Skip to main content
Log in

Multicast throughput for large scale cognitive networks

  • Published:
Wireless Networks Aims and scope Submit manuscript

Abstract

In this paper, we focus on the achievable throughput of cognitive networks consisting of the primary ad hoc network (PaN) and the secondary ad hoc network (SaN). We construct PaN and SaN by placing nodes according to Poisson point processes of density n and m respectively over a unit square region. We directly study the multicast throughput of cognitive network to unify that of unicast and broadcast sessions. In order to ensure the priority of primary users in meanings of throughput, we design a metric called throughput decrement ratio (TDR) to measure the ratio of the throughput of PaN in presence of SaN to that of PaN in absence of SaN. Endowing PaN with the right to determine the threshold of the TDR, we propose multicast schemes based on TDMA and multihop routing for the two networks respectively and derive their achievable multicast throughput depending on the given threshold. Specially, we show when PaN has sparser density than SaN, to be specific, \(n=o\left({\frac{m} {(\log m)^2}}\right),\) and if PaN only cares about the order of its throughput, SaN can simultaneously achieve the same order of the aggregated multicast throughput as it were a stand-alone network in absence of PaN.

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

Similar content being viewed by others

Notes

  1. Knuth’s Notation: The following notations are used throughout this paper. Given non-negative functions f (n) and g(n): f (n) = O(g(n)) means there exist positive constants c and n 0 such that f (n) ≤ cg(n) for all n ≥ n 0; f (n) = Ω(g(n)) means there exist positive constants c and n 0 such that f (n) ≥ cg(n) for all nn 0, namely, g(n) = O(f (n)); f (n) = Θ(g(n)) means that both f (n) = Ω(g(n)) and f (n) = O(g(n)) hold; f (n) = o(g(n)) means that \(\lim\limits_{n\to\infty} {\frac{f (n)} {g(n)}}= 0;\) f (n) = ω(g(n)) means that \(\lim\limits_{n\to\infty} {\frac{g(n)} {f (n)}}= 0,\) namely, f (n) = o(g(n)).

References

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

    Article  MATH  MathSciNet  Google Scholar 

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

  3. Xiang-Yang, L., Tang, S., & Ophir, F. (2007). Multicast capacity for large scale wireless ad hoc networks. In Proceedings of ACM MobiCom.

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

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

    Article  MathSciNet  Google Scholar 

  6. Zheng, R. (2008). Asymptotic bounds of information dissemination in power-constrained wireless networks. IEEE Transactions on Wireless Communications, 7(1), 251–259.

    Article  Google Scholar 

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

    Article  Google Scholar 

  8. Li, S., Liu, Y., & Li, X.-Y. (2008). Capacity of large scale wireless networks under gaussian channel model. In Proceedings of ACM MobiCom.

  9. Wang, C., Li, X., Jiang, C., Tang, S., & Liu, Y. (2009). Scaling laws on multicast capacity of large scale wireless networks. In Proceedings of IEEE INFOCOM.

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

  11. Akyildiz, I. F., Lee, W.-Y., Vuran, M. C., & Mohanty, S. (2006). Next generation/dynamic spectrum access/cognitive radio wireless networks: A survey. Elsevier Computer Networks, 50, 2127–2159.

    Article  MATH  Google Scholar 

  12. Federal Communications Commission Spectrum Policy Task Force. (2002). Report of the spectrum efficiency working group, technical report, FCC.

  13. Devroye, N., Mitran, P., & Tarokh, V. (2006). Achievable rates in cognitive radio channels. IEEE Transactions on Information Theory, 52(5), 1813–1827.

    Article  MathSciNet  Google Scholar 

  14. Mitola, J. (2000). Cognitive radio, Ph.D. thesis, Royal Institute of Technology (KTH).

  15. Jafar, S., & Srinivasa, S. (2007). Capacity limits of cognitive radio with distributed and dynamic spectral activity. IEEE Transactions on Computer, 25(5), 529–537.

