Skip to main content
Log in

Optimality Properties, Closed-Form Parameterizations and Distributed Strategy of the Two-User MISO Interference Channel

  • Published:
Wireless Personal Communications Aims and scope Submit manuscript

Abstract

In this paper, the closed-form parameterizations to the Pareto boundary for the two-user multiple-input single-output interference channel are studied. Firstly, for the equivalent channel model with each transmitter having only two antennas, the weighted sum-rate maximization (WSRMax) problem is reformulated with newly defined angle variables. Then, a centralized weighted leakage-plus-noise-to-signal ratio minimization (WLNSRMin) algorithm is proposed to find a locally optimal weighted sum-rate point. Each step of the algorithm is solved by evaluating closed-form expressions. A distributed algorithm is also given to avoid the exchange of the channel state information (CSI) between transmitters. Numerical results show that the centralized WLNSRMin algorithm converges to a local optimum of the WSRMax problem after a few iterations and the distributed algorithm achieves a performance very close to that of the centralized algorithm with only local CSI.

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

Similar content being viewed by others

References

  1. Jorswieck, E. A., Larsson, E. G., & Danev, D. (2008). Complete characterization of the Pareto boundary for the MISO interference channel. IEEE Transactions on Signal Processing, 56(10), 5292–5296.

    Article  MathSciNet  Google Scholar 

  2. Zhang, R., & Cui, S. (2010). Cooperative interference management with MISO beamforming. IEEE Transactions on Signal Processing, 58(10), 5450–5458.

    Article  MathSciNet  Google Scholar 

  3. Mochaourab, R., & Jorswieck, E. A. (2011). Optimal beamforming in interference networks with perfect local channel information. IEEE Transactions on Signal Processing, 59(3), 1128–1141.

    Article  MathSciNet  Google Scholar 

  4. Zakhour, R., & Gesbert, D. (2009). Coordination on the MISO interference channel using the virtual SINR framework. In Proceedings of the ITG Workshop on Smart Antennas (WSA09), Berlin, (pp. 1–7).

  5. Shang, X., Chen, B., & Poor, H. V. (2011). Multiuser MISO interference channels with single-user detection: Optimality of beamforming and the achievable rate region. IEEE Transactions on Information Theory, 57(7), 4255–4273.

    Article  MathSciNet  MATH  Google Scholar 

  6. Shang, X., & Poor, H. V. (2013). Noisy-interference sum-rate capacity for vector Gaussian interference channels. IEEE Transactions on Information Theory, 59(1), 132–153.

    Article  MathSciNet  MATH  Google Scholar 

  7. Lindblom, J., Karipidis, E., & Larsson, E. G. (2011). Closed-form parameterization of the Pareto boundary for the two-user MISO interference channel. In Proceedings of the IEEE International Conference acoustics, speech and signal processing (ICASSP), Prague, Czech (pp. 3372–3375).

  8. Mochaourab, R., & Jorswieck, E. A. (2011). Walrasian equilibrium in two-user multiple-input single-output interference channels. In Proceedings of the IEEE ICC, Kyoto, Japan (pp. 1–5).

  9. Zeng, Y., Yetis, C. M., Gunawan, E., Guan, Y. L., & Zhang, R. (2013). Transmit optimization with improper Gaussian signaling for interference channels. IEEE Transactions on Signal Processing, 61(11), 2899–2913.

    Article  MathSciNet  Google Scholar 

  10. Annapureddy, V. S., & Veeravalli, V. V. (2011). Sum capacity of MIMO interference channels in the low interference regime. IEEE Transactions on Information Theory, 57(5), 2565–2581.

    Article  MathSciNet  MATH  Google Scholar 

  11. Björnson, E., Jaldén, N., Bengtsson, M., & Ottersten, B. (2011). Optimality properties, distributed strategies, and measurement-based evaluation of coordinated multicell OFDMA transmission. IEEE Transactions on Signal Processing, 59(12), 6086–6101.

