Joint Jammer Mitigation and Data Detection


Abstract

Multi-antenna (or MIMO) processing is a promising solution to the problem of jammer mitigation. Existing methods mitigate the jammer based on an estimate of its spatial signature that is acquired through a dedicated training phase. This strategy has two main drawbacks: (i) it reduces the communication rate since no data can be transmitted during the training phase and (ii) it can be evaded by smart or multi-antenna jammers that do not transmit during the training phase or that dynamically change their subspace through time-varying beamforming. To address these drawbacks, we propose Joint jammer Mitigation and data Detection (JMD), a novel paradigm for MIMO jammer mitigation. The core idea of JMD is to estimate and remove the jammer interference subspace jointly with detecting the legitimate transmit data over multiple time slots. Doing so removes the need for a dedicated and rate-reducing training period while being able to mitigate smart and dynamic multi-antenna jammers. We provide two JMD-type algorithms, SANDMAN and MAED, that differ in the way they estimate the channels of the legitimate transmitters and achieve different complexity-performance tradeoffs. Extensive simulations demonstrate the efficacy of JMD for jammer mitigation.

Jammer mitigation, joint jammer mitigation and data detection (JMD), MAED, MIMO, SANDMAN,

1 Introduction↩︎

Averting the threat of jamming attacks on wireless communication systems is a problem of vital importance as society becomes ever more reliant on wireless communication infrastructure [2][5]. Multi-antenna (or MIMO) processing offers an attractive and effective solution by enabling the mitigation of jammers through spatial filtering [6]. In MIMO-based jammer mitigation, the jammer’s interference is often removed by projecting the receive signal onto the orthogonal complement of the jammer channel’s subspace, which traditionally is estimated in a dedicated training period [7][14]. Such an approach has two main disadvantages: First, a dedicated training period for estimating the jammer’s channel reduces the achievable data rate, since no information can be transmitted during the training period. Second, a training-period-based approach is ineffective against smart jammers that transmit only at specific instances to evade estimation [15] or against dynamic jammers that change their subspace through time-varying multi-antenna transmit beamforming [16]. In order to avoid both of these limitations, we propose a novel paradigm for jammer mitigation with MIMO processing that we call joint jammer mitigation and data detection (JMD).

1.1 State of the Art↩︎

The fundamental challenge of jammer mitigation through spatial filtering is that it requires information about the jammer, such as the subspace spanned by the jammer’s channel [7][10] or the covariance matrix of the jammer’s interference [8], [17], [18]. Existing methods often assume that the jammer transmits permanently and with static beamforming. Such a behavior would enable one to estimate the required quantities during a dedicated training period in which the legitimate transmitters do not transmit [7], [8], [18] or in which they transmit predetermined symbols (pilots) that carry no information [9][13]. Relying on such a static jammer assumption, the receiver can filter the jammer in the subsequent communication period until the wireless channel changes and the jammer training process is started anew. A smart jammer, however, can easily circumvent such mitigation methods by deliberately violating their assumptions: It can pause jamming for the duration of the dedicated training period, so that the receiver learns nothing meaningful [15]. Or, if the jammer has multiple antennas, it can use beamforming to dynamically change its subspace (as well as the interference covariance matrix at the receiver), so that after the training period, the receiver’s filter will no longer match the jammer’s transmit characteristics [16], [19].4

In order to mitigate smart jammers, methods have been suggested that attempt to fool the jammer into transmitting during the training period by distributing and randomizing the timing of the training period [7], [9]. However, such methods may not work against jammers that jam only intermittently at random time instants. Similarly, methods have been proposed to mitigate dynamic multi-antenna jammers by perpetually estimating their instantaneous subspace [16]. However, such methods may be effective only against jammers that change their subspace in a sufficiently slow manner. Furthermore, all training-period based mitigation methods are subject to an inherent trade-off between the time that they dedicate to the attempt of estimating the required jammer characteristics and the time that remains for payload data transmission.