    Google Scholar 

  16. Vu, M., & Tarokh, V. (to appear). Scaling laws of single-hop cognitive networks. IEEE Transactions on Wireless Communications.

  17. Jeon, S.-W., Devroye, N., Vu, M., Chung, S.-Y., & Tarokh, V. (2009). Cognitive networks achieve throughput scaling of a homogeneous network. In Proceedings of IEEE WiOpt.

  18. Liu, B., Liu, Z., & Towsley, D. (2003). On the capacity of hybrid wireless networks. In Proceedings of IEEE INFOCOM.

  19. Mao, X., Li, X.-Y., Tang, S. (2008) Multicast capacity for hybrid wireless networks. In Proceedings of ACM MobiHoc.

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

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

    Article  Google Scholar 

  22. Wang, C., Jiang, C., Li, X.-Y., Tang, S., & Tang, X. (2009). Achievable multicast throughput for homogeneous wireless ad hoc networks. In Proceedings of IEEE WCNC.

  23. Vu, M., Devroye, N. (to appear). On the primary exclusive regions in cognitive networks. IEEE Transactions on Wireless Communications.

  24. Meester, R., & Roy, R. (1996). Continuum percolation. Cambridge: Cambridge University Press.

    MATH  Google Scholar 

  25. Grossglauser, M., & Thiran, P. (2005). Networks out of control: Models and methods for random networks, technical report. School of Computer and Communication Sciences (EPFL).

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

  27. Wang, C., Tang, S., Li, X.-Y., Jiang, C., & Liu, Y. (2009). Multicast throughput of hybrid wireless networks under Gaussian channel model. In Proceedings of IEEE ICDCS.

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

    Article  Google Scholar 

  29. 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 (TON), 15(5), 992–1103.

    Article  Google Scholar 

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

Download references

Acknowledgements

The research of authors are partially supported by NSF CNS-0832120, National Natural Science Foundation of China under No. 90718012, No. 90818023, No. 60828003, the National High Technology Research and Development Program of China (863 Program) under Grants No. 2007AA01Z180, No. 2007AA01Z136, No. 2007AA01Z149, Shanghai International Cooperation Project under Grant No. 075107005, the Natural Science Foundation of Zhejiang Province under Grant No. Z1080979, National Basic Research Program of China (973 Program) under grant No. 2010CB328100, No. 2006CB30300, Hong Kong RGC HKUST 6169/07, the RGC under Grant HKBU 2104/06E, and CERG under Grant PolyU-5232/07E.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Cheng Wang.

Appendix A

Appendix A

1.1 Proofs for lemmas

Proof

Lemma 8: Consider any link \(\langle v^p_i,v^p_j\rangle\in \Upxi_p(\digamma,\tau)\) for any \(\tau \in {\mathcal{T}},\) where vp i and vp j locate on two adjacent cells respectively.

First, we bound I pp (v p i v p j ; τ). We notice that the transmitters in the eight closest cells are located at distance at least \((K-2)\sqrt{a_p}\) (i.e., \(\sqrt{a_p}\) in Fig. 1) from the receiver. The 16 next closest cells are of distance at least \((2K-2)\sqrt{a_p}\) (i.e., \(4\sqrt{a_p}\) in Fig. 1). By extending the sum of the interferences to the whole region, it can be upper bounded as follows:

$$ I_{pp}(v^p_i,v^p_j;\tau) \le \sum\limits_{i = 1}^{\lceil1/{a_p}\rceil} {8iP({a_p})^{\alpha/2} \ell((Ki - 2)\sqrt {a_p} )} \leq P\sum\limits_{i = 1}^n {8i(Ki - 2)^{ - \alpha }}. $$

