Skip to main content
Log in

On 2-D Direction-of-Arrival Estimation Performance for Rank Reduction Estimator in Presence of Unexpected Modeling Errors

  • Published:
Circuits, Systems, and Signal Processing Aims and scope Submit manuscript

Abstract

The rank reduction estimator (RARE) is one kind of autocalibration method used in the presence of sensor errors. It demonstrates high accuracy of direction-of-arrival (DOA) estimation in the absence of multidimensional search or iteration. However, its estimation performance is affected by “unexpected modeling errors.” There is a lack of research regarding the performance of 2-D RARE estimation, although 2-D RARE estimation is extensively employed in applications. This paper presents a theoretical derivation for the closed-form expression of the mean square error (MSE) of 2-D RARE estimation under the influence of small unexpected modeling errors in the first order analysis. First, three definitions of 2-D joint direction-finding success are introduced, in order to establish the criterion for estimate performance. Then corresponding theoretical formulas for three probabilities of direction-finding success are given with the circularly Gaussian assumption of unexpected modeling errors, and their relations are discussed. Finally, the results of simulations utilizing our analysis method are demonstrated, verifying the effectiveness of the MSE expression and the formulas for probabilities of success. Therefore, our first order approximation provides a good prediction of the necessary calibration accuracy in the presence of unexpected modeling errors in order to help RARE meet an expected performance specification.

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
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

References

  1. K.H. Biyari, W.C. Lindsey, Statistical distributions of Hermitian quadratic forms in complex Gaussian variables. IEEE Trans. Inf. Theory 39(3), 1076–1082 (1993)

    Article  MATH  MathSciNet  Google Scholar 

  2. T.N. Dugina, G.V. Martynov, Computing the distribution function of the ratio of quadratic forms in normal variables. J. Math. Sci. 53(6), 628–631 (1991)

    Article  Google Scholar 

  3. A. Ferréol, E. Boyer, P. Larzabal, Low-cost algorithm for some bearing estimation methods in presence of separable nuisance parameters. Electron. Lett. 40(15), 966–967 (2004)

    Article  Google Scholar 

  4. A. Ferréol, E. Boyer, P. Larzabal, Asymptotic performance for subspace bearing methods with separable nuisance parameters in presence of modeling errors, in IEEE International Conference on Acoustics, Speech, and Signal Processing, 2005. Proceedings, vol. 4 (2005), pp. iv/953–iv/956. 2005

    Chapter  Google Scholar 

  5. A. Ferréol, P. Larzabal, M. Viberg, On the resolution probability of MUSIC in presence of modeling errors. IEEE Trans. Signal Process. 56(5), 1945–1953 (2008)

    Article  MathSciNet  Google Scholar 

  6. A. Ferréol, P. Larzabal, M. Viberg, Statistical analysis of the MUSIC algorithm in the presence of modeling errors, taking into account the resolution probability. IEEE Trans. Signal Process. 58(8), 4156–4166 (2010)

    Article  MathSciNet  Google Scholar 

  7. B. Friedlander, A sensitivity analysis of the MUSIC algorithm. IEEE Trans. Acoust. Speech Signal Process. 38, 1740–1751 (1990)

    Article  Google Scholar 

  8. M. Grice, J. Rodenkirch, Direction of arrival estimation using advanced signal processing, in 3rd International Conference on Recent Advances in Space Technologies (2007), pp. 515–522

    Google Scholar 

  9. X.Q. Hu, H. Chen, Y.L. Wang, J.W. Chen, A self-calibration algorithm for cross array in the presence of mutual coupling. Sci. China, Inf. Sci. 54(4), 836–848 (2011)

    Article  MathSciNet  Google Scholar 

  10. E. Hung, Matrix-construction calibration method for antenna arrays. IEEE Trans. Aerosp. Electron. Syst. 36(3), 819–828 (2000)