In light of these considerations, a more principled approach to MIMO-based jammer mitigation is needed. We have recently proposed MAED [15], which, in hindsight, can be viewed as the first instance of our JMD paradigm. MAED unifies not only jammer mitigation and data detection, but also channel estimation (at the cost of increased computational complexity).

1.2 Contributions↩︎

We propose joint jammer mitigation and data detection (JMD), a novel paradigm for jammer mitigation. The core idea is to estimate and remove the subspace of the jammer interference jointly with detecting the data of an entire transmission frame (or coherence interval). JMD removes the need for a dedicated training period, which enables higher data rates because more time is available for data transmission. Beyond that, considering an entire transmission frame at once enables JMD to deal with smart jammers that try to evade mitigation (i) by jamming only at specific instances or (ii) by dynamically changing their subspace through multi-antenna beamforming.

As in [15], we leverage the fact that a jammer cannot leave its subspace within a coherence interval (which holds also for multi-antenna jammers with time-varying beamforming). Going beyond our work in [15], we show that this fact can be exploited to mitigate the jammer’s contamination of the channel estimate even when the channel is estimated independently of the subsequent jammer mitigation and data detection. This insight opens the door to computationally more efficient signal-processing algorithms for jammer mitigation.5 Moreover, our earlier work in [15] only mitigates single-antenna jammers, while this paper shows that JMD is also well suited for the mitigation of multi-antenna jammers—even dynamic ones. We capitalize on these new insights by proposing the JMD-type algorithm SANDMAN (an algorithm that offers all of the jammer mitigation capabilities of MAED [15], but at reduced computational complexity), and by extending MAED to multi-antenna jammers. Furthermore, we present theoretical guarantees for JMD that are stronger than those in [15] and that rely on weaker assumptions. Extensive simulations demonstrate the efficacy of our JMD-type algorithms SANDMAN and MAED for mitigating a wide range of smart and dynamic single- and multi-antenna jammers. This paper therefore provides a full and mature account of the JMD paradigm, while the earlier work in [15] should (in hindsight) be regarded as a precursor that did not yet grasp the full breadth and depth of the JMD paradigm. Compared to the conference version [1], this paper adds most of the theoretical results (and all of the proofs), more comprehensive discussion, and more extensive evaluation.

1.3 Notation↩︎

Matrices and column vectors are represented by boldface uppercase and lowercase letters, respectively. For a matrix \(\mathbf{A}\), the transpose is \(\mathbf{A}^{\text{T}}\), the conjugate transpose is \(\mathbf{A}^{\text{H}}\), the Moore-Penrose pseudoinverse is \(\mathbf{A}^{\dagger}\) (if \(\mathbf{A}\) is full-rank and tall, then \(\mathbf{A}^{\dagger}=(\mathbf{A}^{\text{H}}\mathbf{A})^{-1}\mathbf{A}^{\text{H}}\), and if \(\mathbf{A}\) is full-rank and wide, then \(\mathbf{A}^{\dagger}=\mathbf{A}^{\text{H}}(\mathbf{A}\mathbf{A}^{\text{H}})^{-1}\)), and the Frobenius norm is \(\| \mathbf{A}\|_F\). The columnspace and rowspace of \(\mathbf{A}\) are \(\textit{col}(\mathbf{A})\) and \(\textit{row}(\mathbf{A})\), respectively, and their orthogonal complements are \(\textit{col}(\mathbf{A})^\perp\) and \(\text{row}(\mathbf{A})^\perp\). Horizontal concatenation of two matrices \(\mathbf{A}\) and \(\mathbf{B}\) is denoted by \([\mathbf{A},\mathbf{B}]\); vertical concatenation by \([\mathbf{A};\mathbf{B}]\). The \(N\!\times\!N\) identity matrix is \(\mathbf{I}_N\). The \(\ell_2\)-norm of a vector \(\mathbf{a}\) is \(\|\mathbf{a}\|_2\), and the \(\ell_0\)-“norm” \(\|\mathbf{a}\|_0\) indicates the number of nonzero entries of \(\mathbf{a}\). We use \(\mathcal{C}\mathcal{N}(\mathbf{0},\mathbf{C})\) to denote the \(N\)-dimensional circularly-symmetric complex Gaussian distribution with covariance matrix \(\mathbf{C}\in\mathbb{C}^{N\times N}\). Finally, \([1:N]\) denotes the set of integers from \(1\) through \(N\).