    Article  MathSciNet  Google Scholar 

  12. Shi, Q., Razaviyayn, M., Luo, Z., & He, C. (2011). An iteratively weighted MMSE approach to distributed sum-utility maximization for a MIMO interfering broadcast channel. IEEE Transactions on Signal Processing, 59(9), 4331–4340.

    Article  MathSciNet  Google Scholar 

  13. Christersen, S. S., Agarwal, R., de Carvalho, E., & Cioffi, J. M. (2008). Weighted sum-rate maximization using weighted MMSE for MIMO-BC beamforming design. IEEE Transactions on Wireless Communications, 7(12), 1–7.

    Google Scholar 

  14. Lindblom, J., & Karipidis, E. (2010). Cooperative beamforming for the MISO interference channel. In Proceedings of the European wireless conference (EW10), Lucca (pp. 631–638).

  15. Luenberger, D. G. (2003). Linear and nonlinear programming. Berlin: Springer.

    MATH  Google Scholar 

  16. Zhu, H., & Wang, J. (2012). Chunk-based resource allocation in OFDMA systems—part II: Joint chunk, power and bit allocation. IEEE Transactions on Communications, 60(2), 499–509.

    Article  Google Scholar 

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

    Book  MATH  Google Scholar 

  18. Lee, B. O., Je, H. W., Shin, E.-S., & Lee, K. B. (2009). A novel uplink MIMO transmission scheme in a multicell environment. IEEE Transactions on Wireless Communications, 8(10), 4981–4987.

    Article  Google Scholar 

  19. Karipidis, E., Gründinger, A., Lindblom, J., & Larsson, E. G. (2009). Pareto-optimal beamforming for the MISO interference channel with partial CSI. In Proceedings of the IEEE international workshop on computational advances in multi-sensor adaptive processing (CAMSAP), Aruba, Dutch Antilles (pp. 5–8).

Download references

Acknowledgements

This work was supported in part by National Natural Science Foundation of China (NSFC) (Grant Nos. 61501113, 61271205, 61521061, 61372100, 61571120), and Jiangsu Provincial Natural Science Foundation (Grant Nos. BK20150630, BK20151415).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jiamin Li.

Appendices

Appendix 1: Proof of Proposition 2

Considering the second entry of \(\mathbf{h }_{kk}\), one has

$$\begin{aligned}\frac{\tilde{\mathbf{h }}_{kk}^{H}\tilde{\mathbf{h }}_{kj}^{\perp }}{\Vert \tilde{\mathbf{h }}_{kj}^{\perp }\Vert \Vert \tilde{\mathbf{h }}_{kk}\Vert }& = \frac{\tilde{\mathbf{h }}_{kk}^{H}\left( \tilde{\mathbf{h }}_{kk}-\frac{\tilde{\mathbf{h }}_{kj}^H\tilde{\mathbf{h }}_{kk}}{\Vert \tilde{\mathbf{h }}_{kj}\Vert ^2}\tilde{\mathbf{h }}_{kj}\right) }{\Vert \tilde{\mathbf{h }}_{kj}^{\perp }\Vert \Vert \tilde{\mathbf{h }}_{kk}\Vert }\nonumber \\&\overset{\left( a\right) }{=}\frac{\Vert \tilde{\mathbf{h }}_{kk}\Vert ^2-\frac{\tilde{\mathbf{h }}_{kj}^H\tilde{\mathbf{h }}_{kk}\tilde{\mathbf{h }}_{kk}^H\tilde{\mathbf{h }}_{kj}}{\Vert \tilde{\mathbf{h }}_{kj}\Vert ^2}}{\Vert \tilde{\mathbf{h}}_{kj}^{\perp }\Vert \Vert \tilde{\mathbf{h }}_{kk}\Vert }\nonumber \\ & = \frac{\Vert \tilde{\mathbf{h }}_{kk}\Vert ^2-\frac{|\langle \tilde{\mathbf{h }}_{kk},\tilde{\mathbf{h }}_{kj}\rangle |^2}{\Vert \tilde{\mathbf{h }}_{kj}\Vert ^2}}{\Vert \tilde{\mathbf{h }}_{kj}^{\perp }\Vert \Vert \tilde{\mathbf{h }}_{kk}\Vert }\nonumber \\&\overset{\left( b\right) }{\ge }\frac{\Vert \tilde{\mathbf{h }}_{kk}\Vert ^2-\frac{\Vert \tilde{\mathbf{h }}_{kk}\Vert ^2\Vert \tilde{\mathbf{h }}_{kj}\Vert ^2}{\Vert \tilde{\mathbf{h }}_{kj}\Vert ^2}}{\Vert \tilde{\mathbf{h }}_{kj}^{\perp }\Vert \Vert \tilde{\mathbf{h }}_{kk}\Vert }\nonumber \\& = 0\nonumber \end{aligned}$$
(44)

