Skip to main content
Log in

A characterization of max–min SIR-balanced power allocation with applications

  • Published:
Wireless Networks Aims and scope Submit manuscript

Abstract

We consider a power-controlled wireless network with an established network topology in which the communication links (transmitter–receiver pairs) are corrupted by the co-channel interference and background noise. We have fairly general power constraints since the vector of transmit powers is confined to belong to an arbitrary convex polytope. The interference is completely determined by a so-called gain matrix. Assuming irreducibility of this gain matrix, we provide an elegant characterization of the max–min SIR-balanced power allocation under such general power constraints. This characterization gives rise to two types of algorithms for computing the max–min SIR-balanced power allocation. One of the algorithms is a utility-based power control algorithm to maximize a weighted sum of the utilities of the link SIRs. Our results show how to choose the weight vector and utility function so that the utility-based solution is equal to the solution of the max–min SIR-balancing problem. The algorithm is not amenable to distributed implementation as the weights are global variables. In order to mitigate the problem of computing the weight vector in distributed wireless networks, we point out a saddle point characterization of the Perron root of some extended gain matrices and discuss how this characterization can be used in the design of algorithms in which each link iteratively updates its weight vector in parallel to the power control recursion. Finally, the paper provides a basis for the development of distributed power control and beamforming algorithms to find a global solution of the max–min SIR-balancing problem.

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. \({{\mathbb{R}}_{+}}, {{\mathbb{R}}_{++}}\) are non-negative and positive reals, respectively.

References

  1. Aein, J. M. (1973). Power balancing in systems employing frequency reuse. COMSAT Technical Review 3(2), 277–300.

    Google Scholar 

  2. Alavi, H., & Nettleton, R. (1982). Downstream power control for a spread spectrum cellular mobile radio system. In Proceedings of IEEE Global Communications Conference (GLOBECOM), pp. 84–88.

  3. Bambos, N. (1998). Toward power-sensitive network architectures in wireless communications: Concepts, issues, and design aspects. IEEE Personal Communication Magazine, 5, 50–59.

    Article  Google Scholar 

  4. Berry, R., Liu, P., & Honig, M. (2002). Design and analysis of downlink utility-based schedulers. In: Proceedings of 43rd annual allerton conference on communication, control and computing.

  5. Bertsekas, D., & Gallager, R. (1992). Data networks. Englewood Cliffs: Prentice-Hall.

    MATH  Google Scholar 

  6. Boche, H., & Schubert, M. (2006). Resource allocation in multiantenna systems—Achieving max–min fairness by optimizing a sum of inverse SIRs. IEEE Transactions on Signal Processing, 54(6).

  7. Boche, H., & Stanczak, S. (2005). Log-convexity of the minimum total power in CDMA systems with certain quality-of-service guaranteed. IEEE Transactions on Information Theory, 51(1), 374–381.

    Article  MathSciNet  Google Scholar 

  8. Boche, H., & Stanczak, S. (2006). The Kullback–Leibler divergence and nonnegative matrices. IEEE Transactions on Information Theory, 52(12), 5539–5545.

    Article  MathSciNet  Google Scholar 

  9. Borst, S., & Whiting, P. (2003). Dynamic channel-sensitive scheduling algorithms for wireless data throughput optimization. IEEE Transactions on Vehicular Technology, 52(3), 569–586.

    Article  Google Scholar 

  10. Boyd, S., & Vandenberghe, L. (2004). Convex optimization. Cambridge: Cambridge University Press.

    MATH  Google Scholar 

  11. Chen, L., Low, S., Chiang, M., & Doyle, J. (2006). Optimal cross-layer congestion control, routing and scheduling design in ad hoc wireless networks. In: Proceedings of 25th IEEE Conference on Computer Communications (INFOCOM), Barcelona.

  12. Chiang, M. (2005). Balancing transport and physical layers in wireless multihop networks: Jointly optimal congestion control and power control. IEEE Journal on Selected Areas in Communications, 23(1), 104–116.

    Article  Google Scholar 

  13. Chiang, M., Hande, P., Lan, T., & Tan, C. (2008). Power control in wireless cellular networks. Foundations and Trends in Networking, 2(4), 381–533.