2 Transmission Model↩︎

We focus on mitigating jamming attacks in the flat-fading, block-fading multi-user (MU) MIMO uplink.6 We consider the following transmission model: \[\begin{align} \mathbf{y}_k = \mathbf{H}\mathbf{s}_k + \mathbf{J}\mathbf{w}_k + \mathbf{n}_k. \label{eq:model} \end{align}\tag{1}\] Here, \(\mathbf{y}_k\in\mathbb{C}^B\) is the BS receive vector at sample instant \(k\), \(\mathbf{H}\in \mathbb{C}^{B\times U}\) is the UE channel matrix, \(\mathbf{s}_k\in\mathcal{S}^U\) contains the sample-\(k\) transmit symbols of \(U\) single-antenna UEs with constellation \(\mathcal{S}\) and is assumed to be uniformly distributed over \(\mathcal{S}^U\), \(\mathbf{J}\in\mathbb{C}^{B\times I}\) is the channel matrix of an \(I\)-antenna jammer,7 \(\mathbf{w}_k\in\mathbb{C}^{I}\) is the sample-\(k\) jammer transmit vector, and \(\mathbf{n}_k\sim\mathcal{C}\mathcal{N}(\mathbf{0},N_0\mathbf{I}_B)\) is circularly-symmetric complex Gaussian noise with per-entry variance \(N_0\). We assume that \(B\geq U+I\), i.e., the number of receive antennas is greater than or equal to the number of antennas of the UEs and the jammer combined, and that the matrix \([\mathbf{H},\mathbf{J}]\) has full rank \(U+I\). For simplicity, we take \(\mathcal{S}\) to be QPSK, though other constellations would also be possible [15]. The constellation \(\mathcal{S}\) is scaled to unit symbol power, so that \(\mathop{\mathrm{\mathbb{E}}}_{}\mathopen{}\left[\mathbf{s}_k\mathbf{s}_k^{\text{H}}\right]=\mathbf{I}_U\), and factors related to power control are absorbed into the channel matrix \(\mathbf{H}\). In general, the jammer is a dynamic multi-antenna jammer and is able to dynamically change its jamming activity. Specifically, the jammer transmits \[\begin{align} \mathbf{w}_k = \mathbf{A}_k \tilde{\mathbf{w}}_k, \label{eq:jammer95beamforming} \end{align}\tag{2}\] where, without loss of generality, \(\mathop{\mathrm{\mathbb{E}}}_{}\mathopen{}\left[\tilde{\mathbf{w}}_k\tilde{\mathbf{w}}_k^{\text{H}}\right]=\mathbf{I}_I\) for all \(k\), and \(\mathbf{A}_k\in\mathbb{C}^{I\times I}\) is a beamforming matrix that can change arbitrarily over time (i.e., \(\mathbf{A}_k\) depends on the sample instant \(k\)). In particular, \(\mathbf{A}_k\) can be the all-zero matrix (no jamming at sample instant \(k\)), or some of its rows can be zero (the jammer uses only a subset of its antennas at sample instant \(k\)), or it can be rank-deficient in some other way.