    Article  Google Scholar 

  11. J.P. Imhof, Computing the distribution of quadratic forms in normal variables. Biometrika 48(12), 419–426 (1961)

    MATH  MathSciNet  Google Scholar 

  12. Y.M. Li, M.H. Er, Theoretical analyses of gain and phase error calibration with optimal implementation for linear equispaced array. IEEE Trans. Signal Process. 54(2), 712–723 (2006)

    Article  Google Scholar 

  13. J.L. Liang, X.J. Zeng, W.Y. Wang, H.Y. Chen, L-shaped array-based elevation and azimuth direction finding in the presence of mutual coupling. Signal Process. 91(5), 1319–1328 (2011)

    Article  MATH  Google Scholar 

  14. C. Liu, Z.F. Ye, Y.F. Zhang, Auto-calibration algorithm for mutual coupling of planar array. Signal Process. 90(3), 784–794 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  15. T. Nizar, M.K. Hyuck, L-shaped 2-dimensional arrival angle estimation with propagator method. IEEE Trans. Antennas Propag. 53(5), 1622–1630 (2005)

    Article  Google Scholar 

  16. J.E.F.D. Rio, M.F. Catedra-Perez, The matrix pencil method for two-dimensional direction of arrival estimation employing an L-shaped array. IEEE Trans. Antennas Propag. 45(11), 1693–1694 (1997)

    Article  Google Scholar 

  17. H. Schippers, W. Schaap, J.M.P.C.M. Visser, G. Vos, Conformal phased array antennas on vibrating structures, in The 2nd European Workshop on Conformal Antennas, Bonn (2003)

    Google Scholar 

  18. H. Schippers, J.H. van Tongeren, P. Knott, T. Deloues, P. Lacomme, M.R. Scherbarth, Vibrating antennas and compensation techniques research in NATO/RTO/SET 087/RTG 50, in IEEE Aerospace Conference, Big Sky, MT (2007), pp. 3–10

    Google Scholar 

  19. K. Stavropoulos, A. Manikas, Array calibration in the presence of unknown sensor characteristics and mutual coupling, in Proceedings of the European Signal Processing Conference, EUSIPO 2000 (2000), pp. 1417–1420

    Google Scholar 

  20. P. Stoica, A. Nehorai, MUSIC, maximum likelihood, and Cramer-Rao bound. IEEE Trans. Acoust. Speech Signal Process. 37(5), 720–741 (1989)

    Article  MATH  MathSciNet  Google Scholar 

  21. M. Viberg, A.L. Swindlehurst, Analysis of the combined effects of finite samples and model errors on array processing performance. IEEE Trans. Signal Process. 42(1), 3073–3083 (1994)

    Article  Google Scholar 

  22. D. Wang, Y. Wu, Performance analysis of rank reduction estimator in the presence of unexpected modeling errors. J. Commun. 32(8), 81–90 (2011)

    Google Scholar 

  23. D. Wang, Y. Wu, Statistical characteristics and resolution probability of rank reduction spatial spectrum in the presence of unexpected model errors. J. Appl. Sci. 29(2), 176–184 (2011)

    Google Scholar 

  24. B.-h. Wang, Y.-l. Wang, H. Chen, Array calibration of angularly dependent gain and phase uncertainties with instrumental sensors, in IEEE International Symposium on Phased Array Systems and Technology (2003), pp. 182–186

    Chapter  Google Scholar 

  25. B.H. Wang, Y.L. Wang, H. Chen, Robust DOA estimation and array calibration in the presence of mutual coupling for uniform linear array. Sci. China, Ser. F Inf. Sci. 47(3), 348–361 (2004)

    Article  Google Scholar 

  26. B. Wang, Y. Wang, H. Chen, Y. Guo, Array calibration of angularly dependent gain and phase uncertainties with carry-on instrumental sensors. Sci. China, Ser. F Inf. Sci. 47(6), 777–792 (2004)

    Article  Google Scholar 

  27. A.J. Weiss, B. Friedlander, Mutual coupling effects on phase-only direction finding. IEEE Trans. Antennas Propag. 40(5), 535–541 (1992)

