• Aucun résultat trouvé

Direct derivation of the stochastic CRB of DOA estimation for rectilinear sources

N/A
N/A
Protected

Academic year: 2021

Partager "Direct derivation of the stochastic CRB of DOA estimation for rectilinear sources"

Copied!
6
0
0

Texte intégral

(1)

HAL Id: hal-01629444

https://hal.archives-ouvertes.fr/hal-01629444

Submitted on 6 Nov 2017

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

estimation for rectilinear sources

Habti Abeida, Jean-Pierre Delmas

To cite this version:

Habti Abeida, Jean-Pierre Delmas. Direct derivation of the stochastic CRB of DOA estimation for

rectilinear sources. IEEE Signal Processing Letters, Institute of Electrical and Electronics Engineers,

2017, 24 (10), pp.1522 - 1526. �10.1109/LSP.2017.2744673�. �hal-01629444�

(2)

Direct derivation of the stochastic CRB of DOA

estimation for rectilinear sources

Habti Abeida and Jean Pierre Delmas, Senior member, IEEE

Abstract—Several direction of arrival (DOA) estimation

algorithms have been proposed to exploit the structure of rectilinear or strictly second-order noncircular signals. But until now, only the compact closed-form expressions of the corresponding deterministic Cram´er Rao bound (DCRB) have been derived because it is much easier to derive than the stochastic CRB (SCRB). As this latter bound is asymptotically achievable by the maximum likelihood (ML) estimator, while the DCRB is unattainable, it is important to have a compact closed-form expression for this SCRB to assess the perfor-mance of DOA estimation algorithms for rectilinear signals. The aim of this paper is to derive this expression directly from the Slepian-Bangs formula including in particular the case of prior knowledge of uncorrelated or coherent sources. Some properties of these SCRBs are proved and numerical illustrations are given.

Index Terms—Deterministic and stochastic Cram´er Rao

bound (CRB), direction of arrival (DOA), circular, noncircular, rectilinear, strictly noncircular, Slepian-Bangs formula.

I. INTRODUCTION

Various DOA estimation algorithms such as MUSIC [1], [2], root-MUSIC [3], standard ESPRIT [4] and unitary ESPRIT [5], [6] have been adapted to exploit the structure of rectilinearity or strictly second-order noncircularity of signals, which include commonly used digital modulation schemes such as BPSK and ASK. These algorithms are known to achieve a higher estimation accuracy and can resolve up to twice as many sources compared to the traditional DOA algorithms. To assess the performance of these algorithms, it is necessary to derive the SCRB for rectilinear sources. Nonetheless, only the SCRB for arbitrary noncircular sources [7], [8] and the DCRB for rectilinear sources [9]–[11] are available, among many other bounds (e.g., [12] and references therein). But the first bound does not take into account the prior knowledge of rectilinearity and the second bound, although providing valuable engi-neering insight is unattainable.

As generally the exploitation of prior knowledge usually reduces the estimation error, this paper derives closed-form expressions of the SCRB for arbitrary rectilinear sources and for the specific prior knowledge of uncorrelated, and

Habti Abeida is with Dept. of Electrical Engineering, University of Taif, Al-Haweiah, 21974, Saudi Arabia, e-mail: abeida3@yahoo.fr.

Jean-Pierre Delmas is with Telecom SudParis, UMR CNRS 5157, Universit´e de Paris Saclay, 91011 Evry Cedex, France, e-mail: jean-pierre.delmas@it-sudparis.eu, phone: +(33).1.60.76.46.32.

fully correlated (referred to coherent) sources. Note that explicit expressions of circular and noncircular SCRBs for DOA parameter alone have been derived by two different methods for arbitrary sources. The first one consists of computing the asymptotic covariance matrix of the concen-trated ML estimator [13], which is asymptotically efficient and the other one is obtained directly from the Slepian-Bangs formula [14], [15]. Similarly to [16], we present here a direct derivation of the different rectilinear SCRBs from the extended Slepian-Bangs formula [7] for noncircular Gaussian distributions. Finally some properties of these SCRBs are proved and numerical illustrations are given.