where \(\left( a\right)\) is because \(\tilde{\mathbf{h }}_{kj}^H\tilde{\mathbf{h }}_{kk}\) is a scalar, \(\left( b\right)\) is by the Cauchy–Schwartz inequality \(|\langle \tilde{\mathbf{h }}_{kk},\tilde{\mathbf{h }}_{kj}\rangle |^2\le \Vert \tilde{\mathbf{h }}_{kk}\Vert ^2\Vert \tilde{\mathbf{h }}_{kj}\Vert ^2\), equality holds if and only if \(\tilde{\mathbf{h }}_{kk}\) and \(\tilde{\mathbf{h }}_{kj}\) are linearly dependent. Since the linearly independent channels are considered in this paper, the second entry of \(\mathbf{h }_{kk}\) is a positive real number. Combining this property and Proposition 1, we can rewrite the equivalent channel vectors \(\mathbf{h }_{kk}\) and \(\mathbf{h }_{kj}\) as in (18) and (19) respectively. This completes the proof.

Appendix 2: Proof of Theorem 1

Firstly, we reformulate the optimization problem (17) with the newly defined angle variable \(\varphi _k\). For linearly independent channels, the original non-convex optimization problem (17) is equivalent to the following optimization problem for the same rate weight \(\mu _k\) [5]:

$$\begin{array}{ll} {\mathop {{\mathrm {max}}}\limits _{{\mathbf{S }}_k,\;{\mathbf{S }}_j}}&{}\quad \sum\limits_{k=1}^{2}\mu _k \log _2\left( 1+\frac{u_k^2}{\mathbf{h }_{jk}^{H}{{\mathbf{S }}}_{j}\mathbf{h }_{jk}+\sigma ^2}\right) \\ \text {s.t.}&{}\quad \mathbf{h }_{kk}^{H}{{\mathbf{S }}}_k\mathbf{h }_{kk}=u_k^2, \\ &{}\quad {{\text {tr}}}({{\mathbf{S }}}_k)= P_k,\quad j\ne k,\;\; k, j\in \{1,2\}, \end{array}$$
(45)

where \({{\mathbf{S }}}_k=\mathbf{w }_k\mathbf{w }_k^H\), and the power constraint at the \({\text {BS}}_k\) is \({\text {tr}}({{\mathbf{S }}}_k)= P_k\) which results from that full transmit power should be used to obtain the optimal solution to the WSRMax problem (17) when the channels are linearly independent [3, 11].

  1. (1)

    Useful signal power From the definition (22), the useful signal power term of (45) can be expressed with \(\varphi _k\) as

    $$\begin{aligned} u_k^2=\mathbf{h }_{kk}^H {{\mathbf{S }}}_k \mathbf{h }_{kk}=P_k\sin ^2{\varphi _k}. \end{aligned}$$
    (46)
  1. (2)

    Interference power Assume that the singular-value decompositions (SVDs) of \(\mathbf{h }_{kk}\)

and \(\mathbf{h }_{kj}\) are

$$\begin{aligned} \mathbf{h }_{kk}=\mathbf{U }_{kk}\left[ {\begin{array}{l}{\Vert \mathbf{h }_{kk} \Vert }\\ {0}\end{array}}\right] =\mathbf{U }_{kk}\left[ {\begin{array}{l}{1}\\ {0}\end{array}}\right] , \end{aligned}$$
(47)

