This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the detection of low-rank signal in the presence of spatially uncorrelated noise: a frequency domain approach.

A. Rosuel, , P. Vallet, , P. Loubaton, , and X. Mestre A. Rosuel and P. Loubaton are with Laboratoire d’Informatique Gaspard Monge (CNRS, Univ. Gustave-Eiffel), 5 Bd. Descartes 77454 Marne-la-Vallée (France), {alexis.rosuel, philippe.loubaton}@univ-eiffel.fr P. Vallet is with Laboratoire de l’Intégration du Matériau au Système (CNRS, Univ. Bordeaux, Bordeaux INP), 351, Cours de la Libération 33405 Talence (France), [email protected] X. Mestre is with Centre Tecnològic de Telecomunicacions de Catalunya (CTTC), Av. Carl Friedrich Gauss 08860 Castelldefels, Barcelona (Spain), [email protected] This work was partially supported by project ANR-17-CE40-0003. The material of this paper was partly presented in the conference paper [1]
Abstract

This paper analyzes the detection of a M𝑀M–dimensional useful signal modeled as the output of a M×K𝑀𝐾M\times K MIMO filter driven by a K𝐾K–dimensional white Gaussian noise, and corrupted by a M𝑀M–dimensional Gaussian noise with mutually uncorrelated components. The study is focused on frequency domain test statistics based on the eigenvalues of an estimate of the spectral coherence matrix (SCM), obtained as a renormalization of the frequency-smoothed periodogram of the observed signal. If N𝑁N denotes the sample size and B𝐵B the smoothing span, it is proved that in the high-dimensional regime where M,B,N𝑀𝐵𝑁M,B,N converge to infinity while K𝐾K remains fixed, the SCM behaves as a certain correlated Wishart matrix. Exploiting well-known results on the behaviour of the eigenvalues of such matrices, it is deduced that the standard tests based on linear spectral statistics of the SCM fail to detect the presence of the useful signal in the high-dimensional regime. A new test based on the SCM, which is proved to be consistent, is also proposed, and its statistical performance is evaluated through numerical simulations.

Index Terms:
detection, spectral coherence matrix, periodogram, high-dimensional statistics, Random Matrix Theory

I Introduction

Detecting the presence of an unknown multivariate signal corrupted by noise is one of the fundamental problems in signal processing, which is found in many applications including array and radar processing, wireless communications, radio-astronomy or seismology among others. In a statistical framework, this problem is usually formulated as the following binary hypothesis test, where the objective is to discriminate between the null hypothesis 0subscript0\mathcal{H}_{0} and the alternative hypothesis 1subscript1\mathcal{H}_{1} defined as

0subscript0\displaystyle\mathcal{H}_{0} :(𝐲n)n=(𝐯n)n:absentsubscriptsubscript𝐲𝑛𝑛subscriptsubscript𝐯𝑛𝑛\displaystyle:(\mathbf{y}_{n})_{n\in\mathbb{Z}}=(\mathbf{v}_{n})_{n\in\mathbb{Z}}
1subscript1\displaystyle\mathcal{H}_{1} :(𝐲n)n=(𝐮n)n+(𝐯n)n:absentsubscriptsubscript𝐲𝑛𝑛subscriptsubscript𝐮𝑛𝑛subscriptsubscript𝐯𝑛𝑛\displaystyle:(\mathbf{y}_{n})_{n\in\mathbb{Z}}=(\mathbf{u}_{n})_{n\in\mathbb{Z}}+(\mathbf{v}_{n})_{n\in\mathbb{Z}} (1)

where (𝐲n)nsubscriptsubscript𝐲𝑛𝑛(\mathbf{y}_{n})_{n\in\mathbb{Z}} is the M𝑀M-variate observed signal, and where (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} and (𝐯n)nsubscriptsubscript𝐯𝑛𝑛(\mathbf{v}_{n})_{n\in\mathbb{Z}} represent a non observable signal of interest and the noise respectively, both modeled in this paper as mutually independent zero-mean complex Gaussian stationary time series.

Without further knowledge on the covariance function of (𝐯n)nsubscriptsubscript𝐯𝑛𝑛(\mathbf{v}_{n})_{n\in\mathbb{Z}} and/or (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}}, or access to “noise only” samples, the test problem (I) is ill-posed, even for temporally white time series (𝐯n)nsubscriptsubscript𝐯𝑛𝑛(\mathbf{v}_{n})_{n\in\mathbb{Z}} and (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}}, and one needs to exploit additional information on the covariance structure of the useful signal and noise. One common assumption, widely used in the context of array processing and multi-antenna communications, is to consider that the noise (𝐯n)nsubscriptsubscript𝐯𝑛𝑛(\mathbf{v}_{n})_{n\in\mathbb{Z}} is spatially uncorrelated. Moreover, when the receive antennas are not calibrated, it is reasonable to assume that the spectral densities of the components of the noise may not coincide, see e.g. [2], [3], [4], [5]. This will be the context of the present paper.

A first class of tests is based on the observation that the noise is spatially uncorrelated if and only if the matrices 𝐑𝐯()=𝔼[𝐯n𝐯n]subscript𝐑𝐯𝔼delimited-[]subscript𝐯𝑛superscriptsubscript𝐯𝑛\mathbf{R}_{\mathbf{v}}(\ell)=\mathbb{E}[\mathbf{v}_{n}\mathbf{v}_{n-\ell}^{*}] are diagonal for all \ell\in\mathbb{Z}, whereas if the useful signal (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} is assumed spatially correlated, 𝐑𝐮()=𝔼[𝐮n𝐮n]subscript𝐑𝐮𝔼delimited-[]subscript𝐮𝑛superscriptsubscript𝐮𝑛\mathbf{R}_{\mathbf{u}}(\ell)=\mathbb{E}[\mathbf{u}_{n}\mathbf{u}_{n-\ell}^{*}] is non-diagonal for some \ell\in\mathbb{Z}. Under this assumption, the problem in (I) can be formulated as the following correlation test:

0subscript0\displaystyle\mathcal{H}_{0} :𝐑𝐲()=dg(𝐑𝐲()) for all :absentsubscript𝐑𝐲dgsubscript𝐑𝐲 for all \displaystyle:\mathbf{R}_{\mathbf{y}}(\ell)=\operatorname*{dg}\left(\mathbf{R}_{\mathbf{y}}(\ell)\right)\text{ for all }\ell\in\mathbb{Z}
1subscript1\displaystyle\mathcal{H}_{1} :𝐑𝐲()dg(𝐑𝐲()) for some :absentsubscript𝐑𝐲dgsubscript𝐑𝐲 for some \displaystyle:\mathbf{R}_{\mathbf{y}}(\ell)\neq\operatorname*{dg}\left(\mathbf{R}_{\mathbf{y}}(\ell)\right)\text{ for some }\ell\in\mathbb{Z} (2)

where 𝐑𝐲()=𝔼[𝐲n𝐲n]subscript𝐑𝐲𝔼delimited-[]subscript𝐲𝑛superscriptsubscript𝐲𝑛\mathbf{R}_{\mathbf{y}}(\ell)=\mathbb{E}[\mathbf{y}_{n}\mathbf{y}_{n-\ell}^{*}] and dg(𝐑𝐲())=𝐑𝐲()𝐈Mdgsubscript𝐑𝐲direct-productsubscript𝐑𝐲subscript𝐈𝑀\operatorname*{dg}\left(\mathbf{R}_{\mathbf{y}}(\ell)\right)=\mathbf{R}_{\mathbf{y}}(\ell)\odot\mathbf{I}_{M}, where direct-product\odot is the element-wise (Hadamard) product and 𝐈Msubscript𝐈𝑀\mathbf{I}_{M} the M×M𝑀𝑀M\times M identity matrix. A number of previous works developed lag domains tests that specifically tackle the above problem, see e.g. [6], [7], [8], [9], [10], [11]. Also relevant are the approaches in [2] and [3], where the possible useful signal is supposed to be the output of a filter driven by a low-dimensional white noise sequence.

Our focus here is on another type of formulation, referred to as frequency domain approach, which consists in rewriting problem (I) as

0subscript0\displaystyle\mathcal{H}_{0} :𝐒𝐲(ν)=dg(𝐒𝐲(ν)) for all ν[0,1]:absentsubscript𝐒𝐲𝜈dgsubscript𝐒𝐲𝜈 for all 𝜈01\displaystyle:\mathbf{S}_{\mathbf{y}}(\nu)=\operatorname*{dg}\left(\mathbf{S}_{\mathbf{y}}(\nu)\right)\text{ for all }\nu\in[0,1]
1subscript1\displaystyle\mathcal{H}_{1} :𝐒𝐲(ν)dg(𝐒𝐲(ν)) for some ν[0,1]:absentsubscript𝐒𝐲𝜈dgsubscript𝐒𝐲𝜈 for some 𝜈01\displaystyle:\mathbf{S}_{\mathbf{y}}(\nu)\neq\operatorname*{dg}\left(\mathbf{S}_{\mathbf{y}}(\nu)\right)\text{ for some }\nu\in[0,1] (3)

where 𝐒𝐲(ν)subscript𝐒𝐲𝜈\mathbf{S}_{\mathbf{y}}(\nu) is the M×M𝑀𝑀M\times M spectral density matrix of (𝐲n)nsubscriptsubscript𝐲𝑛𝑛(\mathbf{y}_{n})_{n\in\mathbb{Z}} at frequency ν𝜈\nu, defined by

𝐒𝐲(ν)=k𝐑𝐲(k)e2iπνk.subscript𝐒𝐲𝜈subscript𝑘subscript𝐑𝐲𝑘superscript𝑒2i𝜋𝜈𝑘\mathbf{S}_{\mathbf{y}}(\nu)=\sum_{k\in\mathbb{Z}}\mathbf{R}_{\mathbf{y}}(k)e^{-2\mathrm{i}\pi\nu k}.

This problem is equivalent to testing whether the spectral coherence matrix (see for instance [12, Chapter 7-6], [13, Chapter 5.5])

𝐂𝐲(ν)=dg(𝐒𝐲(ν))12𝐒𝐲(ν)dg(𝐒𝐲(ν))12\displaystyle\mathbf{C}_{\mathbf{y}}(\nu)=\operatorname*{dg}\left(\mathbf{S}_{\mathbf{y}}(\nu)\right)^{-\frac{1}{2}}\mathbf{S}_{\mathbf{y}}(\nu)\operatorname*{dg}\left(\mathbf{S}_{\mathbf{y}}(\nu)\right)^{-\frac{1}{2}} (4)

is equal to 𝐈Msubscript𝐈𝑀\mathbf{I}_{M} for all frequencies ν[0,1]𝜈01\nu\in[0,1]. In this approach, usual test statistics are mostly based on consistent sample estimates of 𝐒𝐲(ν)subscript𝐒𝐲𝜈\mathbf{S}_{\mathbf{y}}(\nu) or 𝐂𝐲(ν)subscript𝐂𝐲𝜈\mathbf{C}_{\mathbf{y}}(\nu) that are compared to a diagonal matrix or to the identity 𝐈Msubscript𝐈𝑀\mathbf{I}_{M} respectively. Previous works that developed this approach include [14], [15], [16], [17]. In particular, [14] considered the following frequency smoothed-periodogram estimator 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu) defined by

𝐒^𝐲(ν)=1B+1b=B2B2𝝃𝐲(ν+bN)𝝃𝐲(ν+bN)subscript^𝐒𝐲𝜈1𝐵1superscriptsubscript𝑏𝐵2𝐵2subscript𝝃𝐲𝜈𝑏𝑁subscript𝝃𝐲superscript𝜈𝑏𝑁\displaystyle\hat{\mathbf{S}}_{\mathbf{y}}(\nu)=\frac{1}{B+1}\sum_{b=-\frac{B}{2}}^{\frac{B}{2}}\boldsymbol{\xi}_{\mathbf{y}}\left(\nu+\frac{b}{N}\right)\boldsymbol{\xi}_{\mathbf{y}}\left(\nu+\frac{b}{N}\right)^{*} (5)

with 𝝃𝐲(ν)=1Nn=0N1𝐲ne2iπnνsubscript𝝃𝐲𝜈1𝑁superscriptsubscript𝑛0𝑁1subscript𝐲𝑛superscript𝑒2𝑖𝜋𝑛𝜈\boldsymbol{\xi}_{\mathbf{y}}(\nu)=\frac{1}{\sqrt{N}}\sum_{n=0}^{N-1}\mathbf{y}_{n}e^{-2i\pi n\nu} the renormalized finite Fourier transform of (𝐲n)n=0,,N1subscriptsubscript𝐲𝑛𝑛0𝑁1(\mathbf{y}_{n})_{n=0,\ldots,N-1}, B𝐵B the smoothing span, assumed to be an even number, and where 𝝃𝐲(ν+bN)subscript𝝃𝐲superscript𝜈𝑏𝑁\boldsymbol{\xi}_{\mathbf{y}}\left(\nu+\frac{b}{N}\right)^{*} is the conjugate transpose of the vector 𝝃𝐲(ν+bN)subscript𝝃𝐲𝜈𝑏𝑁\boldsymbol{\xi}_{\mathbf{y}}\left(\nu+\frac{b}{N}\right). [14] was devoted to the study of the limiting distribution of

log{i=1Pdet(𝐒^𝐲(νi))/m=1Ms^m,m(νi)}superscriptsubscriptproduct𝑖1𝑃detsubscript^𝐒𝐲subscript𝜈𝑖superscriptsubscriptproduct𝑚1𝑀subscript^𝑠𝑚𝑚subscript𝜈𝑖\log\left\{\prod_{i=1}^{P}\mathrm{det}(\hat{\mathbf{S}}_{\mathbf{y}}(\nu_{i}))/\prod_{m=1}^{M}\hat{s}_{m,m}(\nu_{i})\right\}

for some properly defined subset of frequencies (νi)i=1,,Psubscriptsubscript𝜈𝑖𝑖1𝑃(\nu_{i})_{i=1,\ldots,P}, where s^m,m(ν)=(𝐒^𝐲(ν))m,msubscript^𝑠𝑚𝑚𝜈subscriptsubscript^𝐒𝐲𝜈𝑚𝑚\hat{s}_{m,m}(\nu)=\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)_{m,m}. When M=2𝑀2M=2, [16] considered a general kernel estimator of 𝐒𝐲(ν)subscript𝐒𝐲𝜈\mathbf{S}_{\mathbf{y}}(\nu):

𝐒~𝐲(ν)=b=N2N2wN(bN)𝝃𝐲(ν+bN)𝝃𝐲(ν+bN)subscript~𝐒𝐲𝜈superscriptsubscript𝑏𝑁2𝑁2subscript𝑤𝑁𝑏𝑁subscript𝝃𝐲𝜈𝑏𝑁subscript𝝃𝐲superscript𝜈𝑏𝑁\tilde{\mathbf{S}}_{\mathbf{y}}(\nu)=\sum_{b=-\frac{N}{2}}^{\frac{N}{2}}w_{N}\left(\frac{b}{N}\right)\boldsymbol{\xi}_{\mathbf{y}}\left(\nu+\frac{b}{N}\right)\boldsymbol{\xi}_{\mathbf{y}}\left(\nu+\frac{b}{N}\right)^{*}

where wNsubscript𝑤𝑁w_{N} is a weight function satisfying some specific properties, and a test statistic of the form

1Nn=1N|(𝐒~𝐲)12(ν)|2(𝐒~𝐲)11(ν)(𝐒~𝐲)22(ν)1𝑁superscriptsubscript𝑛1𝑁superscriptsubscriptsubscript~𝐒𝐲12𝜈2subscriptsubscript~𝐒𝐲11𝜈subscriptsubscript~𝐒𝐲22𝜈\frac{1}{N}\sum_{n=1}^{N}\frac{|(\tilde{\mathbf{S}}_{\mathbf{y}})_{12}(\nu)|^{2}}{(\tilde{\mathbf{S}}_{\mathbf{y}})_{11}(\nu)(\tilde{\mathbf{S}}_{\mathbf{y}})_{22}(\nu)}

which is proven to be, after proper recentring and renormalization, asymptotically normally distributed. Finally, [15] and [17] considered the more general class of test statistics, defined by:

1/21/2K((𝐒~𝐲)12(ν))𝑑ν and 1/21/2ψ((𝐒~𝐲)12(ν),ν)2𝑑νsuperscriptsubscript1212𝐾subscriptsubscript~𝐒𝐲12𝜈differential-d𝜈 and superscriptsubscript1212superscriptnorm𝜓subscriptsubscript~𝐒𝐲12𝜈𝜈2differential-d𝜈\int_{-1/2}^{1/2}K\left((\tilde{\mathbf{S}}_{\mathbf{y}})_{12}(\nu)\right)d\nu\text{ and }\int_{-1/2}^{1/2}\left\|\psi\left((\tilde{\mathbf{S}}_{\mathbf{y}})_{12}(\nu),\nu\right)\right\|^{2}d\nu

for some well-defined functions K𝐾K and ψ𝜓\psi, and where \|\cdot\| is the Euclidian norm. They proved that these quantities asymptotically follow normal distributions. In the present paper, we focus on the natural estimator (see e.g. [12, Chapter 7-6], [13, Chapter 8-4]) of 𝐂𝐲subscript𝐂𝐲\mathbf{C}_{\mathbf{y}}, defined by

𝐂^𝐲(ν)=dg(𝐒^𝐲(ν))12𝐒^𝐲(ν)dg(𝐒^𝐲(ν))12\displaystyle\hat{\mathbf{C}}_{\mathbf{y}}(\nu)=\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)^{-\frac{1}{2}}\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)^{-\frac{1}{2}} (6)

where 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu) is the frequency-smoothed periodogram estimate defined by (5). Note that adding a weight to the matrices 𝝃𝐲(ν+bN)𝝃𝐲(ν+bN)subscript𝝃𝐲𝜈𝑏𝑁subscript𝝃𝐲superscript𝜈𝑏𝑁\boldsymbol{\xi}_{\mathbf{y}}(\nu+\frac{b}{N})\boldsymbol{\xi}_{\mathbf{y}}(\nu+\frac{b}{N})^{*} leads to a more general class of estimators of 𝐒𝐲(ν)subscript𝐒𝐲𝜈\mathbf{S}_{\mathbf{y}}(\nu). The study of this more general class of estimators involves different techniques and random matrix models than the ones used here, and is therefore out of the scope of this paper.

I-A Low vs High-dimensional regime

The performance of the test statistics developed in the above mentioned previous works is usually studied in the low-dimensional regime where N𝑁N\to\infty and M𝑀M is fixed. It is well known (see for instance [12]) that 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu) and 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) are consistent estimates if B+𝐵B\rightarrow+\infty and BN0𝐵𝑁0\frac{B}{N}\to 0. Under mild assumptions on the memory of the time series (𝐲n)nsubscriptsubscript𝐲𝑛𝑛(\mathbf{y}_{n})_{n\in\mathbb{Z}}, 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) is a consistent and asymptotically normal estimate of 𝐂𝐲(ν)subscript𝐂𝐲𝜈\mathbf{C}_{\mathbf{y}}(\nu), which can in turn be used to study the asymptotic performance of the various tests based on 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu). In practice, the above asymptotic regime allows to predict the actual performance of the tests quite accurately, provided the ratio MN𝑀𝑁\frac{M}{N} is small enough. If this condition is not met, test statistics based on 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) may be of delicate use, as the choice of the smoothing span B𝐵B must meet the constraints BM𝐵𝑀\frac{B}{M} much larger than 1 (because B𝐵B is supposed to converge towards ++\infty) as well as BN𝐵𝑁\frac{B}{N} small enough (because BN𝐵𝑁\frac{B}{N} is supposed to converge towards 00).

Nowadays, in many practical applications involving high-dimensional signals and/or a moderate sample size, the ratio MN𝑀𝑁\frac{M}{N} may not be small enough to be able to choose B𝐵B so as to meet BM𝐵𝑀\frac{B}{M} much larger than 1 and BN𝐵𝑁\frac{B}{N} small enough. Therefore, the results obtained in the low-dimensional regime may fail to provide accurate predictions of the behaviour of the aforementioned test statistics. In this situation, one may rely on the more relevant high-dimensional regime in which M,B,N𝑀𝐵𝑁M,B,N converge to infinity such that MB𝑀𝐵\frac{M}{B} converges to a positive constant while BN𝐵𝑁\frac{B}{N} converges to zero.

In comparison to the low-dimensional regime, the literature concerning correlation tests for the frequency domain in the high-dimensional regime is quite scarce. Recent results obtained in [18] show that under hypothesis 0subscript0\mathcal{H}_{0}, the empirical eigenvalue distribution of the spectral coherence estimate 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu) behaves in the high-dimensional regime as the well-known Marcenko-Pastur distribution [19]. The result of [18] allows to predict the performance under 0subscript0\mathcal{H}_{0} of a large class of test statistics based on

Lφ(ν)=1Mm=1Mφ(λm(𝐂^𝐲(ν)))subscript𝐿𝜑𝜈1𝑀superscriptsubscript𝑚1𝑀𝜑subscript𝜆𝑚subscript^𝐂𝐲𝜈\displaystyle L_{\varphi}(\nu)=\frac{1}{M}\sum_{m=1}^{M}\varphi\left(\lambda_{m}(\hat{\mathbf{C}}_{\mathbf{y}}(\nu))\right)

where λ1(𝐂^𝐲(ν)),,λM(𝐂^𝐲(ν))subscript𝜆1subscript^𝐂𝐲𝜈subscript𝜆𝑀subscript^𝐂𝐲𝜈\lambda_{1}(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)),\ldots,\lambda_{M}(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)) are the eigenvalues of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu), and φ𝜑\varphi belongs to a certain functional class. Such family of statistics Lφsubscript𝐿𝜑L_{\varphi}, called linear spectral statistics (LSS) of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu), include in particular the choice φ(x)=logx𝜑𝑥𝑥\varphi(x)=\log x, i.e. Lφ(ν)=1Mlogdet𝐂^𝐲(ν)subscript𝐿𝜑𝜈1𝑀detsubscript^𝐂𝐲𝜈L_{\varphi}(\nu)=\frac{1}{M}\log\mathrm{det}\hat{\mathbf{C}}_{\mathbf{y}}(\nu) and the choice φ(x)=(x1)2𝜑𝑥superscript𝑥12\varphi(x)=(x-1)^{2}, i.e. Lφ(ν)=1M𝐂^𝐲(ν)𝐈MF2subscript𝐿𝜑𝜈1𝑀subscriptsuperscriptnormsubscript^𝐂𝐲𝜈subscript𝐈𝑀2𝐹L_{\varphi}(\nu)=\frac{1}{M}\|\hat{\mathbf{C}}_{\mathbf{y}}(\nu)-\mathbf{I}_{M}\|^{2}_{F}, where F\|\cdot\|_{F} represents the Frobenius norm.

In this paper, we consider the study of the eigenvalues of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) in the high-dimensional regime under the special alternative 1subscript1\mathcal{H}_{1} for which the useful signal (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} is modeled as the output of a stable MIMO filter driven by a K𝐾K–dimensional white complex Gaussian noise. In the context where the intrinsic dimension K𝐾K is fixed while M,N,B𝑀𝑁𝐵M,N,B\to\infty, it is shown that the empirical eigenvalue distribution of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) still converges to the Marcenko-Pastur distribution, showing that the test statistic based on Lφ(ν)subscript𝐿𝜑𝜈L_{\varphi}(\nu) is unable to discriminate between hypotheses 0subscript0\mathcal{H}_{0} and 1subscript1\mathcal{H}_{1} in the high-dimensional regime. Nevertheless, we also prove that, provided that the signal-to-noise ratio is large enough, the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) asymptotically splits from the support of the Marcenko-Pastur distribution. We can therefore exploit this result to design a new frequency domain test statistic, which is shown to be consistent in the high-dimensional regime. This result is connected to the widely studied spiked models in Random Matrix Theory, defined as low rank perturbations of large random matrices. These models were extensively studied in the context of sample covariance matrices of independent identically distributed high-dimensional vectors, see e.g. [20]. We however notice that papers addressing the behaviour of the corresponding sample correlation matrices are quite scarce, see [21] when the low rank perturbation affects only the first components of the observations.

I-B Related works

Although the asymptotic framework differs from the high-dimensional regime considered here, we also mention the series of studies [22, 23] in the econometrics field, which consider a similar model under 1subscript1\mathcal{H}_{1}. In these works, it is assumed that M,N𝑀𝑁M,N\to\infty so the ratio MN𝑀𝑁\frac{M}{N} remains bounded, while the K𝐾K non-zero eigenvalues of the spectral density 𝐒𝐮(ν)subscript𝐒𝐮𝜈\mathbf{S}_{\mathbf{u}}(\nu) of (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} are assumed to converge towards ++\infty at rate M𝑀M. This last assumption, which ensures that the Signal-to-Noise Ratio (SNR) 𝔼𝐮n2𝔼𝐯n2𝔼superscriptnormsubscript𝐮𝑛2𝔼superscriptnormsubscript𝐯𝑛2\frac{\mathbb{E}\|\mathbf{u}_{n}\|^{2}}{\mathbb{E}\|\mathbf{v}_{n}\|^{2}} remains bounded away from 00 as M𝑀M\to\infty, significantly facilitates the design of consistent detection methods. Nevertheless, while relevant in the domain of econometrics, this assumption may be unrealistic in several applications of array processing, where the challenge is to manage situations in which the SNR converges towards 00 at rate 1M1𝑀\frac{1}{M}. This situation is the one considered in this paper and, in that case, the results of [22, 23] can not be used. We discuss this point further in Section II below.

The rest of the paper is organized as follows. In Section II, we formally introduce the model of signals used in the remainder, as well as the required technical assumptions. In section III, we introduce informally the proposed test statistic, and illustrate its behaviour in order to provide some intuition before a more rigorous presentation. In section IV, we study some approximation results for the spectral coherence 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) which are useful to study the linear spectral statistics considered here. This study is then used in Section V to introduce a new test statistic that is consistent in the high-dimensional regime. Finally Section VI provides some simulations illustrating its performance and comparisons against other relevant approaches.

Notations. For a complex matrix 𝐀𝐀\mathbf{A}, we denote by 𝐀superscript𝐀\mathbf{A}^{*} its conjugate transpose, and by 𝐀2subscriptnorm𝐀2\|\mathbf{A}\|_{2} and 𝐀Fsubscriptnorm𝐀𝐹\|\mathbf{A}\|_{F} its spectral and Frobenius norms respectively. If 𝐀𝐀\mathbf{A} is a n×n𝑛𝑛n\times n complex matrix, we denote by Tr(𝐀)Tr𝐀\mathrm{Tr}\,(\mathbf{A}) its trace, and by λ1(𝐀),,λn(𝐀)subscript𝜆1𝐀subscript𝜆𝑛𝐀\lambda_{1}(\mathbf{A}),\ldots,\lambda_{n}(\mathbf{A}) its eigenvalues; if moreover 𝐀𝐀\mathbf{A} is Hermitian, they are sorted in decreasing order λ1(𝐀)λn(𝐀)subscript𝜆1𝐀subscript𝜆𝑛𝐀\lambda_{1}(\mathbf{A})\geq\ldots\geq\lambda_{n}(\mathbf{A}). The n×n𝑛𝑛n\times n identity matrix is denoted 𝐈nsubscript𝐈𝑛\mathbf{I}_{n}. The expectation of a complex random variable Z𝑍Z is denoted by 𝔼[Z]𝔼delimited-[]𝑍\mathbb{E}[Z]. The complex circular Gaussian distribution with variance σ2superscript𝜎2\sigma^{2} is denoted as 𝒩(0,σ2)subscript𝒩0superscript𝜎2\mathcal{N}_{\mathbb{C}}(0,\sigma^{2}) and a random vector 𝐱𝐱\mathbf{x} of nsuperscript𝑛\mathbb{C}^{n} follows the 𝒩n(𝟎,𝐑)subscript𝒩superscript𝑛0𝐑\mathcal{N}_{\mathbb{C}^{n}}(\mathbf{0},\mathbf{\mathbf{R}}) distribution if 𝐛𝐱𝒩(0,𝐛𝐑𝐛)similar-tosuperscript𝐛𝐱subscript𝒩0superscript𝐛𝐑𝐛\mathbf{b}^{*}\mathbf{x}\sim\mathcal{N}_{\mathbb{C}}(0,\mathbf{b}^{*}\mathbf{R}\mathbf{b}) for all deterministic (column) vector 𝐛𝐛\mathbf{b} and a fixed n×n𝑛𝑛n\times n positive definite matrix 𝐑𝐑\mathbf{R}. Finally, 𝒞1(I)superscript𝒞1𝐼\mathcal{C}^{1}(I) (resp. 𝒞c1(I)subscriptsuperscript𝒞1𝑐𝐼\mathcal{C}^{1}_{c}(I)) represents the set of continuously differentiable functions (resp. continuously differentiable functions with compact support) on an open set I𝐼I.