II. DATA MODEL AND PROBLEM FORMULATION

ConsiderK zero-mean narrowband signals (xt,k)k=1,...,K

impinging on an arbitrary array ofM sensors. These signals

are supposed rectilinear (also called strictly second-order noncircular), i.e., described by the following model:

xt,k = st,keiφk with st,k real-valued, (1)

where the phases φk associated with different propagation

delays are assumed fixed, but unknown during the array observation. The array output at timet is modeled as

yt= Aθ∆φst+ nt, t = 1, . . . , T, (2)

where (yt)t=1,...,T are independent. Aθ def=

[a(θ1), ..., a(θK)] denotes the conventional steering matrix,

∆φ def

= Diag(eiφ1, ..., eiφK) and s

t def

= (st,1, ..., st,K)T.

nt is the additive noise, which is assumed zero-mean

circular complex Gaussian, spatially uncorrelated with E(ntnHt ) = σn2I and independent from st,k.

(st,k)k=1,...,K,t=1,..T are either real-valued deterministic

unknown parameters (in the so-called conditional or deterministic model), or zero-mean real-valued Gaussian distributed with covarianceE(stsTt) = Rs(in the so-called

unconditional or stochastic model).

To derive the CRB from the Slepian-Bangs formula, we have to carefully specify the parameters of the Gaus-sian distribution of (yt)t=1,...,T. Under the deterministic

assumption, yt are circularly Gaussian distributed with

mean (Aθ∆φst)t=1,...,T and covariance σ2nI, which are

parameterized by the real-valued parameter:

(3)

where θ def= (θ1, ..., θK)T, φ def = (φ1, ..., φK)T and ρ def = (sT

1, ..., sTT)T. On the other hand, under the stochastic

assumption, yt is noncircularly Gaussian distributed with

zero-mean, covariance

Ry def

= E(ytyHt ) = Aθ∆φRs∆∗φAHθ + σn2I (4)

and complementary covariance

Cy def= E(ytyTt) = Aθ∆φRs∆φATθ, (5)

which is also generally parameterized by (3), but where ρ is now the K(K + 1)/2 vector made from [Rs]i,j for1 ≤

i ≤ j ≤ K. In the particular case, where prior knowledge of

uncorrelated or full coherent sourcesst,k are incorporated,

the parameter ρ reduces to ρ= (σ2

1, .., σ2k, , .., σ2K)T where σ2 k def = E(s2 t,k) = [Rs]k,k and to ρ = c = (c1, ..., cK)T where Rs= ccT, respectively.

Under the stochastic assumption, (yt)t=1,..,T are

inde-pendent and noncircular Gaussian distributed and therefore the Fisher information matrix (FIM) for the parameter α is given (elementwise) by [7]: FIMi,j = T 2Tr ∂R ˜ y ∂αi R−1y˜ ∂Ry˜ ∂αj R−1y˜  , (6)

where Ry˜ is the covariance of the extended signal eyt def = [yT t, ytH]T given by: R˜y def = E(eyteyHt ) = eARsAeH+ σn2I, (7) where eA def=  Aθ∆φ A∗θ∆∗φ 

= [ea1, ..., eaK] with eak def=

[aT

k)eiφk, aH(θk)e−iφk]T.

The purpose of the next section is to directly derive the SCRB of the parameter θ alone from the FIM (6). Noting that the parameters θk and φk are non-linearly related in

the extended steering vectoreak, closed-form expressions of

the SCRB of the couple ωdef= (θT, φT)T are first derived

through its inverse [CRBsto(ω)]−1 def=



Iθ,θ Iθ,φ

ITθ,φIφ,φ



. Thus the SCRB of θ alone is deduced by

CRBsto(θ) = (Iθ,θ− Iθ,φI−1φ,φITθ,φ)−1. (8)

III. DERIVATION OF THE DIFFERENTCRB Writing the FIM (6) in compact matrix from as:

FIM = T 2  ∂r˜y ∂αT H R−T˜y ⊗ R−1y˜   ∂ry˜ ∂αT  , (9) where ry˜ def

= vec(Ry˜) = ( eA∗ ⊗ eA)vec(Rs) + σ2nvec(I)