In JMD, the receiver will process the receive signal in blocks consisting of \(K\) sample indices (or channel uses). The length \(K\) of these blocks, which we call frames, may be up to the length of a coherence interval, so that the channel matrices \(\mathbf{H}\) and \(\mathbf{J}\) can be assumed to be constant for the duration of a frame. It will be useful to define notation for the receive signal of an entire frame: \[\begin{align} \mathbf{Y}&= \mathbf{H}\mathbf{S}+ \mathbf{J}\mathbf{W}+ \mathbf{N}. \tag{3} \intertext{ Here, \mathbf{Y}=[\mathbf{y}_1,\dots,\mathbf{y}_K]\in\mathbb{C}^{B\times K} is the receive matrix, and the quantities \mathbf{S}\in\mathbb{C}^{U\times K}, \mathbf{W}\in\mathbb{C}^{I\times K}, and \mathbf{N}\in\mathbb{C}^{B\times K} are defined analogously. We divide each frame into a \emph{pilot phase} of length T\geq U and a \emph{data phase} of length D (such that T+D=K). The input-output relations for these phases are denoted by } \mathbf{Y}_T &= \mathbf{H}\mathbf{S}_T + \mathbf{J}\mathbf{W}_T + \mathbf{N}_T \tag{4} \intertext{ and } \mathbf{Y}\!_D &= \mathbf{H}\mathbf{S}_D + \mathbf{J}\mathbf{W}\!_D + \mathbf{N}_D, \tag{5} \end{align}\] respectively. Thus, we have \(\mathbf{Y}=[\mathbf{Y}_T,\mathbf{Y}\!_D]\), \(\mathbf{S}=[\mathbf{S}_T,\mathbf{S}_D]\) with \(\mathbf{S}_T\in\mathbb{C}^{U\times T}\) and \(\mathbf{S}_D\in\mathcal{S}^{U\times D}\), \(\mathbf{W}=[\mathbf{W}_T,\mathbf{W}\!_D]\), and \(\mathbf{N}=[\mathbf{N}_T,\mathbf{N}_D]\). Note that we do not restrict the pilots to the constellation \(\mathcal{S}\). However, we do impose an average power constraint \(\frac{1}{T}\|\mathbf{s}_T\|_2^2=1\) on all rows \(\mathbf{s}_T\) of \(\mathbf{S}_T\). We assume that the columns \(\mathbf{s}_k\) of \(\mathbf{S}_D\) are independent of each other, so that \(\mathbf{S}_D\sim\text{Unif}[\mathcal{S}^{U\times D}]\) (where \(\text{Unif}[\mathcal{A}]\) denotes the uniform distribution over the set \(\mathcal{A}\)). In what follows, we use plain letters for the true transmit signals and channels, letters with a tilde for optimization variables, and letters with a hat for (approximate) solutions to optimization problems. For example, \(\hat{\mathbf{S}}_D\) is an estimate of the data matrix \(\mathbf{S}_D\) obtained by solving an optimization problem with respect to a variable \(\tilde{\mathbf{S}}_D\).