II Model and assumptions

Let us consider a M𝑀M–dimensional observed time series (𝐲n)nsubscriptsubscript𝐲𝑛𝑛(\mathbf{y}_{n})_{n\in\mathbb{Z}} defined as

𝐲n=𝐮n+𝐯nsubscript𝐲𝑛subscript𝐮𝑛subscript𝐯𝑛\displaystyle\mathbf{y}_{n}=\mathbf{u}_{n}+\mathbf{v}_{n} (7)

where (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} represents a useful signal and where (𝐯n)nsubscriptsubscript𝐯𝑛𝑛(\mathbf{v}_{n})_{n\in\mathbb{Z}} represents an additive noise. The useful signal is modeled as the output of an unknown stable MIMO filter (𝐇k)ksubscriptsubscript𝐇𝑘𝑘(\mathbf{H}_{k})_{k\in\mathbb{Z}} driven by a non-observable K𝐾K–dimensional complex Gaussian white noise (ϵn)nsubscriptsubscriptbold-italic-ϵ𝑛𝑛(\boldsymbol{\epsilon}_{n})_{n\in\mathbb{Z}} with 𝔼[ϵnϵn]=𝐈K𝔼delimited-[]subscriptbold-italic-ϵ𝑛superscriptsubscriptbold-italic-ϵ𝑛subscript𝐈𝐾\mathbb{E}[\boldsymbol{\epsilon}_{n}\boldsymbol{\epsilon}_{n}^{*}]=\mathbf{I}_{K}, i.e.

𝐮n=k𝐇kϵnksubscript𝐮𝑛subscript𝑘subscript𝐇𝑘subscriptbold-italic-ϵ𝑛𝑘\displaystyle\mathbf{u}_{n}=\sum_{k\in\mathbb{Z}}\mathbf{H}_{k}\boldsymbol{\epsilon}_{n-k}

with probability one. We notice that K𝐾K represents the number of sources in the context of array processing. (𝐯n)nsubscriptsubscript𝐯𝑛𝑛(\mathbf{v}_{n})_{n\in\mathbb{Z}} is modeled as a M𝑀M–dimensional stationary complex Gaussian time series such that its component time series (v1,n)n,,(vM,n)nsubscriptsubscript𝑣1𝑛𝑛subscriptsubscript𝑣𝑀𝑛𝑛(v_{1,n})_{n\in\mathbb{Z}},\ldots,(v_{M,n})_{n\in\mathbb{Z}} are mutually independent.

For each m=1,,M𝑚1𝑀m=1,\ldots,M, we denote by (rm(k))ksubscriptsubscript𝑟𝑚𝑘𝑘(r_{m}(k))_{k\in\mathbb{Z}} the covariance function of (vm,n)nsubscriptsubscript𝑣𝑚𝑛𝑛(v_{m,n})_{n\in\mathbb{Z}}, i.e. rm(k)=𝔼[vm,nvm,nk¯]subscript𝑟𝑚𝑘𝔼delimited-[]subscript𝑣𝑚𝑛¯subscript𝑣𝑚𝑛𝑘r_{m}(k)=\mathbb{E}[v_{m,n}\overline{v_{m,n-k}}], which verifies the following memory assumption.

Assumption 1.

The covariance coefficients decay sufficiently fast in the lag domain, in the sense that

supm1k(1+|k|)2|rm(k)|<.subscriptsupremum𝑚1subscript𝑘superscript1𝑘2subscript𝑟𝑚𝑘\displaystyle\sup_{m\geq 1}\sum_{k\in\mathbb{Z}}(1+|k|)^{2}|r_{m}(k)|<\infty. (8)

In particular, Assumption 1 implies that the spectral density smsubscript𝑠𝑚s_{m} of (vm,n)nsubscriptsubscript𝑣𝑚𝑛𝑛(v_{m,n})_{n\in\mathbb{Z}}, given by

sm(ν)=krm(k)ei2πνksubscript𝑠𝑚𝜈subscript𝑘subscript𝑟𝑚𝑘superscriptei2𝜋𝜈𝑘\displaystyle s_{m}(\nu)=\sum_{k\in\mathbb{Z}}r_{m}(k)\mathrm{e}^{-\mathrm{i}2\pi\nu k}

verifies

supm1supν[0,1]sm(ν)<.subscriptsupremum𝑚1subscriptsupremum𝜈01subscript𝑠𝑚𝜈\displaystyle\sup_{m\geq 1}\sup_{\nu\in[0,1]}s_{m}(\nu)<\infty.

Assumption 1 is in particular verified as soon as the condition

|rm(k)|C(1+|k|)3+δsubscript𝑟𝑚𝑘𝐶superscript1𝑘3𝛿|r_{m}(k)|\leq\frac{C}{(1+|k|)^{3+\delta}} (9)

holds for each k𝑘k\in\mathbb{Z} and each m1𝑚1m\geq 1, where C𝐶C and δ𝛿\delta are positive constants. As the autocovariance function of ARMA signals decreases exponentially towards 00, Assumption 1 thus holds if the time series (vm)m1subscriptsubscript𝑣𝑚𝑚1(v_{m})_{m\geq 1} are ARMA signals, provided some extra purely technical conditions that allow to manage the supremum over m𝑚m in (8) are met. As the spectral coherence matrix of (𝐯n)nsubscriptsubscript𝐯𝑛𝑛(\mathbf{v}_{n})_{n\in\mathbb{Z}}, involves a renormalization by the inverse of the spectral densities smsubscript𝑠𝑚s_{m}, we also need that smsubscript𝑠𝑚s_{m} does not vanish for each m𝑚m.

Assumption 2.

The spectral densities are uniformly bounded away from zero, that is

infm1minν[0,1]sm(ν)>0.subscriptinfimum𝑚1subscript𝜈01subscript𝑠𝑚𝜈0\displaystyle\inf_{m\geq 1}\min_{\nu\in[0,1]}s_{m}(\nu)>0.

Assumptions 1 and 2 also imply that the total noise power satisfies

0<infM11M𝔼𝐯n22supM11M𝔼𝐯n22<.evaluated-at0brasubscriptinfimum𝑀11𝑀𝔼subscript𝐯𝑛22subscriptsupremum𝑀11𝑀𝔼superscriptsubscriptnormsubscript𝐯𝑛22\displaystyle 0<\inf_{M\geq 1}\frac{1}{M}\mathbb{E}\|\mathbf{v}_{n}\|_{2}^{2}\leq\sup_{M\geq 1}\frac{1}{M}\mathbb{E}\|\mathbf{v}_{n}\|_{2}^{2}<\infty. (10)

The next assumption is related to the signal part (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}}. For each ν[0,1]𝜈01\nu\in[0,1], we denote by 𝐇(ν)𝐇𝜈\mathbf{H}(\nu) the Fourier transform of (𝐇k)ksubscriptsubscript𝐇𝑘𝑘(\mathbf{H}_{k})_{k\in\mathbb{Z}}, i.e.

𝐇(ν)=k𝐇kei2πνk𝐇𝜈subscript𝑘subscript𝐇𝑘superscriptei2𝜋𝜈𝑘\displaystyle\mathbf{H}(\nu)=\sum_{k\in\mathbb{Z}}\mathbf{H}_{k}\mathrm{e}^{-\mathrm{i}2\pi\nu k}

and by 𝐡1(ν),,𝐡M(ν)superscript𝐡1𝜈superscript𝐡𝑀𝜈\mathbf{h}^{1}(\nu),\ldots,\mathbf{h}^{M}(\nu) the rows of 𝐇(ν)𝐇𝜈\mathbf{H}(\nu).

Assumption 3.

The MIMO filter coefficient matrices are such that

supM1k(1+|k|)𝐇k2<subscriptsupremum𝑀1subscript𝑘1𝑘subscriptnormsubscript𝐇𝑘2\displaystyle\sup_{M\geq 1}\sum_{k\in\mathbb{Z}}(1+|k|)\left\|\mathbf{H}_{k}\right\|_{2}<\infty (11)

and

limMmaxm=1,,Mmaxν[0,1]𝐡m(ν)2=0.subscript𝑀subscript𝑚1𝑀subscript𝜈01subscriptnormsuperscript𝐡𝑚𝜈20\displaystyle\lim_{M\to\infty}\max_{m=1,\ldots,M}\max_{\nu\in[0,1]}\left\|\mathbf{h}^{m}(\nu)\right\|_{2}=0. (12)

When K𝐾K is fixed while M𝑀M\to\infty, condition (11) in Assumption 3 implies that the total useful signal power remains bounded, i.e.

𝔼𝐮n22=k𝐇kF2=𝒪(1)𝔼superscriptsubscriptnormsubscript𝐮𝑛22subscript𝑘superscriptsubscriptnormsubscript𝐇𝑘𝐹2𝒪1\displaystyle\mathbb{E}\left\|\mathbf{u}_{n}\right\|_{2}^{2}=\sum_{k\in\mathbb{Z}}\left\|\mathbf{H}_{k}\right\|_{F}^{2}=\mathcal{O}(1) (13)

so that, using (10), the SNR vanishes at rate 1M1𝑀\frac{1}{M}, i.e.

𝔼𝐮n22𝔼𝐯n22=𝒪(1M).𝔼superscriptsubscriptnormsubscript𝐮𝑛22𝔼superscriptsubscriptnormsubscript𝐯𝑛22𝒪1𝑀\displaystyle\frac{\mathbb{E}\|\mathbf{u}_{n}\|_{2}^{2}}{\mathbb{E}\|\mathbf{v}_{n}\|_{2}^{2}}=\mathcal{O}\left(\frac{1}{M}\right). (14)

Likewise, condition (12) in Assumption 3 implies that the SNR per time series vanishes, i.e.

𝔼|um,n|2𝔼|vm,n|2=01𝐡m(ν)22dν01sm(ν)dν=o(1)𝔼superscriptsubscript𝑢𝑚𝑛2𝔼superscriptsubscript𝑣𝑚𝑛2superscriptsubscript01superscriptsubscriptnormsuperscript𝐡𝑚𝜈22differential-d𝜈superscriptsubscript01subscript𝑠𝑚𝜈differential-d𝜈𝑜1\displaystyle\frac{\mathbb{E}|u_{m,n}|^{2}}{\mathbb{E}|v_{m,n}|^{2}}=\frac{\int_{0}^{1}\|\mathbf{h}^{m}(\nu)\|_{2}^{2}\mathrm{d}\nu}{\int_{0}^{1}s_{m}(\nu)\mathrm{d}\nu}=o(1) (15)

as M𝑀M\to\infty. We finally notice that (11) is stronger than (13). While 𝔼𝐮n22=𝒪(1)𝔼superscriptsubscriptnormsubscript𝐮𝑛22𝒪1\mathbb{E}\left\|\mathbf{u}_{n}\right\|_{2}^{2}=\mathcal{O}(1) is a rather fundamental assumption that allows to precise the behaviour of the signal to noise ratio, the extra condition supm1k|k|𝐇k2<subscriptsupremum𝑚1subscript𝑘𝑘subscriptnormsubscript𝐇𝑘2\sup_{m\geq 1}\sum_{k}|k|\left\|\mathbf{H}_{k}\right\|_{2}<\infty is essentially motivated by technical reasons (it is needed to establish Theorem 2). However, it is clearly not restrictive in practice.

Remark 1.

Conditions (11) and (12) in Assumption 3 are especially relevant in the context of array processing, where M𝑀M represents the number of sensors, which may be large [24, 25]. In this context, (14) represents the SNR before matched filtering, while (15) represents the SNR per sensor. The use of spatial filtering techniques, which combine the observations y1,n,,yM,nsubscript𝑦1𝑛subscript𝑦𝑀𝑛y_{1,n},\ldots,y_{M,n} across the M𝑀M sensors, allows to increase the SNR by a factor M𝑀M when the second order statistics of (𝐲n)nsubscriptsubscript𝐲𝑛𝑛(\mathbf{y}_{n})_{n\in\mathbb{Z}} are known, which leads to a SNR after matched filtering of the order of magnitude 𝒪(1)𝒪1\mathcal{O}(1). Thus, despite the apparent low SNR, reliable information on the useful signal (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} can potentially still be extracted from the observed signal (𝐲n)nsubscriptsubscript𝐲𝑛𝑛(\mathbf{y}_{n})_{n\in\mathbb{Z}}.

Let 𝐒𝐲subscript𝐒𝐲\mathbf{S}_{\mathbf{y}} denote the spectral density of (𝐲n)nsubscriptsubscript𝐲𝑛𝑛(\mathbf{y}_{n})_{n\in\mathbb{Z}}, given by

𝐒𝐲(ν)=𝐇(ν)𝐇(ν)+𝐒𝐯(ν)subscript𝐒𝐲𝜈𝐇𝜈𝐇superscript𝜈subscript𝐒𝐯𝜈\displaystyle\mathbf{S}_{\mathbf{y}}(\nu)=\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}+\mathbf{S}_{\mathbf{v}}(\nu)

where 𝐒𝐯(ν)=dg(s1(ν),,sM(ν))subscript𝐒𝐯𝜈dgsubscript𝑠1𝜈subscript𝑠𝑀𝜈\mathbf{S}_{\mathbf{v}}(\nu)=\operatorname*{dg}\left(s_{1}(\nu),\ldots,s_{M}(\nu)\right). To estimate 𝐒𝐲subscript𝐒𝐲\mathbf{S}_{\mathbf{y}}, we consider in this paper a frequency-smoothed periodogram 𝐒^𝐲subscript^𝐒𝐲\hat{\mathbf{S}}_{\mathbf{y}}, which we defined in (5). In the classical low-dimensional regime where B,N𝐵𝑁B,N\to\infty while M,K𝑀𝐾M,K remain fixed, it is well-known [12] that

𝔼[𝐒^𝐲(ν)]=𝐒𝐲(ν)+𝒪(B2N2)𝔼delimited-[]subscript^𝐒𝐲𝜈subscript𝐒𝐲𝜈𝒪superscript𝐵2superscript𝑁2\displaystyle\mathbb{E}[\hat{\mathbf{S}}_{\mathbf{y}}(\nu)]=\mathbf{S}_{\mathbf{y}}(\nu)+\mathcal{O}\left(\frac{B^{2}}{N^{2}}\right)

and

𝔼𝐒^𝐲(ν)𝔼[𝐒^𝐲(ν)]22=𝒪(1B).𝔼superscriptsubscriptnormsubscript^𝐒𝐲𝜈𝔼delimited-[]subscript^𝐒𝐲𝜈22𝒪1𝐵\displaystyle\mathbb{E}\left\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)-\mathbb{E}[\hat{\mathbf{S}}_{\mathbf{y}}(\nu)]\right\|_{2}^{2}=\mathcal{O}\left(\frac{1}{B}\right).

Thus, in this regime, 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu) is a consistent estimator of 𝐒𝐲(ν)subscript𝐒𝐲𝜈\mathbf{S}_{\mathbf{y}}(\nu) as long as B𝐵B\to\infty and BN0𝐵𝑁0\frac{B}{N}\to 0. Likewise, the sample Spectral Coherence Matrix (SCM, not to be confused with the sample covariance matrix, which will not be used in this paper) defined in (6) is a consistent estimator of the true SCM 𝐂𝐲(ν)subscript𝐂𝐲𝜈\mathbf{C}_{\mathbf{y}}(\nu) defined in (4). When M+𝑀M\rightarrow+\infty and MN0𝑀𝑁0\frac{M}{N}\rightarrow 0, it can be shown that, under some additional mild extra assumptions, the consistency of 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu) and 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) in the spectral norm sense still holds provided that B𝐵B is chosen in such a way that BN0𝐵𝑁0\frac{B}{N}\rightarrow 0 and MB0𝑀𝐵0\frac{M}{B}\rightarrow 0. In practice, for finite values of M𝑀M and N𝑁N, the above asymptotic regime will allow to predict the performance of various inference schemes in situations where it is possible to choose B𝐵B in such a way that MB𝑀𝐵\frac{M}{B} and BN𝐵𝑁\frac{B}{N} are both small enough. Nevertheless, when the dimension M𝑀M is large and the sample size N𝑁N is not unlimited, or equivalently if MN𝑀𝑁\frac{M}{N} is not small enough, such a choice of B𝐵B may be impossible. In such a context, it seems more relevant to consider asymptotic regimes for which MN0𝑀𝑁0\frac{M}{N}\rightarrow 0 and MB𝑀𝐵\frac{M}{B} converging towards a positive constant. In the following, we will consider the following asymptotic regime.

Assumption 4.

N=N(M)𝑁𝑁𝑀N=N(M) and B=B(M)𝐵𝐵𝑀B=B(M) are both functions of M𝑀M such that, for some α(0,1)𝛼01\alpha\in(0,1),

M=𝒪(Nα) and MBMc(0,1)𝑀𝒪superscript𝑁𝛼 and 𝑀𝐵𝑀absent𝑐01\displaystyle M=\mathcal{O}\left(N^{\alpha}\right)\text{ and }\frac{M}{B}\xrightarrow[M\to\infty]{}c\in(0,1)

while K𝐾K is fixed with respect to M𝑀M.

As MB𝑀𝐵\frac{M}{B} does not converge towards 00, the consistency of 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu) and 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) is lost. This can be explained in a simple way when 𝐮n=0subscript𝐮𝑛0\mathbf{u}_{n}=0 for each n𝑛n and the signals ((vm,n)n)m1subscriptsubscriptsubscript𝑣𝑚𝑛𝑛𝑚1((v_{m,n})_{n\in\mathbb{Z}})_{m\geq 1} are mutually independent i.i.d. 𝒩c(0,σ2)subscript𝒩𝑐0superscript𝜎2\mathcal{N}_{c}(0,\sigma^{2}) distributed sequences. In this context, for each ν𝜈\nu, the renormalized Fourier transform vectors (𝝃𝐲(ν+b/N))b=B/2,,B/2subscriptsubscript𝝃𝐲𝜈𝑏𝑁𝑏𝐵2𝐵2(\boldsymbol{\xi}_{\mathbf{y}}(\nu+b/N))_{b=-B/2,\ldots,B/2} are mutually independent 𝒩(0,σ2𝐈)subscript𝒩0superscript𝜎2𝐈\mathcal{N}_{\mathbb{C}}(0,\sigma^{2}\mathbf{I}) random vectors. The spectral density estimate 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu) defined by (5) thus coincides with the sample covariance matrix of these (B+1)𝐵1(B+1) M𝑀M–dimensional vectors. If B𝐵B and M𝑀M are of the same order to magnitude, it cannot be expected that 𝐒^𝐲(ν)𝔼(𝐒^𝐲(ν))normsubscript^𝐒𝐲𝜈𝔼subscript^𝐒𝐲𝜈\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)-\mathbb{E}(\hat{\mathbf{S}}_{\mathbf{y}}(\nu))\| converges towards 00 because the true covariance matrix 𝔼(𝐒^𝐲(ν))𝔼subscript^𝐒𝐲𝜈\mathbb{E}(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)) to be estimated depends on 𝒪(M2)𝒪superscript𝑀2\mathcal{O}(M^{2}) parameters, and that the number MB𝑀𝐵MB of available scalar observations used to estimate 𝔼(𝐒^𝐲(ν))𝔼subscript^𝐒𝐲𝜈\mathbb{E}(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)) is also 𝒪(M2)𝒪superscript𝑀2\mathcal{O}(M^{2}). Despite the loss of the convergence of the estimators 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu) and 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu), we will see that one can still rely on the high-dimensional structure of these matrices to design relevant test statistics.

III Informal presentation of the proposed test statistic

Mathematical details will reveal later that for each ν𝜈\nu, 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu) behaves as a spike model covariance matrix, whose eigenvalues are precisely described by [20]. More precisely, we will see that, in some sense, the eigenvalues of 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu) that are due to the noise belong to the interval [λ,λ+]subscript𝜆subscript𝜆[\lambda_{-},\lambda_{+}] where λ=(1c)2subscript𝜆superscript1𝑐2\lambda_{-}=(1-\sqrt{c})^{2} and λ+=(1+c)2subscript𝜆superscript1𝑐2\lambda_{+}=(1+\sqrt{c})^{2}, and that in the presence of signal, some eigenvalues of 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu) may be strictly greater than λ+subscript𝜆\lambda_{+} if an SNR criteria is respected. For the remainder, we define

𝒱N={kN:k=0,,N1}subscript𝒱𝑁conditional-set𝑘𝑁𝑘0𝑁1\displaystyle\mathcal{V}_{N}=\left\{\frac{k}{N}:k=0,\ldots,N-1\right\} (16)

the set of Fourier frequencies. A natural way to test for 0subscript0\mathcal{H}_{0} against 1subscript1\mathcal{H}_{1} is to compute the largest eigenvalue of 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu) over the frequencies of 𝒱Nsubscript𝒱𝑁\mathcal{V}_{N}, and compare it with λ+subscript𝜆\lambda_{+}. This leads to the following test statistic:

Tϵ=𝟙[λ++ϵ,)(maxν𝒱Nλ1(𝐂^𝐲(ν))).subscript𝑇italic-ϵsubscriptdouble-struck-𝟙superscript𝜆italic-ϵsubscript𝜈subscript𝒱𝑁subscript𝜆1subscript^𝐂𝐲𝜈\displaystyle T_{\epsilon}=\mathbb{1}_{[\lambda^{+}+\epsilon,\infty)}\left(\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\right). (17)

We will prove later that, under proper assumption on the SNR, this test statistic is consistent in the present high-dimensional regime. Before describing the mathematical details leading to consider Tϵsubscript𝑇italic-ϵT_{\epsilon}, we now provide some numerical illustrations of its behaviour. The general settings are given as follows. The noise is generated as a Gaussian AR(1) process having spectral density

sm(ν)=1|1θei2πν|2,subscript𝑠𝑚𝜈1superscript1𝜃superscriptei2𝜋𝜈2\displaystyle s_{m}(\nu)=\frac{1}{\left|1-\theta\mathrm{e}^{-\mathrm{i}2\pi\nu}\right|^{2}}, (18)

for all m=1,,M𝑚1𝑀m=1,\ldots,M, with θ=0.5𝜃0.5\theta=0.5, whereas for the useful signal, we also consider an AR(1) process by choosing K=1𝐾1K=1 and

𝐇k=CMβk(1,,1)Tsubscript𝐇𝑘𝐶𝑀superscript𝛽𝑘superscript11𝑇\mathbf{H}_{k}=\sqrt{\frac{C}{M}}\beta^{k}(1,\ldots,1)^{T} (19)

with β=1011𝛽1011\beta=\frac{10}{11} and C𝐶C being a positive constant used to adjust the SNR.

In order to understand how the test statistics Tϵsubscript𝑇italic-ϵT_{\epsilon} discriminates between 0subscript0\mathcal{H}_{0} and 1subscript1\mathcal{H}_{1}, we show in Figure 1 the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) for ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N} in the presence of signal, and compare it to the threshold λ+subscript𝜆\lambda_{+}. We see that for some frequencies ν𝜈\nu around 00, the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) deviates significantly from λ+subscript𝜆\lambda_{+}. As we will see later, it is possible to evaluate the asymptotic behaviour of the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu), and to establish that it converges towards ϕ(SNR(ν))italic-ϕ𝑆𝑁𝑅𝜈\phi(SNR(\nu)) where ϕitalic-ϕ\phi is a certain function, and where SNR(ν)𝑆𝑁𝑅𝜈SNR(\nu) can be interpreted as a signal-to-noise ratio at frequency ν𝜈\nu. ϕ(SNR(ν))italic-ϕ𝑆𝑁𝑅𝜈\phi(SNR(\nu)) is also represented in Figure 1, and it is seen that it is close to the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu). In Figure 2, we compare the empirical distribution of Tϵsubscript𝑇italic-ϵT_{\epsilon} under 0subscript0\mathcal{H}_{0} and 1subscript1\mathcal{H}_{1} over 10000 repetitions. We see that the distribution of our test statistic Tϵsubscript𝑇italic-ϵT_{\epsilon} is able to discriminate the scenarios where the data 𝐲nsubscript𝐲𝑛\mathbf{y}_{n} are generated under 0subscript0\mathcal{H}_{0} or 1subscript1\mathcal{H}_{1}, and that Tϵsubscript𝑇italic-ϵT_{\epsilon} is typically over the threshold λ+subscript𝜆\lambda_{+} under 1subscript1\mathcal{H}_{1}.

Refer to caption
Figure 1: Largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) for ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N} vs the threshold λ+=(1+MB+1)2subscript𝜆superscript1𝑀𝐵12\lambda_{+}=(1+\sqrt{\frac{M}{B+1}})^{2}. M=60𝑀60M=60, c=0.5𝑐0.5c=0.5, N=6000𝑁6000N=6000, θ=0.5𝜃0.5\theta=0.5, C=0.05𝐶0.05C=0.05
Refer to caption
Figure 2: Histogram of Tϵsubscript𝑇italic-ϵT_{\epsilon} under 0subscript0\mathcal{H}_{0} and 1subscript1\mathcal{H}_{1}, over 10000 repetitions. M=40𝑀40M=40, c=0.5𝑐0.5c=0.5, N=1000𝑁1000N=1000, θ=0.5𝜃0.5\theta=0.5, C=0.05𝐶0.05C=0.05

IV Approximation results for 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) in the high-dimensional regime

In this section we present the mathematical details which lead to the test statistic (17). More specifically, we provide useful approximation results for 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu), which basically show that 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) behaves as a certain Wishart matrix in the high-dimensional regime. These approximation results are the keystone for the study of the behaviour of the eigenvalues of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) and the detection test proposed in Section V.

We first study separately the signal-free case (i.e. 𝐲n=𝐯nsubscript𝐲𝑛subscript𝐯𝑛\mathbf{y}_{n}=\mathbf{v}_{n}) as well as the noise-free case (i.e. 𝐲n=𝐮nsubscript𝐲𝑛subscript𝐮𝑛\mathbf{y}_{n}=\mathbf{u}_{n}).