(with⊗ is the Kronecker product), all the first steps of the

proof given in [16] apply. In particular, using the partition

(R−T /2y˜ ⊗ R −1/2 ˜ y ) ∂r ˜ y ∂ωT| ∂r˜y ∂ρT, ∂ry˜ ∂σ2 n def = (G|V u), we can deduce from (9): 2 T [CRBsto(ω)] −1 = GHΠ⊥∆G. (10)

with ∆def= (V u) and Π⊥ ∆

def

= I − Π∆ where Π∆ denotes

the orthonormal projector on the columns of ∆. Starting from (10), where Π⊥

∆ is given by [16, rel. (14], the main

steps of the proof of the following theorem are given in the Appendix:

Theorem 1: The SCRB under the general rectilinear

as-sumption is given by forK < 2M : CRBrec1 sto (ω) = σ2 n T  ( eDHωΠ⊥AeDeω)⊙ 1 1 1 1  ⊗ eH −1 , (11) with Π⊥ e A def

= I − ΠAe, where ΠAe denotes the orthonormal

projector on the columns of eA, eDω def= [ eDθ, eDφ] with

e Dθ def = h∂˜a1 ∂θ1, ..., ∂˜aK ∂θK i , eDφ def = h∂˜a1 ∂φ1, ..., ∂˜aK ∂φK i , eH def= RsAeHR−1˜y ARe s and ⊙ denotes the element by element

matrix product. Applying (8), the SCRB on θ alone is given by: CRBrec1 sto (θ)= σ2 n T  [( eDHθΠ⊥AeDeθ)⊙ eH] −[( eDHθ Π⊥AeDeφ)⊙ eH] × [( eDHφΠ⊥AeDeφ) ⊙ eH]−1[( eDHφΠ⊥AeDeθ) ⊙ eH] −1 .(12)

Including the prior knowledge that the K rectilinear

sources are coherent (where Rs = ccT), which appears in

specular multipath propagation, the main steps of the proof of the following theorem are given in the Appendix:

Theorem 2: The SCRB under the prior knowledge of

fully coherent rectilinear sources is given by forK < 2M : CRBrec2 sto (ω) = σ2 n T κc  ( eDHωΠ⊥AeDeω)⊙ 1 1 1 1  ⊗Rs −1 , (13) whereκc def = cTAeHR−1 ˜ y Ac.e

Including now the prior knowledge that theK rectilinear

sources are uncorrelated, the following theorem (presented in [9]) is proved in the Appendix:

Theorem 3: The SCRB under the prior knowledge of

uncorrelated rectilinear sources is given byK < 2M : CRBrec3 sto (ω)= 2 T  e ∆σD H ωB( ee BHG eB)−1BeHDω∆eσ −1 , (14) where Gdef= RT ˜ y⊗Ry˜+ σ 4 n 2M−Kvec(ΠAe)vec H e A), e∆σ def = Diag(σ2 1, ..., σK2, σ12, ..., σ2K), Dω def = ( eA∗| eA) ◦ ( eD θ| eDφ) + ( eD∗

θ| eD∗φ) ◦ ( eA| eA), eB is any(2M )2× ((2M )2− K) matrix

whose columns span the null space of eA∗◦ eA, where◦ is

the columnwise Kronecker product.

Finally, to make comparisons with the DCRB, we recall its closed-form expression derived in [9] and [10] and then in [11] under more general rectilinear models forK < 2M .

CRBrecdet(ω)= σ2 n T  ( eDHωΠ⊥AeDeω)⊙  1 1 1 1  ⊗Rs,T −1 , (15) where Rs,T def = 1 T PT

t=1stsTt. We also recall the

closed-form expression of the SCRB derived under the assumption of arbitrary noncircular sources in [7], applied to rectilinear

(4)

sources forK < M : CRBncsto(θ) = σ2 n 2T  Re[(DHθ Π⊥AθDθ)⊙(∆∗φH∆e φ)] −1 , (16) where Dθdef= h ∂a1 ∂θ1, ..., ∂aK ∂θK i .

IV. ANALYTICAL AND NUMERICAL COMPARISONS

Considering the comparison of the previously introduced closed-form expressions of the CRB, the main steps of the proof of the following theorem are given in the Appendix:

Theorem 4: Under the general rectilinear assumption, the

DCRB (forT → ∞, i.e., replacing Rs,T by Rs) and SCRB

have the relationships:

CRBrecdet(θ) ≤ CRBrec

1

sto (θ) ≤ CRBncsto(θ). (17)

Note that for a finite value of T , we cannot be sure that CRBrecdet(θ) ≤ CRB

rec1

sto (θ). In fact for low T and high

SNRs, the inequality reverses.