3 Joint Jammer Mitigation and Data Detection↩︎

Existing methods for MIMO jammer mitigation often null a jammer by projecting the receive signal onto the orthogonal complement of the jammer subspace [8][10], [16]. That is, they compute \(\mathbf{P}\mathbf{y}_k\), where \(\mathbf{P}= \mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}\) is the orthogonal projection onto \(\textit{col}(\mathbf{J})^\perp\) and has the property \(\mathbf{P}\mathbf{J}=\mathbf{0}\). After this projection, the transmit data can be detected using the effective channel matrix \(\mathbf{H}_\mathbf{P}\triangleq \mathbf{P}\mathbf{H}\), since \[\begin{align} \mathbf{P}\mathbf{y}_k &= \mathbf{P}\mathbf{H}\mathbf{s}_k + \mathbf{P}\mathbf{J}\mathbf{A}_k \tilde{\mathbf{w}}_k + \mathbf{P}\mathbf{n}_k = \mathbf{P}\mathbf{H}\mathbf{s}_k + \mathbf{P}\mathbf{n}_k \\ &\triangleq \mathbf{H}_\mathbf{P}\mathbf{s}_k + \mathbf{n}_{\mathbf{P},k}. \label{eq:projected95model} \end{align}\tag{6}\] This strategy mitigates the jammer regardless of which vectors \(\tilde{\mathbf{w}}_k\) and which beamforming matrices \(\mathbf{A}_k\) the jammer uses (cf. 2 ), since \(\textit{col}(\mathbf{J}\mathbf{A}_k)\subseteq\textit{col}(\mathbf{J})\) for any \(\mathbf{A}_k\). The problem, however, is that of estimating \(\mathbf{J}\)—or \(\textit{col}(\mathbf{J})\)—when the jammer changes \(\mathbf{A}_k\) dynamically such that \(\textit{col}(\mathbf{J}\mathbf{A}_k)\) depends on \(k\). In that case, the receiver will observe different interference subspaces \(\textit{col}(\mathbf{J}\mathbf{A}_k)\) at different instants, but possibly never the “pure” jammer subspace \(\textit{col}(\mathbf{J})\).

To address this problem, we propose JMD. The key idea is to consider jammer subspace estimation and nulling as well as data detection over an entire transmission frame (or coherence interval) jointly. The jammer subspace is identified with the subspace that is not explainable in terms of UE transmit signals; simultaneously, the UE transmit data are estimated while projecting the receive signals onto the orthogonal complement of the jammer subspace estimate. Mathematically, this can be formulated as the following optimization problem:8 \[\begin{align} \min_{\substack{\tilde{\mathbf{S}}_D \,\in\, \mathcal{S}^{U\times D},\\ \tilde{\mathbf{P}}\,\in\, \mathscr{G}_{B-I}(\mathbb{C}^B)}} \| \tilde{\mathbf{P}}(\mathbf{Y}\!_D -\mathbf{H}\tilde{\mathbf{S}}_D) \|_F^2. \label{eq:jmd95problem} \end{align}\tag{7}\] Here, \(\tilde{\mathbf{S}}_D\) is the potential estimate of the data matrix \(\mathbf{S}_D\) and \(\tilde{\mathbf{P}}= \mathbf{I}_B - \tilde{\mathbf{J}}\tilde{\mathbf{J}}^{\dagger}\) is the projection onto the orthogonal complement of the potential jammer subspace estimate \(\textit{col}(\tilde{\mathbf{J}})\), with \(\tilde{\mathbf{J}}\in\mathbb{C}^{B\times I}\). The range of \(\tilde{\mathbf{P}}\) is the Grassmannian manifold \(\mathscr{G}_{B-I}(\mathbb{C}^B)=\{\mathbf{I}_B-\tilde{\mathbf{J}}\tilde{\mathbf{J}}^{\dagger}:\tilde{\mathbf{J}}\!\in\!\mathbb{C}^{B\times I}\}\), i.e., the set of orthogonal projections onto \((B\!-\!I)\)-dimensional subspaces of \(\mathbb{C}^B\). The intuition behind this problem formulation is as follows: Using 5 , we can rewrite the objective of 7 as \[\begin{align} \|\tilde{\mathbf{P}}( \mathbf{H}(\mathbf{S}_D-\tilde{\mathbf{S}}_D) + \mathbf{J}\mathbf{W}\!_D + \mathbf{N}_D)\|_F^2. \label{eq:pre95intuition} \end{align}\tag{8}\] If we assume that the thermal noise \(\mathbf{N}_D\) is negligible (\(\mathbf{N}_D=\mathbf{0}\)) compared to the legitimate signals \(\mathbf{H}\mathbf{S}_D\) and to the jammer interference \(\mathbf{J}\mathbf{W}\!_D\), then we may further simplify 8 to \[\begin{align} \textstyle \bigg\|\tilde{\mathbf{P}}\begin{bmatrix}\mathbf{H},\mathbf{J}\end{bmatrix} \begin{bmatrix}\mathbf{S}_D-\tilde{\mathbf{S}}_D \\ \mathbf{W}\!_D \end{bmatrix} \bigg\|_{F \,\displaystyle{.}}^2 \label{eq:intuition} \end{align}\tag{9}\] The justification for this low-noise assumption—which we make repeatedly throughout the paper—is that this paper’s focus is jammer mitigation, not data detection under high thermal noise. In other words, we are concerned with scenarios where communication performance is not primarily limited by thermal noise, but by jamming. The objective in 9 is minimized when the argument of the Frobenius norm is the all-zero matrix, and the projector \(\tilde{\mathbf{P}}\) can null a matrix of rank less than or equal to \(I\). If \(\mathbf{J}\mathbf{W}\!_D\) has rank \(I\), this necessitates that \(\tilde{\mathbf{S}}_D=\mathbf{S}_D\) (by assumption, the matrix \([\mathbf{H},\mathbf{J}]\) has full rank \(U+I\); see ). To make this intuition more rigorous, we introduce the notion of eclipsing, which describes a certain relation between the jammer transmit matrix \(\mathbf{W}\!_D\) and the error matrix \(\mathbf{S}_D-\tilde{\mathbf{S}}_D\) corresponding to a potential symbol estimate \(\tilde{\mathbf{S}}_D\). In this section, we define eclipsing while assuming perfect channel state information (CSI). The more realistic case, in which CSI is estimated using pilots, is considered in .