IV-A Signal-free case

Let

𝝃𝐯(ν)=1Nn=0N1𝐯nei2πνnsubscript𝝃𝐯𝜈1𝑁superscriptsubscript𝑛0𝑁1subscript𝐯𝑛superscriptei2𝜋𝜈𝑛\boldsymbol{\xi}_{\mathbf{v}}(\nu)=\frac{1}{\sqrt{N}}\sum_{n=0}^{N-1}\mathbf{v}_{n}\mathrm{e}^{-\mathrm{i}2\pi\nu n}

denote the discrete (time-limited) Fourier transform of (𝐯n)n=0,,N1subscriptsubscript𝐯𝑛𝑛0𝑁1(\mathbf{v}_{n})_{n=0,\ldots,N-1}, and define the M×(B+1)𝑀𝐵1M\times(B+1) matrix 𝚺𝐯(ν)subscript𝚺𝐯𝜈\boldsymbol{\Sigma}_{\mathbf{v}}(\nu) as

𝚺𝐯(ν)=1B+1[𝝃𝐯(νB2N),,𝝃𝐯(ν+B2N)].subscript𝚺𝐯𝜈1𝐵1subscript𝝃𝐯𝜈𝐵2𝑁subscript𝝃𝐯𝜈𝐵2𝑁\displaystyle\boldsymbol{\Sigma}_{\mathbf{v}}(\nu)=\frac{1}{\sqrt{B+1}}\left[\boldsymbol{\xi}_{\mathbf{v}}\left(\nu-\frac{B}{2N}\right),\ldots,\boldsymbol{\xi}_{\mathbf{v}}\left(\nu+\frac{B}{2N}\right)\right].

The following result, derived in [18], reveals an interesting behaviour of the frequency-smoothed periodogram of the noise.

Theorem 1.

Under Assumptions 1, 2 and 4, for all ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N}, there exists an M×(B+1)𝑀𝐵1M\times(B+1) matrix 𝐙(ν)𝐙𝜈\mathbf{Z}(\nu) with i.i.d. 𝒩(0,1)subscript𝒩01\mathcal{N}_{\mathbb{C}}(0,1) entries such that

maxν𝒱N𝚺𝐯(ν)1B+1𝐒𝐯(ν)1/2𝐙(ν)2Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\Sigma}_{\mathbf{v}}(\nu)-\frac{1}{\sqrt{B+1}}\mathbf{S}_{\mathbf{v}}(\nu)^{1/2}\mathbf{Z}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0.

Informally speaking, Theorem 1 shows that the random vectors 1B+1𝝃𝐯(νBN)1𝐵1subscript𝝃𝐯𝜈𝐵𝑁\frac{1}{\sqrt{B+1}}\boldsymbol{\xi}_{\mathbf{v}}\left(\nu-\frac{B}{N}\right),…,1B+1𝝃𝐯(ν+BN)1𝐵1subscript𝝃𝐯𝜈𝐵𝑁\frac{1}{\sqrt{B+1}}\boldsymbol{\xi}_{\mathbf{v}}\left(\nu+\frac{B}{N}\right) asymptotically behave as a family of i.i.d. 𝒩M(𝟎,𝐒𝐯(ν))subscript𝒩superscript𝑀0subscript𝐒𝐯𝜈\mathcal{N}_{\mathbb{C}^{M}}(\mathbf{0},\mathbf{S}_{\mathbf{v}}(\nu)) vectors, for all ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N}. Moreover, if

𝐒^𝐯(ν):=1B+1b=B/2B/2𝝃𝐯(ν+bN)𝝃𝐯(ν+bN)assignsubscript^𝐒𝐯𝜈1𝐵1superscriptsubscript𝑏𝐵2𝐵2subscript𝝃𝐯𝜈𝑏𝑁subscript𝝃𝐯superscript𝜈𝑏𝑁\displaystyle\hat{\mathbf{S}}_{\mathbf{v}}(\nu):=\frac{1}{B+1}\sum_{b=-B/2}^{B/2}\boldsymbol{\xi}_{\mathbf{v}}\left(\nu+\frac{b}{N}\right)\boldsymbol{\xi}_{\mathbf{v}}\left(\nu+\frac{b}{N}\right)^{*}

denotes the frequency-smoothed periodogram of the noise observations (𝐯n)nsubscriptsubscript𝐯𝑛𝑛(\mathbf{v}_{n})_{n\in\mathbb{Z}}, we deduce that 𝐒^𝐯(ν)subscript^𝐒𝐯𝜈\hat{\mathbf{S}}_{\mathbf{v}}(\nu) asymptotically behaves as a complex Gaussian Wishart matrix with covariance matrix 𝐒𝐯(ν)subscript𝐒𝐯𝜈\mathbf{S}_{\mathbf{v}}(\nu), thanks to the following corollary.

Corollary 1.

Under the assumptions of Theorem 1, it holds that

maxν𝒱N𝐒^𝐯(ν)𝐒𝐯(ν)1/2𝐙(ν)𝐙(ν)B+1𝐒𝐯(ν)1/22Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{v}}(\nu)-\mathbf{S}_{\mathbf{v}}(\nu)^{1/2}\frac{\mathbf{Z}(\nu)\mathbf{Z}(\nu)^{*}}{B+1}\mathbf{S}_{\mathbf{v}}(\nu)^{1/2}\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0.
Proof:

The proof is deferred to Appendix D-A. ∎

It is worth noticing that Corollary 1 implies in particular

maxν𝒱Ndg(𝐒^𝐯(ν))𝐒𝐯(ν)Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{v}}(\nu)\right)-\mathbf{S}_{\mathbf{v}}(\nu)\right\|\xrightarrow[M\to\infty]{a.s.}0

and consequently dg(𝐒^(ν))dg^𝐒𝜈\operatorname*{dg}(\hat{\mathbf{S}}(\nu)) is a consistent estimator of the noise spectral density 𝐒𝐯(ν)subscript𝐒𝐯𝜈\mathbf{S}_{\mathbf{v}}(\nu) in the operator norm sense, at each Fourier frequency ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N}. This convergence may be directly obtained using Lemma 1 in Appendix A and we omit the details since this result is well-known.

IV-B Noise-free case

Let

𝝃𝐮(ν)=1Nn=0N1𝐮nei2πνnsubscript𝝃𝐮𝜈1𝑁superscriptsubscript𝑛0𝑁1subscript𝐮𝑛superscriptei2𝜋𝜈𝑛\boldsymbol{\xi}_{\mathbf{u}}(\nu)=\frac{1}{\sqrt{N}}\sum_{n=0}^{N-1}\mathbf{u}_{n}\mathrm{e}^{-\mathrm{i}2\pi\nu n}

and let 𝚺𝐮(ν)subscript𝚺𝐮𝜈\boldsymbol{\Sigma}_{\mathbf{u}}(\nu) be the K×(B+1)𝐾𝐵1K\times(B+1) matrix defined as

𝚺𝐮(ν)=1B+1[𝝃𝐮(νB2N),,𝝃𝐮(ν+B2N)].subscript𝚺𝐮𝜈1𝐵1subscript𝝃𝐮𝜈𝐵2𝑁subscript𝝃𝐮𝜈𝐵2𝑁\displaystyle\boldsymbol{\Sigma}_{\mathbf{u}}(\nu)=\frac{1}{\sqrt{B+1}}\left[\boldsymbol{\xi}_{\mathbf{u}}\left(\nu-\frac{B}{2N}\right),\ldots,\boldsymbol{\xi}_{\mathbf{u}}\left(\nu+\frac{B}{2N}\right)\right].

In the same way, we also denote by 𝝃ϵsubscript𝝃bold-italic-ϵ\boldsymbol{\xi}_{\boldsymbol{\epsilon}} the normalized discrete (time-limited) Fourier transform of (ϵn)n=0,,N1subscriptsubscriptbold-italic-ϵ𝑛𝑛0𝑁1(\boldsymbol{\epsilon}_{n})_{n=0,\ldots,N-1}, and consider the K×(B+1)𝐾𝐵1K\times(B+1) matrix 𝚺ϵ(ν)subscript𝚺bold-italic-ϵ𝜈\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu) defined as 𝚺𝐮(ν)subscript𝚺𝐮𝜈\boldsymbol{\Sigma}_{\mathbf{u}}(\nu). We then have the following important approximation result.

Theorem 2.

Under Assumptions 3 and 4, it holds that

maxν𝒱N𝚺𝐮(ν)𝐇(ν)𝚺ϵ(ν)2Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\Sigma}_{\mathbf{u}}(\nu)-\mathbf{H}(\nu)\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0.
Proof:

The proof is deferred to Appendix B. ∎

As in Theorem 1, Theorem 2 shows that the random vectors 𝝃𝐮(νBN),,𝝃𝐮(ν+BN)subscript𝝃𝐮𝜈𝐵𝑁subscript𝝃𝐮𝜈𝐵𝑁\boldsymbol{\xi}_{\mathbf{u}}\left(\nu-\frac{B}{N}\right),\ldots,\boldsymbol{\xi}_{\mathbf{u}}\left(\nu+\frac{B}{N}\right) asymptotically behave as the i.i.d. vectors 𝐇(ν)𝝃ϵ(νBN),,𝐇(ν)𝝃ϵ(ν+BN)𝐇𝜈subscript𝝃bold-italic-ϵ𝜈𝐵𝑁𝐇𝜈subscript𝝃bold-italic-ϵ𝜈𝐵𝑁\mathbf{H}(\nu)\boldsymbol{\xi}_{\boldsymbol{\epsilon}}\left(\nu-\frac{B}{N}\right),\ldots,\mathbf{H}(\nu)\boldsymbol{\xi}_{\boldsymbol{\epsilon}}\left(\nu+\frac{B}{N}\right), for all ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N}.

Remark 2.

The type of approximation given in Theorem 2 is well-known in the low-dimensional regime in which M,K,B𝑀𝐾𝐵M,K,B are fixed while N𝑁N\to\infty. Indeed, in that case, we have [12, Th. 4.5.2]

maxν[0,1]𝚺𝐮(ν)𝐇(ν)𝚺ϵ(ν)2=𝒪P(log(N)N).subscript𝜈01subscriptnormsubscript𝚺𝐮𝜈𝐇𝜈subscript𝚺bold-italic-ϵ𝜈2subscript𝒪𝑃𝑁𝑁\displaystyle\max_{\nu\in[0,1]}\left\|\boldsymbol{\Sigma}_{\mathbf{u}}(\nu)-\mathbf{H}(\nu)\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)\right\|_{2}=\mathcal{O}_{P}\left(\sqrt{\frac{\log(N)}{N}}\right).

In the high-dimensional regime where M𝑀M and B𝐵B also converge to infinity as described in Assumption 4, the result of Theorem 2 cannot be obtained from [12, Th. 4.5.2] and thus requires a new study.

We also deduce the following approximation result on the frequency-smoothed periodogram of the signal observations (𝐮n)n=0,,N1subscriptsubscript𝐮𝑛𝑛0𝑁1(\mathbf{u}_{n})_{n=0,\ldots,N-1} given by

𝐒^𝐮(ν):=1B+1b=B/2B/2𝝃𝐮(ν+bN)𝝃𝐮(ν+bN).assignsubscript^𝐒𝐮𝜈1𝐵1superscriptsubscript𝑏𝐵2𝐵2subscript𝝃𝐮𝜈𝑏𝑁subscript𝝃𝐮superscript𝜈𝑏𝑁\displaystyle\hat{\mathbf{S}}_{\mathbf{u}}(\nu):=\frac{1}{B+1}\sum_{b=-B/2}^{B/2}\boldsymbol{\xi}_{\mathbf{u}}\left(\nu+\frac{b}{N}\right)\boldsymbol{\xi}_{\mathbf{u}}\left(\nu+\frac{b}{N}\right)^{*}.
Corollary 2.

Under the assumptions of Theorem 2, it holds that

maxν𝒱N𝐒^𝐮(ν)𝐇(ν)𝐇(ν)2Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{u}}(\nu)-\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0.
Proof:

The proof is deferred to Appendix D-B. ∎

As a result of Corollary 2, we deduce that the frequency-smoothed periodogram 𝐒^𝐮(ν)subscript^𝐒𝐮𝜈\hat{\mathbf{S}}_{\mathbf{u}}(\nu) is a consistent estimator of the spectral density 𝐒𝐮(ν)=𝐇(ν)𝐇(ν)subscript𝐒𝐮𝜈𝐇𝜈𝐇superscript𝜈\mathbf{S}_{\mathbf{u}}(\nu)=\mathbf{H}(\nu)\mathbf{H}(\nu)^{*} of (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} in the high-dimensional regime, for each ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N}.

Having characterized the pure noise and pure signal cases, we are now in position to study the high-dimensional behaviour of the spectral coherence matrix 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu).

IV-C The signal-plus-noise case

First, using Corollaries 1 and 2, we deduce the high-dimensional behaviour of the frequency smoothed periodogram 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu). The following results show that, as it could be expected, the frequency smoothed periodogram essentially behaves as a colored Wishart matrix in the large asymptotic regime.

Proposition 1.

For all ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N}, there exists an M×(B+1)𝑀𝐵1M\times(B+1) matrix 𝐗(ν)𝐗𝜈\mathbf{X}(\nu) with i.i.d. 𝒩(0,1)subscript𝒩01\mathcal{N}_{\mathbb{C}}(0,1) entries such that

maxν𝒱N𝐒^𝐲(ν)𝐒𝐲(ν)12𝐗(ν)𝐗(ν)B+1𝐒𝐲(ν)122Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)-\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0. (20)
Proof:

The proof is deferred to Appendix D-C. ∎

We finally consider the study of the spectral coherence 𝐂^𝐲(ν)=dg(𝐒^𝐲(ν))12𝐒^𝐲(ν)dg(𝐒^𝐲(ν)12\hat{\mathbf{C}}_{\mathbf{y}}(\nu)=\operatorname*{dg}(\hat{\mathbf{S}}_{\mathbf{y}}(\nu))^{-\frac{1}{2}}\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\operatorname*{dg}(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)^{-\frac{1}{2}}. From condition (12) in Assumption 3 on the SNR, it turns out that (cf. proof of Theorem 3 below where the result is shown) that

maxν𝒱Ndg(𝐒^𝐲(ν))𝐒𝐯(ν)2Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)-\mathbf{S}_{\mathbf{v}}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0. (21)

This approximation result regarding the normalization term dg(𝐒^𝐲(ν))dgsubscript^𝐒𝐲𝜈\operatorname*{dg}(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)) in the SCM naturally leads to the following theorem, which is the key result of this paper.

Theorem 3.

Under Assumptions 1, 2, 3 and 4,

maxν𝒱N𝐂^𝐲(ν)𝚵(ν)12𝐗(ν)𝐗(ν)B+1𝚵(ν)122Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{C}}_{\mathbf{y}}(\nu)-\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0

where

𝚵(ν)=𝐒𝐯(ν)12𝐇(ν)𝐇(ν)𝐒𝐯(ν)12+𝐈M.𝚵𝜈subscript𝐒𝐯superscript𝜈12𝐇𝜈𝐇superscript𝜈subscript𝐒𝐯superscript𝜈12subscript𝐈𝑀\displaystyle\boldsymbol{\Xi}(\nu)=\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}+\mathbf{I}_{M}.

and 𝐗(ν)𝐗𝜈\mathbf{X}(\nu) is the matrix defined in Proposition 1.

Proof:

The proof is deferred to Appendix C. ∎

Let us make a few important comments regarding the result of Theorem 3.

First, used in conjunction with Weyl’s inequalities [26, Th. 4.3.1], Theorem 3 implies in particular that each eigenvalue of the SCM 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) behaves as its counterpart of the Wishart matrix

𝐖(ν)=𝚵(ν)12𝐗(ν)𝐗(ν)B+1𝚵(ν)12𝐖𝜈𝚵superscript𝜈12𝐗𝜈𝐗superscript𝜈𝐵1𝚵superscript𝜈12\mathbf{W}(\nu)=\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}

for ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N}, that is

maxm=1,,Mmaxν𝒱N|λm(𝐂^𝐲(ν))λm(𝐖(ν))|Ma.s.0.\displaystyle\max_{m=1,\ldots,M}\max_{\nu\in\mathcal{V}_{N}}\Bigl{|}\lambda_{m}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)-\lambda_{m}\left(\mathbf{W}(\nu)\right)\Bigr{|}\xrightarrow[M\to\infty]{a.s.}0. (22)

Second, Theorem 3 has an important consequence regarding the behaviour of linear spectral statistics of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu), that is statistics of the type

Lφ(ν)=1Mm=1Mφ(λm(𝐂^𝐲(ν)))subscript𝐿𝜑𝜈1𝑀superscriptsubscript𝑚1𝑀𝜑subscript𝜆𝑚subscript^𝐂𝐲𝜈\displaystyle L_{\varphi}(\nu)=\frac{1}{M}\sum_{m=1}^{M}\varphi\left(\lambda_{m}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\right) (23)

where φ𝜑\varphi belongs to a certain class of functions.

Corollary 3.

Let φ𝒞1((0,+))𝜑superscript𝒞10\varphi\in\mathcal{C}^{1}\left((0,+\infty)\right). Under Assumptions 1, 2, 3 and 4, we have

maxν𝒱N|Lφ(ν)φ(λ)f(λ)dλ|Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left|L_{\varphi}(\nu)-\int_{\mathbb{R}}\varphi(\lambda)f(\lambda)\mathrm{d}\lambda\right|\xrightarrow[M\to\infty]{a.s.}0

where f𝑓f is the density of the Marcenko-Pastur distribution given by

f(λ)=(λλ)(λ+λ)2πcλ𝟙[λ,λ+](λ)𝑓𝜆𝜆superscript𝜆superscript𝜆𝜆2𝜋𝑐𝜆subscriptdouble-struck-𝟙superscript𝜆superscript𝜆𝜆\displaystyle f(\lambda)=\frac{\sqrt{(\lambda-\lambda^{-})(\lambda^{+}-\lambda)}}{2\pi c\lambda}\mathbb{1}_{[\lambda^{-},\lambda^{+}]}(\lambda)

with λ±=(1±c)2superscript𝜆plus-or-minussuperscriptplus-or-minus1𝑐2\lambda^{\pm}=\left(1\pm\sqrt{c}\right)^{2}.

Proof:

The proof is deferred to Appendix E. ∎

Therefore, Corollary 3 shows that linear spectral statistics of the SCM converge to the same limit regardless of whether the observations contain only pure noise or signal-plus-noise contributions. This shows that any test statistic solely relying on linear spectral statistics of the SCM is unable to distinguish between absence or presence of useful signal, and cannot be consistent in the high-dimensional regime. Nevertheless, in the next section we will see that we can exploit Theorem 3 to build a new test statistic based on the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu), which is proved to be consistent in the high-dimensional regime.

Remark 3.

Corollary 1, Corollary 2 and Theorems 3 may be interpreted in the context of array processing. Indeed, in the time model (7), usually referred to as “wideband”, the signal contribution (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} modeled as a linear process, is in general not confined to a low-dimensional subspace (i.e. with dimension less than M𝑀M). However, in the frequency domain, Corollary 1 and Corollary 2 show that we can retrieve, in the high-dimensional regime, a “narrowband” model, since the useful signal is confined to a K𝐾K–dimensional subspace of Msuperscript𝑀\mathbb{C}^{M}. Thus, standard narrowband techniques used in array processing for detection may be used, see e.g. [27].

V A new consistent test statistic

As we have seen in Theorem 3 and the related comments, the SCM 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) behaves in the high-dimensional regime as a Wishart matrix with scale 𝚵(ν)=𝐒𝐯(ν)12𝐇(ν)𝐇(ν)𝐒𝐯(ν)12+𝐈M𝚵𝜈subscript𝐒𝐯superscript𝜈12𝐇𝜈𝐇superscript𝜈subscript𝐒𝐯superscript𝜈12subscript𝐈𝑀\boldsymbol{\Xi}(\nu)=\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}+\mathbf{I}_{M} being a fixed rank K𝐾K perturbation of the identity matrix. The behaviour of the eigenvalues for each ν𝜈\nu of such matrix model is well-known since [20] (and other related works such as the well-known BBP-phase-transition [28] or [29]), and the rest of this section is devoted to the application of the results from [20] in our frequency-domain detection context. A crucial point is to choose the particular frequency at which the above mentioned results will be used in order to obtain information on the behaviour of maxν𝒱Nλ1(𝐂^𝐲(ν))subscript𝜈subscript𝒱𝑁subscript𝜆1subscript^𝐂𝐲𝜈\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right). For this, we have first to define some notations. We consider the fundamental function ϕitalic-ϕ\phi which already appears in [20]:

ϕ(x)={(x+1)(x+c)x if x>cλ+ if xcitalic-ϕ𝑥cases𝑥1𝑥𝑐𝑥 if 𝑥𝑐superscript𝜆 if 𝑥𝑐\displaystyle\phi(x)=\begin{cases}\frac{(x+1)(x+c)}{x}&\quad\text{ if }x>\sqrt{c}\\ \lambda^{+}&\quad\text{ if }x\leq\sqrt{c}\end{cases}

where we recall that λ+=(1+c)2superscript𝜆superscript1𝑐2\lambda^{+}=(1+\sqrt{c})^{2} (see Corollary 3). We notice that for all x>c𝑥𝑐x>\sqrt{c}, ϕ(x)>ϕ(c)=λ+italic-ϕ𝑥italic-ϕ𝑐superscript𝜆\phi(x)>\phi(\sqrt{c})=\lambda^{+}. Define as γ(ν)𝛾𝜈\gamma(\nu) the maximum eigenvalue of the finite rank perturbation for each ν𝜈\nu, that is

γ(ν)=λ1(𝐒𝐯(ν)12𝐇(ν)𝐇(ν)𝐒𝐯(ν)12)𝛾𝜈subscript𝜆1subscript𝐒𝐯superscript𝜈12𝐇𝜈𝐇superscript𝜈subscript𝐒𝐯superscript𝜈12\displaystyle\gamma(\nu)=\lambda_{1}\left(\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\right) (24)

and let νN𝒱Nsuperscriptsubscript𝜈𝑁subscript𝒱𝑁\nu_{N}^{*}\in\mathcal{V}_{N} such that

νNargmaxν𝒱Nγ(ν).superscriptsubscript𝜈𝑁subscriptargmax𝜈subscript𝒱𝑁𝛾𝜈\displaystyle\nu_{N}^{*}\in\operatorname*{argmax}_{\nu\in\mathcal{V}_{N}}\gamma(\nu).

We remark that γ(νN)𝛾subscriptsuperscript𝜈𝑁\gamma(\nu^{*}_{N}) may be interpreted as a certain SNR metric in the frequency domain. In the following, we study the behaviour of the largest eigenvalue of 𝐂^𝐲(νN)subscript^𝐂𝐲subscriptsuperscript𝜈𝑁\hat{\mathbf{C}}_{\mathbf{y}}(\nu^{*}_{N}), which requires the following additional assumption on γ(νN)𝛾subscriptsuperscript𝜈𝑁\gamma(\nu^{*}_{N}).

Assumption 5.

There exists γ0subscript𝛾0\gamma_{\infty}\geq 0 such that

γ(νN)Mγ.𝑀absent𝛾subscriptsuperscript𝜈𝑁subscript𝛾\displaystyle\gamma(\nu^{*}_{N})\xrightarrow[M\to\infty]{}\gamma_{\infty}.

Theorem 3 implies that the eigenvalues of 𝐂^𝐲(νN)subscript^𝐂𝐲subscriptsuperscript𝜈𝑁\hat{\mathbf{C}}_{\mathbf{y}}(\nu^{*}_{N}) have the same asymptotic behaviour as the corresponding eigenvalues of matrix 𝚵(νN)12𝐗(νN)𝐗(νN)B+1𝚵(νN)12𝚵superscriptsubscriptsuperscript𝜈𝑁12𝐗subscriptsuperscript𝜈𝑁𝐗superscriptsubscriptsuperscript𝜈𝑁𝐵1𝚵superscriptsubscriptsuperscript𝜈𝑁12\boldsymbol{\Xi}(\nu^{*}_{N})^{\frac{1}{2}}\frac{\mathbf{X}(\nu^{*}_{N})\mathbf{X}(\nu^{*}_{N})^{*}}{B+1}\boldsymbol{\Xi}(\nu^{*}_{N})^{\frac{1}{2}}. Under Assumption 5, [28], [20] or [29] immediately imply the following result. Note that since νNsubscriptsuperscript𝜈𝑁\nu^{*}_{N} is unknown in practice, this proposition is an intermediate theoretical result that will justify the detection test statistic introduced below.

Proposition 2.

Under Assumptions 1, 2, 3, 4 and 5, we have

λ1(𝐂^𝐲(νN))Ma.s.ϕ(γ)\displaystyle\lambda_{1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu_{N}^{*})\right)\xrightarrow[M\to\infty]{a.s.}\phi(\gamma_{\infty}) (25)

while

λK+1(𝐂^𝐲(νN))Ma.s.λ+\displaystyle\lambda_{K+1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu_{N}^{*})\right)\xrightarrow[M\to\infty]{a.s.}\lambda^{+} (26)

and

λM(𝐂^𝐲(νN))Ma.s.λ.\displaystyle\lambda_{M}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu_{N}^{*})\right)\xrightarrow[M\to\infty]{a.s.}\lambda^{-}. (27)

Moreover, if γ=0subscript𝛾0\gamma_{\infty}=0,

lim supMmaxν𝒱Nλ1(𝐂^𝐲(ν))λ+a.s.subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆1subscript^𝐂𝐲𝜈superscript𝜆a.s.\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\leq\lambda^{+}\quad\text{a.s.} (28)
Proof:

It just remains to establish (28), see Appendix F. ∎

Since neither the intrinsic dimensionality K𝐾K of the useful signal (𝐮n)nsubscriptsubscript𝐮𝑛𝑛(\mathbf{u}_{n})_{n\in\mathbb{Z}} nor the frequency νNsubscriptsuperscript𝜈𝑁\nu^{*}_{N} are known in practice, we use the largest eigenvalue of the SCM maximized over all Fourier frequencies as a test statistic. This leads to the test statistic Tϵsubscript𝑇italic-ϵT_{\epsilon} defined previously in (17) which we recall here:

Tϵ=𝟙[λ++ϵ,)(maxν𝒱Nλ1(𝐂^𝐲(ν))).subscript𝑇italic-ϵsubscriptdouble-struck-𝟙superscript𝜆italic-ϵsubscript𝜈subscript𝒱𝑁subscript𝜆1subscript^𝐂𝐲𝜈\displaystyle T_{\epsilon}=\mathbb{1}_{[\lambda^{+}+\epsilon,\infty)}\left(\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\right).

It turns out that this test statistics is consistent in the high-dimensional regime, as stated in the following result.

Proposition 3.

Under Assumptions 1, 2, 3, 4 and 5, and if under Hypothesis 1subscript1\mathcal{H}_{1},

γ>csubscript𝛾𝑐\displaystyle\gamma_{\infty}>\sqrt{c}

then for all 0<ϵ<ϕ(γ)λ+0italic-ϵitalic-ϕsubscript𝛾superscript𝜆0<\epsilon<\phi(\gamma_{\infty})-\lambda^{+} and i{0,1}𝑖01i\in\{0,1\},

i(limMTϵ=i)=1subscript𝑖subscript𝑀subscript𝑇italic-ϵ𝑖1\displaystyle\mathbb{P}_{i}\left(\lim_{M\to\infty}T_{\epsilon}=i\right)=1

where isubscript𝑖\mathbb{P}_{i} is the underlying probability measure under Hypothesis isubscript𝑖\mathcal{H}_{i}.