If we consider now the exploitation of prior knowledge, the following theorem is proved in the Appendix:

Theorem 5: Under the prior knowledge that the sources

are rectilinear uncorrelated, CRBrec3

sto (θ) is reduced w.r.t.

CRBrec1

sto (θ). In contrast, the exploitation of fully coherency

of the sources does not reduceCRBrec1

sto (θ): CRBrec3 sto(θ)≤CRBrec 1 sto(θ), CRBrec 2 sto(θ)=CRBrec 1 sto(θ). (18)

Note that similar properties have been proved for circular sources in [17] for uncorrelated sources and in [18] for fully coherent sources.

Finally, in the case of a single rectilinear source, we have proved after tedious algebraic manipulations that the SCRBs ofθ1 alone deduced from (11), (13) and (14) reduce to:

CRBrecsto(θ1) = 1 2T a′H 1 Π⊥a1a ′ 1 σn2 σ2 s  1 + σ 2 n 2σ2 ska1k2  , (19)

where a1def= a(θ1) and a

1 def

= da(θ1)/dθ1. Comparing (19)

to the SCRB derived in [7] under the general noncircular assumption, we see that CRBrecsto(θ1) = CRBncsto(θ1), i.e.,

the SCRB is not reduced by exploiting the rectilinear prior knowledge.

To illustrate, the difference between the different CRBs, we consider now the case of two equal-power rectilinear sources of signal-to-noise ratio 10 log10(σ2s/σn2) = 10 dB

and correlationρ, impinging on an ULA of M = 6 sensors

with half-wavelength spacing. Figs 1-3 exhibit different ratios of CRB. Fig.1 shows that there are significant gaps between the DSCB and the SCRB for closely spaced and strongly correlated rectilinear sources. More generally, ex-tensive numerical comparisons have shown that this gap increases for low SNR, low phase and DOA separations and high source correlation, but this gap will always vanish for high SNR. This proves that the conclusions based on the DCRB may be very optimistic for not too high

SNR. Fig. 2 highlights that the exploitation of the prior of rectilinearity greatly reduces the estimation error for closely spaced and uncorrelated sources. Finally, Fig. 3 proves that the joint exploitation of the prior of uncorrelatedness and rectilinearity greatly reduces the estimation error for closely spaced sources with different phases.

0.05 0.1 0.15 0.2 0.25 0.3 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 DOA separation (rd) ratio of CRB ρ=0 ρ=0.5 ρ=0.7 ρ=0.9 ρ=1

Fig. 1. RatioCRBrec

det(θ1)/CRB rec1 sto (θ1) for φ1−φ2= 0.1rd. 0.05 0.1 0.15 0.2 0.25 0.3 10−2 10−1 100 DOA separation (rd) ratio of CRB ρ=0 ρ=0.9 ρ=0.95 ρ=0.99 ρ=1

Fig. 2. RatioCRBrec1

sto (θ1)/CRBncsto(θ1) for φ1−φ2= 0.1rd.

0.05 0.1 0.15 0.2 0.25 0.3 10−4 10−3 10−2 10−1 DOA separation (rd) ratio of CRB ∆φ=0 rd ∆φ=0.05rd ∆φ=0.2rd ∆φ=0.5rd

Fig. 3. RatioCRBrec3

sto (θ1)/CRBncsto(θ1) for ρ = 0.

V. APPENDIX

Proof of theorem 1: (Detailed proofs of Theorem 1 are

avail-able at [23]) Since Rsis a(K × K) real symmetric matrix,

(5)

where DK is a so-called duplication matrix, and hence [16,

rel.(19)] becomes

V= (R−T /2y˜ A˜∗⊗ R−1/2y˜ A)D˜ K def

= WDK.