    Article  Google Scholar 

  14. Deutsch, E., & Neumann, M. (1984). Derivatives of the Perron root at an essentially nonnegative matrix and the group inverse of an M-matrix. Journal of Mathematical Analysis and Applications, I-29(102), 1–29.

    Article  MathSciNet  Google Scholar 

  15. ElBatt, T., & Ephremides, A. (2004). Joint scheduling and power control for wireless ad hoc networks. IEEE Transactions on Wireless Communication, 3(1), 74–85.

    Article  Google Scholar 

  16. Feistel, A., Stanczak, S., & Kaliszan, M. (2009). Resource allocation in multiantenna systems under general power constraints. In: Proceedings of international ITG Workshop on Smart Antennas 2009 (WSA). Berlin, Germany.

  17. Foschini, G., & Miljanic, Z. (1993). A simple distributed autonomous power control algorithm and its convergence. IEEE Transactions on Vehicular Technology, 42(4), 641–646.

    Article  Google Scholar 

  18. Friedland, S., & Karlin, S. (1975). Some inequalities for the spectral radius of non-negative matrices and applications. Duke Mathematical Journal, 42(3), 459–490.

    Article  MATH  MathSciNet  Google Scholar 

  19. Gerlach, D., & Paulraj, A. (1996). Base station transmitting antenna arrays for multipath environments. Signal Processing (Elsevier Science), 54, 59–73.

    Article  MATH  Google Scholar 

  20. Hande, P., Rangan, S., Chiang, M., & Wu, X. (2008). Distributed uplink power control for optimal sir assignment in cellular data networks. IEEE/ACM Transactions on Networking, 16(6), 1420–1433.

    Article  Google Scholar 

  21. He, B., Wang, M., & Li, E. (1997). A new distributed power balancing algorithm for CDMA cellular systems. In: Proceedings of IEEE International Symposium on Circuits and Systems (ISCAS), Vol. 3, pp. 1768–1771.

  22. Hongyu, W., Aiging, H., Rong, H., & Weikang, G. (2000). Balanced distributed power control. In: IEEE international symposium on Personal, Indoor and Mobile Radio Communications (PIMRC), Vol. 2, pp. 1415–1419.

  23. Horn, R., & Johnson, C. (1985). Matrix analysis. Cambridge: Cambridge University Press.

    MATH  Google Scholar 

  24. Huang, J., Subramanian, V. G., Agrawal, R., & Berry, R. (2009). Joint scheduling and resource allocation in uplink ofdm systems for broadband wireless access networks. IEEE Journal on Selected Areas in Communications, 27(2), 288–296.

    Article  Google Scholar 

  25. Kandukuri, S., & Boyd, S. (2002). Optimal power control in interference-limited fading wireless channels with outage-probability specifications. IEEE Transactions on Wireless Communication, 1(1), 46–55.

    Article  Google Scholar 

  26. Kelly, F., Maulloo, A., & Tan, D. (1998). Rate control for communication networks: Shadow prices, proportional fairness and stability. Journal of the Operational Research Society, 49(3), 237–252.

    MATH  Google Scholar 

  27. Lee, J. W., Mazumdar, R., & Shroff, N. (2006). Oportunistic power scheduling for dynamic multi-server wireless systems. IEEE Transactions on Wireless Communication, 5(6), 1506–1515.

    Article  Google Scholar 

  28. Mahdavi-Doost, H., Ebrahimi, M., & Khandani, A. K. (2007). Characterization of rate region in interference channels with constrained power. In: Proceedings of IEEE International Symposium on Information Theory (ISIT). Nice, France.

  29. Meyer, C. D. (2000). Matrix analysis and applied linear algebra. Philadelphia: SIAM.

    MATH  Google Scholar 

  30. Meyerhoff, H. J. (1974). Method for computing the optimum power balance in multibeam satellites. COMSAT Techical Review, 4(1), 139–146.

    Google Scholar 

  31. Mo, J., & Walrand, J. (2000). Fair end-to-end window-based congestion control. IEEE/ACM Transactions on Networking, 8(5), 556–567.