Proof:

Under Hypothesis 0subscript0\mathcal{H}_{0}, since γ=0subscript𝛾0\gamma_{\infty}=0, we directly apply (28) in Proposition 2 to obtain that for all ϵ>0italic-ϵ0\epsilon>0, Tϵ=0subscript𝑇italic-ϵ0T_{\epsilon}=0 with probability one, for all large M𝑀M. Under Hypothesis 1subscript1\mathcal{H}_{1}, we get

lim infMmaxν𝒱Nλ1(𝐂^𝐲(ν))limMλ1(𝐂^𝐲(νN))=ϕ(γ)subscriptlimit-infimum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆1subscript^𝐂𝐲𝜈subscript𝑀subscript𝜆1subscript^𝐂𝐲subscriptsuperscript𝜈𝑁italic-ϕsubscript𝛾\displaystyle\liminf_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\geq\lim_{M\to\infty}\lambda_{1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu^{*}_{N})\right)=\phi(\gamma_{\infty})

with probability one. Since by assumption, ϕ(γ)>λ++ϵitalic-ϕsubscript𝛾superscript𝜆italic-ϵ\phi(\gamma_{\infty})>\lambda^{+}+\epsilon, we deduce that Tϵ=1subscript𝑇italic-ϵ1T_{\epsilon}=1 with probability one for all large M𝑀M. ∎

VI Simulations

In this section, we provide some numerical illustrations of the approximation results of Section IV. We will consider the case where the rank K𝐾K of the signal is equal to one and then the case where K𝐾K is strictly greater than one.

VI-A Case K=1𝐾1K=1

As in the numerical simulation presented in Section III, each component of the noise 𝐯nsubscript𝐯𝑛\mathbf{v}_{n} is generated as a Gaussian AR(1) process with θ=0.5𝜃0.5\theta=0.5. The expression of its spectral density smsubscript𝑠𝑚s_{m} for all m=1,,M𝑚1𝑀m=1,\ldots,M is still given in (18). The useful signal is generated as an AR(1) process with K=1𝐾1K=1, 𝐇ksubscript𝐇𝑘\mathbf{H}_{k} defined by (19) and β=1011𝛽1011\beta=\frac{10}{11}. C𝐶C is again a positive constant used to tune the SNR. Note that, in this context, the SNR γ(ν)𝛾𝜈\gamma(\nu) at frequency ν𝜈\nu defined in (24) takes the form

γ(ν)=C|1θei2πν1βei2πν|2.𝛾𝜈𝐶superscript1𝜃superscriptei2𝜋𝜈1𝛽superscriptei2𝜋𝜈2\displaystyle\gamma(\nu)=C\left|\frac{1-\theta\mathrm{e}^{-\mathrm{i}2\pi\nu}}{1-\beta\mathrm{e}^{-\mathrm{i}2\pi\nu}}\right|^{2}.

Figures 3 and 4 illustrate the signal-free case C=0𝐶0C=0, and where (N,M,B)=(20000,100,200)𝑁𝑀𝐵20000100200(N,M,B)=(20000,100,200). In Figure 3, we plot the histogram of the eigenvalues of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) for ν=0𝜈0\nu=0.

Refer to caption
Figure 3: Eigenvalue distribution of 𝐂^𝐲(0)subscript^𝐂𝐲0\hat{\mathbf{C}}_{\mathbf{y}}(0) vs the density of the Marcenko-Pastur distribution with parameter c=1/2𝑐12c=1/2.

As predicted by Corollary 3 in the signal-free case, the empirical eigenvalue distribution of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) is well approximated by the Marcenko-Pastur distribution with shape parameter c=0.5M/(B+1)𝑐0.5𝑀𝐵1c=0.5\approx M/(B+1). Figure 4 further illustrates this convergence, where the cumulative distribution function (cdf) of the Marcenko-Pastur distribution is plotted against the two following quantities:

Fmin(t)subscript𝐹𝑡\displaystyle F_{\min}(t) =minν𝒱N1Mλi(𝐂^(ν))<tδλi(𝐂^(ν))absentsubscript𝜈subscript𝒱𝑁1𝑀subscriptsubscript𝜆𝑖^𝐂𝜈𝑡subscript𝛿subscript𝜆𝑖^𝐂𝜈\displaystyle=\min_{\nu\in\mathcal{V}_{N}}\frac{1}{M}\sum_{\lambda_{i}(\hat{\mathbf{C}}(\nu))<t}\delta_{\lambda_{i}(\hat{\mathbf{C}}(\nu))}
Fmax(t)subscript𝐹𝑡\displaystyle F_{\max}(t) =maxν𝒱N1Mλi(𝐂^(ν))<tδλi(𝐂^(ν)).absentsubscript𝜈subscript𝒱𝑁1𝑀subscriptsubscript𝜆𝑖^𝐂𝜈𝑡subscript𝛿subscript𝜆𝑖^𝐂𝜈\displaystyle=\max_{\nu\in\mathcal{V}_{N}}\frac{1}{M}\sum_{\lambda_{i}(\hat{\mathbf{C}}(\nu))<t}\delta_{\lambda_{i}(\hat{\mathbf{C}}(\nu))}.

These two functions represent the maximum deviations (from above and below) over the frequencies ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N} of the empirical spectral distribution of 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu) against the Marcenko-Pastur distribution. As suggested by the uniform convergence in the frequency domain in Corollary 3, the Marcenko-Pastur approximation in the high-dimensional regime is reliable over the whole set of Fourier frequencies. Note that the statement of Corollary 3 does not exactly match the setting used in Figure 4, as the test function used here is not in 𝒞1((0,+))superscript𝒞10\mathcal{C}^{1}((0,+\infty)).

Refer to caption
Figure 4: Uniform convergence of the eigenvalue distribution of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) over ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N} toward the Marcenko-Pastur distribution with parameter c=1/2𝑐12c=1/2.

To illustrate the signal-plus-noise case and the results of Corollary 3 and Proposition 2, we plot in Figure 5, the histogram of the eigenvalues of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) for ν=0𝜈0\nu=0, with γ(0)=2.9𝛾02.9\gamma(0)=2.9. We see that the largest eigenvalue deviates from the right edge (1+c)2superscript1𝑐2(1+\sqrt{c})^{2} and is located around the value ϕ(γ(0))=4.5italic-ϕ𝛾04.5\phi\left(\gamma(0)\right)=4.5, as predicted by Proposition 2, while all the other eigenvalues spread as the Marcenko-Pastur distribution, as predicted by Corollary 3.

Refer to caption
Figure 5: Eigenvalue distribution of 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu) vs Marcenko-Pastur distribution with parameter c=1/2𝑐12c=1/2 in the signal case.

In order to compare the test statistic (17) with other frequency domain methods based on the SCM, we consider:

  • the new test statistic (17), denoted as LE (for largest eigenvalue),

  • two tests based on LSS of the SCM given by

    Tϵ=𝟙[ϵ,+)(maxν𝒱N|Lφ(ν)φ(λ)f(λ)dλ|)subscriptsuperscript𝑇italic-ϵsubscriptdouble-struck-𝟙italic-ϵsubscript𝜈subscript𝒱𝑁subscript𝐿𝜑𝜈subscript𝜑𝜆𝑓𝜆differential-d𝜆\displaystyle T^{\prime}_{\epsilon}=\mathbb{1}_{[\epsilon,+\infty)}\left(\max_{\nu\in\mathcal{V}_{N}}\left|L_{\varphi}(\nu)-\int_{\mathbb{R}}\varphi(\lambda)f(\lambda)\mathrm{d}\lambda\right|\right)

    where Lφsubscript𝐿𝜑L_{\varphi} and density f𝑓f are defined in (23) and Corollary 3 respectively, and with φ(x)=(x1)2𝜑𝑥superscript𝑥12\varphi(x)=(x-1)^{2} for the Frobenius norm test (denoted as LSS Frob.) and φ(x)=log(x)𝜑𝑥𝑥\varphi(x)=\log(x) for the logdet test (denoted as LSS logdet),

  • a test statistic based on the largest off-diagonal entry of the SCM:

    Tϵ′′=𝟙[ϵ,+)(maxν𝒱Nmaxi,j=1,,Mi<j|[𝐂^𝐲(ν)]i,j|)subscriptsuperscript𝑇′′italic-ϵsubscriptdouble-struck-𝟙italic-ϵsubscript𝜈subscript𝒱𝑁subscriptformulae-sequence𝑖𝑗1𝑀𝑖𝑗subscriptdelimited-[]subscript^𝐂𝐲𝜈𝑖𝑗\displaystyle T^{{}^{\prime\prime}}_{\epsilon}=\mathbb{1}_{[\epsilon,+\infty)}\left(\max_{\nu\in\mathcal{V}_{N}}\max_{\begin{subarray}{c}i,j=1,\ldots,M\\ i<j\end{subarray}}\left|[\hat{\mathbf{C}}_{\mathbf{y}}(\nu)]_{i,j}\right|\right)

    denoted as MCC (for Maximum of Cross Coherence),

and where ϵ>0italic-ϵ0\epsilon>0 is some threshold. In Table I, we provide, via Monte-Carlo simulations (100001000010000 draws), the power of each of the four tests, calibrated so that the empirical type I error is equal to 0.050.050.05. The results are provided for various values of (N,M,B)𝑁𝑀𝐵(N,M,B) chosen so that M{20,40,,180}𝑀2040180M\in\{20,40,\ldots,180\}, N=M2𝑁superscript𝑀2N=M^{2} and B=2M𝐵2𝑀B=2M. We set the SNR in the frequency domain as maxν𝒱Nγ(ν)=2MB=1.41subscript𝜈subscript𝒱𝑁𝛾𝜈2𝑀𝐵1.41\max_{\nu\in\mathcal{V}_{N}}\gamma(\nu)=2\sqrt{\frac{M}{B}}=1.41.

The LE test presents the best detection performance among the four candidates, whereas the MCC test does not seem to be adapted to the detection of this alternative. While it is proved in Corollary 3 that the test statistics based on the LSS of 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu) can not asymptotically distinguish between 0subscript0\mathcal{H}_{0} and 1subscript1\mathcal{H}_{1}, they remain sensible to a large variation of a single eigenvalue for finite values of M𝑀M. Consider for instance the Frobenius LSS test, where the test statistic is based on :

maxν𝒱N|1Mm=1M(λm(𝐂^(ν))1)2(λ1)2f(λ)𝑑λ|subscript𝜈subscript𝒱𝑁1𝑀superscriptsubscript𝑚1𝑀superscriptsubscript𝜆𝑚^𝐂𝜈12superscript𝜆12𝑓𝜆differential-d𝜆\max_{\nu\in\mathcal{V}_{N}}\left|\frac{1}{M}\sum_{m=1}^{M}(\lambda_{m}(\hat{\mathbf{C}}(\nu))-1)^{2}-\int(\lambda-1)^{2}f(\lambda)d\lambda\right|

where an explicit computation shows that (λ1)2f(λ)𝑑λ=csuperscript𝜆12𝑓𝜆differential-d𝜆𝑐\int(\lambda-1)^{2}f(\lambda)d\lambda=c. An 𝒪(1)𝒪1\mathcal{O}(1) variation of λ1(𝐂^(ν))subscript𝜆1^𝐂𝜈\lambda_{1}(\hat{\mathbf{C}}(\nu)), the largest eigenvalue of 𝐂^(ν)^𝐂𝜈\hat{\mathbf{C}}(\nu), will lead to a variation of order 𝒪(1M)𝒪1𝑀\mathcal{O}(\frac{1}{M}) of the above term. Therefore, the power of a LSS based test asymptotically converge towards zero, while having non-zero power for finite values of M𝑀M as it is visible on the results of Table I.

TABLE I: Power comparison, K=1, γ(νN)=212𝛾superscriptsubscript𝜈𝑁212\gamma(\nu_{N}^{*})=2\sqrt{\frac{1}{2}}, type I error = 5%
LSS Frob. LSS logdet MCC LE
N M B
400 20 40 0.09 0.07 0.06 0.15
1600 40 80 0.15 0.08 0.06 0.37
3600 60 120 0.19 0.08 0.06 0.68
6400 80 160 0.25 0.08 0.06 0.87
10000 100 200 0.26 0.07 0.06 0.96
14400 120 240 0.25 0.06 0.06 0.99
19600 140 280 0.28 0.06 0.06 1.00
25600 160 320 0.30 0.06 0.06 1.00
32400 180 360 0.31 0.06 0.06 1.00

VI-B Case K>1𝐾1K>1

We eventually consider a model which have the flexibility to consider a signal with an arbitrary value of K1𝐾1K\geq 1. We assume that matrices (𝐇l)l0subscriptsubscript𝐇𝑙𝑙0(\mathbf{H}_{l})_{l\geq 0} verify 𝐇l=0subscript𝐇𝑙0\mathbf{H}_{l}=0 if l>L𝑙𝐿l>L for a certain integer L𝐿L, and that the sequence of M×K𝑀𝐾M\times K matrices (𝐇l)0lLsubscriptsubscript𝐇𝑙0𝑙𝐿(\mathbf{H}_{l})_{0\leq l\leq L} is defined by:

𝐇l=(C1𝐰l,1,,CK𝐰l,K)subscript𝐇𝑙subscript𝐶1subscript𝐰𝑙1subscript𝐶𝐾subscript𝐰𝑙𝐾\mathbf{H}_{l}=(C_{1}\mathbf{w}_{l,1},\ldots,C_{K}\mathbf{w}_{l,K})

where the vectors ((𝐰l,k)l=0,,L)k=1,,Ksubscriptsubscriptsubscript𝐰𝑙𝑘𝑙0𝐿𝑘1𝐾((\mathbf{w}_{l,k})_{l=0,\ldots,L})_{k=1,\ldots,K} are generated as independent realisations of M𝑀M–dimensional vectors uniformly distributed on the unit sphere of Msuperscript𝑀\mathbb{C}^{M} and where the C1C2CKsubscript𝐶1subscript𝐶2subscript𝐶𝐾C_{1}\geq C_{2}\geq\ldots\geq C_{K} are positive constants used to tune the SNR of each of the K𝐾K sources at the desired level. Moreover, as the K𝐾K columns of each matrix 𝐇lsubscript𝐇𝑙\mathbf{H}_{l} coincide with the realisations of mutually independent random vectors, the columns of 𝐇(ν)𝐇𝜈\mathbf{H}(\nu) are easily seen to be nearly orthogonal and to nearly share the same norm for each ν𝜈\nu if M𝑀M is large enough. More precisely, for each ν𝜈\nu, it holds that 𝐇(ν)𝐇(ν)(L+1)Diag(C1,,CK)𝐇superscript𝜈𝐇𝜈𝐿1Diagsubscript𝐶1subscript𝐶𝐾\mathbf{H}(\nu)^{*}\mathbf{H}(\nu)\rightarrow(L+1)\,\mathrm{Diag}(C_{1},\ldots,C_{K}) when M+𝑀M\rightarrow+\infty. As the spectral densities of the components of the noise all coincide with s(ν)=1|1θe2iπν|2𝑠𝜈1superscript1𝜃superscript𝑒2𝑖𝜋𝜈2s(\nu)=\frac{1}{|1-\theta e^{-2i\pi\nu}|^{2}}, the non-zero eigenvalues of 𝐒𝐯(ν)12𝐇(ν)𝐇(ν)𝐒𝐯(ν)12subscript𝐒𝐯superscript𝜈12𝐇𝜈𝐇superscript𝜈subscript𝐒𝐯superscript𝜈12\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}} converge towards the ((L+1)Ck/s(ν))k=1,,Ksubscript𝐿1subscript𝐶𝑘𝑠𝜈𝑘1𝐾\left((L+1)C_{k}/s(\nu)\right)_{k=1,\ldots,K} when M𝑀M increases. Therefore, the signal obtained by this model satisfies Assumption 3. Rather than just providing the performance of the test Tϵsubscript𝑇italic-ϵT_{\epsilon} based on the maximum of the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) proposed in this paper, we compare in the following Tϵsubscript𝑇italic-ϵT_{\epsilon} with TK,ϵsubscript𝑇𝐾italic-ϵT_{K,\epsilon} defined by

TK,ϵ=𝟙[Kλ++ϵ,)(maxν𝒱Nk=1Kλk(𝐂^𝐲(ν)))subscript𝑇𝐾italic-ϵsubscriptdouble-struck-𝟙𝐾superscript𝜆italic-ϵsubscript𝜈subscript𝒱𝑁superscriptsubscript𝑘1𝐾subscript𝜆𝑘subscript^𝐂𝐲𝜈T_{K,\epsilon}=\mathbb{1}_{[K\lambda^{+}+\epsilon,\infty)}\left(\max_{\nu\in\mathcal{V}_{N}}\sum_{k=1}^{K}\lambda_{k}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\right)

which depends on the K𝐾K largest eigenvalues of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) rather than on the largest one. It is easy to generalize Proposition 2 and Proposition 3 in order to study the asymptotic properties of TK,ϵsubscript𝑇𝐾italic-ϵT_{K,\epsilon}. More precisely, for each k=1,,K𝑘1𝐾k=1,\ldots,K, we define γk(ν)subscript𝛾𝑘𝜈\gamma_{k}(\nu) by

γk(ν)=λk(𝐒𝐯(ν)12𝐇(ν)𝐇(ν)𝐒𝐯(ν)12)subscript𝛾𝑘𝜈subscript𝜆𝑘subscript𝐒𝐯superscript𝜈12𝐇𝜈𝐇superscript𝜈subscript𝐒𝐯superscript𝜈12\displaystyle\gamma_{k}(\nu)=\lambda_{k}\left(\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\right) (29)

and denote νK,Nsuperscriptsubscript𝜈𝐾𝑁\nu_{K,N}^{*} one of the frequency such that maxν𝒱Nk=1Kγk(ν)=k=1Kγk(νK,N)subscript𝜈subscript𝒱𝑁superscriptsubscript𝑘1𝐾subscript𝛾𝑘𝜈superscriptsubscript𝑘1𝐾subscript𝛾𝑘superscriptsubscript𝜈𝐾𝑁\max_{\nu\in\mathcal{V}_{N}}\sum_{k=1}^{K}\gamma_{k}(\nu)=\sum_{k=1}^{K}\gamma_{k}(\nu_{K,N}^{*}). γk(ν)subscript𝛾𝑘𝜈\gamma_{k}(\nu) can of course be seen as a generalization of γ(ν)𝛾𝜈\gamma(\nu) defined by (24). Then, under the extra assumption that for k=1,,K𝑘1𝐾k=1,\ldots,K, γk(νK,N)subscript𝛾𝑘superscriptsubscript𝜈𝐾𝑁\gamma_{k}(\nu_{K,N}^{*}) converges towards a finite limit γk,subscript𝛾𝑘\gamma_{k,\infty} (a condition which holds in the context of the present experiment because it is easily seen that γk(νK,N)(L+1)(1+θ)2Cksubscript𝛾𝑘superscriptsubscript𝜈𝐾𝑁𝐿1superscript1𝜃2subscript𝐶𝑘\gamma_{k}(\nu_{K,N}^{*})\rightarrow(L+1)\,(1+\theta)^{2}C_{k}), λk(𝐂^𝐲(νK,N))subscript𝜆𝑘subscript^𝐂𝐲superscriptsubscript𝜈𝐾𝑁\lambda_{k}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu_{K,N}^{*})\right) converges towards λ+superscript𝜆\lambda^{+} if γk,csubscript𝛾𝑘𝑐\gamma_{k,\infty}\leq\sqrt{c} and towards ϕ(γk,)>λ+italic-ϕsubscript𝛾𝑘superscript𝜆\phi(\gamma_{k,\infty})>\lambda^{+} if γk,>csubscript𝛾𝑘𝑐\gamma_{k,\infty}>\sqrt{c}. It is easy to check that if γ=γ1,>csubscript𝛾subscript𝛾1𝑐\gamma_{\infty}=\gamma_{1,\infty}>\sqrt{c}, then the statistics TK,ϵsubscript𝑇𝐾italic-ϵT_{K,\epsilon} also leads to a consistent test provided 0<ϵ<ϕ(γ)λ+0italic-ϵitalic-ϕsubscript𝛾subscript𝜆0<\epsilon<\phi(\gamma_{\infty})-\lambda_{+}. While in practice the number of sources K𝐾K is unknown, it is interesting to evaluate the performance provided by TK,ϵsubscript𝑇𝐾italic-ϵT_{K,\epsilon} which can be considered as an ideal reference. Intuitively, TK,ϵsubscript𝑇𝐾italic-ϵT_{K,\epsilon} could lead to a better performance than Tϵsubscript𝑇italic-ϵT_{\epsilon} when γk,>csubscript𝛾𝑘𝑐\gamma_{k,\infty}>\sqrt{c} for k=1,,K𝑘1𝐾k=1,\ldots,K, because, in this context, if ν^K,Nsuperscriptsubscript^𝜈𝐾𝑁\hat{\nu}_{K,N}^{*} is a frequency that maximises k=1K λk(𝐂^𝐲(ν))superscriptsubscript𝑘1𝐾 subscript𝜆𝑘subscript^𝐂𝐲𝜈\sum_{k=1}^{K} \lambda_{k}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right), then lim infM+λk(𝐂^𝐲(νK,N))>λ+subscriptlimit-infimum𝑀subscript𝜆𝑘subscript^𝐂𝐲superscriptsubscript𝜈𝐾𝑁superscript𝜆\liminf_{M\rightarrow+\infty}\lambda_{k}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu_{K,N}^{*})\right)>\lambda^{+}. Therefore, the K𝐾K largest eigenvalues of 𝐂^𝐲(ν^K,N)subscript^𝐂𝐲superscriptsubscript^𝜈𝐾𝑁\hat{\mathbf{C}}_{\mathbf{y}}(\hat{\nu}_{K,N}^{*}) bring useful information to the detection of the useful signal.

In order to evaluate numerically the compared performance of Tϵsubscript𝑇italic-ϵT_{\epsilon} and TK,ϵsubscript𝑇𝐾italic-ϵT_{K,\epsilon} when K𝐾K is known, we first consider the case K=2𝐾2K=2, L=3𝐿3L=3, and where (γ1+γ2)(ν2,N)=3csubscript𝛾1subscript𝛾2superscriptsubscript𝜈2𝑁3𝑐(\gamma_{1}+\gamma_{2})(\nu_{2,N}^{*})=3\sqrt{c}. Concerning the value of (C1,C2)subscript𝐶1subscript𝐶2(C_{1},C_{2}), we consider the two following cases: C1C2=1subscript𝐶1subscript𝐶21\frac{C_{1}}{C_{2}}=1 and C1C2=4subscript𝐶1subscript𝐶24\frac{C_{1}}{C_{2}}=4. This corresponds respectively to the case where both sources contributes exactly the same on each sensor, and where the first source contributes much more than the second one. Tables II, III report the power of the proposed test (LE(1) represents Tϵsubscript𝑇italic-ϵT_{\epsilon} and LE(2) represents T2,ϵsubscript𝑇2italic-ϵT_{2,\epsilon}) against the LSS tests and the MCC test, with a type I error fixed at 5%. When C1C2=4subscript𝐶1subscript𝐶24\frac{C_{1}}{C_{2}}=4, it can be expected that the most powerful source is dominant, and that γ2(ν2,N)<csubscript𝛾2superscriptsubscript𝜈2𝑁𝑐\gamma_{2}(\nu_{2,N}^{*})<\sqrt{c}. Therefore, λ2(𝐂^𝐲(ν))subscript𝜆2subscript^𝐂𝐲𝜈\lambda_{2}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right) is likely to stay close to λ+superscript𝜆\lambda^{+} for each ν𝜈\nu, so that the use of T2,ϵsubscript𝑇2italic-ϵT_{2,\epsilon} should not bring any extra performance. This intuition is confirmed by Table II. When C1C2=1subscript𝐶1subscript𝐶21\frac{C_{1}}{C_{2}}=1, γ1(ν2,N)subscript𝛾1superscriptsubscript𝜈2𝑁\gamma_{1}(\nu_{2,N}^{*}) and γ2(ν2,N)subscript𝛾2superscriptsubscript𝜈2𝑁\gamma_{2}(\nu_{2,N}^{*}) should be both close to 32c32𝑐\frac{3}{2}\sqrt{c}, thus suggesting that the two largest eigenvalues of 𝐂^𝐲subscript^𝐂𝐲\hat{\mathbf{C}}_{\mathbf{y}} at the maximizing frequency ν^2,Nsuperscriptsubscript^𝜈2𝑁\hat{\nu}_{2,N}^{*} should also nearly coincide, and should escape from [λ,λ+]superscript𝜆superscript𝜆[\lambda^{-},\lambda^{+}]. While the second eigenvalue brings here some information, Table III tends to indicate that Tϵsubscript𝑇italic-ϵT_{\epsilon} has better performance than T2,ϵsubscript𝑇2italic-ϵT_{2,\epsilon}. In the next experiment, (γ1+γ2)(ν2,N)=2csubscript𝛾1subscript𝛾2superscriptsubscript𝜈2𝑁2𝑐(\gamma_{1}+\gamma_{2})(\nu_{2,N}^{*})=2\sqrt{c}. For C1C2=4subscript𝐶1subscript𝐶24\frac{C_{1}}{C_{2}}=4, the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) is likely to be still dominant for each ν𝜈\nu, and Table IV confirms the better performance of Tϵsubscript𝑇italic-ϵT_{\epsilon}. When C1C2=1subscript𝐶1subscript𝐶21\frac{C_{1}}{C_{2}}=1, γ1(ν2,N)subscript𝛾1superscriptsubscript𝜈2𝑁\gamma_{1}(\nu_{2,N}^{*}) and γ2(ν2,N)subscript𝛾2superscriptsubscript𝜈2𝑁\gamma_{2}(\nu_{2,N}^{*}) should be both close to the detectability threshold c𝑐\sqrt{c}, and Table V this time shows that the use of T2,ϵsubscript𝑇2italic-ϵT_{2,\epsilon} leads to some improvement. For comparison, we also report the results of Tϵsubscript𝑇italic-ϵT_{\epsilon} for C2=0subscript𝐶20C_{2}=0 in Table VI.