and

$$\begin{aligned} \mathbf{h }_{kj}=\mathbf{U }_{kj}\left[ {\begin{array}{l}{\Vert \mathbf{h }_{kj} \Vert }\\ {0}\end{array}}\right] =\mathbf{U }_{kj}\left[ {\begin{array}{l}{d}\\ {0}\end{array}}\right] , \end{aligned}$$
(48)

where \(\mathbf{U }_{kk}\) and \(\mathbf{U }_{kj}\) are \(2\times 2\) unitary matrices. Define

$$\begin{aligned}\hat{\mathbf{h }}_{kj}\,\triangleq\,\mathbf{U }_{kk}^H\mathbf{h }_{kj}=\left[ {\begin{array}{l}{\alpha }\\ {\beta }\end{array}}\right] ,\end{aligned}$$
(49)
$$\begin{aligned}&\hat{\mathbf{S }}_k\,\triangleq\,{\mathbf{U}} _{kk}^H{\mathbf{S}} _k{\mathbf{U}} _{kk}=\left[ {\begin{array}{*{20}c}{\hat{S}_{11}}&{}{\hat{S}_{21}^H}\\ {\hat{S}_{21}}&{}{\hat{S}_{22}}\end{array}}\right], \end{aligned}$$
(50)

where all \(\alpha ,\, \beta ,\, \hat{S}_{11},\, \hat{S}_{21}\) and \(\hat{S}_{22}\) are scalars. Considering the useful signal power term in (46), we have

$$\begin{aligned} {\mathbf{h}} _{kk}^H{\mathbf{S}}_k {\mathbf{h}} _{kk}=\left[ {\begin{array}{*{20}l}{1}&{0}\end{array}}\right] {\mathbf{U}}_{kk}^H{\mathbf{S}} _k{\mathbf{U}} _{kk}\left[ {\begin{array}{*{20}c}{1}\\ {0}\end{array}}\right] =\hat{S}_{11}= u_k^2. \end{aligned}$$
(51)

Substituting (49), (50) and (51) into the interference power term of (45), we have

$$\begin{aligned} {\mathbf{h}} _{kj}^H{\mathbf{S}} _k {\mathbf{h}} _{kj}& = \left[ {\begin{array}{*{20}l}{\alpha ^H}&{\beta ^H}\end{array}}\right] {\mathbf{U}} _{kk}^H{\mathbf{S}} _k{\mathbf{U}} _{kk}\left[ {\begin{array}{*{20}l}{\alpha }\\ {\beta }\end{array}}\right] \nonumber \\& = |\alpha |^2 u_k^2 + \hat{S}_{21}\alpha \beta ^H+\hat{S}_{21}^H\alpha ^H\beta + \hat{S}_{22}|\beta |^2 \nonumber \\&\overset{(a)}{=} |\alpha |^2 u_k^2 + \hat{S}_{21}\alpha \beta ^H+\hat{S}_{21}^H\alpha ^H\beta +|\beta |^2(P_k - u_k^2), \end{aligned}$$
(52)

where (a) results from \({\text {tr}}({{\mathbf{S }}}_k)={\text {tr}}(\hat{{{\mathbf{S }}}}_{k})=\hat{S}_{11}+\hat{S}_{22}=P_k\). With the definition of

$$\begin{aligned} \tilde{S}_{21}\,\triangleq\,\frac{\beta ^H}{|\beta |}\hat{S}_{21}, \end{aligned}$$
(53)

we have

$$\begin{aligned} \mathbf{h }_{kj}^H{{\mathbf{S }}}_k \mathbf{h }_{kj}& = |\alpha |^2 u_k^2 + \hat{S}_{21}\alpha \beta ^H+\hat{S}_{21}^H\alpha ^H\beta +|\beta |^2(P_k - u_k^2) \nonumber \\& = |\alpha |^2u_k^2 + |\beta |\left( \tilde{S}_{21} \alpha + \alpha ^H\tilde{S}_{21}^H\right) + |\beta |^2(P_k - u_k^2) \nonumber \\&\overset{(a)}{\ge }|\alpha |^2u_k^2 -2|\beta ||\tilde{S}_{21}||\alpha | + |\beta |^2(P_k - u_k^2) \nonumber \\&\overset{(b)}{\ge }|\alpha |^2u_k^2 -2|\beta ||\alpha | u_k\sqrt{P_k-u_k^2} + |\beta |^2(P_k - u_k^2)2 \nonumber \\& = \left( |\alpha |u_k - |\beta |\sqrt{P_k- u_k^2} \right) ^2, \end{aligned}$$
(54)