Then it follows from [19, Theorem 7.38], and some simple algebraic manipulations using [19, Theorem 7.34, rel.(b)] that [16, rel.(20)] becomes

Π⊥V= I − W(U ⊗ U)NKWH,

where U def= ˜AHR−1y˜ A and N˜ K is an (K × K) matrix

defined in [19, Theorem 7.34]. By evaluating the derivatives in G and u, and through some further algebra, one finds

uHΠ⊥Vgk= 0,

where gk is the kth column of G given by gk= vec(Zk+

ZHk ) and where Zk def= R−1/2y˜ Ar˜ s,k˜a

H k R −1/2 ˜ y , ˜a ′ k def = d˜ak/dwk and rs,k is the kth column of Rs. This identity

allows us to rewrite the individual elements of (10) as

2 T  CRB−1sto(ω)  k,l= g H kΠ⊥Vgl.

By further calculations we get

 CRB−1sto(ω)  k,l= T σ2 n Re(˜a′kA˜˜a′l)(rT s,kA˜HR−1y˜ Ar˜ s,l)  . (20) Finally, we can write (20) in matrix form as in (11).

Proof of theorem 2: (Detailed proofs of Theorem 1 are

available at [23]) We follow the steps similar to those in the proof of theorem 1. Since Rs= ccT and its derivative

w.r.t. c is given by Dc def

= c ⊗ I + I ⊗ c = 2NK(c ⊗ I), it

follows then that V has the form V= WDc. After some

algebraic manipulation using [19, Theorem 7.34, rel.(d)], it follows that Π⊥V def = I − V(VHV)−1VH= I − V1V¯−1VH1, with V1 def = WNK(c ⊗ I) and ¯V def = 12(κcU+ ucuTc) where uc def = Uc and κc def

= cTUc. Thanks to the matrix

inversion lemma, we have ¯V−1 = 2 κc(U

−1 1 2κccc

T).

Through some further algebra using u = vec(R−1y˜ ) and gk= vec(Zk+ZHk) where Zk def= R−1/2y˜ Ac˜˜ a

H

k R −1/2 ˜ y , one

finds that uHV1= cTU where ¯¯ U def = ˜AHR−2y˜ A, u˜ Hgk = 2cka˜ ′H k R −2 ˜ y Ac and V˜ H1gk = ck(νc(k)uc+ κcA˜HR−1˜y ˜a ′ k) where νc(k) def= ˜a ′H

k R−1˜y Ac. By further calculations we˜

arrive at uHΠ⊥Vgk = 0. This identity allows us to rewrite

the individual elements of (10) as

2 T  CRB−1sto(ω)  k,l=g H k Π⊥Vgl=2κcT σ2 n ckcl(˜a ′H k Π⊥A˜˜a ′ l),

which can also be written in the matrix form (13).

Proof of theorem 3: Noting that (7) becomes Ry˜ =

PK

k=1σk2eakeaHk + σn2I, all the steps of the proof given

in [17, Appendix A] apply to the parameter ω with the

FIM (6) associated with the noncircular zero-mean Gaussian distribution of yt.

Proof of theorem 4: Using Rs≥ RsAeHR−1y˜ ARe sthanks to

[20, rel. B.6.37] and noting that 1 1 1 1  is positive semidef-inite, we have 1 1 1 1  ⊗(Rs− eH) ≥ 0 from [19, th. 7.10].

Then, noting that eDHωΠ⊥ e

ADeω is positive definite, we have

( eDHωΠ⊥AeDeω) ⊙  1 1 1 1  ⊗(Rs− eH)  ≥ 0 thanks to [20,

rel. R.19, p.358], and CRBrecdet(θ) ≤ CRB rec1

sto (θ) directly

follows.

Consider the parametrization αnc= (θT, ρTnc, σn2)T

associ-ated with the assumption of arbitrary noncircular sources in [7], where ρnc is the 2K2 + K vector made from

[Re(Rx)]i,j, [Im(Rx)]i,j,[Re(Cx)]i,j and[Re(Cx)]i,j for

1 ≤ i < j ≤ K, and [Rx]i,i, [Re(Cx)]i,i and

[Im(Cx)]i,i for 1 ≤ i ≤ K, where Rx def

= E(xtxHt ) and

Cx def= E(xtxTt) with xt def= (xt,1, ..., xt,K)T. Consider

now the one to one mapping between αnc and α′nc =

(θT, φT, ρT 1, ρ

T

, σ2

n)T where φ = (φ1, .., φk, .., φK)T is

defined by2φk def= ∠[Cx]k,k/[Rx]k,k, ρ1is the vector made

from [Re(Rs)]i,j for1 ≤ i ≤ j ≤ K, and ρ

is the vector gathering[Im(Rs)]i,j,[Im(Cs)]i,j for1 ≤ i < j ≤ K and

[Re(Cs)]i,j for1 ≤ i ≤ j ≤ K.