TABLE II: Power comparison, C1C2=4subscript𝐶1subscript𝐶24\frac{C_{1}}{C_{2}}=4, (γ1+γ2)(ν2,N)=312subscript𝛾1subscript𝛾2superscriptsubscript𝜈2𝑁312(\gamma_{1}+\gamma_{2})(\nu_{2,N}^{*})=3\sqrt{\frac{1}{2}}, type I error = 5%
LSS Fr. LSS ld MCC LE(1) LE(2)
N M B
100 10 20 0.31 0.18 0.16 0.42 0.37
400 20 40 0.79 0.39 0.45 0.94 0.89
900 30 60 0.94 0.49 0.53 1.00 0.99
1600 40 80 0.98 0.50 0.55 1.00 1.00
2500 50 100 0.99 0.52 0.55 1.00 1.00
3600 60 120 1.00 0.51 0.43 1.00 1.00
4900 70 140 1.00 0.55 0.37 1.00 1.00
6400 80 160 1.00 0.54 0.28 1.00 1.00
TABLE III: Power comparison, C1C2=1subscript𝐶1subscript𝐶21\frac{C_{1}}{C_{2}}=1, (γ1+γ2)(ν2,N)=312subscript𝛾1subscript𝛾2superscriptsubscript𝜈2𝑁312(\gamma_{1}+\gamma_{2})(\nu_{2,N}^{*})=3\sqrt{\frac{1}{2}}, type I error = 5%
LSS Fr. LSS ld MCC LE(1) LE(2)
N M B
100 10 20 0.38 0.22 0.16 0.48 0.46
400 20 40 0.58 0.30 0.30 0.75 0.73
900 30 60 0.67 0.30 0.28 0.91 0.89
1600 40 80 0.74 0.29 0.18 0.96 0.97
2500 50 100 0.79 0.30 0.16 0.99 0.99
3600 60 120 0.79 0.24 0.13 1.00 1.00
4900 70 140 0.85 0.28 0.12 1.00 1.00
TABLE IV: Power comparison, C1C2=4subscript𝐶1subscript𝐶24\frac{C_{1}}{C_{2}}=4, (γ1+γ2)(ν2,N)=212subscript𝛾1subscript𝛾2superscriptsubscript𝜈2𝑁212(\gamma_{1}+\gamma_{2})(\nu_{2,N}^{*})=2\sqrt{\frac{1}{2}}, type I error = 5%
LSS Fr. LSS ld MCC LE(1) LE(2)
N M B
100 10 20 0.15 0.10 0.10 0.21 0.20
400 20 40 0.33 0.15 0.12 0.55 0.50
900 30 60 0.39 0.15 0.17 0.75 0.71
1600 40 80 0.52 0.16 0.14 0.94 0.90
2500 50 100 0.54 0.15 0.14 0.98 0.97
3600 60 120 0.56 0.13 0.13 1.00 0.99
4900 70 140 0.55 0.13 0.10 1.00 1.00
6400 80 160 0.62 0.11 0.10 1.00 1.00
TABLE V: Power comparison, C1C2=1subscript𝐶1subscript𝐶21\frac{C_{1}}{C_{2}}=1, (γ1+γ2)(ν2,N)=212subscript𝛾1subscript𝛾2superscriptsubscript𝜈2𝑁212(\gamma_{1}+\gamma_{2})(\nu_{2,N}^{*})=2\sqrt{\frac{1}{2}}, type I error = 5%
LSS Fr. LSS ld MCC LE(1) LE(2)
N M B
400 20 40 0.17 0.11 0.08 0.27 0.27
1600 40 80 0.18 0.10 0.08 0.45 0.48
3600 60 120 0.15 0.07 0.07 0.58 0.62
6400 80 160 0.16 0.07 0.08 0.69 0.75
10000 100 200 0.13 0.05 0.07 0.76 0.83
14400 120 240 0.10 0.03 0.07 0.82 0.86
19600 140 280 0.09 0.04 0.07 0.86 0.89
25600 160 320 0.10 0.03 0.06 0.89 0.93
32400 180 360 0.09 0.03 0.06 0.87 0.93
TABLE VI: Power comparison, C2=0subscript𝐶20C_{2}=0, γ(νN)=212𝛾superscriptsubscript𝜈𝑁212\gamma(\nu_{N}^{*})=2\sqrt{\frac{1}{2}}, type I error = 5%
LSS Fr. LSS ld MCC LE(1) LE(2)
N M B
100 10 20 0.19 0.12 0.12 0.26 0.22
400 20 40 0.43 0.19 0.14 0.66 0.59
900 30 60 0.51 0.19 0.19 0.88 0.83
1600 40 80 0.62 0.20 0.15 0.97 0.95
2500 50 100 0.65 0.18 0.17 0.99 0.99
3600 60 120 0.68 0.16 0.12 1.00 1.00
4900 70 140 0.71 0.16 0.13 1.00 1.00
6400 80 160 0.75 0.17 0.12 1.00 1.00

This discussion tends to indicate that, even when K>1𝐾1K>1 is assumed known, the use of the maximum over 𝒱Nsubscript𝒱𝑁\mathcal{V}_{N} of the largest eigenvalue of 𝐂^𝐲(ν)subscript^𝐂𝐲𝜈\hat{\mathbf{C}}_{\mathbf{y}}(\nu) does not introduce any significant loss of performance.

VII Conclusion

In this paper, we have studied the statistical behaviour of certain frequency-domain detection test statistics, based on the eigenvalues of a sample estimate of the SCM, in the high-dimensional regime in which both the dimension M𝑀M of the underlying signals and the number of samples N𝑁N converge to infinity at certain rates. In particular, we have proved various approximation results showing that the sample SCM asymptotically behaves as a Wishart matrix. These results have been exploited to prove that test statistics based on LSS of the sample SCM are not consistent in the high-dimensional regime. A new test statistic relying on the largest eigenvalue of the sample SCM has also been proposed and proved to be consistent in the high-dimensional regime. Finally, numerical results have demonstrated that this new test statistic provides reasonable performance and outperforms other standard test statistics in situations where the dimension M𝑀M and the number of samples N𝑁N are large.

References

  • [1] A. Rosuel, P. Vallet, P. Loubaton, and X. Mestre, “On the frequency domain detection of high dimensional time series,” in IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP).   IEEE, 2020, pp. 8782–8786.
  • [2] D. Ramírez, G. Vazquez-Vilar, R. López-Valcarce, J. Vía, and I. Santamaría, “Detection of rank-p signals in cognitive radio networks with uncalibrated multiple antennas,” IEEE Trans. Signal Process., vol. 59, no. 8, pp. 3764–3774, 2011.
  • [3] J. Sala-Alvarez, G. Vázquez-Vilar, R. López-Valcarce, S. Sedighi, and A. Taherpour, “Multiantenna glr detection of rank-one signals with known power spectral shape under spatially uncorrelated noise,” IEEE Transactions on Signal Processing, vol. 64, no. 23, pp. 6269–6283, 2016.
  • [4] A.-J. Boonstra and A.-J. Van der Veen, “Gain calibration methods for radio telescope arrays,” IEEE Trans. Signal Process., vol. 51, no. 1, pp. 25–38, 2003.
  • [5] A. Leshem and A.-J. Van der Veen, “Multichannel detection of gaussian signals with uncalibrated receivers,” IEEE Signal Process. Lett., vol. 8, no. 4, pp. 120–122, 2001.
  • [6] L. D. Haugh, “Checking the independence of two covariance-stationary time series: a univariate residual cross-correlation approach,” Journal of the American Statistical Association, vol. 71, no. 354, pp. 378–385, 1976.
  • [7] Y. Hong, “Testing for independence between two covariance stationary time series,” Biometrika, vol. 83, no. 3, pp. 615–625, 1996.
  • [8] W. Li and Y. Hui, “Robust residual cross correlation tests for lagged relations in time series,” Journal of Statistical Computation and Simulation, vol. 49, no. 1-2, pp. 103–109, 1994.
  • [9] K. El Himdi, R. Roy, and P. Duchesne, “Tests for non-norrelation of two multivariate time series: A nonparametric approach,” Lecture Notes-Monograph Series, vol. 42, pp. 397–416, 2003.
  • [10] D. Ramirez, J. Via, I. Santamaria, and L. Scharf, “Detection of spatially correlated gaussian time series,” IEEE Trans. Signal Process., vol. 58, no. 10, pp. 5006–5015, 2010.
  • [11] N. Klausner, M. Azimi-Sadjadi, and L. Scharf, “Detection of spatially correlated time series from a network of sensor arrays,” IEEE Trans. Signal Process., vol. 62, no. 6, pp. 1396–1407, 2014.
  • [12] D. Brillinger, Time series: data analysis and theory.   Classics in Applied Mathematics, Siam, 2001, vol. 36.
  • [13] L. Koopmans, The spectral analysis of time series.   Probability and Mathematical Statistics, Academic Press, 1995, vol. 22.
  • [14] G. Wahba, “Some tests of independence for stationary multivariate time series,” Journal of the Royal Statistical Society: Series B (Methodological), vol. 33, no. 1, pp. 153–166, 1971.
  • [15] M. Taniguchi, M. L. Puri, and M. Kondo, “Nonparametric approach for non-gaussian vector stationary processes,” Journal of Multivariate Analysis, vol. 56, no. 2, pp. 259–283, 1996.
  • [16] M. Eichler, “A frequency-domain based test for non-correlation between stationary time series,” Metrika, vol. 65, no. 2, pp. 133–157, 2007.
  • [17] ——, “Testing nonparametric and semiparametric hypotheses in vector stationary processes,” Journal of Multivariate Analysis, vol. 99, no. 5, pp. 968–1009, 2008.
  • [18] P. Loubaton and A. Rosuel, “Large random matrix approach for testing independence of a large number of gaussian time series,” arXiv preprint arXiv:2007.08806, 2020.
  • [19] V. Marcenko and L. Pastur, “Distribution of eigenvalues for some sets of random matrices,” Mathematics of the USSR-Sbornik, vol. 1, p. 457, 1967.
  • [20] J. Baik and J. Silverstein, “Eigenvalues of large sample covariance matrices of spiked population models,” J. Multivariate Anal., vol. 97, no. 6, pp. 1382–1408, 2006.
  • [21] D. Morales-Jimenez, I. Johnstone, M. MacKay, and J. Yang, “Asymptotics of eigenstructure of sample correlation matrices for high-dimensional spiked models,” To appear in Statistica Sinica, 2019, preprint arXiv:1810.10214v3.
  • [22] M. Forni, M. Hallin, M. Lippi, and L. Reichlin, “The generalized dynamic-factor model: Identification and estimation,” Rev. Econ. Stat., vol. 82, no. 4, pp. 540–554, 2000.
  • [23] ——, “The generalized dynamic factor model consistency and rates,” J. Econom., vol. 119, no. 2, pp. 231–255, 2004.
  • [24] P. Vallet, P. Loubaton, and X. Mestre, “Improved Subspace Estimation for Multivariate Observations of High Dimension: The Deterministic Signal Case,” IEEE Trans. Inf. Theory, vol. 58, no. 2, 2012.
  • [25] P. Vallet, X. Mestre, and P. Loubaton, “Performance Analysis of an Improved MUSIC DoA Estimator,” IEEE Trans. Signal Process., vol. 63, no. 23, pp. 6407–6422, Dec. 2015.
  • [26] R. Horn and C. Johnson, Matrix analysis.   Cambridge university press, 2005.
  • [27] M. Wax and T. Kailath, “Detection of signals by information theoretic criteria,” IEEE Transactions on acoustics, speech, and signal processing, vol. 33, no. 2, pp. 387–392, 1985.
  • [28] J. Baik, G. B. Arous, S. Péché et al., “Phase transition of the largest eigenvalue for nonnull complex sample covariance matrices,” Annals of Probability, vol. 33, no. 5, pp. 1643–1697, 2005.
  • [29] F. Benaych-Georges and R. R. Nadakuditi, “The eigenvalues and eigenvectors of finite, low rank perturbations of large random matrices,” Advances in Mathematics, vol. 227, no. 1, pp. 494–521, 2011.
  • [30] M. Rudelson and R. Vershynin, “Hanson-wright inequality and sub-gaussian concentration,” Electron. Commun. Probab., vol. 18, 2013.
  • [31] U. Haagerup and S. Thorbjørnsen, “Random matrices with complex gaussian entries,” Expo. Math., vol. 21, no. 4, pp. 293–337, 2003.
  • [32] T. Tao, Topics in random matrix theory.   American Mathematical Soc., 2012, vol. 132.
  • [33] A. Guionnet and O. Zeitouni, “Concentration of the spectral measure for large matrices,” Electron. Commun. Probab., vol. 5, pp. 119–136, 2000. [Online]. Available: https://doi.org/10.1214/ECP.v5-1026
  • [34] W. Hachem, O. Khorunzhiy, P. Loubaton, J. Najim, and L. Pastur, “A new approach for capacity analysis of large dimensional multi-antenna channels,” IEEE Trans. Inf. Theory, vol. 54, no. 9, pp. 3987–4004, 2008.

Appendix A Useful results

In this section, we recall some useful results which will be constantly used in the proofs developed in the following sections.

The first result is based on a Chernoff bound for the χ2superscript𝜒2\chi^{2} distribution, and is also a special case of the well-known Hanson-Wright inequality describing the concentration of sub-Gaussian quadratic forms around their means (see [30]).

Lemma 1.

Let 𝐳𝒩n(𝟎,𝐈n)similar-to𝐳subscript𝒩superscript𝑛0subscript𝐈𝑛\mathbf{z}\sim\mathcal{N}_{\mathbb{C}^{n}}(\mathbf{0},\mathbf{I}_{n}) and 𝚵𝚵\boldsymbol{\Xi} a deterministic n×n𝑛𝑛n\times n complex matrix. Then there exists a constant κ>0𝜅0\kappa>0 independent of n𝑛n and 𝚵𝚵\boldsymbol{\Xi} such that for all t0𝑡0t\geq 0,

(|𝐳𝚵𝐳𝔼[𝐳𝚵𝐳]|>t)2exp(κmin{t2𝚵F2,t𝚵2}).superscript𝐳𝚵𝐳𝔼delimited-[]superscript𝐳𝚵𝐳𝑡2exp𝜅superscript𝑡2superscriptsubscriptnorm𝚵𝐹2𝑡subscriptnorm𝚵2\mathbb{P}\left(\left|\mathbf{z}^{*}\boldsymbol{\Xi}\mathbf{z}-\mathbb{E}[\mathbf{z}^{*}\boldsymbol{\Xi}\mathbf{z}]\right|>t\right)\leq 2\ \mathrm{exp}\left(-\kappa\min\left\{\frac{t^{2}}{\|\boldsymbol{\Xi}\|_{F}^{2}},\frac{t}{\|\boldsymbol{\Xi}\|_{2}}\right\}\right).

The second following result describes the behaviour of the largest and smallest eigenvalues of a standard Wishart matrix.

Lemma 2 ([31, Proof of Lemma 7.3]).

Let 𝐙𝐙\mathbf{Z} be a M×(B+1)𝑀𝐵1M\times(B+1) matrix with i.i.d. 𝒩(0,1)subscript𝒩01\mathcal{N}_{\mathbb{C}}\left(0,1\right) entries. Then under Assumption 4, there exists a constant C>0𝐶0C>0 independent of M,B𝑀𝐵M,B such that for all t>0𝑡0t>0,

(λ1(𝐙𝐙B+1)>(1+MB+1)2+t)(B+1)exp(C(B+1)t2)subscript𝜆1superscript𝐙𝐙𝐵1superscript1𝑀𝐵12𝑡𝐵1exp𝐶𝐵1superscript𝑡2\mathbb{P}\left(\lambda_{1}\left(\frac{\mathbf{Z}\mathbf{Z}^{*}}{B+1}\right)>\left(1+\sqrt{\frac{M}{B+1}}\right)^{2}+t\right)\leq(B+1)\mathrm{exp}\left(-C(B+1)t^{2}\right)

and

(λM(𝐙𝐙B+1)<(1MB+1)2t)(B+1)exp(C(B+1)t2).subscript𝜆𝑀superscript𝐙𝐙𝐵1superscript1𝑀𝐵12𝑡𝐵1exp𝐶𝐵1superscript𝑡2\mathbb{P}\left(\lambda_{M}\left(\frac{\mathbf{Z}\mathbf{Z}^{*}}{B+1}\right)<\left(1-\sqrt{\frac{M}{B+1}}\right)^{2}-t\right)\leq(B+1)\mathrm{exp}\left(-C(B+1)t^{2}\right).

We will mainly use Lemma 2 as follows; let (𝐙(ν))ν𝒱Nsubscript𝐙𝜈𝜈subscript𝒱𝑁(\mathbf{Z}(\nu))_{\nu\in\mathcal{V}_{N}} be a family of M×(B+1)𝑀𝐵1M\times(B+1) random matrices such that 𝐙(ν)𝐙𝜈\mathbf{Z}(\nu) has i.i.d. 𝒩(0,1)subscript𝒩01\mathcal{N}_{\mathbb{C}}(0,1) (recall the definition of the index set 𝒱Nsubscript𝒱𝑁\mathcal{V}_{N} in (16)), then from the union bound

(maxν𝒱Nλ1(𝐙(ν)𝐙(ν)B+1)>(1+MB+1)2+t)subscript𝜈subscript𝒱𝑁subscript𝜆1𝐙𝜈𝐙superscript𝜈𝐵1superscript1𝑀𝐵12𝑡\displaystyle\mathbb{P}\left(\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\frac{\mathbf{Z}(\nu)\mathbf{Z}(\nu)^{*}}{B+1}\right)>\left(1+\sqrt{\frac{M}{B+1}}\right)^{2}+t\right)
ν𝒱N(λ1(𝐙(ν)𝐙(ν)B+1)>(1+MB+1)2+t)absentsubscript𝜈subscript𝒱𝑁subscript𝜆1𝐙𝜈𝐙superscript𝜈𝐵1superscript1𝑀𝐵12𝑡\displaystyle\leq\sum_{\nu\in\mathcal{V}_{N}}\mathbb{P}\left(\lambda_{1}\left(\frac{\mathbf{Z}(\nu)\mathbf{Z}(\nu)^{*}}{B+1}\right)>\left(1+\sqrt{\frac{M}{B+1}}\right)^{2}+t\right)
exp(C(B+1)t2+log(N(B+1))).absentexp𝐶𝐵1superscript𝑡2𝑁𝐵1\displaystyle\leq\mathrm{exp}\Bigl{(}-C(B+1)t^{2}+\log(N(B+1))\Bigr{)}.

Using Assumption 4 and Borel-Cantelli lemma, we deduce that

lim supMmaxν𝒱Nλ1(𝐙(ν)𝐙(ν)B+1)(1+c)2subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆1𝐙𝜈𝐙superscript𝜈𝐵1superscript1𝑐2\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\frac{\mathbf{Z}(\nu)\mathbf{Z}(\nu)^{*}}{B+1}\right)\leq(1+\sqrt{c})^{2}

with probability one.

Appendix B Proof of Theorem 2

B-A Reduction to K=1𝐾1K=1

First, note that we may assume K=1𝐾1K=1 without loss of generality. Indeed, consider the decomposition

𝐮n==1K𝐮n()subscript𝐮𝑛superscriptsubscript1𝐾superscriptsubscript𝐮𝑛\displaystyle\mathbf{u}_{n}=\sum_{\ell=1}^{K}\mathbf{u}_{n}^{(\ell)}

where 𝐮n()=k=0+𝐡,kϵnk()superscriptsubscript𝐮𝑛superscriptsubscript𝑘0subscript𝐡𝑘superscriptsubscriptitalic-ϵ𝑛𝑘\mathbf{u}_{n}^{(\ell)}=\sum_{k=0}^{+\infty}\mathbf{h}_{\ell,k}\epsilon_{n-k}^{(\ell)} and where 𝐡,ksubscript𝐡𝑘\mathbf{h}_{\ell,k} and ϵn()superscriptsubscriptitalic-ϵ𝑛\epsilon_{n}^{(\ell)} are the \ell-th column of 𝐇ksubscript𝐇𝑘\mathbf{H}_{k} and the \ell-th entry of ϵnsubscriptbold-italic-ϵ𝑛\boldsymbol{\epsilon}_{n} respectively. Moreover, Assumption 3 implies that

supM1k(1+|k|)𝐡,k2<subscriptsupremum𝑀1subscript𝑘1𝑘subscriptnormsubscript𝐡𝑘2\displaystyle\sup_{M\geq 1}\sum_{k\in\mathbb{Z}}(1+|k|)\|\mathbf{h}_{\ell,k}\|_{2}<\infty

From the fact that K𝐾K is fixed with respect to N𝑁N (Assumption 4) and

maxν𝒱N𝚺𝐮(ν)𝐇(ν)𝚺ϵ(ν)2=1Kmaxν𝒱N𝚺𝐮()(ν)𝐡(ν)𝚺ϵ()(ν)2subscript𝜈subscript𝒱𝑁subscriptnormsubscript𝚺𝐮𝜈𝐇𝜈subscript𝚺bold-italic-ϵ𝜈2superscriptsubscript1𝐾subscript𝜈subscript𝒱𝑁subscriptnormsubscript𝚺superscript𝐮𝜈subscript𝐡𝜈subscript𝚺superscriptitalic-ϵ𝜈2\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\Sigma}_{\mathbf{u}}\left(\nu\right)-\mathbf{H}\left(\nu\right)\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}\left(\nu\right)\right\|_{2}\leq\sum_{\ell=1}^{K}\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\Sigma}_{\mathbf{u}^{(\ell)}}\left(\nu\right)-\mathbf{h}_{\ell}\left(\nu\right)\boldsymbol{\Sigma}_{\epsilon^{(\ell)}}\left(\nu\right)\right\|_{2}

where 𝚺𝐮()(ν)subscript𝚺superscript𝐮𝜈\boldsymbol{\Sigma}_{\mathbf{u}^{(\ell)}}\left(\nu\right), 𝐡(ν)subscript𝐡𝜈\mathbf{h}_{\ell}\left(\nu\right), 𝚺ϵ()(ν)subscript𝚺superscriptitalic-ϵ𝜈\boldsymbol{\Sigma}_{\epsilon^{(\ell)}}\left(\nu\right) are defined as 𝚺𝐮(ν)subscript𝚺𝐮𝜈\boldsymbol{\Sigma}_{\mathbf{u}}\left(\nu\right), 𝐇(ν)𝐇𝜈\mathbf{H}\left(\nu\right), 𝚺ϵ(ν)subscript𝚺bold-italic-ϵ𝜈\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}\left(\nu\right) respectively, Theorem 2 is proved if we can show that

maxν𝒱N𝚺𝐮()(ν)𝐡(ν)𝚺ϵ()(ν)2Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\Sigma}_{\mathbf{u}^{(\ell)}}\left(\nu\right)-\mathbf{h}_{\ell}\left(\nu\right)\boldsymbol{\Sigma}_{\epsilon^{(\ell)}}\left(\nu\right)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0

for all =1,,K1𝐾\ell=1,\ldots,K. Therefore, we assume for the remainder of the proof that

𝐮n=k𝐡kϵnk,subscript𝐮𝑛subscript𝑘subscript𝐡𝑘subscriptitalic-ϵ𝑛𝑘\displaystyle\mathbf{u}_{n}=\sum_{k\in\mathbb{Z}}\mathbf{h}_{k}\epsilon_{n-k},

where

  • (𝐡k)ksubscriptsubscript𝐡𝑘𝑘(\mathbf{h}_{k})_{k\in\mathbb{Z}} is a filter, with 𝐡kMsubscript𝐡𝑘superscript𝑀\mathbf{h}_{k}\in\mathbb{C}^{M} and such that

    supM1k(1+|k|)𝐡k2<.subscriptsupremum𝑀1subscript𝑘1𝑘subscriptnormsubscript𝐡𝑘2\displaystyle\sup_{M\geq 1}\sum_{k\in\mathbb{Z}}(1+|k|)\left\|\mathbf{h}_{k}\right\|_{2}<\infty. (30)
  • (ϵn)nsubscriptsubscriptitalic-ϵ𝑛𝑛(\epsilon_{n})_{n\in\mathbb{Z}} is a scalar standard complex Gaussian white noise.

B-B Reduction to B=1𝐵1B=1

Let 𝐡(ν)=k𝐡kei2πνk𝐡𝜈subscript𝑘subscript𝐡𝑘superscriptei2𝜋𝜈𝑘\mathbf{h}(\nu)=\sum_{k\in\mathbb{Z}}\mathbf{h}_{k}\mathrm{e}^{-\mathrm{i}2\pi\nu k} and

ξϵ(ν)=1Nn=0N1ϵnei2πνn.subscript𝜉italic-ϵ𝜈1𝑁superscriptsubscript𝑛0𝑁1subscriptitalic-ϵ𝑛superscriptei2𝜋𝜈𝑛\xi_{\epsilon}(\nu)=\frac{1}{\sqrt{N}}\sum_{n=0}^{N-1}\epsilon_{n}\mathrm{e}^{-\mathrm{i}2\pi\nu n}.

From (30) and Assumption 4, a first-order Taylor expansion of b𝐡(ν+bN)maps-to𝑏𝐡𝜈𝑏𝑁b\mapsto\mathbf{h}\left(\nu+\frac{b}{N}\right) at 00 leads to

supν[0,1]maxb{B2,,B2}𝐡(ν)𝐡(ν+bN)2subscriptsupremum𝜈01subscript𝑏𝐵2𝐵2subscriptnorm𝐡𝜈𝐡𝜈𝑏𝑁2\displaystyle\sup_{\nu\in[0,1]}\max_{b\in\{-\frac{B}{2},\ldots,\frac{B}{2}\}}\left\|\mathbf{h}(\nu)-\mathbf{h}\left(\nu+\frac{b}{N}\right)\right\|_{2} =𝒪(BN)=𝒪(1N1α).absent𝒪𝐵𝑁𝒪1superscript𝑁1𝛼\displaystyle=\mathcal{O}\left(\frac{B}{N}\right)=\mathcal{O}\left(\frac{1}{N^{1-\alpha}}\right).

Moreover, from Lemma 1 applied to the random vector

𝐳=(ξϵ(νB2N),,ξϵ(ν+B2N))T𝒩B+1(𝟎,𝐈B+1)𝐳superscriptsubscript𝜉italic-ϵ𝜈𝐵2𝑁subscript𝜉italic-ϵ𝜈𝐵2𝑁𝑇similar-tosubscript𝒩superscript𝐵10subscript𝐈𝐵1\displaystyle\mathbf{z}=\left(\xi_{\epsilon}\left(\nu-\frac{B}{2N}\right),\ldots,\xi_{\epsilon}\left(\nu+\frac{B}{2N}\right)\right)^{T}\sim\mathcal{N}_{\mathbb{C}^{B+1}}\left(\mathbf{0},\mathbf{I}_{B+1}\right)

and matrix 𝚵=𝐈B+1B+1𝚵subscript𝐈𝐵1𝐵1\boldsymbol{\Xi}=\frac{\mathbf{I}_{B+1}}{B+1}, there exists some constant κ𝜅\kappa independent of M𝑀M such that for all t2𝑡2t\geq 2,

(maxν𝒱N1B+1b=B/2B/2|ξϵ(ν+bN)|2>t)N(1B+1b=B/2B/2|ξϵ(bN)|2>t)Nexp(κB)subscript𝜈subscript𝒱𝑁1𝐵1superscriptsubscript𝑏𝐵2𝐵2superscriptsubscript𝜉italic-ϵ𝜈𝑏𝑁2𝑡𝑁1𝐵1superscriptsubscript𝑏𝐵2𝐵2superscriptsubscript𝜉italic-ϵ𝑏𝑁2𝑡𝑁exp𝜅𝐵\displaystyle\mathbb{P}\left(\max_{\nu\in\mathcal{V}_{N}}\frac{1}{B+1}\sum_{b=-B/2}^{B/2}\left|\xi_{\epsilon}\left(\nu+\frac{b}{N}\right)\right|^{2}>t\right)\leq N\mathbb{P}\left(\frac{1}{B+1}\sum_{b=-B/2}^{B/2}\left|\xi_{\epsilon}\left(\frac{b}{N}\right)\right|^{2}>t\right)\leq N\mathrm{exp}\left(-\kappa B\right)

and Borel-Cantelli lemma together with Assumption 4 imply

maxν𝒱N1B+1b=B/2B/2|ξϵ(ν+bN)|2=𝒪(1)subscript𝜈subscript𝒱𝑁1𝐵1superscriptsubscript𝑏𝐵2𝐵2superscriptsubscript𝜉italic-ϵ𝜈𝑏𝑁2𝒪1\displaystyle\max_{\nu\in\mathcal{V}_{N}}\frac{1}{B+1}\sum_{b=-B/2}^{B/2}\left|\xi_{\epsilon}\left(\nu+\frac{b}{N}\right)\right|^{2}=\mathcal{O}\left(1\right)

with probability one. Defining

𝚺ϵ(ν)=1B+1(ξϵ(νB2N),,ξϵ(ν+B2N))subscript𝚺italic-ϵ𝜈1𝐵1subscript𝜉italic-ϵ𝜈𝐵2𝑁subscript𝜉italic-ϵ𝜈𝐵2𝑁\displaystyle\boldsymbol{\Sigma}_{\epsilon}(\nu)=\frac{1}{\sqrt{B+1}}\left(\xi_{\epsilon}\left(\nu-\frac{B}{2N}\right),\ldots,\xi_{\epsilon}\left(\nu+\frac{B}{2N}\right)\right)

as well as

𝚽(ν)=1B+1[ϕ(νB2N),,ϕ(ν+B2N)]𝚽𝜈1𝐵1bold-italic-ϕ𝜈𝐵2𝑁bold-italic-ϕ𝜈𝐵2𝑁\displaystyle\boldsymbol{\Phi}(\nu)=\frac{1}{\sqrt{B+1}}\left[\boldsymbol{\phi}\left(\nu-\frac{B}{2N}\right),\ldots,\boldsymbol{\phi}\left(\nu+\frac{B}{2N}\right)\right]

with ϕ(ν)=𝐡(ν)ξϵ(ν)bold-italic-ϕ𝜈𝐡𝜈subscript𝜉italic-ϵ𝜈\boldsymbol{\phi}(\nu)=\mathbf{h}(\nu)\xi_{\epsilon}(\nu), we therefore have the control

maxν𝒱N𝐡(ν)𝚺ϵ(ν)𝚽(ν)2subscript𝜈subscript𝒱𝑁subscriptnorm𝐡𝜈subscript𝚺italic-ϵ𝜈𝚽𝜈2absent\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\mathbf{h}(\nu)\boldsymbol{\Sigma}_{\epsilon}(\nu)-\boldsymbol{\Phi}(\nu)\right\|_{2}\leq supν[0,1]maxb{B2,,B2}𝐡(ν)𝐡(ν+bN)2maxν𝒱N1B+1b=B/2B/2|ξϵ(ν+bN)|2subscriptsupremum𝜈01subscript𝑏𝐵2𝐵2subscriptnorm𝐡𝜈𝐡𝜈𝑏𝑁2subscript𝜈subscript𝒱𝑁1𝐵1superscriptsubscript𝑏𝐵2𝐵2superscriptsubscript𝜉italic-ϵ𝜈𝑏𝑁2\displaystyle\sup_{\nu\in[0,1]}\max_{b\in\{-\frac{B}{2},\ldots,\frac{B}{2}\}}\left\|\mathbf{h}(\nu)-\mathbf{h}\left(\nu+\frac{b}{N}\right)\right\|_{2}\sqrt{\max_{\nu\in\mathcal{V}_{N}}\frac{1}{B+1}\sum_{b=-B/2}^{B/2}\left|\xi_{\epsilon}\left(\nu+\frac{b}{N}\right)\right|^{2}}
=𝒪(1N1α) a.s.absent𝒪1superscript𝑁1𝛼 a.s.\displaystyle=\mathcal{O}\left(\frac{1}{N^{1-\alpha}}\right)\text{ a.s.}
Ma.s.0.\displaystyle\xrightarrow[M\to\infty]{a.s.}0.