    Article  Google Scholar 

  32. Montalbano, G., Ghauri, I., & Slock, D. T. M. (1998). Spatio-temporal array processing for CDMA/SDMA downlink transmission. In: Proceedings of Asilomar conference on signals, systems and computers, Monterey, CA, USA, pp. 1337–1341.

  33. Neely, M., Lee, C.P., & Modiano, E. (2005). Fairness and optimal stochastic control for heterogeneous networks. In: Proceedings of 23rd IEEE Conference on Computer Communications (INFOCOM), Miami, FL, USA.

  34. Neely, M. J. (2003). Dynamic power allocation and routing for satellite and wireless networks with time varying channels. Ph. D. thesis, Massachusetts Institute of Technology, LIDS, Cambridge, MA, USA.

  35. Palomar, D. P., & Chiang, M. (2006). A tutorial on decomposition methods for network utility maximization. IEEE Journal on Selected Areas in Communications, 24(8), 1439–1451.

    Article  Google Scholar 

  36. Papandriopoulos, J., Evans, J., & Dey, S. (2005). Optimal power control for rayleigh-faded multiuser systems with outage constraints. IEEE Transactions on Wireless Communication, 4(6), 2705–2715.

    Article  Google Scholar 

  37. Papandriopoulos, J., Evans, J., & Dey, S. (2006). Outage-based optimal power control for generalized multiuser fading channels. IEEE Transactions on Communication, 54(4), 693–703.

    Article  Google Scholar 

  38. Rockafellar, R. (1971) Saddle points and convex analysis. In: H. Kuhn, & G. Szego (Eds.), Differential games and related topics (pp. 109–127). Amsterdam, The Netherlands: North-Holland.

  39. Rockafellar, R. T. (1972). Convex analysis. New Jersey: Princeton University Press.

    Google Scholar 

  40. Rudin, W. (1976). Principles of mathematical analysis. New York: McGraw-Hill.

    MATH  Google Scholar 

  41. Schubert, M., & Boche, H. (2006). Qos-based resource allocation and transceiver optimization. Foundation and Trends in Communications and Information Theory, 2(6).

  42. Seneta, E. (1981). Non-negative matrices and Markov chains. Berlin: Springer.

    MATH  Google Scholar 

  43. Stanczak, S., & Boche, H. (2006). Strict log-convexity of the minimum power vector. In: Proceedings of IEEE International Symposium on Information Theory (ISIT), Seattle, WA, USA.

  44. Stanczak, S., Kaliszan, M., Bambos, N., & Wiczanowski, M. (2009). A characterization of max–min SIR-balanced power allocation with applications. Technical Report. http://www.arxiv.org/abs/0901.0824v1, Submitted on 7 Jan 2009.

  45. Stanczak, S., Wiczanowski, M., & Boche, H. (2006). Theory and algorithms for resource allocation in wireless networks. Lecture notes in computer science (LNCS 4000)). Berlin: Springer.

    Google Scholar 

  46. Stanczak, S., Wiczanowski, M., & Boche, H. (2007). Distributed utility-based power control: Objectives and algorithms. IEEE Transactions on Signal Processing, 55(10), 5058–5068.

    Article  Google Scholar 

  47. Stolyar, A. (2005). Maximizing queueing network utility subject to stability. Queueing Systems, 50(4), 401–457.

    Article  MATH  MathSciNet  Google Scholar 

  48. Subramanian, A., & Sayed, A. H. (2005). Joint rate and power control algorithms for wireless networks. IEEE Transactions on Signal Processing, 53(11), 4204–4214.

    Article  MathSciNet  Google Scholar 

  49. Tan, C., Chiang, M., & Srikant, R. (2009). Fast algorithms and performance bounds for sum rate maximization in wireless networks. In: Proceedings of 28th IEEE Conference on Computer Communications (INFOCOM). Rio de Janeiro, Brazil.

  50. Wiczanowski, M., Stanczak, S., & Boche, H. (2008). Providing quadratic convergence of decentralized power control in wireless networks: The method of min-max functions. IEEE Transactions on Signal Processing, 56(8), 4053–4068.

