Skip to main content
Log in

Asymptotic throughput for large-scale wireless networks with general node density

  • Published:
Wireless Networks Aims and scope Submit manuscript

Abstract

We study the asymptotic throughput for a large-scale wireless ad hoc network consisting of n nodes under the generalized physical model. We directly compute the throughput of multicast sessions to unify the unicast and broadcast throughputs. We design two multicast schemes based on the so-called ordinary arterial road system and parallel arterial road system, respectively. Correspondingly, we derive the achievable multicast throughput by taking account of all possible cases of n s  = ω(1) and 1 ≤ n d  ≤ n − 1, rather than only the cases of \(n_s=\Uptheta(n)\) as in most related works, where n s and n d denote the number of sessions and the number of destinations of each session, respectively. Furthermore, we consider the network with a general node density \(\lambda \in [1,n]\), while the models in most related works are either random dense network (RDN) or random extended network (REN) where the density is λ = n and λ = 1, respectively, which further enhances the generality of this work. Particularly, for the special case of our results by letting λ = 1 and \(n_s=\Uptheta(n)\), we show that for some regimes of n d , the multicast throughputs achieved by our schemes are better than those derived by the well-known percolation-based schemes.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. We use the term θ(n):[θ1(n), θ2(n)] to represent that \(\theta(n)=\Upomega(\theta_1(n))\) and θ(n) = O2(n)), use θ(n):(θ1(n), θ2(n)] to represent that θ(n) = ω(θ1(n)) and θ(n) = O2(n)), and use θ(n):[θ1(n), θ2(n)) to represent that \(\theta(n)=\Upomega(\theta_1(n))\) and θ(n) = o2(n)).

References

  1. Gupta, P., & Kumar, P. R. (2000). The capacity of wireless networks. IEEE Transactions on Information Theory 46(2), 388–404.

    Article  MathSciNet  MATH  Google Scholar 

  2. Zheng, R. (2008). Asymptotic bounds of information dissemination in power-constrained wireless networks. IIEEE Transactions on Wireless communication, 7(1), 251–259. doi:10.1109/TWC.2008.060474.

    Article  Google Scholar 

  3. Xie, L., & Kumar, P. (2004). A network information theory for wireless communication: scaling laws and optimal operation. IEEE Transactions on Information Theory, 50(5), 748–767.

    Article  MathSciNet  Google Scholar 

  4. ÖzgÜr, A., LÉvÊque, O., & Tse, D. (2007). Hierarchical cooperation achieves optimal capacity scaling in Ad Hoc networks. IIEEE Transactions on Information Theory, 53(10), 3549–3572.

    Article  Google Scholar 

  5. Agarwal, A., & Kumar, P. R. (2004). Capacity bounds for Ad hoc and hybrid wireless networks. ACM SIGCOMM Computer Communication Review, 34(3), 71–81.

    Google Scholar 

  6. Li, X.-Y., Liu, Y., Li, S., & Tang, S. (2010). Multicast capacity of wireless Ad hoc networks under Gaussian channel model. IEEE/ACM Transactions on Networking, 18(4), 1145–1157.

    Google Scholar 

  7. Keshavarz-Haddad, A., & Riedi, R. (2008). Multicast capacity of large homogeneous multihop wireless networks. In Proceedings of the IEEE WiOpt.

  8. Franceschetti, M., Dousse, O., Tse, D., & Thiran, P. (2007). Closing the gap in the capacity of wireless networks via percolation theory. IIEEE Transactions on Information Theory, 53(3), 1009–1018.

    Article  MathSciNet  Google Scholar 

  9. Meester, R., & Roy, R. (1996). Continuum Percolation. Cambridge: Cambridge University Press.

    Book  MATH  Google Scholar 

  10. Keshavarz-Haddad, A., Ribeiro, V., & Riedi, R. (2006). Broadcast capacity in multihop wireless networks. In Proceedings of the ACM MobiCom.

  11. Shakkottai, X., Liu, S., & Srikant, R. (2007). The multicast capacity of large multihop wireless networks. In Proceedings of the ACM MobiHoc.

  12. Keshavarz-Haddad, A., & Riedi, R. (2007). Bounds for the capacity of wireless multihop networks imposed by topology and demand. In Proceedings of the MobiHoc.

  13. Hu, C., Wang, X., Nie, D., & Zhao, J. (2009). Multicast scaling laws with hierarchical cooperation. In Proceedings of the IEEE INFOCOM.