Finally, since the spectral norm of a matrix is bounded by its Frobenius norm,

maxν𝒱N𝚺𝐮(ν)𝚽(ν)2subscript𝜈subscript𝒱𝑁subscriptnormsubscript𝚺𝐮𝜈𝚽𝜈2\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\Sigma}_{\mathbf{u}}(\nu)-\boldsymbol{\Phi}(\nu)\right\|_{2} 1B+1b=B/2B/2𝝃𝐮(ν+bN)ϕ(ν+bN)22absent1𝐵1superscriptsubscript𝑏𝐵2𝐵2superscriptsubscriptnormsubscript𝝃𝐮𝜈𝑏𝑁bold-italic-ϕ𝜈𝑏𝑁22\displaystyle\leq\sqrt{\frac{1}{B+1}\sum_{b=-B/2}^{B/2}\left\|\boldsymbol{\xi}_{\mathbf{u}}\left(\nu+\frac{b}{N}\right)-\boldsymbol{\phi}\left(\nu+\frac{b}{N}\right)\right\|_{2}^{2}}
maxν𝒱N𝝃𝐮(ν)ϕ(ν)2.absentsubscript𝜈subscript𝒱𝑁subscriptnormsubscript𝝃𝐮𝜈bold-italic-ϕ𝜈2\displaystyle\leq\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\xi}_{\mathbf{u}}\left(\nu\right)-\boldsymbol{\phi}\left(\nu\right)\right\|_{2}.

Theorem 1 is proven if we show that

maxν𝒱N𝝃𝐮(ν)𝐡(ν)𝝃ϵ(ν)2Na.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\xi}_{\mathbf{u}}\left(\nu\right)-\mathbf{h}\left(\nu\right)\boldsymbol{\xi}_{\epsilon}\left(\nu\right)\right\|_{2}\xrightarrow[N\to\infty]{a.s.}0.

B-C Periodization

For all integer n𝑛n, let [n]delimited-[]𝑛[n] denotes the integer contained in {0,,N1}0𝑁1\{0,\ldots,N-1\} such that [n]n(modN)delimited-[]𝑛annotated𝑛pmod𝑁[n]\equiv n\pmod{N} and define

𝐮~n=k𝐡kϵ[nk]subscript~𝐮𝑛subscript𝑘subscript𝐡𝑘subscriptitalic-ϵdelimited-[]𝑛𝑘\displaystyle\tilde{\mathbf{u}}_{n}=\sum_{k\in\mathbb{Z}}\mathbf{h}_{k}\epsilon_{[n-k]}

where (𝐮~n)nsubscriptsubscript~𝐮𝑛𝑛(\tilde{\mathbf{u}}_{n})_{n\in\mathbb{Z}} represents the circular convolution between (𝐡k)ksubscriptsubscript𝐡𝑘𝑘(\mathbf{h}_{k})_{k\in\mathbb{Z}} and (ϵn)nsubscriptsubscriptitalic-ϵ𝑛𝑛(\epsilon_{n})_{n\in\mathbb{Z}}. If 𝝃𝐮~(ν)=1Nn=0N1𝐮~nei2πnνsubscript𝝃~𝐮𝜈1𝑁superscriptsubscript𝑛0𝑁1subscript~𝐮𝑛superscriptei2𝜋𝑛𝜈\boldsymbol{\xi}_{\tilde{\mathbf{u}}}(\nu)=\frac{1}{\sqrt{N}}\sum_{n=0}^{N-1}\tilde{\mathbf{u}}_{n}\mathrm{e}^{-\mathrm{i}2\pi n\nu}, then the equality

𝝃𝐮~(ν)=𝐡(ν)ξϵ(ν)subscript𝝃~𝐮𝜈𝐡𝜈subscript𝜉italic-ϵ𝜈\displaystyle\boldsymbol{\xi}_{\tilde{\mathbf{u}}}(\nu)=\mathbf{h}(\nu)\xi_{\epsilon}(\nu)

holds for all ν𝒱N𝜈subscript𝒱𝑁\nu\in\mathcal{V}_{N}. It is straightforward to check that

𝝃𝐮~(ν)𝝃𝐮(ν)=𝜹(ν)+𝜹ˇ(ν)subscript𝝃~𝐮𝜈subscript𝝃𝐮𝜈𝜹𝜈ˇ𝜹𝜈\displaystyle\boldsymbol{\xi}_{\tilde{\mathbf{u}}}(\nu)-\boldsymbol{\xi}_{\mathbf{u}}(\nu)=\boldsymbol{\delta}(\nu)+\check{\boldsymbol{\delta}}(\nu)

where

𝜹(ν)𝜹𝜈\displaystyle\boldsymbol{\delta}(\nu) =1Nk=1N1𝐡kp=1k(ϵ[p]ϵp)ei2πν(kp)absent1𝑁superscriptsubscript𝑘1𝑁1subscript𝐡𝑘superscriptsubscript𝑝1𝑘subscriptitalic-ϵdelimited-[]𝑝subscriptitalic-ϵ𝑝superscriptei2𝜋𝜈𝑘𝑝\displaystyle=\frac{1}{\sqrt{N}}\sum_{k=1}^{N-1}\mathbf{h}_{k}\sum_{p=1}^{k}\left(\epsilon_{[-p]}-\epsilon_{-p}\right)\mathrm{e}^{-\mathrm{i}2\pi\nu(k-p)}
+1Nk=N+𝐡kp=0N1(ϵ[pk]ϵpk)ei2πνp1𝑁superscriptsubscript𝑘𝑁subscript𝐡𝑘superscriptsubscript𝑝0𝑁1subscriptitalic-ϵdelimited-[]𝑝𝑘subscriptitalic-ϵ𝑝𝑘superscriptei2𝜋𝜈𝑝\displaystyle+\frac{1}{\sqrt{N}}\sum_{k=N}^{+\infty}\mathbf{h}_{k}\sum_{p=0}^{N-1}\left(\epsilon_{[p-k]}-\epsilon_{p-k}\right)\mathrm{e}^{-\mathrm{i}2\pi\nu p}

and

𝜹ˇ(ν)=ˇ𝜹𝜈absent\displaystyle\check{\boldsymbol{\delta}}(\nu)= 1Nk=1N1𝐡kp=1k(ϵ[N+p1]ϵN+p1)ei2πν(N1+pk)1𝑁superscriptsubscript𝑘1𝑁1subscript𝐡𝑘superscriptsubscript𝑝1𝑘subscriptitalic-ϵdelimited-[]𝑁𝑝1subscriptitalic-ϵ𝑁𝑝1superscriptei2𝜋𝜈𝑁1𝑝𝑘\displaystyle\frac{1}{\sqrt{N}}\sum_{k=1}^{N-1}\mathbf{h}_{-k}\sum_{p=1}^{k}\left(\epsilon_{[N+p-1]}-\epsilon_{N+p-1}\right)\mathrm{e}^{-\mathrm{i}2\pi\nu(N-1+p-k)}
+1Nk=N+𝐡kp=0N1(ϵ[p+k]ϵp+k)ei2πνp1𝑁superscriptsubscript𝑘𝑁subscript𝐡𝑘superscriptsubscript𝑝0𝑁1subscriptitalic-ϵdelimited-[]𝑝𝑘subscriptitalic-ϵ𝑝𝑘superscriptei2𝜋𝜈𝑝\displaystyle+\frac{1}{\sqrt{N}}\sum_{k=N}^{+\infty}\mathbf{h}_{-k}\sum_{p=0}^{N-1}\left(\epsilon_{[p+k]}-\epsilon_{p+k}\right)\mathrm{e}^{-\mathrm{i}2\pi\nu p}

Theorem 2 is proved if we can show that

maxν𝒱N𝜹(ν)2Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\delta}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0 (31)

and

maxν𝒱N𝜹ˇ(ν)2Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\check{\boldsymbol{\delta}}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0. (32)

In the remainder, we only prove (31) and omit the details for (32) whose treatment is similar. To that end, we define

𝜹1(ν)=1Nk=1N1𝐡kp=1k(ϵ[p]ϵp)ei2πν(kp)subscript𝜹1𝜈1𝑁superscriptsubscript𝑘1𝑁1subscript𝐡𝑘superscriptsubscript𝑝1𝑘subscriptitalic-ϵdelimited-[]𝑝subscriptitalic-ϵ𝑝superscriptei2𝜋𝜈𝑘𝑝\displaystyle\boldsymbol{\delta}_{1}(\nu)=\frac{1}{\sqrt{N}}\sum_{k=1}^{N-1}\mathbf{h}_{k}\sum_{p=1}^{k}\left(\epsilon_{[-p]}-\epsilon_{-p}\right)\mathrm{e}^{-\mathrm{i}2\pi\nu(k-p)}
𝜹2(ν)=1Nk=N+𝐡kp=0N1(ϵ[pk]ϵpk)ei2πνp.subscript𝜹2𝜈1𝑁superscriptsubscript𝑘𝑁subscript𝐡𝑘superscriptsubscript𝑝0𝑁1subscriptitalic-ϵdelimited-[]𝑝𝑘subscriptitalic-ϵ𝑝𝑘superscriptei2𝜋𝜈𝑝\displaystyle\boldsymbol{\delta}_{2}(\nu)=\frac{1}{\sqrt{N}}\sum_{k=N}^{+\infty}\mathbf{h}_{k}\sum_{p=0}^{N-1}\left(\epsilon_{[p-k]}-\epsilon_{p-k}\right)\mathrm{e}^{-\mathrm{i}2\pi\nu p}.

B-D Control of 𝛅1(ν)subscript𝛅1𝜈\boldsymbol{\delta}_{1}(\nu)

For p{1,,N1}𝑝1𝑁1p\in\{1,\ldots,N-1\}, let

zp(ν)=(ϵ[p]ϵp)ei2πνp=(ϵNpϵp)ei2πνp.subscript𝑧𝑝𝜈subscriptitalic-ϵdelimited-[]𝑝subscriptitalic-ϵ𝑝superscriptei2𝜋𝜈𝑝subscriptitalic-ϵ𝑁𝑝subscriptitalic-ϵ𝑝superscriptei2𝜋𝜈𝑝\displaystyle z_{p}(\nu)=\left(\epsilon_{[-p]}-\epsilon_{-p}\right)\mathrm{e}^{\mathrm{i}2\pi\nu p}=\left(\epsilon_{N-p}-\epsilon_{-p}\right)\mathrm{e}^{\mathrm{i}2\pi\nu p}.

Then z1(ν),,zN1(ν)subscript𝑧1𝜈subscript𝑧𝑁1𝜈z_{1}(\nu),\ldots,z_{N-1}(\nu) are i.i.d. 𝒩(0,2)subscript𝒩02\mathcal{N}_{\mathbb{C}}(0,2) and by rearranging the sums in 𝜹1(ν)subscript𝜹1𝜈\boldsymbol{\delta}_{1}(\nu), we have

𝜹1(ν)=p=1N1zp(ν)𝐠p(ν)subscript𝜹1𝜈superscriptsubscript𝑝1𝑁1subscript𝑧𝑝𝜈subscript𝐠𝑝𝜈\displaystyle\boldsymbol{\delta}_{1}(\nu)=\sum_{p=1}^{N-1}z_{p}(\nu)\mathbf{g}_{p}(\nu)

with

𝐠p(ν)=1Nk=pN1𝐡kei2πkν.subscript𝐠𝑝𝜈1𝑁superscriptsubscript𝑘𝑝𝑁1subscript𝐡𝑘superscriptei2𝜋𝑘𝜈\displaystyle\mathbf{g}_{p}(\nu)=\frac{1}{\sqrt{N}}\sum_{k=p}^{N-1}\mathbf{h}_{k}\mathrm{e}^{-\mathrm{i}2\pi k\nu}.

Therefore, 𝜹1(ν)𝒩M(𝟎,𝐆(ν))similar-tosubscript𝜹1𝜈subscript𝒩superscript𝑀0𝐆𝜈\boldsymbol{\delta}_{1}(\nu)\sim\mathcal{N}_{\mathbb{C}^{M}}\left(\mathbf{0},\mathbf{G}(\nu)\right) with

𝐆(ν)=2p=1N1𝐠p(ν)𝐠p(ν).𝐆𝜈2superscriptsubscript𝑝1𝑁1subscript𝐠𝑝𝜈subscript𝐠𝑝superscript𝜈\displaystyle\mathbf{G}(\nu)=2\sum_{p=1}^{N-1}\mathbf{g}_{p}(\nu)\mathbf{g}_{p}(\nu)^{*}.

Moreover,

𝔼𝜹1(ν)22=Tr𝐆(ν)2Np=1N1(k=pN1𝐡k22+2pk<kN1𝐡k2𝐡k2)𝔼superscriptsubscriptnormsubscript𝜹1𝜈22Tr𝐆𝜈2𝑁superscriptsubscript𝑝1𝑁1superscriptsubscript𝑘𝑝𝑁1superscriptsubscriptnormsubscript𝐡𝑘222subscript𝑝𝑘superscript𝑘𝑁1subscriptnormsubscript𝐡𝑘2subscriptnormsubscript𝐡superscript𝑘2\displaystyle\mathbb{E}\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2}=\mathrm{Tr}\,\mathbf{G}(\nu)\leq\frac{2}{N}\sum_{p=1}^{N-1}\left(\sum_{k=p}^{N-1}\left\|\mathbf{h}_{k}\right\|_{2}^{2}+2\sum_{p\leq k<k^{\prime}\leq N-1}\left\|\mathbf{h}_{k}\right\|_{2}\left\|\mathbf{h}_{k^{\prime}}\right\|_{2}\right)

and a straightforward rearrangement together with (30) leads to

maxν[0,1]𝔼𝜹1(ν)22subscript𝜈01𝔼superscriptsubscriptnormsubscript𝜹1𝜈22\displaystyle\max_{\nu\in[0,1]}\mathbb{E}\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2} 2Nk=1N1k𝐡k22+4N1k<kN1kk𝐡k2𝐡k2absent2𝑁superscriptsubscript𝑘1𝑁1𝑘superscriptsubscriptnormsubscript𝐡𝑘224𝑁subscript1𝑘superscript𝑘𝑁1𝑘superscript𝑘subscriptnormsubscript𝐡𝑘2subscriptnormsubscript𝐡superscript𝑘2\displaystyle\leq\frac{2}{N}\sum_{k=1}^{N-1}k\left\|\mathbf{h}_{k}\right\|_{2}^{2}+\frac{4}{N}\sum_{1\leq k<k^{\prime}\leq N-1}\sqrt{k}\sqrt{k^{\prime}}\left\|\mathbf{h}_{k}\right\|_{2}\left\|\mathbf{h}_{k^{\prime}}\right\|_{2}
=2N(k=1N1k𝐡k2)2absent2𝑁superscriptsuperscriptsubscript𝑘1𝑁1𝑘subscriptnormsubscript𝐡𝑘22\displaystyle=\frac{2}{N}\left(\sum_{k=1}^{N-1}\sqrt{k}\left\|\mathbf{h}_{k}\right\|_{2}\right)^{2}
=𝒪(1N).absent𝒪1𝑁\displaystyle=\mathcal{O}\left(\frac{1}{N}\right).

where we used that kkk𝑘𝑘superscript𝑘k\leq\sqrt{k}\sqrt{k^{\prime}} for kksuperscript𝑘𝑘k^{\prime}\geq k. Additionally,

maxν[0,1]𝐆(ν)2maxν[0,1]Tr𝐆(ν)=𝒪(1N)subscript𝜈01subscriptnorm𝐆𝜈2subscript𝜈01Tr𝐆𝜈𝒪1𝑁\displaystyle\max_{\nu\in[0,1]}\left\|\mathbf{G}(\nu)\right\|_{2}\leq\max_{\nu\in[0,1]}\mathrm{Tr}\,\mathbf{G}(\nu)=\mathcal{O}\left(\frac{1}{N}\right)

and

maxν[0,1]𝐆(ν)FMmaxν[0,1]𝐆(ν)2=𝒪(MN).subscript𝜈01subscriptnorm𝐆𝜈𝐹𝑀subscript𝜈01subscriptnorm𝐆𝜈2𝒪𝑀𝑁\displaystyle\max_{\nu\in[0,1]}\left\|\mathbf{G}(\nu)\right\|_{F}\leq\sqrt{M}\max_{\nu\in[0,1]}\left\|\mathbf{G}(\nu)\right\|_{2}=\mathcal{O}\left(\frac{\sqrt{M}}{N}\right).

Using Lemma 1, there exists a constant κ>0𝜅0\kappa>0 independent of M,(𝐡k)k𝑀subscriptsubscript𝐡𝑘𝑘M,(\mathbf{h}_{k})_{k\in\mathbb{Z}} such that for all t>0𝑡0t>0,

(maxν𝒱N|𝜹1(ν)22𝔼𝜹1(ν)22|>t)2Nmaxν𝒱Nexp(κmin(t2𝐆(ν)F2,t𝐆(ν)2)).subscript𝜈subscript𝒱𝑁superscriptsubscriptnormsubscript𝜹1𝜈22𝔼superscriptsubscriptnormsubscript𝜹1𝜈22𝑡2𝑁subscript𝜈subscript𝒱𝑁exp𝜅superscript𝑡2superscriptsubscriptnorm𝐆𝜈𝐹2𝑡subscriptnorm𝐆𝜈2\displaystyle\mathbb{P}\left(\max_{\nu\in\mathcal{V}_{N}}\left|\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2}-\mathbb{E}\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2}\right|>t\right)\leq 2N\max_{\nu\in\mathcal{V}_{N}}\mathrm{exp}\left(-\kappa\min\left(\frac{t^{2}}{\left\|\mathbf{G}(\nu)\right\|_{F}^{2}},\frac{t}{\left\|\mathbf{G}(\nu)\right\|_{2}}\right)\right).

Applying Assumption 4 and Borel-Cantelli lemma, it follows that

maxν𝒱N|𝜹1(ν)22𝔼𝜹1(ν)22|Na.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left|\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2}-\mathbb{E}\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2}\right|\xrightarrow[N\to\infty]{a.s.}0.

Finally, we deduce that

maxν𝒱N𝜹1(ν)22subscript𝜈subscript𝒱𝑁superscriptsubscriptnormsubscript𝜹1𝜈22\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2} maxν𝒱N𝔼𝜹1(ν)22+maxν𝒱N|𝜹1(ν)22𝔼𝜹1(ν)22|Na.s.0.\displaystyle\leq\max_{\nu\in\mathcal{V}_{N}}\mathbb{E}\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2}+\max_{\nu\in\mathcal{V}_{N}}\left|\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2}-\mathbb{E}\left\|\boldsymbol{\delta}_{1}(\nu)\right\|_{2}^{2}\right|\xrightarrow[N\to\infty]{a.s.}0.

B-E Control of 𝛅2(ν)subscript𝛅2𝜈\boldsymbol{\delta}_{2}(\nu)

We first split 𝜹2(ν)subscript𝜹2𝜈\boldsymbol{\delta}_{2}(\nu) in the following two parts

𝜹2(ν)=𝜹2,1(ν)+𝜹2,2(ν)subscript𝜹2𝜈subscript𝜹21𝜈subscript𝜹22𝜈\displaystyle\boldsymbol{\delta}_{2}(\nu)=\boldsymbol{\delta}_{2,1}(\nu)+\boldsymbol{\delta}_{2,2}(\nu)

where

𝜹2,1(ν)=1Nk=N+𝐡kp=0N1ϵ[pk]ei2πpνsubscript𝜹21𝜈1𝑁superscriptsubscript𝑘𝑁subscript𝐡𝑘superscriptsubscript𝑝0𝑁1subscriptitalic-ϵdelimited-[]𝑝𝑘superscriptei2𝜋𝑝𝜈\displaystyle\boldsymbol{\delta}_{2,1}(\nu)=\frac{1}{\sqrt{N}}\sum_{k=N}^{+\infty}\mathbf{h}_{k}\sum_{p=0}^{N-1}\epsilon_{[p-k]}\mathrm{e}^{-\mathrm{i}2\pi p\nu}
𝜹2,2(ν)=1Nk=N+𝐡kp=0N1ϵpkei2πpν.subscript𝜹22𝜈1𝑁superscriptsubscript𝑘𝑁subscript𝐡𝑘superscriptsubscript𝑝0𝑁1subscriptitalic-ϵ𝑝𝑘superscriptei2𝜋𝑝𝜈\displaystyle\boldsymbol{\delta}_{2,2}(\nu)=\frac{1}{\sqrt{N}}\sum_{k=N}^{+\infty}\mathbf{h}_{k}\sum_{p=0}^{N-1}\epsilon_{p-k}\mathrm{e}^{-\mathrm{i}2\pi p\nu}.

We remark that 𝜹2,1(ν)subscript𝜹21𝜈\boldsymbol{\delta}_{2,1}(\nu) only involves the N𝑁N i.i.d. random variables ϵ0,,ϵN1subscriptitalic-ϵ0subscriptitalic-ϵ𝑁1\epsilon_{0},\ldots,\epsilon_{N-1} and that

𝜹2,1(ν)=p=0N1ϵp𝐠~p(ν)subscript𝜹21𝜈superscriptsubscript𝑝0𝑁1subscriptitalic-ϵ𝑝subscript~𝐠𝑝𝜈\displaystyle\boldsymbol{\delta}_{2,1}(\nu)=\sum_{p=0}^{N-1}\epsilon_{p}\tilde{\mathbf{g}}_{p}(\nu)

with 𝐠~p(ν)subscript~𝐠𝑝𝜈\tilde{\mathbf{g}}_{p}(\nu) defined as

𝐠~p(ν)=1Nk=N+𝐡kei2πν[p+k].subscript~𝐠𝑝𝜈1𝑁superscriptsubscript𝑘𝑁subscript𝐡𝑘superscriptei2𝜋𝜈delimited-[]𝑝𝑘\displaystyle\tilde{\mathbf{g}}_{p}(\nu)=\frac{1}{\sqrt{N}}\sum_{k=N}^{+\infty}\mathbf{h}_{k}\mathrm{e}^{-\mathrm{i}2\pi\nu[p+k]}.

It is clear that

maxp=1,,Nmaxν[0,1]𝐠~p(ν)2subscript𝑝1𝑁subscript𝜈01subscriptnormsubscript~𝐠𝑝𝜈2\displaystyle\max_{p=1,\ldots,N}\max_{\nu\in[0,1]}\|\tilde{\mathbf{g}}_{p}(\nu)\|_{2} 1Nk=N+𝐡k21N3/2k=N+k𝐡k2absent1𝑁superscriptsubscript𝑘𝑁subscriptnormsubscript𝐡𝑘21superscript𝑁32superscriptsubscript𝑘𝑁𝑘subscriptnormsubscript𝐡𝑘2\displaystyle\leq\frac{1}{\sqrt{N}}\sum_{k=N}^{+\infty}\left\|\mathbf{h}_{k}\right\|_{2}\leq\frac{1}{N^{3/2}}\sum_{k=N}^{+\infty}k\left\|\mathbf{h}_{k}\right\|_{2}

and from (30),

maxp=1,,Nmaxν[0,1]𝐠~p(ν)2=o(1N3/2)subscript𝑝1𝑁subscript𝜈01subscriptnormsubscript~𝐠𝑝𝜈2𝑜1superscript𝑁32\displaystyle\max_{p=1,\ldots,N}\max_{\nu\in[0,1]}\|\tilde{\mathbf{g}}_{p}(\nu)\|_{2}=o\left(\frac{1}{N^{3/2}}\right)

Thus 𝜹2,1(ν)𝒩M(𝟎,𝐆~(ν))similar-tosubscript𝜹21𝜈subscript𝒩superscript𝑀0~𝐆𝜈\boldsymbol{\delta}_{2,1}(\nu)\sim\mathcal{N}_{\mathbb{C}^{M}}\left(\mathbf{0},\tilde{\mathbf{G}}(\nu)\right) with 𝐆~(ν)=p=0N1𝐠~p(ν)𝐠~p(ν)~𝐆𝜈superscriptsubscript𝑝0𝑁1subscript~𝐠𝑝𝜈subscript~𝐠𝑝superscript𝜈\tilde{\mathbf{G}}(\nu)=\sum_{p=0}^{N-1}\tilde{\mathbf{g}}_{p}(\nu)\tilde{\mathbf{g}}_{p}(\nu)^{*} and

maxν[0,1]Tr𝐆~(ν)=o(1N2)subscript𝜈01Tr~𝐆𝜈𝑜1superscript𝑁2\displaystyle\max_{\nu\in[0,1]}\mathrm{Tr}\,\tilde{\mathbf{G}}(\nu)=o\left(\frac{1}{N^{2}}\right)

as M𝑀M\to\infty. Using Lemma 1 as for the control of 𝜹1(ν)subscript𝜹1𝜈\boldsymbol{\delta}_{1}(\nu) in the previous section, we end up with

maxν𝒱N𝜹2,1(ν)2Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\delta}_{2,1}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0.