    Article  Google Scholar 

  28. S.J. Wijnholds, A.J. Veen, Multisource self-calibration for sensor arrays. IEEE Trans. Signal Process. 57(9), 3512–3522 (2009)

    Article  MathSciNet  Google Scholar 

  29. B. Wu, H. Chen, J.-d. Li, Direction of arrival (DOA) estimation under non-Toeplitz mutual coupling matrix of ULA. Radar Sci. Technol. 7(5), 358–364 (2009)

    MathSciNet  Google Scholar 

  30. L. Yang, K.C. Ho, Alleviating sensor position error in source location using calibration emitters at inaccurate locations. IEEE Trans. Signal Process. 58(1), 67–83 (2010)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the Editor-in-Chief, Prof. M.N.S. Swamy, and the Associate Editor for their helpful suggestions in revising and improving our paper.

This work was supported by the National Natural Science Foundation of China under Grant 61201381 and the Future Development Foundation of Zhengzhou Information Science and Technology College under Grant YP12JJ202057.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jie-xin Yin.

Appendices

Appendix A

In this appendix, we prove (21). Before the proof, we will introduce Lemma A.1 and Theorem A.1.

Lemma A.1

A sequence composed of 1,2,…,K can be sorted from 1 to K through several exchanges between two integers, and the number of exchange times has the same parity as the inverse number of the original sequence.

Theorem A.1

Introduce the exchange operator J transforming a sequence composed of 1,2,…,K to the sorted sequence from 1 to K. When the sorted sequence is transformed by J, its inverse number has the same parity as that of the original sequence.

Proof

The number of exchange times of J has the same parity as the inverse number of the original sequence according to Lemma A.1. Besides, τ(1,2,…,K), the inverse number of the sorted sequence equals zero. One exchange leads to the change of the parity of the inverse number. Thus the inverse number of the sorted sequence after transformation by J will have the same parity as the number of exchange times of J. Based on these equivalence relations, the conclusion in Theorem A.1 is proven.

Proof of (21)

We only prove the first equation here; it is easy to extend the proof to the second and third ones.

First, we introduce \(\tilde{\mathbf{T}}_{k}(\theta_{m},\varphi_{m}) \in \mathbf{C}^{M \times K}\), k=1,2,…,K as T(θ m ,φ m ) recomposed with the kth column replaced by \(\mathbf{t}_{k}^{(\theta \theta )}(\theta_{m},\varphi_{m}) \), its second partial derivative with respect to azimuth at the true DOAs (θ m ,φ m ). The rank defect of both \(\tilde{\mathbf{T}}_{k}^{H}(\theta_{m},\varphi_{m})\boldsymbol{\pi}\mathbf{ T}(\theta_{m},\varphi_{m}) \) and its transposed matrix follows from the subspace orthogonality principle, which is shown as

(46)

Equation (46) provides an important proof for the latter discussion.

Present \(w_{ij}(\theta,\varphi,\boldsymbol{\varepsilon} ) = [ W(\theta,\varphi,\boldsymbol{\varepsilon} ) ]_{ij} = \mathbf{t}_{i}^{\mathbf{H}}(\theta,\varphi )\hat{\boldsymbol{\pi}}\mathbf{ t}_{j}(\theta,\varphi ) \), and expand D θθ (θ m ,φ m ,0) defined in (16):

(47)

in which \(w_{ij}^{(\theta \theta )}(\theta_{m},\varphi_{m},\mathbf{0}) = \frac{\partial^{2}w_{ij}(\theta,\varphi,\boldsymbol{\varepsilon} )}{\partial \theta^{2}} \vert _{(\theta,\varphi,\boldsymbol{\varepsilon} ) = (\theta_{m},\varphi_{m},\mathbf{0})} = f(\mathbf{t}_{i}^{(\theta \theta )},\mathbf{t}_{j}) + 2f(\mathbf{t}_{i}^{(\theta )},\mathbf{t}_{j}^{(\theta )}) + f(\mathbf{t}_{i},\mathbf{t}_{j}^{(\theta \theta )}) \) where f(•,•) is defined in (23). \(\mathbf{t}_{i}^{(\theta )} \) is short for the first partial derivative of t i (θ m ,φ m ) as to azimuth and \(\mathbf{t}_{i}^{(\theta \theta )} \) is short for \(\mathbf{t}_{i}^{(\theta \theta )}(\theta_{m},\varphi_{m}) \).