For K ≥ 3 and α > 2, and by the Cauchy Test, we have

$$ \lim_{n\to \infty}I_{pp}(v^p_i,v^p_j;\tau )\leq 8P \int_{1}^{+\infty}{{\frac{x} {(Kx-2)^{\alpha}}}}dx = 8P\cdot C_1(K,\alpha), $$

where the constant C 1(K, α) (depending on K and α) that is defined in Eq. (16). It is clear that \(\lim\limits_{n\to \infty}{I_{pp}(v^p_i,v^p_j;\tau)}\) gets the minimum value when v p j locates in the cell on the corner. Thus, we obtain, for any \(\langle v^p_i,v^p_j\rangle{\in}\Upxi_p(\digamma,\tau),\) it holds that

$$ \lim_{n\to \infty}{I_{pp}(v^p_i,v^p_j;\tau)}\geq {\frac{1} {4}}\cdot8P\cdot C_1(K,\alpha)=2P\cdot C_1(K,\alpha). $$
(18)

Second, we upper bound I sp (v p i , v p j ; τ). The preservation region centered on v p j consists of K 2 secondary cells, hence, for any slot τ, there must exist one cell out of the K 2 cells that were to be scheduled in τ except it locates in the preservation region, we denote that cell as c τ. We consider those cells containing the nodes in \({\mathcal{V}}_s(v^p_i,\tau)\). It can be seen that the secondary users in the eight closest cells to c τ are far away from v p j with distance at least of \(\sqrt{a_s}(K-1)/2\). The secondary users in the 16 next closest cells to c τ are at Euclidean distance at least \(\sqrt{a_s}(K+(K-1)/2)\) between v p j . Thus, the sum of the interferences can be upper-bounded as follows:

$$ \begin{aligned} I_{sp}(v^p_i,v^p_j;\tau) \le & \sum\limits_{i = 1}^{\lceil1/{a_s}\rceil} {8i\cdot P\cdot\theta(n)\cdot {a_s}^{{\frac{\alpha} {2}}} \cdot \ell\left(\left(K(i-1)+{\frac{K-1} {2}}\right)\sqrt {a_s} \right)} \\ \leq\;2P\cdot\theta(n)\cdot\sum\limits_{i = 1}^m {4i\left(Ki - {\frac{K+1} {2}}\right)^{ - \alpha }}. \\ \end{aligned} $$

Recall that \(n=o\left({\frac{m} {(\log m)^2}}\right),\) we have

$$ \lim\limits_{n\to \infty} \sum\limits_{i = 1}^m {4i\left(Ki - {\frac{K+1} {2}}\right)^{ - \alpha}} \leq \int_{1}^{+\infty}{{\frac{4x} {(Kx-{\frac{k+1} {2}})^{\alpha}}}}d x. $$

The right integral can be calculated to be a constant C 2(K, α) depending on K and α that is described by Eq. (16). Hence, we have

$$ \lim_{n\to \infty} I_{sp}(v^p_i,v^p_j;\tau)\le 2P \cdot\theta\cdot C_2(K, \alpha). $$
(19)

Third, we lower bound the signal received from the transmitter S(v p i v p j ;τ). Since the node only transmits to a destination located in the adjacent cell, i.e., \( |v^p_i-v^p_j|\leq\sqrt {5a_p}.\) Thus, the strength of the signal is bounded as:

$$ S(v^p_i,v^p_j;\tau) \ge P \cdot ({a_p})^{\alpha/2} \cdot (\sqrt{5a_p})^{-\alpha }=5^{ -{\frac{\alpha} {2}}}\cdot P. $$
(20)

Finally, taking the limit of the SINR, we can obtain

$$ R_p(v^p_i,v^p_j;\tau) \geq \log \left(1 + {\frac{5^{ -{\frac{\alpha} { 2}}}} {N_0/P + 8C_1(K,\alpha)+8\theta(n)C_2(K,\alpha)}}\right). $$