We now consider the term 𝜹2,2(ν)subscript𝜹22𝜈\boldsymbol{\delta}_{2,2}(\nu), which involves the sequence of random variables (ϵn)n1subscriptsubscriptitalic-ϵ𝑛𝑛1(\epsilon_{-n})_{n\geq 1}. For all kN𝑘𝑁k\geq N, set

𝝌k=1N𝐡kp=0N1ϵpkei2πpνsubscript𝝌𝑘1𝑁subscript𝐡𝑘superscriptsubscript𝑝0𝑁1subscriptitalic-ϵ𝑝𝑘superscriptei2𝜋𝑝𝜈\displaystyle\boldsymbol{\chi}_{k}=\frac{1}{\sqrt{N}}\mathbf{h}_{k}\sum_{p=0}^{N-1}\epsilon_{p-k}\mathrm{e}^{-\mathrm{i}2\pi p\nu}

and consider the sequence (𝝌p)kNsubscriptsubscript𝝌𝑝𝑘𝑁(\boldsymbol{\chi}_{p})_{k\geq N}. Using Assumption 3,

k=N+𝝌k2superscriptsubscript𝑘𝑁subscriptnormsubscript𝝌𝑘2\displaystyle\sum_{k=N}^{+\infty}\|\boldsymbol{\chi}_{k}\|_{2} k=N+k𝐡k2|1Np=0N1ϵpkei2πpν|absentsuperscriptsubscript𝑘𝑁𝑘subscriptnormsubscript𝐡𝑘21𝑁superscriptsubscript𝑝0𝑁1subscriptitalic-ϵ𝑝𝑘superscriptei2𝜋𝑝𝜈\displaystyle\leq\sum_{k=N}^{+\infty}\sqrt{k}\|\mathbf{h}_{k}\|_{2}\left|\frac{1}{N}\sum_{p=0}^{N-1}\epsilon_{p-k}\mathrm{e}^{-\mathrm{i}2\pi p\nu}\right|
(supkN|1Np=0N1ϵpkei2πpν|)k=N+k𝐡k2absentsubscriptsupremum𝑘𝑁1𝑁superscriptsubscript𝑝0𝑁1subscriptitalic-ϵ𝑝𝑘superscriptei2𝜋𝑝𝜈superscriptsubscript𝑘𝑁𝑘subscriptnormsubscript𝐡𝑘2\displaystyle\leq\left(\sup_{k\geq N}\left|\frac{1}{N}\sum_{p=0}^{N-1}\epsilon_{p-k}\mathrm{e}^{-\mathrm{i}2\pi p\nu}\right|\right)\sum_{k=N}^{+\infty}\sqrt{k}\|\mathbf{h}_{k}\|_{2}
<+ a.s.absent a.s.\displaystyle<+\infty\text{ a.s.}

since for any k𝑘k, by the gaussianity of the ϵpksubscriptitalic-ϵ𝑝𝑘\epsilon_{p-k}, supν𝒱N|1Np=0N1ϵpkei2πpν|subscriptsupremum𝜈subscript𝒱𝑁1𝑁superscriptsubscript𝑝0𝑁1subscriptitalic-ϵ𝑝𝑘superscriptei2𝜋𝑝𝜈\sup_{\nu\in\mathcal{V}_{N}}\left|\frac{1}{N}\sum_{p=0}^{N-1}\epsilon_{p-k}\mathrm{e}^{-\mathrm{i}2\pi p\nu}\right| converges almost surely towards 00 as N+𝑁N\to+\infty by the law of the large numbers, so it remains almost surely bounded for any finite N𝑁N. This implies that the family (𝝌k)kNsubscriptsubscript𝝌𝑘𝑘𝑁(\boldsymbol{\chi}_{k})_{k\geq N} is a.s. absolutely summable. Therefore, we can rearrange the series defining 𝜹2,2(ν)subscript𝜹22𝜈\boldsymbol{\delta}_{2,2}(\nu) and write

𝜹2,2(ν)=p=1+ϵp𝐠ˇp(ν)subscript𝜹22𝜈superscriptsubscript𝑝1subscriptitalic-ϵ𝑝subscriptˇ𝐠𝑝𝜈\displaystyle\boldsymbol{\delta}_{2,2}(\nu)=\sum_{p=1}^{+\infty}\epsilon_{-p}\check{\mathbf{g}}_{p}(\nu)

with probability one, where this time 𝐠p(ν)subscript𝐠𝑝𝜈\mathbf{g}_{p}(\nu) is defined for all p1𝑝1p\geq 1 as

𝐠ˇp(ν)=subscriptˇ𝐠𝑝𝜈absent\displaystyle\check{\mathbf{g}}_{p}(\nu)=
{1Nk=0p1𝐡k+Nei2π(N+kp)νif p{1,,N}1Nk=0N1𝐡p+kei2πkνif pN+1..cases1𝑁superscriptsubscript𝑘0𝑝1subscript𝐡𝑘𝑁superscriptei2𝜋𝑁𝑘𝑝𝜈if 𝑝1𝑁1𝑁superscriptsubscript𝑘0𝑁1subscript𝐡𝑝𝑘superscriptei2𝜋𝑘𝜈if 𝑝𝑁1\displaystyle\quad\begin{cases}\frac{1}{\sqrt{N}}\sum_{k=0}^{p-1}\mathbf{h}_{k+N}\ \mathrm{e}^{-\mathrm{i}2\pi(N+k-p)\nu}&\text{if }p\in\{1,\ldots,N\}\\ \frac{1}{\sqrt{N}}\sum_{k=0}^{N-1}\mathbf{h}_{p+k}\ \mathrm{e}^{-\mathrm{i}2\pi k\nu}&\text{if }p\geq N+1.\end{cases}.

Again,

supp1maxν[0,1]𝐠ˇp(ν)2=o(1N)subscriptsupremum𝑝1subscript𝜈01subscriptnormsubscriptˇ𝐠𝑝𝜈2𝑜1𝑁\displaystyle\sup_{p\geq 1}\max_{\nu\in[0,1]}\left\|\check{\mathbf{g}}_{p}(\nu)\right\|_{2}=o\left(\frac{1}{N}\right) (33)

𝜹2,2(ν)𝒩M(𝟎,𝐆ˇ(ν))similar-tosubscript𝜹22𝜈subscript𝒩superscript𝑀0ˇ𝐆𝜈\boldsymbol{\delta}_{2,2}(\nu)\sim\mathcal{N}_{\mathbb{C}^{M}}\left(\mathbf{0},\check{\mathbf{G}}(\nu)\right), where

𝐆ˇ(ν)=p=1+𝐠ˇp(ν)𝐠ˇp(ν)ˇ𝐆𝜈superscriptsubscript𝑝1subscriptˇ𝐠𝑝𝜈subscriptˇ𝐠𝑝superscript𝜈\displaystyle\check{\mathbf{G}}(\nu)=\sum_{p=1}^{+\infty}\check{\mathbf{g}}_{p}(\nu)\check{\mathbf{g}}_{p}(\nu)^{*}

and such that Tr𝐆ˇ(ν)=o(1N)Trˇ𝐆𝜈𝑜1𝑁\mathrm{Tr}\,\check{\mathbf{G}}(\nu)=o\left(\frac{1}{N}\right). Thus, using Lemma 1 also yields

maxν𝒱N𝜹2,2(ν)2Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\delta}_{2,2}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0.

This concludes the proof of Theorem 2.

Appendix C Proof of Theorem 3

To prove Theorem 3, we need as a preliminary step to study the behaviour of the renormalization by dg(𝐒^𝐲(ν))12\operatorname*{dg}(\hat{\mathbf{S}}_{\mathbf{y}}(\nu))^{-\frac{1}{2}} in the SCM.

Lemma 3.

Under Assumptions 1, 3 and 4, we have

maxν𝒱Ndg(𝐒^𝐲(ν))𝐒𝐯(ν)2Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)-\mathbf{S}_{\mathbf{v}}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0 (34)

as well as

maxν𝒱Ndg(𝐒^𝐲(ν))12𝐒𝐯(ν)122Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)^{-\frac{1}{2}}-\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0 (35)
Proof:

To prove (34), we establish successively

maxν𝒱Ndg(𝐒^𝐲(ν))dg(𝐒𝐲(ν))2Ma.s.0\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)-\operatorname*{dg}\left(\mathbf{S}_{\mathbf{y}}(\nu)\right)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0 (36)

as well as

maxν𝒱Ndg(𝐒𝐲(ν))𝐒𝐯(ν)2Ma.s.0\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\mathbf{S}_{\mathbf{y}}(\nu)\right)-\mathbf{S}_{\mathbf{v}}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0 (37)

Using (20), we have the bound

maxν𝒱Ndg(𝐒^𝐲(ν))dg(𝐒𝐲(ν))2Δ1+Δ2,subscript𝜈subscript𝒱𝑁subscriptnormdgsubscript^𝐒𝐲𝜈dgsubscript𝐒𝐲𝜈2subscriptΔ1subscriptΔ2\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)-\operatorname*{dg}\left(\mathbf{S}_{\mathbf{y}}(\nu)\right)\right\|_{2}\leq\Delta_{1}+\Delta_{2},

with

Δ1subscriptΔ1\displaystyle\Delta_{1} =maxν𝒱N𝐒^𝐲(ν)𝐒𝐲(ν)12𝐗(ν)𝐗(ν)B+1𝐒𝐲(ν)122absentsubscript𝜈subscript𝒱𝑁subscriptnormsubscript^𝐒𝐲𝜈subscript𝐒𝐲superscript𝜈12𝐗𝜈𝐗superscript𝜈𝐵1subscript𝐒𝐲superscript𝜈122\displaystyle=\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)-\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\right\|_{2}
Ma.s.0,\displaystyle\xrightarrow[M\to\infty]{a.s.}0,

and

Δ2=maxν𝒱Nmaxm=1,,M|[𝐒𝐲(ν)12(𝐗(ν)𝐗(ν)B+1𝐈M)𝐒𝐲(ν)12]m,m|.subscriptΔ2subscript𝜈subscript𝒱𝑁subscript𝑚1𝑀subscriptdelimited-[]subscript𝐒𝐲superscript𝜈12𝐗𝜈𝐗superscript𝜈𝐵1subscript𝐈𝑀subscript𝐒𝐲superscript𝜈12𝑚𝑚\displaystyle\Delta_{2}=\max_{\nu\in\mathcal{V}_{N}}\max_{m=1,\ldots,M}\left|\left[\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}-\mathbf{I}_{M}\right)\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\right]_{m,m}\right|.

Denoting 𝐮m(ν)=𝐒𝐲(ν)12𝐞msubscript𝐮𝑚𝜈subscript𝐒𝐲superscript𝜈12subscript𝐞𝑚\mathbf{u}_{m}(\nu)=\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\mathbf{e}_{m}, where 𝐞msubscript𝐞𝑚\mathbf{e}_{m} is the m𝑚m-th vector of the canonical basis of Msuperscript𝑀\mathbb{C}^{M}, as well as 𝐱1(ν),,𝐱B+1(ν)subscript𝐱1𝜈subscript𝐱𝐵1𝜈\mathbf{x}_{1}(\nu),\ldots,\mathbf{x}_{B+1}(\nu) the i.i.d. 𝒩M(𝟎,𝐈M)subscript𝒩superscript𝑀0subscript𝐈𝑀\mathcal{N}_{\mathbb{C}^{M}}(\mathbf{0},\mathbf{I}_{M}) column vectors of 𝐗(ν)𝐗𝜈\mathbf{X}(\nu), we have for all t>0𝑡0t>0,

(Δ2>t)ν𝒱Nm=1M(|1B+1b=1B+1|𝐮m(ν)𝐱b(ν)|2𝐮m(ν)22|>t).subscriptΔ2𝑡subscript𝜈subscript𝒱𝑁superscriptsubscript𝑚1𝑀1𝐵1superscriptsubscript𝑏1𝐵1superscriptsubscript𝐮𝑚superscript𝜈subscript𝐱𝑏𝜈2superscriptsubscriptnormsubscript𝐮𝑚𝜈22𝑡\displaystyle\mathbb{P}\left(\Delta_{2}>t\right)\leq\sum_{\nu\in\mathcal{V}_{N}}\sum_{m=1}^{M}\mathbb{P}\left(\left|\frac{1}{B+1}\sum_{b=1}^{B+1}\left|\mathbf{u}_{m}(\nu)^{*}\mathbf{x}_{b}(\nu)\right|^{2}-\|\mathbf{u}_{m}(\nu)\|_{2}^{2}\right|>t\right).

From Assumption 1, Assumption 2 and condition (11) from Assumption 3, we have

0<infM1minm=1,,Mminν[0,1]𝐮m(ν)2supM1maxm=1,,Mmaxν[0,1]𝐮m(ν)2<.0subscriptinfimum𝑀1subscript𝑚1𝑀subscript𝜈01subscriptnormsubscript𝐮𝑚𝜈2subscriptsupremum𝑀1subscript𝑚1𝑀subscript𝜈01subscriptnormsubscript𝐮𝑚𝜈2\displaystyle 0<\inf_{M\geq 1}\min_{m=1,\ldots,M}\min_{\nu\in[0,1]}\left\|\mathbf{u}_{m}(\nu)\right\|_{2}\leq\sup_{M\geq 1}\max_{m=1,\ldots,M}\max_{\nu\in[0,1]}\left\|\mathbf{u}_{m}(\nu)\right\|_{2}<\infty.

Setting in the statement of Lemma 1

𝐳=(𝐱1(ν)T,,𝐱B+1(ν)T)T𝒩M(B+1)(𝟎,𝐈M(B+1))𝐳superscriptsubscript𝐱1superscript𝜈𝑇subscript𝐱𝐵1superscript𝜈𝑇𝑇similar-tosubscript𝒩superscript𝑀𝐵10subscript𝐈𝑀𝐵1\displaystyle\mathbf{z}=\left(\mathbf{x}_{1}(\nu)^{T},\ldots,\mathbf{x}_{B+1}(\nu)^{T}\right)^{T}\sim\mathcal{N}_{\mathbb{C}^{M(B+1)}}(\mathbf{0},\mathbf{I}_{M(B+1)})

and 𝚵𝚵\boldsymbol{\Xi} as the M(B+1)×M(B+1)𝑀𝐵1𝑀𝐵1M(B+1)\times M(B+1) block-diagonal matrix

𝚵=𝐈B+1(𝐮m(ν)𝐮m(ν))B+1𝚵tensor-productsubscript𝐈𝐵1subscript𝐮𝑚𝜈subscript𝐮𝑚superscript𝜈𝐵1\displaystyle\boldsymbol{\Xi}=\frac{\mathbf{I}_{B+1}\otimes\left(\mathbf{u}_{m}(\nu)\mathbf{u}_{m}(\nu)^{*}\right)}{B+1}

with tensor-product\otimes denoting the Kronecker product, we obtain

(Δ2>t)2MNmaxν𝒱Nexp(Cmin{Bt2𝐮m(ν)24,Bt𝐮m(ν)22})subscriptΔ2𝑡2𝑀𝑁subscript𝜈subscript𝒱𝑁exp𝐶𝐵superscript𝑡2superscriptsubscriptnormsubscript𝐮𝑚𝜈24𝐵𝑡superscriptsubscriptnormsubscript𝐮𝑚𝜈22\displaystyle\mathbb{P}\left(\Delta_{2}>t\right)\leq 2MN\max_{\nu\in\mathcal{V}_{N}}\mathrm{exp}\left(-C\min\left\{\frac{Bt^{2}}{\|\mathbf{u}_{m}(\nu)\|_{2}^{4}},\frac{Bt}{\|\mathbf{u}_{m}(\nu)\|_{2}^{2}}\right\}\right)

where C>0𝐶0C>0 is a constant independent of M𝑀M, which in turn implies that

Δ2Ma.s.0\displaystyle\Delta_{2}\xrightarrow[M\to\infty]{a.s.}0

and that (36) holds. In order to check (37), we use Assumption 3 eq. (12) to get that

maxν𝒱Ndg(𝐇(ν)𝐇(ν))2subscript𝜈subscript𝒱𝑁subscriptnormdg𝐇𝜈𝐇superscript𝜈2\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}\right)\right\|_{2} =maxν𝒱Nmaxm=1,,M𝐡m(ν)22M0absentsubscript𝜈subscript𝒱𝑁subscript𝑚1𝑀superscriptsubscriptnormsubscript𝐡𝑚𝜈22𝑀absent0\displaystyle=\max_{\nu\in\mathcal{V}_{N}}\max_{m=1,\ldots,M}\left\|\mathbf{h}_{m}(\nu)\right\|_{2}^{2}\xrightarrow[M\to\infty]{}0

and from the fact that

dg(𝐒𝐲(ν))=dg(𝐇(ν)𝐇(ν))+𝐒𝐯(ν)dgsubscript𝐒𝐲𝜈dg𝐇𝜈𝐇superscript𝜈subscript𝐒𝐯𝜈\displaystyle\operatorname*{dg}\left(\mathbf{S}_{\mathbf{y}}(\nu)\right)=\operatorname*{dg}\left(\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}\right)+\mathbf{S}_{\mathbf{v}}(\nu)

we obtain (37) and, in turn, (34).

To prove (35), we write (using that |ab|<|ab|𝑎𝑏𝑎𝑏|\sqrt{a}-\sqrt{b}|<\sqrt{|a-b|} for a,b>0𝑎𝑏0a,b>0)

maxν𝒱Ndg(𝐒^𝐲(ν))12𝐒𝐯(ν)122maxν𝒱Nmaxm=1,,M|[𝐒^𝐲(ν)]m,msm(ν)|[𝐒^𝐲(ν)]m,msm(ν).\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)^{-\frac{1}{2}}-\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\right\|_{2}\leq\max_{\nu\in\mathcal{V}_{N}}\max_{m=1,\ldots,M}\sqrt{\frac{|[\hat{\mathbf{S}}_{\mathbf{y}}(\nu)]_{m,m}-s_{m}(\nu)|}{[\hat{\mathbf{S}}_{\mathbf{y}}(\nu)]_{m,m}\ s_{m}(\nu)}}.

From Assumption 2, there exists ϵ>0italic-ϵ0\epsilon>0 such that

infM1minm=1,,Mminν𝒱Nsm(ν)ϵ>0.subscriptinfimum𝑀1subscript𝑚1𝑀subscript𝜈subscript𝒱𝑁subscript𝑠𝑚𝜈italic-ϵ0\displaystyle\inf_{M\geq 1}\min_{m=1,\ldots,M}\min_{\nu\in\mathcal{V}_{N}}s_{m}(\nu)\geq\epsilon>0.

Using (34) and denoting

Δ=maxν𝒱Ndg(𝐒^𝐲(ν))𝐒𝐯(ν)2Δsubscript𝜈subscript𝒱𝑁subscriptnormdgsubscript^𝐒𝐲𝜈subscript𝐒𝐯𝜈2\displaystyle\Delta=\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)-\mathbf{S}_{\mathbf{v}}(\nu)\right\|_{2}

we have that

maxν𝒱Ndg(𝐒^𝐲(ν))12𝐒𝐯(ν)122\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)^{-\frac{1}{2}}-\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\right\|_{2} Δϵ(ϵΔ)absentΔitalic-ϵitalic-ϵΔ\displaystyle\leq\sqrt{\frac{\Delta}{\epsilon\left(\epsilon-\Delta\right)}}

with probability one for all large M𝑀M, which proves (35). ∎

We also need the following lemma on the boundedness of matrix 𝐒^𝐲(ν)subscript^𝐒𝐲𝜈\hat{\mathbf{S}}_{\mathbf{y}}(\nu).

Lemma 4.

Under Assumptions 1, 3 and 4, we have

lim supMmaxν𝒱M𝐒^𝐲(ν)2<subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑀subscriptnormsubscript^𝐒𝐲𝜈2\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{M}}\left\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right\|_{2}<\infty

with probability one.

Proof:

From (20), we have

lim supMmaxν𝒱N𝐒^𝐲(ν)2lim supMmaxν𝒱N𝐒𝐲(ν)2maxν𝒱N𝐗(ν)𝐗(ν)B+12.subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscriptnormsubscript^𝐒𝐲𝜈2subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscriptnormsubscript𝐒𝐲𝜈2subscript𝜈subscript𝒱𝑁subscriptnorm𝐗𝜈𝐗superscript𝜈𝐵12\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right\|_{2}\leq\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\left\|\mathbf{S}_{\mathbf{y}}(\nu)\right\|_{2}\max_{\nu\in\mathcal{V}_{N}}\left\|\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right\|_{2}.

From Assumptions 1 and 3, it is clear that

supM1maxν𝒱N𝐒𝐲(ν)2<.subscriptsupremum𝑀1subscript𝜈subscript𝒱𝑁subscriptnormsubscript𝐒𝐲𝜈2\displaystyle\sup_{M\geq 1}\max_{\nu\in\mathcal{V}_{N}}\left\|\mathbf{S}_{\mathbf{y}}(\nu)\right\|_{2}<\infty.

Finally, from Lemma 2 and the remarks below this lemma, we have

lim supMmaxν𝒱N𝐗(ν)𝐗(ν)B+12<subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscriptnorm𝐗𝜈𝐗superscript𝜈𝐵12\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\left\|\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right\|_{2}<\infty

with probability one, and Lemma 4 is proved. ∎

Equipped with Lemmas 3 and 4, we are now in position to prove Theorem 3. Define

Δ~=maxν𝒱Ndg(𝐒^𝐲(ν))12𝐒𝐯(ν)122\displaystyle\tilde{\Delta}=\max_{\nu\in\mathcal{V}_{N}}\left\|\operatorname*{dg}\left(\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right)^{-\frac{1}{2}}-\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\right\|_{2}

and recall the definition of the random matrix 𝐗(ν)𝐗𝜈\mathbf{X}(\nu) in (20). Let us write

𝐂^𝐲(ν)𝚵(ν)12𝐗(ν)𝐗(ν)B+1𝚵(ν)12=Ψ1(ν)+Ψ2(ν)subscript^𝐂𝐲𝜈𝚵superscript𝜈12𝐗𝜈𝐗superscript𝜈𝐵1𝚵superscript𝜈12subscriptΨ1𝜈subscriptΨ2𝜈\displaystyle\hat{\mathbf{C}}_{\mathbf{y}}(\nu)-\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}=\Psi_{1}(\nu)+\Psi_{2}(\nu)

where the two error terms are defined by:

Ψ1(ν)subscriptΨ1𝜈\displaystyle\Psi_{1}(\nu) =𝐂^𝐲(ν)𝐒𝐯(ν)12𝐒^𝐲(ν)𝐒𝐯(ν)12absentsubscript^𝐂𝐲𝜈subscript𝐒𝐯superscript𝜈12subscript^𝐒𝐲𝜈subscript𝐒𝐯superscript𝜈12\displaystyle=\hat{\mathbf{C}}_{\mathbf{y}}(\nu)-\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}
Ψ2(ν)=𝐒𝐯(ν)12𝐒^𝐲(ν)𝐒𝐯(ν)12𝚵(ν)12𝐗(ν)𝐗(ν)B+1𝚵(ν)12subscriptΨ2𝜈subscript𝐒𝐯superscript𝜈12subscript^𝐒𝐲𝜈subscript𝐒𝐯superscript𝜈12𝚵superscript𝜈12𝐗𝜈superscript𝐗𝜈𝐵1𝚵superscript𝜈12\displaystyle\Psi_{2}(\nu)=\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\mathbf{S}_{\mathbf{v}}(\nu)^{-\frac{1}{2}}-\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}\frac{\mathbf{X}(\nu)\mathbf{X}^{*}(\nu)}{B+1}\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}

which satisfies:

maxν𝒱NΨ1(ν)2Δ~maxν𝒱N𝐒^𝐲(ν)2(Δ~+2minν𝒱NλM(𝐒𝐯(ν)))subscript𝜈subscript𝒱𝑁subscriptnormsubscriptΨ1𝜈2~Δsubscript𝜈subscript𝒱𝑁subscriptnormsubscript^𝐒𝐲𝜈2~Δ2subscript𝜈subscript𝒱𝑁subscript𝜆𝑀subscript𝐒𝐯𝜈\max_{\nu\in\mathcal{V}_{N}}\left\|\Psi_{1}(\nu)\right\|_{2}\leq\tilde{\Delta}\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)\right\|_{2}\left(\tilde{\Delta}+\frac{2}{\sqrt{\min_{\nu\in\mathcal{V}_{N}}\lambda_{M}\left(\mathbf{S}_{\mathbf{v}}(\nu)\right)}}\right)

and

maxν𝒱NΨ2(ν)2maxν𝒱N𝐒^𝐲(ν)𝐒𝐲(ν)12𝐗(ν)𝐗(ν)B+1𝐒𝐲(ν)122minν𝒱NλM(𝐒𝐯(ν)).subscript𝜈subscript𝒱𝑁subscriptnormsubscriptΨ2𝜈2subscript𝜈subscript𝒱𝑁subscriptnormsubscript^𝐒𝐲𝜈subscript𝐒𝐲superscript𝜈12𝐗𝜈𝐗superscript𝜈𝐵1subscript𝐒𝐲superscript𝜈122subscript𝜈subscript𝒱𝑁subscript𝜆𝑀subscript𝐒𝐯𝜈\max_{\nu\in\mathcal{V}_{N}}\left\|\Psi_{2}(\nu)\right\|_{2}\leq\frac{\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)-\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\mathbf{S}_{\mathbf{y}}(\nu)^{\frac{1}{2}}\right\|_{2}}{\min_{\nu\in\mathcal{V}_{N}}\lambda_{M}(\mathbf{S}_{\mathbf{v}}(\nu))}.

From Assumption 2, we have

infM1minν𝒱NλM(𝐒𝐯(ν))>0.subscriptinfimum𝑀1subscript𝜈subscript𝒱𝑁subscript𝜆𝑀subscript𝐒𝐯𝜈0\displaystyle\inf_{M\geq 1}\min_{\nu\in\mathcal{V}_{N}}\lambda_{M}\left(\mathbf{S}_{\mathbf{v}}(\nu)\right)>0.

Using Lemmas 3 and 4, we directly deduce that

maxν𝒱NΨ1(ν)2Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\Psi_{1}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0.

Likewise, using (20), we deduce that

maxν𝒱NΨ2(ν)2Ma.s.0,\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\Psi_{2}(\nu)\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0,

which concludes the proof of Theorem 3.

Appendix D Proof of Corollary 1, Corollary 2 and Proposition 1

D-A Proof of Corollary 1

Write 𝐒^𝐯(ν)=𝚺𝐯(ν)𝚺𝐯(ν)subscript^𝐒𝐯𝜈subscript𝚺𝐯𝜈subscript𝚺𝐯superscript𝜈\hat{\mathbf{S}}_{\mathbf{v}}(\nu)=\boldsymbol{\Sigma}_{\mathbf{v}}(\nu)\boldsymbol{\Sigma}_{\mathbf{v}}(\nu)^{*}, and denote

Δ𝐯(ν)=𝚺𝐯(ν)1B+1𝐒𝐯(ν)1/2𝐙(ν)2.subscriptΔ𝐯𝜈subscriptnormsubscript𝚺𝐯𝜈1𝐵1subscript𝐒𝐯superscript𝜈12𝐙𝜈2\displaystyle\Delta_{\mathbf{v}}(\nu)=\left\|\boldsymbol{\Sigma}_{\mathbf{v}}(\nu)-\frac{1}{\sqrt{B+1}}\mathbf{S}_{\mathbf{v}}(\nu)^{1/2}\mathbf{Z}(\nu)\right\|_{2}.