Definition 1 (Eclipsing with perfect CSI). The jammer is eclipsed in a frame if there exists a matrix \(\tilde{\mathbf{S}}_D\in\mathcal{S}^{U\times D}\!\setminus\!\{\mathbf{S}_D\}\) so that the virtual interference matrix \(\mathbf{\Sigma}(\tilde{\mathbf{S}}_D)\triangleq[\mathbf{E}(\tilde{\mathbf{S}}_D); \mathbf{W}\!_D]\) is a matrix of rank less than or equal to \(I\), where \(\mathbf{E}(\tilde{\mathbf{S}}_D) \triangleq \mathbf{S}_D-\tilde{\mathbf{S}}_D\) is the error matrix.

To say that a jammer is eclipsed therefore means that there exists an incorrect (i.e., distinct from \(\mathbf{S}_D\)) estimate \(\tilde{\mathbf{S}}_D\) of the data matrix \(\mathbf{S}_D\) such that the virtual interference matrix \(\mathbf{\Sigma}(\tilde{\mathbf{S}}_D)\), which appears in 9 , has rank less than or equal to \(I\). Our first theoretical result is that, under the assumption of negligible thermal noise, eclipsing is the only aspect that might hinder successful jammer nulling and data recovery:

Theorem 1. If the thermal noise is zero (\(\mathbf{N}_D=\boldsymbol{0}\)), and if the jammer is not eclipsed, then the problem in 7 has the unique solution \(\{\hat{\mathbf{P}},\hat{\mathbf{S}}_D\} = \{\mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}, \mathbf{S}_D\}\).