    Article  MathSciNet  Google Scholar 

  51. Yang, W., & Xu, G. (1998). Optimal downlink power assignment for smart antenna systems. In: Proceedings of IEEE International Conference on Acoustics, Speech, and Signal Processing (ICASSP). Seattle, WA, USA.

  52. Yates, R. (1995). A framework for uplink power control in cellular radio systems. IEEE Journal on Selected Areas in Communications, 13(7), 1341–1347.

    Article  MathSciNet  Google Scholar 

  53. Yates, R., & Huang, C. (1995). Integrated power control and base station assignment. EEE Transactions on Vehicular Technology, 44(3), 638–644.

    Article  Google Scholar 

  54. Zander, J. (1992). Distributed cochannel interference control in cellular radio systems. IEEE Transactions on Vehicular Technology, 41, 305–311.

    Article  Google Scholar 

  55. Zander, J. (1992). Performance of optimum transmitter power control in cellular radio systems. IEEE Transactions on Vehicular Technology, 41(1), 57–62.

    Article  Google Scholar 

  56. Zander, J., & Kim, S. L. (2001). Radio resource management for wireless networks. Boston, London: Artech House.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michał Kaliszan.

Additional information

The work was partly presented at the 2009 IEEE Information Symposium on Information Theory (ISIT), Seoul, Korea.

This work was supported in part by the European Union under grant IST-034413, in part by the German Ministry of Education and Research (BMBF) under grants 01BU0680 and 01BN0712C, and in part by the German Ministry of Economics and Technology under grant 20K0805B.

Appendix 1

Appendix 1

1.1 1.1 Auxiliary results

For completeness and reader’s convenience, we restate here a useful result after [45, Theorem A.35], [42, Theorem 2.1], which we refer to in some proofs below.

Theorem 6

LetXbe an arbitrary non-negative matrix, and let α > 0 be any scalar. A necessary and sufficient condition for a solutionp ≥ 0, p ≠ 0, to

$$ (\alpha{\bf I}-{\bf X}){\bf p}={\bf b} $$
(25)

to exist for anyb > 0 is that α > r = ρ(X). In this case, there is only one solutionp, which is strictly positive and given byp = (αI − X)−1b.

1.2 1.2 Proof of Lemma 1

By Definition 1 and (3), we have \(t'\leq\min_{k\in{\mathcal{K}}}(\hbox{SIR}_k({\bf p})/\gamma_k),{\bf p}\in\hbox{P}\), if and only if p is a max–min SIR-balanced power vector. Therefore, in a matrix form, we have \({\varvec{\Upgamma}}{\bf z}\leq(\frac{1}{t^{\prime}}{\bf I}-{\varvec{\Upgamma}}{\bf V}){\bf p},\,{\bf p}\in\hbox{P}\) if and only if p is a solution to (2). Since a solution exists and is positive, it follows from Theorem 6 that

$$ \rho({\varvec{\Upgamma}}{\bf V})<1/t^{\prime} , $$
(26)

where ρ(·) is used to denote the spectral radius. So, by Theorem 6 we know that there exists a unique positive vector \(\bar{\bf p}^{\prime}\in\hbox{P}\) given by (6) and this vector is a solution to the max–min SIR-balancing problem (2). Writing (6) as a system of K SIR equations we get \(\hbox{SIR}_k(\bar{{\bf p}}')/\gamma_k=t'\) for each \(k\in{\mathcal{K}}\) which completes the proof.