Using the fact that for any two matrices 𝐀,𝐁𝐀𝐁\mathbf{A},\mathbf{B} of appropriate dimensions, we have

𝐀𝐀𝐁𝐁=(𝐀𝐁)(𝐀𝐁)+(𝐀𝐁)𝐁+𝐁(𝐀𝐁)superscript𝐀𝐀superscript𝐁𝐁𝐀𝐁superscript𝐀𝐁𝐀𝐁superscript𝐁𝐁superscript𝐀𝐁\mathbf{A}\mathbf{A}^{*}-\mathbf{B}\mathbf{B}^{*}=(\mathbf{A}-\mathbf{B})(\mathbf{A}-\mathbf{B})^{*}+(\mathbf{A}-\mathbf{B})\mathbf{B}^{*}+\mathbf{B}(\mathbf{A}-\mathbf{B})^{*}

and

𝐀𝐁2𝐀2𝐁2subscriptnorm𝐀𝐁2subscriptnorm𝐀2subscriptnorm𝐁2\|\mathbf{A}\mathbf{B}\|_{2}\leq\|\mathbf{A}\|_{2}\|\mathbf{B}\|_{2}

we see that

𝐒^𝐯(ν)1B+1𝐒𝐯(ν)1/2𝐙(ν)𝐙(ν)𝐒𝐯(ν)1/22Δ𝐯(ν)(Δ𝐯(ν)+2𝐒𝐯(ν)2B+1𝐙(ν)2).subscriptnormsubscript^𝐒𝐯𝜈1𝐵1subscript𝐒𝐯superscript𝜈12𝐙𝜈𝐙superscript𝜈subscript𝐒𝐯superscript𝜈122subscriptΔ𝐯𝜈subscriptΔ𝐯𝜈2subscriptnormsubscript𝐒𝐯𝜈2𝐵1subscriptnorm𝐙𝜈2\displaystyle\left\|\hat{\mathbf{S}}_{\mathbf{v}}(\nu)-\frac{1}{B+1}\mathbf{S}_{\mathbf{v}}(\nu)^{1/2}\mathbf{Z}(\nu)\mathbf{Z}(\nu)^{*}\mathbf{S}_{\mathbf{v}}(\nu)^{1/2}\right\|_{2}\leq\Delta_{\mathbf{v}}(\nu)\left(\Delta_{\mathbf{v}}(\nu)+2\sqrt{\frac{\|\mathbf{S}_{\mathbf{v}}(\nu)\|_{2}}{B+1}}\|\mathbf{Z}(\nu)\|_{2}\right).

Assumption 1 implies that

supM1maxν[0,1]𝐒𝐯(ν)2<subscriptsupremum𝑀1subscript𝜈01subscriptnormsubscript𝐒𝐯𝜈2\displaystyle\sup_{M\geq 1}\max_{\nu\in[0,1]}\|\mathbf{S}_{\mathbf{v}}(\nu)\|_{2}<\infty

while from Lemma 2 from Appendix A, since 𝐙(ν)𝐙𝜈\mathbf{Z}(\nu) has i.i.d. complex Gaussian entries,

lim supMmaxν𝒱N𝐙(ν)2B+1<subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscriptnorm𝐙𝜈2𝐵1\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\frac{\left\|\mathbf{Z}(\nu)\right\|_{2}}{\sqrt{B+1}}<\infty (38)

with probability one. This concludes the proof of Corollary 1.

D-B Proof of Corollary 2

The proof of Corollary 2 is similar to the one of Corollary 1. Denoting Δ𝐮(ν)=𝚺𝐮(ν)𝐇(ν)𝚺ϵ(ν)2subscriptΔ𝐮𝜈subscriptnormsubscript𝚺𝐮𝜈𝐇𝜈subscript𝚺bold-italic-ϵ𝜈2\Delta_{\mathbf{u}}(\nu)=\left\|\boldsymbol{\Sigma}_{\mathbf{u}}(\nu)-\mathbf{H}(\nu)\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)\right\|_{2}, and noticing that supM1maxν[0,1]𝐇(ν)2<subscriptsupremum𝑀1subscript𝜈01subscriptnorm𝐇𝜈2\sup_{M\geq 1}\max_{\nu\in[0,1]}\left\|\mathbf{H}(\nu)\right\|_{2}<\infty from Assumption 3 eq. (11), we obtain that

maxν𝒱N𝐒^𝐮(ν)𝐇(ν)𝚺ϵ(ν)𝚺ϵ(ν)𝐇(ν)2maxν𝒱NΔ𝐮(ν)(Δ𝐮(ν)+2𝐇(ν)2𝚺ϵ(ν)2),Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{u}}(\nu)-\mathbf{H}(\nu)\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)^{*}\mathbf{H}(\nu)^{*}\right\|_{2}\leq\max_{\nu\in\mathcal{V}_{N}}\Delta_{\mathbf{u}}(\nu)\left(\Delta_{\mathbf{u}}(\nu)+2\left\|\mathbf{H}(\nu)\right\|_{2}\left\|\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)\right\|_{2}\right),\xrightarrow[M\to\infty]{a.s.}0.

Since K𝐾K is fixed with respect to M𝑀M from Assumption 4, we also have

maxν𝒱N𝚺ϵ(ν)𝚺ϵ(ν)𝐈M2Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)^{*}-\mathbf{I}_{M}\right\|_{2}\xrightarrow[M\to\infty]{a.s.}0 (39)

using Lemma 1, which proves Corollary 2.

D-C Proof of Proposition 1

To prove Proposition 1, let us write

𝐘(ν)=𝐇(ν)𝚺ϵ(ν)+𝐒𝐯(ν)1/2𝐙(ν)B+1.𝐘𝜈𝐇𝜈subscript𝚺bold-italic-ϵ𝜈subscript𝐒𝐯superscript𝜈12𝐙𝜈𝐵1\displaystyle\mathbf{Y}(\nu)=\mathbf{H}(\nu)\boldsymbol{\Sigma}_{\boldsymbol{\epsilon}}(\nu)+\frac{\mathbf{S}_{\mathbf{v}}(\nu)^{1/2}\mathbf{Z}(\nu)}{\sqrt{B+1}}.

Then, from (38) and (39), we have

lim supMmaxν𝒱N𝐘(ν)2<subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscriptnorm𝐘𝜈2\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\left\|\mathbf{Y}(\nu)\right\|_{2}<\infty

with probability one, which implies the following convergence

maxν𝒱N𝐒^𝐲(ν)𝐘(ν)𝐘(ν)2maxν𝒱N(Δ𝐮(ν)+Δ𝐯(ν))(Δ𝐮(ν)+Δ𝐯(ν)+2𝐘(ν)2)Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left\|\hat{\mathbf{S}}_{\mathbf{y}}(\nu)-\mathbf{Y}(\nu)\mathbf{Y}(\nu)^{*}\right\|_{2}\leq\max_{\nu\in\mathcal{V}_{N}}\left(\Delta_{\mathbf{u}}(\nu)+\Delta_{\mathbf{v}}(\nu)\right)\left(\Delta_{\mathbf{u}}(\nu)+\Delta_{\mathbf{v}}(\nu)+2\left\|\mathbf{Y}(\nu)\right\|_{2}\right)\xrightarrow[M\to\infty]{a.s.}0.

Finally, since the columns of B+1𝐘(ν)𝐵1𝐘𝜈\sqrt{B+1}\mathbf{Y}(\nu) are i.i.d. 𝒩M(𝟎,𝐒𝐲(ν))subscript𝒩superscript𝑀0subscript𝐒𝐲𝜈\mathcal{N}_{\mathbb{C}^{M}}(\mathbf{0},\mathbf{S}_{\mathbf{y}}(\nu)) with 𝐒𝐲(ν)=𝐇(ν)𝐇(ν)+𝐒𝐯(ν)subscript𝐒𝐲𝜈𝐇𝜈𝐇superscript𝜈subscript𝐒𝐯𝜈\mathbf{S}_{\mathbf{y}}(\nu)=\mathbf{H}(\nu)\mathbf{H}(\nu)^{*}+\mathbf{S}_{\mathbf{v}}(\nu), it follows that

𝐘(ν)=𝐒𝐲(ν)1/2𝐗(ν)B+1𝐘𝜈subscript𝐒𝐲superscript𝜈12𝐗𝜈𝐵1\displaystyle\mathbf{Y}(\nu)=\mathbf{S}_{\mathbf{y}}(\nu)^{1/2}\frac{\mathbf{X}(\nu)}{\sqrt{B+1}}

for some M×(B+1)𝑀𝐵1M\times(B+1) matrix 𝐗(ν)𝐗𝜈\mathbf{X}(\nu) having i.i.d. 𝒩(0,1)subscript𝒩01\mathcal{N}_{\mathbb{C}}(0,1) entries and the proof of Proposition 1 is complete.

Appendix E Proof of Corollary 3

We first prove that all the eigenvalues of the SCM asymptotically concentrate in a compact set with probability one for all large M𝑀M. Indeed, considering matrix

𝐖(ν)=𝚵(ν)12𝐗(ν)𝐗(ν)B+1𝚵(ν)12𝐖𝜈𝚵superscript𝜈12𝐗𝜈𝐗superscript𝜈𝐵1𝚵superscript𝜈12\mathbf{W}(\nu)=\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\boldsymbol{\Xi}(\nu)^{\frac{1}{2}}

defined through Theorem 3 and using Lemma 2 in conjunction with Borel-Cantelli lemma, we deduce that there exists constants C1,C2subscript𝐶1subscript𝐶2C_{1},C_{2} such that

lim infMminν𝒱NλM(𝐖(ν))C1(1c)2subscriptlimit-infimum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆𝑀𝐖𝜈subscript𝐶1superscript1𝑐2\displaystyle\liminf_{M\to\infty}\min_{\nu\in\mathcal{V}_{N}}\lambda_{M}\left(\mathbf{W}(\nu)\right)\geq C_{1}(1-\sqrt{c})^{2}

and

lim supMmaxν𝒱Nλ1(𝐖(ν))C2(1+c)2subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆1𝐖𝜈subscript𝐶2superscript1𝑐2\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\mathbf{W}(\nu)\right)\leq C_{2}(1+\sqrt{c})^{2} (40)

with probability one, where C1,C2subscript𝐶1subscript𝐶2C_{1},C_{2} verify, thanks to Assumption 3,

0<C1<1=infM1minν𝒱NλM(𝚵(ν))0subscript𝐶11subscriptinfimum𝑀1subscript𝜈subscript𝒱𝑁subscript𝜆𝑀𝚵𝜈\displaystyle 0<C_{1}<1=\inf_{M\geq 1}\min_{\nu\in\mathcal{V}_{N}}\lambda_{M}\left(\boldsymbol{\Xi}(\nu)\right)

and

supM1maxν𝒱NλM(𝚵(ν))<C2<.subscriptsupremum𝑀1subscript𝜈subscript𝒱𝑁subscript𝜆𝑀𝚵𝜈subscript𝐶2\displaystyle\sup_{M\geq 1}\max_{\nu\in\mathcal{V}_{N}}\lambda_{M}\left(\boldsymbol{\Xi}(\nu)\right)<C_{2}<\infty.

Using (22), we obtain similarly

lim infMminν𝒱NλM(𝐂^𝐲(ν))C1(1c)2subscriptlimit-infimum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆𝑀subscript^𝐂𝐲𝜈subscript𝐶1superscript1𝑐2\liminf_{M\to\infty}\min_{\nu\in\mathcal{V}_{N}}\lambda_{M}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\geq C_{1}(1-\sqrt{c})^{2}

and

lim supMmaxν𝒱Nλ1(𝐂^𝐲(ν))C2(1+c)2subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆1subscript^𝐂𝐲𝜈subscript𝐶2superscript1𝑐2\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\leq C_{2}(1+\sqrt{c})^{2}

with probability one. Let 0<ϵ<C12(1c)20italic-ϵsubscript𝐶12superscript1𝑐20<\epsilon<\frac{C_{1}}{2}(1-\sqrt{c})^{2} and h𝒞c1()superscriptsubscript𝒞𝑐1h\in\mathcal{C}_{c}^{1}(\mathbb{R}) such that

h(λ)={1 if λ[C1(1c)2ϵ,C2(1+c)2+ϵ]0 if λ[C1(1c)22ϵ,C2(1+c)2+2ϵ].𝜆cases1 if 𝜆subscript𝐶1superscript1𝑐2italic-ϵsubscript𝐶2superscript1𝑐2italic-ϵ0 if 𝜆subscript𝐶1superscript1𝑐22italic-ϵsubscript𝐶2superscript1𝑐22italic-ϵ\displaystyle h(\lambda)=\begin{cases}1&\text{ if }\lambda\in\left[C_{1}(1-\sqrt{c})^{2}-\epsilon,C_{2}(1+\sqrt{c})^{2}+\epsilon\right]\\ 0&\text{ if }\lambda\not\in\left[C_{1}(1-\sqrt{c})^{2}-2\epsilon,C_{2}(1+\sqrt{c})^{2}+2\epsilon\right]\end{cases}.

Then it follows that

maxν𝒱N|Lφ(ν)Lφh(ν)|Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left|L_{\varphi}(\nu)-L_{\varphi h}(\nu)\right|\xrightarrow[M\to\infty]{a.s.}0.

Thus, without loss of generality, we may assume for the remainder of the proof that φ𝒞c1((0,+))𝜑subscriptsuperscript𝒞1𝑐0\varphi\in\mathcal{C}^{1}_{c}\left((0,+\infty)\right). Using (22), we deduce that

maxν𝒱N1Mm=1M|φ(λm(𝐂^𝐲(ν)))φ(λm(𝐖(ν)))|Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\frac{1}{M}\sum_{m=1}^{M}\left|\varphi\left(\lambda_{m}\left(\hat{\mathbf{C}}_{\mathbf{y}}(\nu)\right)\right)-\varphi\left(\lambda_{m}\left(\mathbf{W}(\nu)\right)\right)\right|\xrightarrow[M\to\infty]{a.s.}0.

Next, consider the two functions

m^(z,ν)=1Mm=1M1λm(𝐖(ν))z=dμ^(λ,ν)λz^𝑚𝑧𝜈1𝑀superscriptsubscript𝑚1𝑀1subscript𝜆𝑚𝐖𝜈𝑧subscriptd^𝜇𝜆𝜈𝜆𝑧\displaystyle\hat{m}(z,\nu)=\frac{1}{M}\sum_{m=1}^{M}\frac{1}{\lambda_{m}\left(\mathbf{W}(\nu)\right)-z}=\int_{\mathbb{R}}\frac{\mathrm{d}\hat{\mu}(\lambda,\nu)}{\lambda-z}

and

m~(z,ν)=1Mm=1M1λm(𝐗(ν)𝐗(ν)B+1)z=dμ~(λ,ν)λz~𝑚𝑧𝜈1𝑀superscriptsubscript𝑚1𝑀1subscript𝜆𝑚𝐗𝜈𝐗superscript𝜈𝐵1𝑧subscriptd~𝜇𝜆𝜈𝜆𝑧\displaystyle\tilde{m}(z,\nu)=\frac{1}{M}\sum_{m=1}^{M}\frac{1}{\lambda_{m}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)-z}=\int_{\mathbb{R}}\frac{\mathrm{d}\tilde{\mu}(\lambda,\nu)}{\lambda-z}

defined for all z+:={ζ:Im(ζ)>0}𝑧superscriptassignconditional-set𝜁Im𝜁0z\in\mathbb{C}^{+}:=\{\zeta\in\mathbb{C}:\mathrm{Im}(\zeta)>0\}, and where for all Borel set A𝐴A\subset\mathbb{R},

μ^(A,ν)=1Mm=1Mδλm(𝐖(ν))(A)^𝜇𝐴𝜈1𝑀superscriptsubscript𝑚1𝑀subscript𝛿subscript𝜆𝑚𝐖𝜈𝐴\displaystyle\hat{\mu}(A,\nu)=\frac{1}{M}\sum_{m=1}^{M}\delta_{\lambda_{m}\left(\mathbf{W}(\nu)\right)}(A)

and

μ~(A,ν)=1Mm=1Mδλm(𝐗(ν)𝐗(ν)B+1)(A)~𝜇𝐴𝜈1𝑀superscriptsubscript𝑚1𝑀subscript𝛿subscript𝜆𝑚𝐗𝜈𝐗superscript𝜈𝐵1𝐴\displaystyle\tilde{\mu}(A,\nu)=\frac{1}{M}\sum_{m=1}^{M}\delta_{\lambda_{m}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)}(A)

denote the empirical eigenvalue distributions of matrices 𝐖(ν)𝐖𝜈\mathbf{W}(\nu) and 𝐗(ν)𝐗(ν)B+1𝐗𝜈𝐗superscript𝜈𝐵1\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1} respectively, and δxsubscript𝛿𝑥\delta_{x} is the Dirac measure at point x𝑥x. Functions zm^(z,ν)maps-to𝑧^𝑚𝑧𝜈z\mapsto\hat{m}(z,\nu) and zm~(z,ν)maps-to𝑧~𝑚𝑧𝜈z\mapsto\tilde{m}(z,\nu) coincide with the Stieltjes transforms of measures μ^(.,ν)\hat{\mu}(.,\nu) and μ~(.,ν)\tilde{\mu}(.,\nu) respectively (see [32] for a review of the main properties of the Stieltjes transform). Since

m^(z,ν)m~(z,ν)=1MTr((𝐖(ν)z𝐈)1(𝐗(ν)𝐗(ν)B+1z𝐈)1)^𝑚𝑧𝜈~𝑚𝑧𝜈1𝑀Trsuperscript𝐖𝜈𝑧𝐈1superscript𝐗𝜈𝐗superscript𝜈𝐵1𝑧𝐈1\displaystyle\hat{m}(z,\nu)-\tilde{m}(z,\nu)=\frac{1}{M}\mathrm{Tr}\,\left(\left(\mathbf{W}(\nu)-z\mathbf{I}\right)^{-1}-\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}-z\mathbf{I}\right)^{-1}\right)

and using the fact that 𝐀1𝐁1=𝐀1(𝐁𝐀)𝐁1superscript𝐀1superscript𝐁1superscript𝐀1𝐁𝐀superscript𝐁1\mathbf{A}^{-1}-\mathbf{B}^{-1}=\mathbf{A}^{-1}(\mathbf{B}-\mathbf{A})\mathbf{B}^{-1} for non-singular matrices 𝐀,𝐁𝐀𝐁\mathbf{A},\mathbf{B}, we have

|m^(z,ν)m~(z,ν)|1|Im(z)|2KM𝐇(ν)22𝐒𝐯(ν)12𝐗(ν)𝐗(ν)B+12^𝑚𝑧𝜈~𝑚𝑧𝜈1superscriptIm𝑧2𝐾𝑀superscriptsubscriptnorm𝐇𝜈22subscriptnormsubscript𝐒𝐯superscript𝜈12subscriptnorm𝐗𝜈𝐗superscript𝜈𝐵12\displaystyle\left|\hat{m}(z,\nu)-\tilde{m}(z,\nu)\right|\leq\frac{1}{|\mathrm{Im}(z)|^{2}}\frac{K}{M}\left\|\mathbf{H}(\nu)\right\|_{2}^{2}\left\|\mathbf{S}_{\mathbf{v}}(\nu)^{-1}\right\|_{2}\left\|\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right\|_{2}

it follows from Assumptions 2, 3, 4 and Lemma 2 that

maxν𝒱N|m^(z,ν)m~(z,ν)|Ma.s.0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left|\hat{m}(z,\nu)-\tilde{m}(z,\nu)\right|\xrightarrow[M\to\infty]{a.s.}0 (41)

for all z+𝑧superscriptz\in\mathbb{C}^{+}. In the following, we fix a realization in an event of probability one for which (41) holds for all z+𝑧superscriptz\in\mathbb{C}^{+} and consider

νargmaxν𝒱N|φ(λ)dμ^(λ,ν)φ(λ)dμ~(λ,ν)|.superscript𝜈subscriptargmax𝜈subscript𝒱𝑁subscript𝜑𝜆differential-d^𝜇𝜆𝜈subscript𝜑𝜆differential-d~𝜇𝜆𝜈\displaystyle\nu^{*}\in\operatorname*{argmax}_{\nu\in\mathcal{V}_{N}}\left|\int_{\mathbb{R}}\varphi(\lambda)\mathrm{d}\hat{\mu}(\lambda,\nu)-\int_{\mathbb{R}}\varphi(\lambda)\mathrm{d}\tilde{\mu}(\lambda,\nu)\right|.

Then |m^(z,ν)m~(z,ν)|0^𝑚𝑧superscript𝜈~𝑚𝑧superscript𝜈0\left|\hat{m}(z,\nu^{*})-\tilde{m}(z,\nu^{*})\right|\to 0 as M𝑀M\to\infty, for all z+𝑧superscriptz\in\mathbb{C}^{+}. From the fact that the pointwise convergence on +superscript\mathbb{C}^{+} of a sequence of Stieltjes transforms is equivalent to the weak convergence of the related sequence of probability measures (see e.g. [32, Ex.2.4.10]), we deduce that

maxν𝒱N|1Mm=1M(φ(λm(𝐖(ν)))φ(λm(𝐗(ν)𝐗(ν)B+1)))|=|φ(λ)dμ^(λ,ν)φ(λ)dμ~(λ,ν)|M0.subscript𝜈subscript𝒱𝑁1𝑀superscriptsubscript𝑚1𝑀𝜑subscript𝜆𝑚𝐖𝜈𝜑subscript𝜆𝑚𝐗𝜈𝐗superscript𝜈𝐵1subscript𝜑𝜆differential-d^𝜇𝜆superscript𝜈subscript𝜑𝜆differential-d~𝜇𝜆superscript𝜈𝑀absent0\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left|\frac{1}{M}\sum_{m=1}^{M}\left(\varphi\left(\lambda_{m}\left(\mathbf{W}(\nu)\right)\right)-\varphi\left(\lambda_{m}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)\right)\right)\right|=\left|\int_{\mathbb{R}}\varphi(\lambda)\mathrm{d}\hat{\mu}(\lambda,\nu^{*})-\int_{\mathbb{R}}\varphi(\lambda)\mathrm{d}\tilde{\mu}(\lambda,\nu^{*})\right|\xrightarrow[M\to\infty]{}0.

To conclude the proof of Corollary 3, it remains to prove that

maxν𝒱N|1Mm=1Mφ(λm(𝐗(ν)𝐗(ν)B+1))φ(λ)f(λ)dλ|Ma.s.0.\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left|\frac{1}{M}\sum_{m=1}^{M}\varphi\left(\lambda_{m}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)\right)-\int_{\mathbb{R}}\varphi(\lambda)f(\lambda)\mathrm{d}\lambda\right|\xrightarrow[M\to\infty]{a.s.}0.

Consider the decomposition

maxν𝒱N|1Mm=1Mφ(λm(𝐗(ν)𝐗(ν)B+1))φ(λ)f(λ)dλ|Δ1+Δ2,subscript𝜈subscript𝒱𝑁1𝑀superscriptsubscript𝑚1𝑀𝜑subscript𝜆𝑚𝐗𝜈𝐗superscript𝜈𝐵1subscript𝜑𝜆𝑓𝜆differential-d𝜆subscriptΔ1subscriptΔ2\displaystyle\max_{\nu\in\mathcal{V}_{N}}\left|\frac{1}{M}\sum_{m=1}^{M}\varphi\left(\lambda_{m}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)\right)-\int_{\mathbb{R}}\varphi(\lambda)f(\lambda)\mathrm{d}\lambda\right|\leq\Delta_{1}+\Delta_{2},

where

Δ1=maxν𝒱N|1Mm=1M(φ(λm(𝐗(ν)𝐗(ν)B+1))𝔼[φ(λm(𝐗(ν)𝐗(ν)B+1))])|subscriptΔ1subscript𝜈subscript𝒱𝑁1𝑀superscriptsubscript𝑚1𝑀𝜑subscript𝜆𝑚𝐗𝜈𝐗superscript𝜈𝐵1𝔼delimited-[]𝜑subscript𝜆𝑚𝐗𝜈𝐗superscript𝜈𝐵1\displaystyle\Delta_{1}=\max_{\nu\in\mathcal{V}_{N}}\Biggl{|}\frac{1}{M}\sum_{m=1}^{M}\Biggl{(}\varphi\left(\lambda_{m}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)\right)-\mathbb{E}\left[\varphi\left(\lambda_{m}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)\right)\right]\Biggr{)}\Biggr{|}

and

Δ2=|1Mm=1M𝔼[φ(λm(𝐗(0)𝐗(0)B+1))]φ(λ)f(λ)dλ|.subscriptΔ21𝑀superscriptsubscript𝑚1𝑀𝔼delimited-[]𝜑subscript𝜆𝑚𝐗0𝐗superscript0𝐵1subscript𝜑𝜆𝑓𝜆differential-d𝜆\displaystyle\Delta_{2}=\left|\frac{1}{M}\sum_{m=1}^{M}\mathbb{E}\left[\varphi\left(\lambda_{m}\left(\frac{\mathbf{X}(0)\mathbf{X}(0)^{*}}{B+1}\right)\right)\right]-\int_{\mathbb{R}}\varphi(\lambda)f(\lambda)\mathrm{d}\lambda\right|.

Using the concentration inequality of [33, Cor. 1.8(b)], it is straightforward to show that

Δ1Ma.s.0.\displaystyle\Delta_{1}\xrightarrow[M\to\infty]{a.s.}0.

Moreover, using again the properties of the Stieltjes transform, it can be deduced from e.g. [34] that

Δ2M0.𝑀absentsubscriptΔ20\displaystyle\Delta_{2}\xrightarrow[M\to\infty]{}0.

This concludes the proof of Corollary 3.

Appendix F Proof of Proposition 2

Convergences (25), (26) and (27) are straightforward consequences of (22) and the results of [20, Th. 1.1] on the behaviour of the largest eigenvalues for the so-called multiplicative spike model random matrices. To prove (28), we use the bound

λ1(𝐖(ν))λ1(𝐗(ν)𝐗(ν)B+1)λ1(𝚵(ν))subscript𝜆1𝐖𝜈subscript𝜆1𝐗𝜈𝐗superscript𝜈𝐵1subscript𝜆1𝚵𝜈\displaystyle\lambda_{1}\left(\mathbf{W}(\nu)\right)\leq\lambda_{1}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)\lambda_{1}\left(\boldsymbol{\Xi}(\nu)\right)

Then, from the fact that γ=0subscript𝛾0\gamma_{\infty}=0 and Lemma 2, we finally obtain

lim supMmaxν𝒱Nλ1(𝐖(ν))subscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆1𝐖𝜈\displaystyle\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\mathbf{W}(\nu)\right) lim supMmaxν𝒱Nλ1(𝐗(ν)𝐗(ν)B+1)absentsubscriptlimit-supremum𝑀subscript𝜈subscript𝒱𝑁subscript𝜆1𝐗𝜈𝐗superscript𝜈𝐵1\displaystyle\leq\limsup_{M\to\infty}\max_{\nu\in\mathcal{V}_{N}}\lambda_{1}\left(\frac{\mathbf{X}(\nu)\mathbf{X}(\nu)^{*}}{B+1}\right)
(1+c)2.absentsuperscript1𝑐2\displaystyle\leq\left(1+\sqrt{c}\right)^{2}.

The proof is concluded by invoking again convergence (22).