  14. Moscibroda, T. (2007). The worst-case capacity of wireless sensor networks. In Proceedings of the ACM/IEEE IPSN.

  15. Huang, W., Wang, X., & Zhang, Q. (2010). Capacity scaling in mobile wireless ad hoc network with infrastructure support. In to appear in Proceedings of the IEEE ICDCS.

  16. Zheng, R. (2006). Information dissemination in power-constrained wireless networks. In Proceedings of the IEEE INFOCOM.

  17. Liu, B., Thiran, P., & Towsley, D. (2007). Capacity of a wireless ad hoc network with infrastructure. In Proceedings of the ACM Mobihoc.

  18. Dousse, O., & Thiran, P. (2004). Connectivity vs capacity in dense ad hoc networks. In Proceedings of the IEEE INFOCOM.

  19. Penrose, M. (1997). The longest edge of the random minimal spanning tree. Annals of Applied Probability, 7 , 340–361.

    Article  MathSciNet  MATH  Google Scholar 

  20. Wang, C., Li, X.-Y., Tang, S., & Jiang, C. (2010). Multicast capacity scaling for cognitive networks: General extended primary network. In Proceedings of the IEEE MASS, pp. 262–271.

  21. Chau, C., Chen, M., & Liew, S. (2009). Capacity of large-scale csma wireless networks. In Proceedings of the ACM MobiCom.

  22. Li, X.-Y. (2009). Multicast capacity of wireless ad hoc networks. IEEE/ACM Transactions on Networking, 17(3) , 950–961.

    Article  Google Scholar 

  23. Grimmett, G. (1999). Percolation. Berlin: Springer.

    MATH  Google Scholar 

  24. Wang, C., Li, X.-Y., Jiang, C., Tang, S., & Liu, Y., Multicast throughput for hybrid wireless networks under gaussian channel model. IEEE Transactions on Mobile Computing (PrePrints), 10(6), 839–852.

  25. Li., X.-Y., Tang, S., & Ophir, F. (2007). Multicast capacity for large scale wireless ad hoc networks, In Proceedings of the ACM MobiCom.

  26. Xie, L., & Kumar, P. (2006). On the path-loss attenuation regime for positive cost and linear scaling of transport capacity in wireless networks. IEEE/ACM Transactions on Networking, 14, 2313–2328.

    Article  MathSciNet  Google Scholar 

  27. Grossglauser, M., & Tse, D. (2002). Mobility increases the capacity of ad hoc wireless networks. IEEE/ACM Transactions on Networking, 10 (4), 477–486.

    Article  Google Scholar 

  28. Sharma, G., Mazumdar, R., & Shroff, N. (2007). Delay and capacity trade-offs in mobile ad hoc networks: A global perspective. IEEE/ACM Transactions on Networking, 15(5), 981–992.

    Article  Google Scholar 

  29. Garetto, M., Giaccone, P., & Leonardi E. (2009). Capacity scaling in ad hoc networks with heterogeneous mobile nodes: the super-critical regime. IEEE/ACM Transactions on Networking (TON), 17(5), 1522–1535.

    Article  Google Scholar 

  30. Garetto, M., Giaccone, P., & Leonardi, E. (2009). Capacity scaling in ad hoc networks with heterogeneous mobile nodes: the subcritical regime. IEEE/ACM Transactions on Networking, 17(6), 1888–1901.

    Article  Google Scholar 

  31. Tavli, B. (2006). Broadcast capacity of wireless networks. IEEE Communications Letters, 10(2), 68–69.

    Article  Google Scholar 

  32. Keshavarz-Haddad, A., & Riedi, R. (2007). On the broadcast capacity of multihop wireless networks: Interplay of power, density and interference, In Proceedings of the IEEE SECON ’07, pp. 314–323.

  33. Jacquet, P., & Rodolakis, G. (2005). Multicast scaling properties in massively dense ad hoc networks. In: Proceedings of the 11th international conference on parrallel and distribution systems - workshops, pp. 93–99.

  34. Mitzenmacher, M., & Upfal, E. (2005). Probability and computing: Randomized algorithms and probabilistic analysis. Cambridge, UK: Cambridge University Press.