Now we reduce the higher order terms in (47):

(48)

where \(\tilde{w}_{ij}^{(\theta \theta )}(\theta_{m},\varphi_{m},\mathbf{0}) \) denotes \(2f(\mathbf{t}_{i}^{(\theta )},\mathbf{t}_{j}^{(\theta )}). Z_{1},Z_{2} \) can be left out because

(49)

Equation (49) is validated based on (46). Theorem A.1 and (46) yield

(50)

Discarding Z 1,Z 2, Eq. (48) becomes

(51)

Finally, substituting (51) back into (47), we have

(52)

where the notation is the same as in (21). Up to this point, the first equation in (21) is proven.

Appendix B

In this appendix, we prove (24). Based on the subspace orthogonality principle [20], the following equation can be established if ε m =0, regardless of the value of other unexpected modeling errors ε n (nm):

$$ \hat{\boldsymbol{\pi}}\mathbf{ T}(\theta_{m},\varphi_{m}) \boldsymbol{\rho} (\theta_{m},\varphi_{m},\boldsymbol{ \eta}_{m}) = \mathbf{0} $$
(53)

Now \(\mathbf{W}(\theta_{m},\varphi_{m},\boldsymbol{\varepsilon} ) = \mathbf{T}(\theta_{m},\varphi_{m})^{H}\hat{\boldsymbol{\pi}}\mathbf{ T}(\theta_{m},\varphi_{m}) \) is still the rank defect which is given by

$$ D(\theta_{m},\varphi_{m},\boldsymbol{\varepsilon} ) \vert _{\boldsymbol{\varepsilon}_{m} = \mathbf{0}} = \det \bigl[\mathbf{W}(\theta_{m}, \varphi_{m},\boldsymbol{\varepsilon} )\bigr] \big \vert _{\boldsymbol{\varepsilon}_{m} = \mathbf{0}} = 0 $$
(54)

Thus

$$ D_{\theta} (\theta_{m},\varphi_{m}, \boldsymbol{\varepsilon} ) \vert _{\boldsymbol{\varepsilon}_{m} = \mathbf{0}} = 0, \qquad D_{\varphi} ( \theta_{m},\varphi_{m},\boldsymbol{\varepsilon} ) \vert _{\boldsymbol{\varepsilon}_{m} = \mathbf{0}} = 0 $$
(55)

Because D θ (θ,φ,ε) and D φ (θ,φ,ε) are constant on (θ m ,φ m ,ε m =0) according to (55), the gradient vectors for each unexpected modeling errors except ε m are zeros on (θ m ,φ m ,ε m =0):

$$ \frac{\partial D_{\theta} (\theta_{m},\varphi_{m},\boldsymbol{\varepsilon} ) \vert _{\boldsymbol{\varepsilon}_{m} = \mathbf{0}}}{\partial \boldsymbol{\varepsilon}_{n}^{(x)}} = \mathbf{0},\qquad \frac{\partial D_{\varphi} (\theta_{m},\varphi_{m},\boldsymbol{\varepsilon} ) \vert _{\boldsymbol{\varepsilon}_{m} = \mathbf{0}}}{\partial \boldsymbol{\varepsilon}_{n}^{(x)}} = \mathbf{0}, \quad x \in \{ r,i\} $$
(56)

It is proved that the operation order of derivation with respect to \(\boldsymbol{\varepsilon}_{n}^{(x)},x \in \{ r,i \} \) and assignment of θ,φ can be changed, and we have

(57)

Therefore, \(\mathbf{D}_{\theta \boldsymbol{\varepsilon}_{{n}}^{(x)}}(\theta_{m},\varphi_{m},\mathbf{0}) = \mathbf{0}\), \(\mathbf{D}_{\varphi \boldsymbol{\varepsilon}_{{n}}^{(x)}}(\theta_{m},\varphi_{m},\mathbf{0}) = \mathbf{0}\), x∈{r,i} in terms of nm. The proof of (24) is completed.