By Eqs. (18) and (19), \( \Updelta(\digamma, K, P)\leq{\theta}\cdot{\frac{ C_2(K,\alpha)} {C_1(K,\alpha)}} \). Let \(\lim\limits_{n\to \infty}\theta(n)=\theta \leq \Updelta \cdot {\frac{C_1(K,\alpha)} { C_2(K,\alpha)}},\)we can get \(\Updelta(\digamma, K, P){\leq}\triangle\). Hence,

$$ R_p(v^p_i,v^p_j;\tau) \geq \log \left(1 + {\frac{5^-{\frac{\alpha}{2}}} {N_0/P + 8(1+\Updelta)C_1(K,\alpha)}} \right). $$

Proof

Lemma 9 Given a primary cell c t *, we define the number of multicast sessions (flows) that are routed through the nodes inside c t * as a random variable X t , and we finally consider the uniform upper bound of X t , denoted by X, for every primary cell.

Define event A(kt): Multicast session \({\mathcal{M}}_{k}\) passes through the cell c t *. For any link \(v_iv_j \in \hbox{EST}({\mathcal{U}}_k)\), Define event A h ij (kt): \({\mathcal{HP}}_{ij}\) passes through c t *; and define event A v ij (kt): \({\mathcal{VP}}_{ij}\) passes through c t *. Obviously, we have

$$ {\Pr}(A(k,t) )=\Pr\left(\bigcup\limits_{e_{ij} \in \Pi_k} (A^h_{ij}(k,t)\cup A^v_{ij}(k,t) )\right). $$

By union bounds, it holds that

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

First, we give the bound of \(\Pr(A(k,t))\). Based on the cell c t *, we construct the region \({\mathcal{S}}^h_{ij}(k,t)\) of area \(S^h_{ij}(k,t)=\sqrt{a_p}(2 | {{\mathcal{HP}}_{ij} } | + \sqrt {a_p}),\) as in Fig. 3. Similarly, we construct the region \({\mathcal{S}}^v_{ij}(k,t)\) area of \(S^v_{ij}(k,t)=\sqrt{a_p}(2 | {{\mathcal{VP}}_{ij} } | + \sqrt {a_p})\). The following two propositions obviously hold.

Fig. 3
figure 3

The construction of the region \({\mathcal{S}}^h_{ij}(k,t)\). Here |·| represents the Euclid length of a line segment or the Euclid distance between two nodes

Proposition 1

The Poisson pointv i locates in the region\({\mathcal{S}}^h_{ij}(k,t)\)if the eventA h ij (kt) happens.

Proposition 2

The Poisson pointv j locates in the region\({\mathcal{S}}^v_{ij}(k,t)\)if the eventA v ij (kt) happens.

Define event B h ij (kt) (or B v ij (kt)): A Poisson node locates in a region of area S h ij (kt) (or S v ij (kt)). Then, by Propositions 1 and 2, we have

$$ \Pr(A_{ij}^h (k,t))\leq \Pr(B_{ij}^h (k,t)), \Pr(A_{ij}^v (k,t))\leq \Pr(B_{ij}^v (k,t)). $$

Moreover, define event B(kt): A Poisson node locates in a region of area \( S(k,t)=\min\{1, \tilde{S}(k,t) \},\)where

$$ \tilde{S}(k,t)=\sum \limits_{e_{ij} \in \Pi_k}(S^h_{ij}(k,t)+ S^v_{ij}(k,t)). $$

Hence, \({\Pr}(A(k,t) ){\leq}{\Pr}(B(k,t) )\).

For A(kt) and B(kt), define their indicator variables as \({\mathcal{I}}(A(K,t))\) and \({\mathcal{I}}(B(K,t)),\) where \({\mathcal{I}}(A)\) takes value 1 if event A happens, otherwise 0. Hence, \(X_t=\sum\nolimits_{k=1}^{n_s}{\mathcal{I}}(A(K,t))\). Define \(Y_t=\sum\nolimits_{k=1}^{n_s}{\mathcal{I}}(B(K,t)),\) then it represents the number of nodes in the region of area S(kt) according to a p.p.p of density n s . So it follows a Poisson distribution of mean λ = n s S(kt).