Let FIM1, FIMnc and FIM′ncbe the FIM associated with

the stochastic parametrizations (3), αnc and α′nc,

respec-tively. From the one to one mapping αnc ↔ α′nc, we get

[FIM′

nc]θ = [FIMnc]θ (where [.]θ denotes the submatrix

determined by the firstK rows and columns). Furthermore FIM1= [FIM′nc]1 where[.]1 denotes the principal

subma-trix determined by the rows and columns induced by the parametrization of FIM1. Consequently, taking the inverse,

we get[FIM1]−1≤ [FIM′−1nc]1 [22, th. 7.7.8]. Then taking

the principal submatrix determined by the firstK rows and

columns, we getCRBrec1

sto(θ) ≤ CRBncsto(θ).

Proof of theorem 5: Let FIM3be the FIM associated with the

parametrizations (3) for which ρ= ([Rs]k,k;1≤k≤K)T. For

rectilinear uncorrelated sources,FIM3= [FIM1]3, where[.]3

denotes the principal submatrix determined by the rows and columns induced by the parametrization of FIM3. Taking

the inverse, we get [FIM3]−1≤ [FIM−11 ]3 [22, th. 7.7.8].

Then taking the principal submatrix determined by the first

K rows and columns, we get CRBrec3

sto(θ) ≤ CRBrec

1

sto(θ).

Finally, by noting that for coherent rectilinear sources,

e

H = κcccT in (11) givesCRBrecsto2(ω) = CRB rec1

sto(ω), It

follows then thatCRBrec2

sto(θ) = CRBrec

1

sto(θ).

VI. CONCLUSION

Closed-form expressions of the SCRB of DOA estimation for rectilinear sources have been directly derived form the extended Slepian-Bangs formula, including the case of prior knowledge of uncorrelated or coherent rectilinear sources. Analytical and numerical comparisons with the DCRB and the SCRB for noncircular sources have shown in particular that the DCRB may be very optimistic.

(6)

REFERENCES

[1] P. Gounon, C. Adnet, and J. Galy, ”Localization angulaire de signaux non circulaires,” Traitement du Signal, vol. 15, no. 1, pp. 17-23, 1998. [2] H. Abeida and J.P. Delmas, ”MUSIC-like estimation of direction of arrival for non-circular sources,” IEEE Trans. Signal Process., vol. 54, no. 7, pp. 2678-2690, July 2006.

[3] P. Charg´e, Y. Wang, and J. Saillard, ”A non-circular sources direction finding method using polynomial rooting,” Signal Process., vol. 81, pp. 1765-1770, 2001.

[4] A. Zoubir, P. Charg´e, and Y. Wang, ”Non circular sources localization with ESPRIT,” in Proc. Eur. Conf. Wireless Technology (ECWT), Munich, Germany, Oct. 2003.

[5] M. Haardt and F. Roemer, ”Enhancements of unitary esprit for noncircular sources,” Proc. ICASSP, vol. 2,pp. 101-104, Montreal, QC, Canada, May 2004.

[6] J. Steinwandt, F. Roemer, M. Haardt, and G. Del Galdo, ”R-dimensional ESPRIT-type algorithms for strictly second-order non-circular sources and their performance analysis,” IEEE Trans. Signal

Processing, vol. 62, no. 18, pp. 4824-4838, Sept. 2014.

[7] J.P. Delmas and H. Abeida, ”Stochastic Cramer-Rao bound for non-circular signals with application to DOA estimation,” IEEE Trans.

Signal Process., vol. 52, no. 11, pp. 3192-3199, Nov. 2004.

[8] H. Abeida and J.P. Delmas, ”Efficiency of subspace-based DOA estimators,” Signal Processing, vol. 87, no. 9, pp. 2075-2084, Sept. 2007.

[9] H. Abeida and J.P. Delmas, ”Bornes de Cramer Rao d´eterministe et stochastique de DOA de signaux rectilignes non corr´el´es,” Proc.