Appendix C

In this appendix, we prove (26).

We choose

$$\mathbf{D}_{\boldsymbol{\theta} \boldsymbol{\varepsilon}_{n}^{(r)}} (\theta_{m},\varphi_{m},\mathbf{0}) = \sum_{(i_{1},i_{2}, \ldots, i_{K})} ( - 1)^{\tau (i_{1},i_{2}, \ldots,i_{K})} \sum_{k = 1}^{2K} \boldsymbol{\varGamma}_{n}^{(r)}\bigl[\boldsymbol{\varSigma}_{k}^{\theta} (i_{1},i_{2}, \ldots, i_{K})\bigr] $$

to prove; it is also easy to extend the proof to others.

(58)

where the notation is the same as in (26). The proof is finished.

Appendix D

In this appendix, we prove Proposition 2. To begin with, we give Lemma D.1, which is useful for the proof of the proposition.

Lemma D.1

[1]

Assume xχ 2(n), i.e., x is chi-square distributed with n degrees of freedom. Its characteristic function is thus given by M x (t)=(1−2it)n/2.

Proof of Proposition 2

Introduce vector \(\boldsymbol{\delta}_{0} = \mathbf{R}_{\boldsymbol{\delta}}^{ - 1/2}\boldsymbol{\delta} \); thus δ 0 is Gaussian distributed with mean zero, and its variance matrix is the identity matrix. Then y is rewritten as

$$ y = \boldsymbol{\delta}_{0}^{T}\mathbf{R}_{\boldsymbol{\delta}}^{1/2T} \mathbf{QR}_{\boldsymbol{\delta}}^{1/2}\boldsymbol{\delta}_{0} = \sum_{n = 1}^{D} \lambda_{n}\bigl( \mathbf{u}_{n}^{T}\boldsymbol{\delta}_{0} \bigr)^{2} $$
(59)

where u n indicates eigenvectors of \(\mathbf{R}_{\boldsymbol{\delta}}^{1/2T}\mathbf{QR}_{\boldsymbol{\delta}}^{1/2} \). Each eigenvector is orthogonal to others, so we have \(E [ (\mathbf{u}_{n}^{T}\boldsymbol{\delta}_{0})(\mathbf{u}_{m}^{T}\boldsymbol{\delta}_{0})^{T} ] = \mathbf{u}_{n}^{T}\mathbf{u}_{m} = 0 \) when nm which implies that Gaussian random variable \(\mathbf{u}_{n}^{T}\boldsymbol{\delta}_{0} \) is statistically independent of \(\mathbf{u}_{m}^{T}\boldsymbol{\delta}_{0} \). Moreover, \((\mathbf{u}_{n}^{T}\boldsymbol{\delta}_{0})^{2} \) from n=1 to D (the rank of Q) are independently chi-square distributed with one degree of freedom. Using Lemma D.1, the characteristic function of \(x_{n} = \lambda_{n}(\mathbf{u}_{n}^{T}\boldsymbol{\delta}_{0})^{2} \) can be expressed as \(M_{x_{n}}(t) = (1 - 2it\lambda_{n})^{ - 1/2} \). If one random variable can be expressed as the sum of several independent variables, its characteristic function equals the product of those specified characteristic functions. Consequently, the characteristic function of y follows, as shown in (36).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Yin, Jx., Wu, Y. & Wang, D. On 2-D Direction-of-Arrival Estimation Performance for Rank Reduction Estimator in Presence of Unexpected Modeling Errors. Circuits Syst Signal Process 33, 515–547 (2014). https://doi.org/10.1007/s00034-013-9654-8

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00034-013-9654-8

Keywords

Navigation