Next, we consider the upper bound of S(kt). Recall that

$$ \tilde{S}(k,t) =\sum\limits_{e_{ij} \in \Pi_k}(2| {{\mathcal{HP}}_{ij} } | + \sqrt {a_p}+ 2| {{\mathcal{VP}}_{ij} } | + \sqrt {a_p})\sqrt {a_p}. $$

Since \( | {{\mathcal{HP}}_{ij} } | + | {{\mathcal{VP}}_{ij} } | \le \sqrt 2 | {v_i-v_j } | \) and by Lemma 5, we have

$$\tilde{S}(k,t) \le 2\sqrt {a_{p} } \sum\limits_{{e_{{ij}} \in \Pi _{k} }} ( \sqrt 2 |v_{i} - v_{j} | + \sqrt {a_{p} } ) \le 2(n_{d} a_{p} + 4\sqrt {n_{d} a_{p} } ).$$

From the latest inequality, we obtain the uniform upper bound of S(kt), independent of k, denoted as S. That is,

$$ S(k,t)\leq S= \min \{2(n_d {a_p}+ 4\sqrt {n_d a_p}), 1\}. $$

Hence, the upper bound of Y t follows Poisson with λ = n s S.

When n s S = ω(log n), by union bounds and Lemma 4, we get

$$ \begin{aligned} \Pr(X \ge 2n_s S ) \le& \left({\frac{n} {\log n}}\right)\cdot\Pr(X_t\ge 2 n_s S ) \le \left({\frac{n} {\log n}}\right)\cdot\Pr(Y_t \ge 2n_s S )\\ \leq& \left({\frac{n} {\log n}}\right)\cdot{\frac{{ e^ {- n_s S}(e n_s S)^{2 n_s S}}} {(2 n_s S)^{(2 n_s S)}}} = \left({\frac{n} {\log n}}\right)\cdot({\frac{4} {e}})^{ - n_s S } = o({\frac{1} {\log n}})\to 0, (n\to\infty). \end{aligned} $$

When n s S = O(logn), we can assume that \(\lim\nolimits_{n\to \infty }{\frac{n_s S} {\log n}}=f,\) where f ≥ 0 is a constant. Let z = e 2max{f, 1}, by union bounds and Lemma 1, we have

$$ \begin{aligned} \Pr(X \ge z\log n ) \le& \left({\frac{n} {z\log n}}\right)\cdot\Pr(X_t\ge z\log n ) \le \left({\frac{n} {\log n}}\right)\cdot {\Pr}(Y_t\geq z\log n ) \\ \leq& \left({\frac{1} {\log n}}\right)\cdot\left({\frac{e^{{\frac{(1-f)} {z}}+1}f} {z}}\right)^{z\log n} \leq (e^{{\frac{1} {e^2}}-1})^{z\log n} \to 0, \hbox{as} n\to \infty. \\ \end{aligned} $$

To sum up the two cases, we can prove the lemma.

Proof

Lemma 10 For any link \(\langle v_i^s, v_j^s\rangle \in \Upxi_s(\bar{\digamma}), v_j^s\) must locate in a primary cell, and it may be served if it is out of the preservation regions. However, there is possibly a time slot τ0 in which the distance from a primary node \(v_0^p \in {\mathcal{V}}_p(v_i^s, \tau_0)\) to v s j is so close that a fatal interference is imposed on v s j . The secondary transmission scheduling scheme can prevent this scenario from happening. Since for \({\mathcal{N}}_s(m)\) the same packet is transmitted along M time slots, we can guarantee that there exists a time slot τ out of M time slots in which the minimum distance to v s j from all \(v^p \in {\mathcal{V}}_p(v_i^s, \tau)\) is at least of \((M-{\frac{5} {2}})\sqrt{a_p}\). Hence,

$$\begin{aligned} I_{{ps}} (v_{i}^{s} ,v_{j}^{s} ;\tau ) \le & \sum\limits_{{i = 1}}^{\infty } {8iP(a_{p} )^{{{\frac{\alpha }{2}}}} \ell ((Ki - 2)\sqrt {a_{p} } )} \\ & + P(a_{p} )^{{{\frac{\alpha }{2}}}} \ell (\left( {M - {\frac{5}{2}}} \right)\sqrt {a_{p} } ). \\ \end{aligned}$$