    Book  MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the anonymous reviewers for their constructive comments. The research of authors is partially supported by the National Basic Research Program of China (973 Program) under grants No. 2010CB328101 and No. 2010CB334707, the Program for Changjiang Scholars and Innovative Research Team in University, the Shanghai Key Basic Research Project under grant No. 10DJ1400300, the Expo Science and Technology Specific Projects of China under grant No. 2009BAK43B37, the NSF CNS-0832120 and CNS-1035894, the National Natural Science Foundation of China under grant No. 60828003, the Program for Zhejiang Provincial Key Innovative Research Team, and the Program for Zhejiang Provincial Overseas High-Level Talents (One-hundred Talents Program).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Cheng Wang.

Proofs of some lemmas and theorems

Proofs of some lemmas and theorems

1.1 Proof of Lemma 4

We first recall a useful lemma about the tails of Chernoff bound from [34].

Lemma 14

([34]) Let X be a Poisson random variable with parameter μ. Then

$$ \Pr(X \ge x) \le { e^{- \mu }{\cdot}(e{\cdot}\mu)^x} /{x^x}, \quad{\rm for}\, x > \mu \\ $$
(13)
$$ \Pr(X \leq x) \le { e^{- \mu }{\cdot}(e{\cdot}\mu)^x} /{x^x},\quad{\rm for}\, 0<x < \mu $$
(14)

Next, we begin to prove Lemma 4.

Proof

Define the maximum and minimum number of nodes for all \({\frac{n}{\lambda {\cdot} S }}\) subregions as \(\overline{\mu}\) and \(\underline{\mu}\), respectively.

  • Case 1 : When \(S {\cdot} \lambda =\Upomega(1)\) and \(S {\cdot} \lambda =o(\log n)\).

  • According to Eq. (14) and union bound, it follows that

    $$ {\Pr}(\overline{\mu}\geq \Updelta_1 {\cdot} \log n) \leq {\frac{n}{\lambda {\cdot} S }} {\cdot} \left( {\frac{e^{1-{\frac{S \lambda}{\Updelta_1 \log n}}}{\cdot} S \lambda}{\Updelta_1 \log n}}\right)^{\Updelta_1 \log n} \to 0, $$

    for \(\Updelta_1\) in Table 2, as \(n \to \infty\), which proves the upper bounds. The lower bounds is straightforward.

  • Case 2 : When \(S {\cdot} \lambda =\omega( \log n)\).

  • According to Eq. (14) and union bound, it follows that

    $$ {\Pr}(\underline {\mu}\leq (1-\Updelta_2) {\cdot} S \lambda) \leq {\frac{n}{\lambda {\cdot} S }} {\cdot} \left( {\frac{e^{-\Updelta_2}}{(1-\Updelta_2)^{(1-\Updelta_2)}}}\right)^{S \lambda} \to 0, $$

    for \(\Updelta_2\) in Table 2, which proves the lower bounds. Similarly,

    $$ {\Pr}(\overline{\mu}\geq (1+\Updelta_3) {\cdot} S \lambda) \leq {\frac{n}{\lambda {\cdot} S }} {\cdot} \left( {\frac{e^{\Updelta_3}}{(1+\Updelta_3)^{(1+\Updelta_3)}}}\right)^{S \lambda} \to 0, $$

    for \(\Updelta_3\) in Table 2, as \(n \to \infty\), which proves the upper bounds.

  • Case 3 : When \(S{\cdot}\lambda : z {\cdot} \log n \), for any \(z\in (0, \infty)\).

  • Similarly, we have

    $$ {\Pr}(\underline {\mu}\leq (1-\Updelta_4) {\cdot} z {\cdot} \log n) \leq {\frac{n}{\lambda {\cdot} S }} {\cdot} \left( {\frac{e^{-\Updelta_4}}{(1-\Updelta_4)^{(1-\Updelta_4)}}}\right)^{z {\cdot} \log n} \to 0, $$

    for \(\Updelta_4\) in Table.2, which proves the lower bounds; and

    $$ {\Pr}(\overline{\mu}\geq (1+\Updelta_5) {\cdot} z {\cdot} \log n) \leq {\frac{n}{\lambda {\cdot} S }} {\cdot} \left( {\frac{e^{\Updelta_5}}{(1+\Updelta_5)^{(1+\Updelta_5)}}}\right)^{z {\cdot} \log n} \to 0, $$

    for \(\Updelta_5\) in Table.2, which proves the upper bounds.