GRETSI, Troyes, Sept. 2007,

[10] F. Roemer and M. Haardt, ”Deterministic Cramer Rao bounds for strict sense noncircular sources,” Proc. Internat. ITG/IEEE Worshop

on Smart Antennas (WSA’07), Vienne, Austria, Feb. 2007.

[11] J. Steinwandt, F. Roemer, M. Haardt, and G. Del Galdo, ”Determinis-tic Cramer-Rao bound for strictly non-circular sources and analy”Determinis-tical analysis of the achievable gains,” IEEE Trans. Signal Process., vol. 64, no. 17, pp. 4417-4431, Sept. 2016.

[12] D.T. Vu, A. Renaux, R. Boyer, and S. Marcos ”Some results on the Weiss-Weinstein bound for conditional and unconditional signal models in array processing,” Signal Processing, vol. 95, pp. 126-148, Feb. 2014.

[13] P. Stoica and A. Nehorai, ”Performance study of conditional and unconditional direction of arrival estimation,” IEEE Trans. Signal

Process., vol. 38, no. 10, pp. 1783-1795, Oct. 1990.

[14] D. Slepian, “Estimation of signal parameters in the presence of noise,”

in Trans. IRE Prof. Grop Inform. Theory PG IT-3, pp. 68-89, 1954.

[15] W.J. Bangs “Array processing with generalized beamformers,” Ph.D.

thesis Yale University, New Haven, CT, 1971.

[16] P. Stoica, A.G. Larsson, and A.B. Gershman, “The stochastic CRB for array processing: a textbook derivation,” IEEE Signal Process.

letters, vol. 8, no. 5, pp. 148-150, May 2001.

[17] M. Jansson, B. G ¨oransson, and B. Ottersten, “Subspace method for direction of arrival estimation of uncorrelated emitter signals,” IEEE

Trans. Signal Process., vol. 47, no. 4, pp. 945-956, April 1999.

[18] J. Sheinvald, M. Wax, and A.J. Weiss, “On maximum-likelihood localization of coherent signals,” IEEE Transactions on Signal

Pro-cessing, vol. 44, no. 10, pp. 2475-2482, Oct. 1996.

[19] J. R. Schott, Matrix analysis for statistics, Wiley, New York, 1997. [20] P. Stoica and R. Moses, Spectral analysis of signals, Prentice-Hall,

Upper Saddle River, NJ, 2005.

[21] P. Stoica and A. Nehorai, ”MUSIC, maximum likelihood, and Cramer-Rao bound: Further results and comparisons,” IEEE Trans.

Signal Process., vol. 38, no. 12, pp. 2140-2150, Dec. 1990.

[22] R.A. Horn and C.R. Johnson, Matrix analysis, Cambridge University Press, 1996.

[23] H. Abeida and J.P. Delmas, Detailed proofs of Theorems 1 and 2 are available at http://www-public.int-evry.fr/∼delmas/

Figure

Fig. 3. Ratio CRB rec sto 3 (θ 1 )/CRB nc sto (θ 1 ) for ρ = 0.

Références

Documents relatifs

Results on the secondary foams obtained after aging the primary foam samples for various times followed by rapidly heating the sample to 120 ° C are shown in Figure 3 where reduction

To improve their performances, DOA estimation schemes can be obtained in the framework of inverse problems2. For that, subspace based method (high-resolution) such as

Il s'agit de reproduire la page de recherche d'un produit de l'exemple 10 (paragraphe 4) en utilisant cette fois DOM pour manipuler le résultat de la requête ajax et pour mettre à

This paper focuses on optimization of some design parameters of such antennas, including the rotation angle, the rotation step number and distances of two dipole couples, in order

In this paper, we derived the deterministic Cram´er-Rao bound in a closed-form expression for a single near-field time-varying narrow- band source for the model parameters

In this paper, we derive a closed- form expression of the WWB for 3D source localization using an arbitrary planar antenna array in the case of a deterministic known signal..

We have based our derivation on the popular method of direction estima- tion (MODE) [12] from which we have deduced the optimal Prior- knowLEDGE (PLEDGE) algorithm for

In this paper, an approach is proposed which uses the classical 2D algorithm to estimate range, Doppler and number of targets and then FDA (Frequency Diversity Array) is used to