1.3 1.3 Proof of Lemma 2

By the definition of F in (5), for any q ∈ F, there exists p(q) ∈ P+ given by

$$ {\bf p}({\bf q})=({\bf I}-{\bf G}({\bf q}){\varvec{\Upgamma}}{\bf V})^{-1}{\bf G}({\bf q}){\varvec{\Upgamma}}{\bf z},\;\quad{\bf q}\in\hbox{F} $$
(27)

where \({\bf G}({\bf q}):=\hbox{diag}(g(q_1),\ldots,g(q_K))\) with g(x) defined by (A.3). From Theorem 6 we know that \(\rho({\bf G}({\bf q}){\varvec{\Upgamma}}{\bf V})<1\) and p(q) is unique so that p(q) defines a function (map) from F into P+. Furthermore, the Neumann series \(\sum_{l=0}^\infty({\bf G}({\bf q}){\varvec{\Upgamma}}{\bf V})^{l}\) exists and converges (see for instance [29, p. 618] and [45, Theorem A.11]) so that

$$ {\bf p}({\bf q})=\sum_{l=0}^\infty({\bf G}({\bf q}){\varvec{\Upgamma}}{\bf V})^{l}{\bf G}({\bf q}){\varvec{\Upgamma}}{\bf z}. $$
(28)

Now, considering strict monotonicity and continuity of \(g:\hbox{Q}\to{{\mathbb{R}}_{++}}\) shows that for every p ∈ P, there is exactly one q ∈ F such that p = p(q). Combining this with the above let us conclude that there is a one-to-one correspondence between F and P. This map is continuous and defined by (27). So, by [40, Theorem 4.22], F is a connected set. Moreover, as g is strictly increasing, the right-hand side of (28) shows that if q ∈ F and q′ ≤ qq′ ∈ QK, then we must have q′ ∈ F, which implies downward comprehensivity. Finally, with (A.3), the convexity property follows from [7] or [45, Sect. 5.3].

1.4 1.4 Proof of Proposition 1

Let (A.3) be satisfied, and let \(\bar{\bf p}\in{\hbox{P}_{+}}\) be any solution to (2) or, equivalently, (4). Let \(\bar{q}_k=\phi(\hbox{SIR}_k(\bar{\bf p})/\gamma_k),k\in{\mathcal{K}}\) . By Lemma 2, F is a convex downward comprehensive set and, by (A.2), (4) and (5), \(\bar{\bf q}=(\bar{q}_1,\ldots,\bar{q}_K)\in\partial\hbox{F}\) is its boundary point since at least one power constraint is active in the optimum (see Lemma 4). Thus, by irreducibility of V and Lemma 3, there exists w > 0 such that \({\bf w}^T(\bar{{\bf q}}-{\bf u})\geq 0\) for all u ∈ F. Due to positivity of w, this implies that for any \({\bf u}\in\hbox{F},{\bf u}\neq\bar{{\bf q}}\), there is at least one index \(i=i(\bar{{\bf q}}, {\bf u})\in{\mathcal{K}}\) such that \(\bar{q}_{i}>u_{i}\). One particular solution to (2) is \(\bar{{\bf p}}^{\prime}\) given by (6), therefore for any \({\bf u}\in\hbox{F},{\bf u}\neq\bar{\bf q}^{\prime}\) there is \(i=i(\bar{\bf q}^{\prime},{\bf u})\in{\mathcal{K}}\) such that \(\bar{q}_{i}^{\prime}>u_i\) where \(\bar{q}^{\prime}_k=\phi(\hbox{SIR}_k(\bar{\bf p}^{\prime})/\gamma_k),k\in{\mathcal{K}}\). On the other hand, however, we have \(\bar{\bf q}^{\prime}\leq\bar{\bf q}\), since by Observation 1 and Definition 1 there holds \(\bar{q}^{\prime}_1=\cdots=\bar{q}^{\prime}_K= \min_{k\in{\mathcal{K}}}\phi(\hbox{SIR}_k(\bar{{\bf p}})/\gamma_k)\). Combining both inequalities shows that \(\bar{{\bf q}}=\bar{\bf q}^{\prime}\), and hence, by bijectivity, we obtain \(\bar{\bf p}=\bar{\bf p}'\), which is unique by Theorem 6.