1.2 Proof of Lemma 8

To simplify the description, we denote \({v_{\mathcal{S},k}}\), \({\mathcal{M}_{\mathcal{S},k}}\), \({\mathcal{U}_{\mathcal{S},k}}\), and \({\mathcal{D}_{\mathcal{S},k}}\) by v k , \(\mathcal{M}_{ k}\), \(\mathcal{U}_{ k}\), and \(\mathcal{D}_{ k}\), respectively, without confusion. Furthermore, we denote the set of all links in \(\hbox{EST}(\mathcal{U}_{ k})\) as \(\Uppi_k\).

Given a cell c * t , we define the number of multicast sessions that are routed through the station inside c * t as a random variable Z t , and finally we consider the uniform upper bound of Z t for every cell, denoted as Z. Define an event B(kt): Multicast session \(\mathcal{M}_{k}\) passes through the cell c * t . For any link \(v_iv_j \in \hbox{EST}(U_k)\), define an event B h ij (kt): \(\mathcal{L}^h_{ij}\) passes through c * t ; and define an event B v ij (kt): \(\mathcal{L}^v_{ij}\) passes through c * t . Then,

$$ {\Pr}(B(k,t) )\le \sum\nolimits_{e_{ij} \in \Uppi_k} (\Pr(B_{ij}^h (k,t)) + \Pr(B_{ij}^v (k,t)) ) $$

Based on c * t , we construct the region \(\mathcal{D}^h_{ij}(k,t)\) as illustrated in Fig. 6, then the following proposition is true.

Fig. 6
figure 6

Construction of the region \(\mathcal{D}^h_{ij}(k,t)\). Here \(|{\cdot}|\) represents the Euclidean length of a line segment or the Euclidean distance between two nodes

Proposition 3

The Poisson node v i is located in the region \(\mathcal{D}^h_{ij}(k,t)\) if the event B h ij (kt)happens.

Similarly, we can construct the region \(\mathcal{D}^v_{ij}(k,t)\), and we have

Proposition 4

The Poisson node v j is located in the region \(\mathcal{D}^v_{ij}(k,t)\) if the event B v ij (kt)happens.

Define the number of nodes in the region of area D(kt) as a random variable ϑ t , where \(D(k,t) = \min\{n/\lambda, \tilde{D}(k,t)\}\), and

$$ \tilde{D}(k,t) = \sum\limits_{e_{ij} \in \Uppi_k}(\|{\mathcal{D}}^h_{ij}(k,t)\|+\|{\mathcal{D}}^v_{ij}(k,t)\|) \le 2\left( {\frac{n_d {\cdot} 8\log n}{\lambda}}+ 2\sqrt {{\frac{2n_d n}{\lambda}}}{\cdot} 2\sqrt{{\frac{2\log n}{\lambda}}}\right) $$

Hence, \( D(k,t)\leq D= \min \{{\frac{16}{\lambda}}{\cdot}(n_d {\cdot}\log n + \sqrt{n_d{\cdot} \log n {\cdot} n}, \,{\frac{n}{\lambda}}\}. \) Let \(\Upgamma_o:=D {\cdot} \lambda\). Then, Obviously, ϑ t follows a Poisson distribution with the mean of at most \(\mu_o = {\frac{n_s }{n}} {\cdot} \Upgamma_o\). From Proposition 3 and Proposition 4, we have that Z t  ≤ ϑ t .

By Lemma 4, according to different cases of \({\frac{n_s }{n}} {\cdot} \Upgamma_o\), we prove this lemma.

1.3 Proof of Lemma 12

From Lemma 10, we get that the rate along the arterial roads can be achieved of order

$$ {\rm R}(n)=\left\{ \begin{array}{lll} \Upomega\left({\frac{\lambda^{{\frac{\alpha}{2}}}}{(\log n)^{{\frac{\alpha}{2}}}}}\right) & {when} & \lambda :[1, (\log n)^{1-{\frac{2}{\alpha}}}] \\ \Upomega\left({\frac{1}{\log n}} \right) & {when} & \lambda :[(\log n)^{1-{\frac{2}{\alpha}}}, n] \\ \end{array}\right.. $$

Thus, we only need prove that for every station the maximum relay burden is w.h.p. of order

$$ {\rm L}(n)=\left\{ \begin{array}{ll} O\left({\frac{n_s{\cdot}\Upgamma_{p}}{n}}\right) & {\rm when}\,n_s{\cdot} \Upgamma_{p} = \Upomega(n\log n) \\ O(\log n) & {\rm when}\,n_s{\cdot} \Upgamma_{p} = O(n\log n)\\ \end{array}\right. $$
(15)

where \(\Upgamma_{p}=\left\{ \begin{array}{ll} \Uptheta({\frac{\sqrt{n n_d}}{\sqrt{\log n}}}) & {when\,} n_d=O({\frac{n}{\log n}})\\ \Uptheta(n_d)& {when\,} n_d=\Upomega({\frac{n}{\log n}}).\\ \end{array}\right.\)

Given a station v * t passed by a horizontal arterial road \({\mathbf h}\) or a vertical arterial road \({\mathbf v}\), we define the number of multicast sessions routed through v * t during Phase 2, 3 and 4 as a random variable X t . Note that we finally need the uniform upper bound of X t for all stations, denoted by X.