So, we have

$$ I_{ps}(v^p_i,v^p_j;\tau )< 8P\cdot C_1+ (M-{\frac{5} {2}})^{-\alpha}\cdot P, $$

where C1 = C1(K, α) is the constant defined in Eq. (16). Similar to Lemma 8, we can obtain that I ps (vs i vs j ) < 8P · θ(n) · C1. Since

$$ S(v^s_i,v^s_j;\tau )\geq P\cdot \theta(n) \cdot {a_s}^{{\frac{\alpha} {2}}}(5 a_s)^{{\frac{\alpha} {2}}})=P \cdot \theta(n) \cdot 5^{-{\frac{\alpha} {2}}}, $$

when \(0<\lim\nolimits_{n {\to}\infty}\theta(n)<\infty,\) we have

$$ R_s(v^s_i,v^s_j;\tau)\geq {\frac{1} {MK^2}}\cdot\log \left( 1+{\frac{P\cdot \theta \cdot (\sqrt{5})^{-\alpha}}{N_0+P\cdot (6C_1(1+ \theta)+({\frac{2} {2M-5}})^{\alpha})}}\right ). $$

If θ(n) = o(1), we have

$$ R_s(v^s_i,v^s_j;\tau)\geq {\frac{P\cdot \theta(n) \cdot (\sqrt{5})^{-\alpha}} {MK^2N_0+P\cdot (6C_1(1+ \theta)+({\frac{2} {2M-5}})^{\alpha})}}. $$

Combining two cases in terms of θ(n), we complete the proof.

Proof

Lemma 12 Define the number of non-served sources as a random variable ξ s ′. According to Lemma 11, it follows a Poisson distribution of mean

$$ \lambda^{\prime}_s \leq m_s\cdot S^{\prime}_{max}=K^2\mu \cdot m_s \cdot n\cdot {\frac{\log m} {m}}. $$

When \(m_s \cdot n\cdot {\frac{\log m} {m}}=\omega(1),\) let \( {\rho}^{\prime}_s(m)=2K^2 \cdot \mu \cdot n\cdot {\frac{\log m} {m}}\). By Lemma 4, we get

$$ \Pr(\xi^{\prime}_s \geq {\rho}^{\prime}_{s}(m)\cdot m_s) \leq ({e}/{4})^{ K^2\mu \cdot m_sn\cdot {\frac{\log m} {m}} } \to 0. $$

When \(m_s \cdot n\cdot {\frac{\log m} {m}}=O(1),\) let \({\rho}^{\prime}_s(m)=(m_s)^{-\kappa_1},\) where 0 < κ1 < 1 is a constant. By Lemma 4, there exists a constant κ2 such that

$$ \Pr(\xi^{\prime}_s \geq (m_s)^{1-\kappa_1}) \leq \kappa_2 \cdot (e/(m_s)^{1-\kappa_1})^{(m_s)^{1-\kappa_1}} \to 0. $$

Combining the two cases in terms of the relations among m s , m and n, we complete the proof.