1.5 1.5 Proof of Lemma 4

Part (1) should be obvious since if we had \(g_n(\bar{{\bf p}})<1\) for all \(n\in{\mathcal{N}}\), then it would be possible to increase \(\min_{k\in{\mathcal{K}}}\hbox{SIR}_k(\bar{{\bf p}})/\gamma_k\) by allocating the power vector \(c\bar{{\bf p}}\in{\hbox{P}_{+}}\) with \(c=1/\max_{n\in{\mathcal{N}}}g_n(\bar{{\bf p}})>1\). In order to show part (ii), note that if (A.2) and (A.3) hold, then, by Lemma 2, F is a convex downward comprehensive set. Moreover, \(\bar{{\bf p}}\in{\hbox{P}_{+}}\) given by (4) corresponds to a point \(\bar{{\bf q}}\in\partial\hbox{F}\), with \(\bar{q}_k=\phi(\hbox{SIR}_k(\bar{{\bf p}})/\gamma_k),k\in{\mathcal{K}}\). Thus, by irreducibility of V, it follows from Lemma 3 that \(\bar{{\bf q}}\) is a maximal point of F, and hence \(\bar{{\bf q}}\leq{\bf q}\) for any q ∈ F implies that \({\bf q}=\bar{{\bf q}}\) [10]. That is, there is no vector in F that is larger in all components than \(\bar{{\bf q}}\). On the other hand, by the discussion in Sect. 3, \(\bar{{\bf q}}\) is a point where the hyperplane in the direction of the vector \((1/K,\ldots,1/K)\) intersects the boundary of F. As a result, \(\bar{q}_1=\cdots=\bar{q}_K\), which together with the maximality property and strict monotonicity of ϕ, shows that \(\hbox{SIR}_k(\bar{{\bf p}})/\gamma_k=\beta\) for each \(k\in{\mathcal{K}}\) where β is positive due to (i). If V is irreducible, the uniqueness of \(\bar{{\bf p}}\) follows from Proposition 1.

1.6 1.6 Proof of Lemma 6

Let \(n\in{\mathcal{N}}\) be arbitrary. First we prove part (i). Since \(1/P_n{\bf z}{\bf c}_n^T\geq 0\) and V is irreducible, we can conclude from (15) that B (n) ≥ 0 is irreducible as well. Thus, by the Perron–Frobenius theorem for irreducible matrices [29, 23], there exists a positive vector p which is an eigenvector of B (n) associated with ρ(B (n)), and there are no nonnegative eigenvectors of B (n) associated with ρ(B (n)) other than p and its positive multiples. Among all the positive eigenvectors, there is exactly one eigenvector p > 0 such that g n (p) = c 1. This proves part (i). In order to prove (ii), note that if A (n) was irreducible, then we could invoke the Perron–Frobenius theorem and proceed essentially as in part (i) to conclude (ii) (with the uniqueness property resulting from the normalization of the eigenvector so that its last component is equal to c 2 > 0). In order to show that A (n) is irreducible, let \({\mathcal{G}}({\bf A}^{(n)})\) be the associated directed graph of \(\{1,\ldots,K+1\}\) nodes [29]. Since \({\varvec{\Upgamma}}{\bf V}\) is irreducible, it follows that the subgraph \({\mathcal{G}}({\varvec{\Upgamma}}{\bf V})\) is strongly connected [29]. Furthermore, as the vector \({\varvec{\Upgamma}}{\bf z}\) is positive, we can conclude from (13) that there is a directed edge leading from node K + 1 to each node n < K + 1 belonging to the subgraph \({\mathcal{G}}({\varvec{\Upgamma}}{\bf V})\). Finally, note that as \({\varvec{\Upgamma}}{\bf V}\) is irreducible, each row of \({\varvec{\Upgamma}}{\bf V}\) has at least one positive entry. Hence, the vector \(1/P_n{\bf c}_n^T{\varvec{\Upgamma}}{\bf V}\) has at least one positive entry as well, from which and (13) it follows that there is a directed edge leading from a node belonging to \({\mathcal{G}}({\varvec{\Upgamma}}{\bf V})\) to node K + 1. So, \({\mathcal{G}}({\bf A}^{(n)})\) is strongly connected, and thus A (n) is irreducible.