Define an event A(kt): During Phase 2, 3 and 4, multicast session \(\mathcal{M}_k\) passes through v * t . Furthermore, for each edge \(e_{ij} \in \Uppi_k\), define the event A h ij (kt) (or A v ij (kt)): During Phase 2 (or Phase 4), \(\mathcal{M}_k\) passes through v * t . Hence, we have

$$ {\Pr}(A(k,t) )\le \sum\nolimits_{e_{ij} \in \Uppi_k} (\Pr(A_{ij}^h (k,t) ) + \Pr(A_{ij}^v (k,t)) ) $$

For each link \(v_iv_j \in \Uppi_k\), construct the region \(\mathcal{S}^h_{ij}(k,t)\) as in Fig. 7, where \(|{\cdot}|_{\rm h}\) and \(|{\cdot}|_{\rm v}\) represent the horizontal and vertical Euclidean distance, respectively, that are defined as: In the 2-dimension plane, for any two nodes u 1 and u 2 with the coordinates (x 1, y 1) and (x 2, y 2) respectively, |u 1 u 2|h = |x 1 − x 2| and |u 1 u 2|v = |y 1 − y 2|. Therefore,

Fig. 7
figure 7

Construction of the region \(\mathcal{S}^h_{ij}(k,t)\). Here \((\Uppsi^h)^{-1}({\mathbf h})\) is the horizontal slice that corresponds to the arterial road \({\mathbf h}\). \(\mathcal{S}^h_{ij}(k,t)\) is the shadowed region of sides l × 2L h ij , where \( l ={\frac{\sqrt{2}}{\sqrt{\lambda {\cdot} \log n}}}\) and \(L_{ij}^h = | v_iv_j |_{\rm h} + 2\sqrt{2\log n /\lambda}\)

Proposition 5

If A h ij (kt) happens, then the Poisson point v i is located in the region \(\mathcal{S}^h_{ij}(k,t)\).

By a similar method, based on the arterial road \({\mathbf v}\) and the vertical slice \((\Uppsi^v)^{-1}({\mathbf v})\), we construct the region \(\mathcal{S}^v_{ij}(k,t)\) of height l and width 2L v ij , where

$$ L_{ij}^v = | v_iv_j |_{\rm v} + 2\sqrt{2\log n /\lambda}. $$

Hence, we have the following result.

Proposition 6

If A v ij (kt) happens, then the Poisson point v j is located in the region \(\mathcal{S}^v_{ij}(k,t)\).

Define Y t as the number of Poisson nodes in the region of area S(kt) according to a p.p.p of density \({\frac{n_s}{n}}{\cdot} \lambda\), where

$$ S(k,t) = \min\{n, \tilde{S}(k,t)\}, \, \tilde{S}(k,t) = \sum \nolimits_{e_{ij} \in \Uppi_k}(\|{\mathcal{S}}^h_{ij}(k,t)\|+\|{\mathcal{S}}^v_{ij}(k,t)\|)\} $$

Then, Y t follows Poisson distribution with \(\mu_p = {\frac{n_s}{n}}{\cdot} \lambda {\cdot} S(k,t)\). According to Proposition 5 and Proposition 6, we can obtain that X t  ≤ Y t . Then, we have

$$ \tilde{S}(k,t) = 2\sum\nolimits_{e_{ij} \in \Uppi_k}( | v_iv_j |_{\rm h} + | v_iv_j |_{\rm v} + 2\sqrt{2\log n /\lambda}){\cdot}{l} $$