where (a) results from \(\tilde{S}_{21}\alpha + \alpha ^H\tilde{S}_{21}^H \ge -2|\tilde{S}_{21}||\alpha |\) since \(\left( \tilde{S}_{21}\alpha + \alpha ^H\tilde{S}_{21}^H \right) ^2=\left( 2\text {Re}(\tilde{S}_{21}\alpha ) \right) ^2 \le 4|\tilde{S}_{21}|^2|\alpha |^2\), and (b) is due to \(\hat{S}_{11}\hat{S}_{22}= u_k^2(P_k-u_k^2) \ge |\hat{S}_{21}|^2=|\tilde{S}_{21}|^2\) since \(\hat{{{\mathbf{S }}}}_k\succeq 0\). The equality of (a) holds if and only if \(\tilde{S}_{21}\alpha\) is non-positive real number, i.e.,

$$\begin{aligned} \tilde{S}_{21}\alpha = (\tilde{S}_{21}\alpha )^H = -|\tilde{S}_{21}||\alpha |. \end{aligned}$$
(55)

The equality of (b) holds under the condition

$$\begin{aligned} \tilde{S}_{21} = \frac{\beta ^H \alpha ^H}{|\beta ||\alpha |}u_k\sqrt{P_k - u_k^2}, \end{aligned}$$
(56)

where the choice of \(\tilde{S}_{21}\) ensures that (55) is satisfied. Under the conditions of (55) and (56), the minimum of the interference power term in (54) can be given by

$$\begin{aligned} \left( \mathbf{h }_{kj}^H{{\mathbf{S }}}_k \mathbf{h }_{kj} \right) _{\text {min}}&\overset{(a)}{=}\,d_k^2\left( \sqrt{a_k^2+b_k^2} u_k - \sqrt{1-(a_k^2+b_k^2)}\sqrt{P_k - u_k^2} \right) ^2 \nonumber \\&\overset{(b)}{=}\,P_k d_k^2 \left( \cos {\theta _k} \frac{u_k}{\sqrt{P_k}} - \sin {\theta _k} \sqrt{\frac{P_k - u_k^2}{P_k}} \right) ^2 \nonumber \\&\overset{(c)}{=}\,P_k d_k^2 \left( \cos {\theta _k}\sin {\varphi _k} - \sin {\theta _k}\cos {\varphi _k} \right) ^2 \nonumber \\ &= P_k \Vert \mathbf{h }_{kj} \Vert ^2 \sin ^2{\left( \varphi _k - \theta _k\right) } \end{aligned}$$
(57)

where (a) results from \(\alpha = d_k(a_k-i b_k)\) and \(|\beta | = \sqrt{\Vert \mathbf{h }_{kj} \Vert ^2 - |\alpha |^2}\) which are obtained from (49), (b) and (c) result from (20) and (22), respectively.

  1. (3)

    Value range of \(\varphi _k\) The value range of \(u_k\) is

    $$\begin{aligned} u_k\in \left[ 0,\sqrt{P_k}\Vert \mathbf{h }_{kk}\Vert \right] =\left[ 0,\sqrt{P_k}\right] , \end{aligned}$$
    (58)

which results from \(\Vert \mathbf{h }_{kk}\Vert =1\). Combined with the definition (22), it can be seen that the value range of \(\varphi _k\) is

$$\begin{aligned} \varphi _k\in \left[ 0,\pi /2\right] . \end{aligned}$$
(59)