1.7 1.7 Proof of Theorem 1

(i)→(ii): By Lemma 5, \(\bar{{\bf p}}\in{\hbox{P}_{+}}\) satisfies (14) for some β > 0. Thus, by Lemma 6, part (i) implies part (ii). (ii)→(iii): Given any \(n\in{\mathcal{N}}_0(\bar{{\bf p}})\), it follows from (14) that \(\rho({\bf B}^{(n)})\bar{{\bf p}}={\bf B}^{(n)}\bar{{\bf p}}\) with \(g_n(\bar{{\bf p}})=1\) is equivalent to \(\rho({\bf B}^{(n)})\bar{{\bf p}}={\varvec{\Upgamma}}{\bf V}\bar{{\bf p}}+{\varvec{\Upgamma}}{\bf z}\), which in turn can be rewritten to give (12) with \(\beta=\rho({\bf B}^{(n)})\) and \(\bar{\underline{\bf p}}=(\bar{{\bf p}},1)\). Since \(\bar{{\bf p}}\) is positive, so is also \(\bar{\underline{\bf p}}\). Thus, \(\bar{\underline{\bf p}}\) with \(\bar{\underline{p}}_{K+1}=1\) is a positive right eigenvector of A (n) and the associated eigenvalue is equal to ρ(B (n)) > 0. So, considering part (ii) of Lemma 6, \(\bar{\underline{\bf p}}\) is unique and therefore we can conclude that (iii) follows from (ii). (iii)→(i): The solution to the problem (8) always exists. By Lemma 5 and Proposition 1, the irreducibility of V implies that the unique solution \(\bar{{\bf p}}\) satisfies (12) with \(\bar{\underline{\bf p}}=(\bar{{\bf p}},1)\) for each \(n\in{\mathcal{N}}_0(\bar{{\bf p}})\). On the other hand, the normalization \(\bar{\underline{p}}_{K+1}=1\) guarantees, by Lemma 6, that for each \(n\in{\mathcal{N}}_0(\bar{{\bf p}})\), there exists exactly one positive vector \(\bar{\underline{\bf p}}\) with \(\bar{\underline{p}}_{K+1}=1\) such that (12) is satisfied. It corresponds thus to the unique solution of (8) which proves the last missing implication.

1.8 1.8 Proof of Lemma 7

By (5) with (A.2), we have q ∈ F if and only if there is p ∈ P such that ϕ\((\hbox{SIR}_k({\bf p})/\gamma_k)\) ≥ q k for each \(k\in{\mathcal{K}}\). Thus, q ∈ F if and only if \(1/\lambda:=\max_{{\bf p}\in\hbox{P}}\min_{k\in{\mathcal{K}}}(\hbox{SIR}_k({\bf p})/\gamma_kg(q_k))\geq 1\) where the maximum always exists. Comparing the left hand side of the inequality above with (2) shows that the only difference to the original problem formulation is that γ k is substituted by γ k g(q k ) or, equivalently, \({\varvec{\Upgamma}}\) by \({\bf G}({\bf q}){\varvec{\Upgamma}}\), which is positive definite as well. Thus, by (9), Theorem 1 and Theorem 2, we have q ∈ F if and only if \(\lambda=\max_{n\in{\mathcal{N}}}\lambda_n({\bf q})\leq 1\). Moreover, p(q) given by (22) is the unique power vector such that \( q_k = \phi(\text{SIR}_k(\mathbf{p}(\mathbf{q}))/\gamma_k) \text{ for each } k\in{\mathcal{K}}\). Since the Neumann series converges for any q ∈ F, we have \({\bf p}({\bf q})=\sum_{l=0}^\infty({\bf G}({\bf q}){\varvec{\Upgamma}}{\bf V})^l{\bf G}({\bf q}){\varvec{\Upgamma}}{\bf z}\). Now as \({\bf G}({\bf q}){\varvec{\Upgamma}}{\bf z}\) is positive and \({\bf G}({\bf q}){\varvec{\Upgamma}}{\bf V}\) is irreducible, we can conclude from [45, Lemma A.22] and (A.2) that each entry of p(q) is strictly increasing in each entry of q. Thus, as F is downward comprehensive and \({\bf q}\notin\partial\hbox{F}\) holds if and only if all power constraints are inactive, for every q ∈ int(F), there is \(\tilde{{\bf q}}\in\partial\hbox{F}\) such that \(\tilde{{\bf q}}={\bf q}+{\bf u}\) for some u > 0. By irreducibility of B (n), this implies that \(\lambda_n({\bf q})<\lambda_n({\bf q}+{\bf u})=\lambda_n(\tilde{{\bf q}})\leq 1\) for each \(n\in{\mathcal{N}}\) So, if \(\max_{n\in{\mathcal{N}}}\lambda_n({\bf q})=1\), then q ∈ ∂F. Conversely, if q ∈ ∂F, we must have \(\max_{n\in{\mathcal{N}}}\lambda_n({\bf q})=1\) since otherwise there would exist \(\tilde{{\bf q}}\notin\hbox{F}\) such that \(\max_{n\in{\mathcal{N}}}\lambda_n(\tilde{{\bf q}})=1\), which would contradict Theorem 2. This completes the proof.