Since \( | v_iv_j |_{\rm h} + | v_iv_j |_{\rm v} \le \sqrt 2 | {v_i v_j } | \) and by Lemma 6,

$$ \tilde{S}(k,t) \leq {\frac{2\sqrt{2}}{\sqrt{\lambda\log n}}} {\cdot} \sum\nolimits_{e_{ij} \in \Uppi_k}( \sqrt{2}{\cdot}| v_iv_j |+ 2\sqrt{{\frac{2\log n}{\lambda}}})\leq {\frac{8}{\lambda}} {\cdot} n_d +{\frac{8\sqrt{2}}{\lambda}}{\cdot} \sqrt {{\frac{n_d n}{\log n}}} $$

Then, we have

$$ S(k,t)\leq {\frac{8}{\lambda}}{\cdot} (n_d + \sqrt {{\frac{2 {\cdot} n_d {\cdot} n}{\log n}}}):= {\frac{8}{\lambda}}{\cdot} \Upgamma_{p} $$

Then, \(\mu_p = {\frac{n_s}{n}}{\cdot} \lambda {\cdot} S(k,t)\leq {\frac{n_s}{n}}{\cdot} 8{\cdot} \Upgamma_{p}\).

From Lemma 4, according to different cases of \({\frac{n_s}{n}}{\cdot} 8{\cdot} \Upgamma_{p}\), we get the Eq. (15). Hence, combining Lemma 10, we prove this lemma.

1.4 Proof of Lemma 13

For any given multicast session \(\mathcal{M}_k\), denote the set of leaf nodes of \(\hbox{EST}(\mathcal{U}_k)\) by \(\mathcal{\tilde{U}}_k\). Then, in Phase 1, only the nodes in the set \(\mathcal{U}_k-\mathcal{\tilde{U}}_k\) intend to access into the corresponding arterial roads via a single hop; while in Phase 5, only the nodes in the set \(\mathcal{U}_k-\{v_k\}\) need to receive the data from the corresponding arterial roads via a single hop. Since it is not always true that \(|\mathcal{U}_k-\mathcal{\tilde{U}}_k|=|\mathcal{U}_k-\{v_k\}|\), where \(|{\cdot}|\) represents the cardinality of a discrete set, we learn the fact that unlike the unicast case, Phase 1 is not symmetrical with Phase 5 due to the tree structure of each EST. However, for \(|\mathcal{U}_k-\mathcal{\tilde{U}}_k|\leq|\mathcal{U}_k-\{v_k\}|\), the load in Phase 5 is no more than that in Phase 1. According to Lemma 11, by parallel transmission scheduling, each cell can sustain a total rate of order

$$ \left\{ \begin{array}{lll} \Upomega\left({\frac{\lambda^{{\frac{\alpha}{2}}}}{(\log n)^{{\frac{\alpha}{2}}-1}}}\right) & {when} & \lambda :[1, (\log n)^{1-{\frac{2}{\alpha}}}] \\ \Upomega(1) & {when} & \lambda :[(\log n)^{1-{\frac{2}{\alpha}}}, n] \\ \end{array}\right.. $$
(16)

By bottleneck-principle, we only need to examine the throughput in Phase 5 out of those two phases. In Phase 5, we only need to prove that the bandwidth as in Eq. (16) bears the load of order

$$ \left\{ \begin{array}{ll} O({\frac{n_s {\cdot} n_d {\cdot} \log n}{n}}) & {\rm when}\,n_s {\cdot} n_d = \Upomega(n) \\ O(\log n) & {\rm when}\,n_s {\cdot} n_d = O(n)\\ \end{array}\right. $$
(17)

Similar to the proof of Lemma 12, since

$$ |\cup_{v_k\in {\mathcal{S}}}{({\mathcal{U}}_k-\{v_k\})}|\leq n_s{\cdot} (n_d-1), $$

an upper bound of the load of each group of \(\Uptheta(\log n)\) arterial roads, with a total rate of \(\Upomega((\log n)^{1-{\frac{\alpha}{2}}})\), follows a Poisson distribution of the mean

$$ \mu'_p= {\frac{n_s}{n}}{\cdot} \lambda {\cdot} (n_d-1)(2\sqrt{2\log n/\lambda})^2 \leq {\frac{8 {\cdot} n_s {\cdot} n_d {\cdot} \log n}{n}}, $$

where the area of each cell is \((2\sqrt{2\log n/\lambda})^2\). Hence, according to Lemma 4, we prove that Eq. (17). Then, we can complete the proof.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wang, C., Jiang, C., Li, XY. et al. Asymptotic throughput for large-scale wireless networks with general node density. Wireless Netw 19, 559–575 (2013). https://doi.org/10.1007/s11276-012-0485-5

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11276-012-0485-5

Keywords

Navigation