Proof

Lemma 13 According to Lemma 11, the random variable ξd k follows a Poisson distribution of mean \(K^2\mu \cdot m_d\cdot n\cdot {\frac{\log m} {m}}\). By using Lemma 4 and union bounds, we separately commence our analysis in two cases according to the relations among m d , m and n. When \(m_d \cdot n \cdot {\frac{\log m} {m}}=\omega(\log m_s),\) it holds that,

$$ \Pr(\xi^d \geq 2K^2\mu m_d\cdot n\cdot {\frac{\log m} {m}}) \leq m_s\cdot \left({\frac{e} {4}}\right)^{ K^2\mu m_d\cdot n\cdot {\frac{m} {\log m}} } \to 0. $$

When \(m_d \cdot n\cdot {\frac{\log m} {m}}=O(\log m_s),\) and

$$ \lim\limits_{n\to \infty }{\frac{m_d \cdot n\cdot \log m} {m\cdot\log m_s}}=f_1, f_1\geq 0 $$

Let z1 = 3 + log(1 + K2μf1), we have

$$ \Pr(\xi^d \geq e^{z_1}\log m_s) \le m_s\Pr(\Upxi^d_k \geq z_1\log m_s) \le m_s^{1+e^{z_1}-K^2\mu f_1-z_1e^{z_1}+e^{z_1}\log (1+K^2\mu f_1)} \to 0. $$

Combining the two cases, we obtain that

$$ \rho^{\prime}_d(m)= \left\{ \begin{array}{lll} O\left(n \cdot {\frac{\log m} {m}}\right) & \hbox{when} & m_d \cdot n \cdot {\frac{\log m} {m}}=\Upomega(\log m_s) \\ O\left({\frac{\log m_s} {m_d}}\right) & \hbox{when} & m_d \cdot n\cdot {\frac{\log m} {m}}=O(\log m_s)\\ \end{array} \right. $$

Thus, we have that ρ d ′(m) = o(1) when m d  = ω(log m s ).

Proof

Lemma 14 Define the number of the failed multicast sessions as a random variable γ. Note that γ follows a Poisson distribution of mean

$$ \lambda^{\prime\prime}_s\leq m_d\cdot m_s \cdot K^2\cdot\mu \cdot n\cdot \log m/m. $$

When \(m_d\cdot m_s \cdot n\cdot {\frac{\log m} {m}}=\omega(1),\) let \({\rho}^{\prime\prime}_s(m)=2m_d \cdot K^2\cdot\mu \cdot n\cdot {\frac{\log m} {m}}\). By Lemma 4, we have

$$ \Pr(\gamma \geq {\rho}^{\prime\prime}_s(m) \cdot m_s )\leq (e/4)^{m_dm_s \cdot K^2\cdot\mu \cdot n\cdot {\frac{\log m} {m}})}\to 0. $$

When \(m_d\cdot m_s \cdot n\cdot {\frac{\log m} {m}}=O(1),\) let \({\rho}^{\prime\prime}_s(m)=(m_s)^{-\kappa_3},\) where 0 < κ3 < 1 is a constant. By Lemma 4, there exists a constant κ4 such that

$$ \Pr(\gamma \geq {\rho}^{\prime\prime}_s(m) \cdot m_s) \leq \kappa_4 \cdot (e/(m_s)^{1-\kappa_3})^{(m_s)^{1-\kappa_3}} \to 0, {\rm as\ } m, n \to \infty. $$

Combining the two cases, the lemma can be proved.

Proof

Lemma 16 As in N p (n), to simplify the description, we define a sequence of sets of directed edges:

$$ \Uppi^{\prime}_k=\{e_{ij} | \langle v_i, v_j\rangle \in \hbox{EST}({\mathcal{U}}^{\prime}_k )\}, {\ for\ } k=1, 2, \ldots, m_s. $$

Given a secondary cell \( \bar{c}^{*}_{t},\) we define the number of multicast flows that will be routed through the nodes inside this cell as a random variable Z t , and we finally consider the uniform upper bound Z of Z t for every secondary cell.