1.9 1.9 Proof of Theorem 3

Let \(\tilde{{\bf q}}\in\partial\hbox{F}\) and \(n_0=\arg\max_{n\in{\mathcal{N}}}\lambda_n({\bf q})\) be arbitrary and note that F is a convex set. So, by Lemma 3, there is w > 0 such that \(\tilde{{\bf q}}\) maximizes \({\bf q}\mapsto{\bf w}^T{\bf q}\) over F. Lemma 7 implies that this convex problem can be stated as \(\tilde{{\bf q}}=\arg\max_{{\bf q}}{\bf w}^T{\bf q}\) subject to \(\lambda_{n_0}({\bf q})=1,{\bf q}\in\hbox{Q}^K\). Due to (A.2), the spectral radius is continuously differentiable on QK. Thus, the Karush–Kuhn–Tucker conditions [10], which are necessary and sufficient for optimality (due to the convexity property), imply that w is parallel with \(\nabla\lambda_{n_0}({\bf q})\). Now, by [14], we have \(\frac{\partial\lambda_{n_0}({\bf q})}{\partial q_k}= y_kg^{\prime}(q_k)\sum_{l\in{\mathcal{K}}} b_{k,\,l}^{(n_0)}x_l =\frac{g^{\prime}(q_k)}{g(q_k)}y_k\sum_{l\in{\mathcal{K}}} g(q_k)b_{k,\,l}^{(n_0)}x_l =\lambda_{n_0}({\bf q})\frac{g^{\prime}(q_k)} {g(q_k)}y_k x_k =\frac{g^{\prime}(q_k)}{g(q_k)}y_k x_k\) for each \(k\in{\mathcal{K}}\) where y and x are left and right positive eigenvectors of \({\bf G}({\bf q}){\bf B}^{(n_0)}\) associated with \(\lambda_{n_0}({\bf q})\), which, by irreducibility, are unique up to positive multiples.

1.10 1.10 Proof of Corollary 1

As V is irreducible, Observation 1 and Proposition 1 (see also the proof) imply that \(\bar{{\bf p}}\) corresponds to a point \(\bar{{\bf q}}\in\partial\hbox{F}\). Since \({\bf q}^\ast({\bf w})\in\partial\hbox{F}\) for any w > 0, it follows from Theorem 3 that \(\bar{{\bf q}}={\bf q}^\ast({\bf w})\) if w is has the form (23). Now by Observation 1, we have \(\bar{q}_1=\cdots=\bar{q}_K\). Thus, as both g and its derivative g′ are strictly monotonic (by (A.2) and (A.3)), we must have \({\bf u}(\bar{{\bf q}})=a{\bf 1},a>0\) and \({\bf G}(\bar{\bf q})=1/\rho({\bf B}^{(n_0)}){\bf I}\). Thus, the corollary follows from Theorem 3.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Stańczak, S., Kaliszan, M. & Bambos, N. A characterization of max–min SIR-balanced power allocation with applications. Wireless Netw 16, 2335–2347 (2010). https://doi.org/10.1007/s11276-010-0261-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11276-010-0261-3

Keywords

Navigation