb. 2019.

[21] J. Rachad, R. Nasri, and L. Decreusefond, “How to dimension 5G net-work when users are distributed on roads modeled by Poisson line pro-cess,” arXiv preprint, 2018, available online: arxiv.org/abs/1805.06637.

[22] C. Choi and F. Baccelli, “Poisson Cox point processes for vehicularnetworks,” IEEE Trans. Veh. Technol., vol. 67, no. 10, pp. 10 160–10 165,Oct. 2018.

[23] F. Voss, C. Gloaguen, F. Fleischer, and V. Schmidt, “Distributionalproperties of Euclidean distances in wireless networks involving roadsystems,” IEEE J. Select. Areas Commun., vol. 27, no. 7, pp. 1047–1055, Sep. 2009.

[24] V. V. Chetlur, S. Guha, and H. S. Dhillon, “Characterization of V2Vcoverage in a network of roads modeled as Poisson line process,” inProc., IEEE Int. Conf. on Commun. (ICC), 2018, pp. 1–6.

[25] C. Choi and F. Baccelli, “An analytical framework for coverage incellular networks leveraging vehicles,” IEEE Trans. Commun., vol. 66,no. 10, pp. 4950–4964, Oct. 2018.

[26] S. Guha, “Cellular-assisted vehicular communications: A stochasticgeometric approach,” Master’s thesis, Virginia Tech, 2016.

[27] M. N. Sial, Y. Deng, J. Ahmed, A. Nallanathan, and M. Dohler,“Stochastic geometry modeling of cellular V2X communication onshared uplink channels,” arXiv preprint, 2018, available online:arxiv.org/abs/1804.08409.

[28] S. N. Chiu, D. Stoyan, W. S. Kendall, and J. Mecke, Stochastic geometryand its applications. John Wiley & Sons, 2013.

[29] H. S. Dhillon and J. G. Andrews, “Downlink rate distribution inheterogeneous cellular networks under generalized cell selection,” IEEEWireless Commun. Lett., vol. 3, no. 1, pp. 42–45, Feb. 2014.

[30] V. Shivaldova, A. Paier, D. Smely, and C. F. Mecklenbrauker, “Onroadside unit antenna measurements for vehicle-to-infrastructure com-munications,” in Proc., IEEE PIMRC, Sep. 2012, pp. 1295–1299.

[31] M. Haenggi, Stochastic Geometry for Wireless Networks. CambridgeUniversity Press, 2013.

[32] J. G. Andrews, A. K. Gupta, and H. S. Dhillon, “A primer on cellularnetwork analysis using stochastic geometry,” arXiv preprint, 2016,available online: arxiv.org/abs/1604.03183.

[33] P. D. Mankar, P. Parida, H. S. Dhillon, and M. Haenggi, “Distance fromthe nucleus to a uniformly random point in the typical and the Croftoncells of the Poisson-Voronoi tessellation,” arXiv preprint, 2019, availableonline: arxiv.org/abs/1907.03635.

[34] S. Singh, H. S. Dhillon, and J. G. Andrews, “Offloading in heterogeneousnetworks: Modeling, analysis, and design insights,” IEEE Trans. WirelessCommun., vol. 12, no. 5, pp. 2484–2497, May 2013.

[35] G. P. Patil, “Weighted distributions,” Wiley StatsRef: Statistics ReferenceOnline, 2014.

[36] V. V. Chetlur and H. S. Dhillon, “Matlab code for the computation ofcoverage probability and rate coverage in a C-V2X network,” 2019,available at: https://github.com/stochastic-geometry/Coverage-and-Rate-CV2X.

[37] F. Baccelli and X. Zhang, “A correlated shadowing model for urbanwireless networks,” in Proc., IEEE Conf. on Comput. Commun. (INFO-COM), 2015, pp. 801–809.

[38] V. V. Chetlur, H. S. Dhillon, and C. P. Dettmann, “Characterizing shortestpaths in road systems modeled as Manhattan Poisson line processes,”arXiv preprint, 2018, available online: arxiv.org/abs/1811.11332.

Vishnu Vardhan Chetlur received the B. E. (Hons.)degree in Electronics and Communications Engi-neering from the Birla Institute of Technology andScience (BITS) Pilani, India, in 2013. After his grad-uation, he worked as a design engineer at RedpineSignals Inc. for two years. He is currently a Ph.D.student at Virginia Tech, where his research interestsinclude wireless communication, smart cities, andstochastic geometry. He graduated top of his classin the department of Electrical Engineering at BITSand was awarded the institute Silver medal for being

ranked second in the whole institute. He was also a recipient of the Prattscholarship at Virginia Tech and BITS merit scholarship for his excellence inacademics.

Harpreet S. Dhillon (S’11–M’13–SM’19) receivedthe B.Tech. degree in electronics and communicationengineering from IIT Guwahati in 2008, the M.S. de-gree in electrical engineering from Virginia Tech in2010, and the Ph.D. degree in electrical engineeringfrom the University of Texas at Austin in 2013.

After serving as a Viterbi Postdoctoral Fellow atthe University of Southern California for a year, hejoined Virginia Tech in 2014, where he is currentlyan Associate Professor of electrical and computerengineering and the Elizabeth and James E. Turner

Jr. ’56 Faculty Fellow. His research interests include communication theory,wireless networks, stochastic geometry, and machine learning. He is a Clar-ivate Analytics Highly Cited Researcher and has coauthored five best paperaward recipients including the 2014 IEEE Leonard G. Abraham Prize, the2015 IEEE ComSoc Young Author Best Paper Award, and the 2016 IEEEHeinrich Hertz Award. He was named the 2017 Outstanding New AssistantProfessor, the 2018 Steven O. Lane Junior Faculty Fellow, and the 2018College of Engineering Faculty Fellow by Virginia Tech. His other academichonors include the 2008 Agilent Engineering and Technology Award, the UTAustin MCD Fellowship, and the 2013 UT Austin Wireless Networking andCommunications Group leadership award. He currently serves as an Editorfor the IEEE TRANSACTIONS ON WIRELESS COMMUNICATIONS, the IEEETRANSACTIONS ON GREEN COMMUNICATIONS AND NETWORKING, andthe IEEE WIRELESS COMMUNICATIONS LETTERS.