Combining (45), (46), (57) and (59), the WSRMax problem (17) can be reformulated with the angle variables \(\varphi _k\) as in (23). Next, we prove that the optimal beamforming vector \(\mathbf{w }_k^{*}\) has the form as in (24). From (47), we have

$$\begin{aligned} u_{11}=a_k+ib_k,\quad {\text {and}}\quad u_{21}= c_k, \end{aligned}$$
(60)

where \(u_{ij}=\left( \mathbf{U }_{kk}\right) _{ij}\). Due to the properties of unitary matrix, the unitary matrix \(\mathbf{U }_{kk}\) can be given by

$$\begin{aligned} \mathbf{U }_{kk}=\left[ {\begin{array}{ll}{a_k+ib_k}&{}{-c_k}\\ {c_k}&{}{a_k-ib_k}\end{array}}\right] . \end{aligned}$$
(61)

From (53) and (56), we obtain

$$\begin{aligned} \hat{S}_{21}=\frac{\alpha ^H }{|\alpha |}u_k\sqrt{P_k - u_k^2}. \end{aligned}$$
(62)

Combining (50), (51), (52) and (62), we have

$$\begin{aligned} \hat{{{\mathbf{S }}}}_{k}=\left[ {\begin{array}{ll}{u_k^2}&{}{\frac{\alpha }{|\alpha |}u_k\sqrt{P_k - u_k^2}}\\ {\frac{\alpha ^H }{|\alpha |}u_k\sqrt{P_k - u_k^2}}&{}{P_k-u_k^2}\end{array}}\right] . \end{aligned}$$
(63)

Then, \({{\mathbf{S }}}_k\) can be obtained from (50), (61) and (63) as

$$\begin{aligned} {{\mathbf{S }}}_{k}=\mathbf{U }_{kk}\hat{{{\mathbf{S }}}}_{k}\mathbf{U }_{kk}^H \overset{(a)}{=} P_k \left[ {\begin{array}{ll}{\sin ^2{(\varphi _k-\theta _k)} }&{}{B\left( \cos {\eta _k}+i\sin {\eta _k} \right) }\\ {B\left( \cos {\eta _k}-i\sin {\eta _k} \right) }&{}{\cos ^2{(\varphi _k-\theta _k )}} \end{array}}\right] , \end{aligned}$$
(64)

where \(B=\sin {(\varphi _k-\theta _k)}\cos {(\varphi _k-\theta _k)},\, (a)\) results from \(\cos {\theta _k}=\sqrt{a_k^2+b_k^2},\, \sin {\theta _k}=c_k,\, \sin {\eta _k}=b_k/\sqrt{a_k^2+b_k^2},\, \cos {\eta _k}=a_k/\sqrt{a_k^2+b_k^2},\, \sin {\varphi _k}=u_k/\sqrt{P_k}\) and \(\cos {\varphi _k}=\sqrt{(P_k-u_k^2)/P_k}\), which can be obtained from (20), (21) and (22), respectively. Moreover, since it has been proved in [1] that for the two-user MISO IC any point on the Pareto boundary is achievable with the weighted sum of the MRT and ZF beamforming vectors with some set of real-valued parameters which are in the range of [0, 1], and the MRT and ZF beamforming vectors at \({\text {BS}}_k\) can be given by

$$\begin{aligned} {\mathbf{w }}_k^{\text {MRT}} = \frac{{\mathbf{h} }_{kk}}{\Vert {\mathbf{h }}_{kk}\Vert }\;=\; \left[ a_k+ib_k,\; c_k\right] ^T,\end{aligned}$$
(65)
$${\mathbf{w }}_k^{\text {ZF}} = \frac{\varPi _{\mathbf{h}_{kj}}^{\perp } {\mathbf{h }_{kk}}}{\Vert \varPi _{\mathbf{h }_{kj}}^{\perp } {\mathbf{h }}_{kk}\Vert }\;= \; \left[ 0, \; 1\right]^T,$$
(66)

where \(\varPi _{{\mathbf{h }}_{kj}}^{\perp }\,\triangleq\,{\mathbf{I}} -{\mathbf{h}}_{kj}\left( {\mathbf{h }}_{kj}^H {\mathbf{h }}_{kj} \right) ^{-1} {\mathbf{h }}_{kj}^H\) is the orthogonal projection onto the orthogonal complement of \(\mathbf{h }_{kj}\), the optimal beamforming vectors can be expressed as the following form without loss of optimality