Define event D(kt): Multicast session \({\mathcal{M}}_{k}\) passes through the cell \( \bar{c}^*_{t} \). For any link \(v_iv_j \in \hbox{EST}({\mathcal{U}}^{\prime}_k),\) Define event D h ij (kt): \(\overline{{\mathcal{HP}}}_{ij}\) passes through \( \bar{c}^*_{t} ;\) and event D v ij (kt): \(\overline{{\mathcal{VP}}}_{ij}\) passes through \( \bar{c}^*_{t} \). Obviously, we have

$$ {\Pr}(D(k,t) ) \leq\sum\limits_{e_{ij} \in \Pi^{\prime}_k} (\Pr(D_{ij}^h (k,t)) + \Pr(D_{ij}^v (k,t)). $$

We firstly give the bound of \(\Pr(D(k,t) )\). Based on the cell \(\bar{c}^{*}_{t},\) we construct the region \({\bar{\mathcal{S}}}^{h}_{ij}(k,t)\) of area

$$ \bar{S}^h_{ij}(k,t)=(2\mu K+1)\cdot \sqrt{a_s}\cdot(2 | {{\mathcal{HP}}_{ij} } | + (K+2)\sqrt {a_s}), $$

as in Fig. 4. Similarly, we construct the region \({\bar{\mathcal{S}}}^{v}_{ij}(k,t)\) of area

$$ \bar{S}^v_{ij}(k,t)=(2\mu K+1)\cdot \sqrt{a_s}\cdot(2 | {{\mathcal{VP}}_{ij} } | + (K+2)\sqrt {a_s}). $$

The following two propositions are obviously true.

Fig. 4
figure 4

The construction of the region \(\bar{{\mathcal{S}}}^h_{ij}(k,t)\). Here, \(l_0=\mu K\sqrt{a_s}\) because the maximum size of any clusters is μ. Obviously, \(l=(2\mu K+1)\cdot \sqrt{a_s}\)

Proposition 3

The Poisson pointv i locates in the region\({\bar{\mathcal{S}}}^{h}_{ij}(k,t)\)if the eventD h ij (kt) happens.

Proposition 4

The Poisson pointv j locates in the region\({\bar{\mathcal{S}}}^{v}_{ij}(k,t)\)if the eventD v ij (kt) happens.

Then, we get that \(Z_t=\sum\nolimits_{k=1}^{m_s}{\mathcal{I}}(D(K,t))\) follows a Poisson distribution of mean \(\bar{\lambda}\leq m_s\cdot \bar{S},\) where

$$ \bar{S}=\min\left\{1, \sum\limits_{e_{ij} \in \Pi^{\prime}_k}{(\bar{S}^h_{ij}(k,t)+\bar{S}^v_{ij}(k,t))} \right\}. $$

By Lemma 5, we have

$$\begin{aligned} \bar{S} \le & 2(2\mu K + 1)\cdot\sqrt {a_{s} } \cdot\sum\limits_{{e_{{ij}} \in \Pi _{k}^{\prime } }} {\left( {\sqrt 2 |v_{i} v_{j} | + (K + 2)\sqrt {a_{s} } } \right)} \\ \le & 2(2\mu K + 1)\cdot(4\sqrt {m_{d} a_{s} } + (K + 2)m_{d} a_{s} ) = \hat{S}. \\ \end{aligned}$$

When \(m_s \hat{S}=\omega(\log n),\) by union bounds and Lemma 4, we can obtain

$$ \Pr(Z \ge 2m_s \hat{S} ) \leq (m / \log m)(4/e)^{ - m_s \hat{S} } = o(1/\log m) \to 0. $$

When \(m_s \hat{S}=O(\log m)\) and \(\lim\nolimits_{n\to \infty }{\frac{m_s \hat{S}} {\log m}}=\bar{f}\geq 0\). Let \(\bar{z}=e^2\max\{\bar{f},1\},\)we obtain that

$$ \Pr(Z \ge \bar{z}\log n ) \leq (e^{{\frac{1} {e^2}}-1})^{\bar{z}\log n} \to 0. $$

Summing up these two cases, we complete the proof.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wang, C., Jiang, C., Li, XY. et al. Multicast throughput for large scale cognitive networks. Wireless Netw 16, 1945–1960 (2010). https://doi.org/10.1007/s11276-010-0237-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11276-010-0237-3

Keywords

Navigation