$$\begin{aligned} \mathbf{w }_k^{*} = \sqrt{P_k}\left[ x_k+iy_k,\; z_k \right] ^T, \end{aligned}$$
(67)

where \(x_k,\, y_k\) and \(z_k\) are real numbers and \(z_k > 0\) which results from \(c_k> 0\). Therefore, the signal correlation matrix \({{\mathbf{S }}}_k^{*}\) can be given with \(\mathbf{w }_k^{*}\) by

$$\begin{aligned} {{\mathbf{S }}}_k^{*} = \mathbf{w }_k^{*} \left( \mathbf{w }_k^{*}\right) ^H = P_k \left[ {\begin{array}{ll}{x_k^2+y_k^2}&{}{x_k z_k+i y_k z_k}\\ {x_k z_k-i y_k z_k}&{}{z_k^2}\end{array}}\right] . \end{aligned}$$
(68)

Combining (64) and (68), we have

$$\begin{aligned} \left\{ \begin{array}{llll} &{}x_k^2+y_k^2 = \sin ^2{(\varphi _k-\theta _k)}\\ &{}x_k z_k = \sin {(\varphi _k-\theta _k)}\cos {(\varphi _k-\theta _k)}\cos {\eta _k} \\ &{}y_k z_k = \sin {(\varphi _k-\theta _k)}\cos {(\varphi _k-\theta _k)}\sin {\eta _k} \\ &{}z_k^2 = \cos ^2{(\varphi _k - \theta _k)}\\ \end{array} \right. \end{aligned}$$
(69)

The remaining challenge is to solve the Eq. (69). From [14, 19], the range of useful signal power \(u_k^2\) obtained by the optimal beamforming vector is given by

$$\begin{aligned} u_k^2 \in \left[ P_k \left| \mathbf{h }_{kk}^H \mathbf{w }_k^{\text {ZF}}\right| ^2,\; P_k \left| \mathbf{h }_{kk}^H \mathbf{w }_k^{\text {MRT}}\right| ^2 \right] =\left[ P_k c_k^2,\; P_k \right] . \end{aligned}$$
(70)

Substituting (70) into (22) yields the range of the optimal angle variable \(\varphi _k^{*}\),

$$\begin{aligned} \varphi _k^{*} \in \left[ {\text {atan}}\left( \sqrt{\frac{c_k^2}{1-c_k^2} }\right) , \; \frac{\pi }{2} \right]\,\overset{(a)}{=}\,\left[ {\text {asin}}(c_k), \; \frac{\pi }{2} \right] , \end{aligned}$$
(71)

where (a) results from \(\Vert \mathbf{h }_{kk}\Vert ^2 = a_k^2 + b_k^2 + c_k^2 =1\). Since \(\theta _k = {\text {asin}}{(c_k)}\) which can be obtained from (20), we have \(0\le \varphi _k^{*} - \theta _k \le \frac{\pi }{2} - {\text {asin}}(c_k)\). Hence, \(\cos (\varphi _k^{*}-\theta _k)\ge 0\). Recalling that \(z_k> 0\), by solving the Eq. (70), we obtain

$$\begin{aligned} \left\{ \begin{array}{lll} &{x_{k}} = \sin {(\varphi _k^{*}-\theta _k)}\cos {\eta _k} \\ &{y_{k}} = \sin {(\varphi _k^{*}-\theta _k)}\sin {\eta _k} \\ &{z_{k}} = \cos {(\varphi _k^{*}-\theta _k)} \end{array} \right. \end{aligned}$$
(72)

Consequently, the beamforming vector \(\mathbf{w }_k^{*}\) can be expressed with \(\varphi _k^{*}\) as in (24).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, J., Wang, D., Zhu, P. et al. Optimality Properties, Closed-Form Parameterizations and Distributed Strategy of the Two-User MISO Interference Channel. Wireless Pers Commun 96, 833–853 (2017). https://doi.org/10.1007/s11277-017-4204-5

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11277-017-4204-5

Keywords

Navigation