(All proofs are in the appendix.) The question is therefore: What are the chances that a jammer is—or is not—eclipsed? This question turns out to be, in general, quite complex. For instance, an immediate consequence of is that any jammer with \(\textit{rank}(\mathbf{W}\!_D)<I\) is eclipsed. The “problem” with such a jammer that only jams with rank smaller than \(I\) is that a projection \(\tilde{\mathbf{P}}\in\mathscr{G}_{B-I}(\mathbb{C}^B)\) can null more dimensions than are occupied by interference, and so can also null any data detection errors \(\mathbf{S}_D-\tilde{\mathbf{S}}_D\) of rank up to \(I-\textit{rank}(\mathbf{W}\!_D)\), which jeopardizes accurate data detection. In that sense, it would therefore be preferable to identify the parameter \(I\) not with the number of physical jammer antennas \(I\), but with the effective rank of the jammer interference, \(\textit{rank}(\mathbf{W}\!_D)\). (Any receive interference \(\mathbf{J}\mathbf{W}\!_D\) with \(\textit{rank}(\mathbf{W}\!_D)<I\) can be rewritten also as a product \(\mathbf{J}'\mathbf{W}_{\!D}'\) for some \(\mathbf{J}'\in\mathbb{C}^{B\times\textit{rank}(\mathbf{W}\!_D)}\).) Nevertheless, we have decided to identify \(I\) with the number of jammer antennas because this is more natural from an operational point of view. We note, however, that our working assumption will therefore be that the jammer transmits with full antenna rank, i.e., \(\textit{rank}(\mathbf{W}\!_D)=I\). As a consequence of this difficulty, we can only give strong theoretical guarantees for the probability of eclipsing for single-antenna jammers.9 But we emphasize that our empirical results demonstrate that JMD works well also for multi-antenna jammers; see . For single-antenna jammers, for which we write \(\mathbf{J}\mathbf{W}\!_D\) as \(\mathbf{j}\mathbf{w}_D^{\text{T}}\), the probability of eclipsing turns out to depend on the number of symbols that the jammer jams (i.e., on \(\|\mathbf{w}_D\|_0\)). A jammer that never jams is guaranteed to eclipse; a jammer that jams perpetually eclipses only with minuscule probability:

Theorem 2. The probability that a single-antenna jammer with transmit signal \(\mathbf{w}_D\) eclipses is upper bounded by \[\begin{align} p_e(\mathbf{w}_D) = \begin{cases} 1 &\text{if } \mathbf{w}_D = \mathbf{0} \\ 1-\left(1-\frac{2^{\|\mathbf{w}_D\|_0}-1}{4^{\|\mathbf{w}_D\|_0-1}}\right)^U & \text{else.} \end{cases} \label{eq:csi95eclipsing} \end{align}\qquad{(1)}\] If \(\|\mathbf{w}_D\|_0\gg0\), then \(p_e(\mathbf{w}_D) \approx \tilde{p}_e(\mathbf{w}_D)\triangleq4U\cdot2^{-\|\mathbf{w}_D\|_0}\).

depicts this probability as a function of the number of jammed symbols \(\|\mathbf{w}_D\|_0\) for different numbers of UEs \(U\).

4 Channel Estimation↩︎

In the previous section, we have assumed that the BS knows the UE channel matrix \(\mathbf{H}\) perfectly. In practice, however, \(\mathbf{H}\) has to be estimated. This is typically done using pilots. In the presence of jamming, the obtained estimate can be contaminated [24][27]. We now propose two different approaches for jammer-resilient channel estimation in combination with JMD.

4.1 Linear Channel Estimation↩︎

The first approach consists of simply using a classical linear channel estimator such as least squares (LS) or linear minimum mean square error (LMMSE). It may not be immediately apparent why these estimates should be jammer-resilient. In fact, there is nothing inherently jammer-resilient about them; they become jammer-resilient only when combined with JMD.

The underlying mechanism works as follows: Let the pilot phase be given as in 4 , with a square (\(T=U\)) or wide (\(T>U\)) pilot matrix, and let the corresponding data phase be given as in 5 . If a linear channel estimator is used, then the channel estimation error due to the jammer interference is contained in \(\textit{col}(\mathbf{J})\). For instance, in the case of LS channel estimation, we have \[\begin{align} \hat{\mathbf{H}}&= \mathbf{Y}_T \mathbf{S}_T^{\dagger} = \mathbf{H}+ \underbrace{\mathbf{J}\mathbf{W}_T\mathbf{S}_T^{\dagger}}_{\in\,\textit{col}(\mathbf{J})} + \mathbf{N}_T\mathbf{S}_T^{\dagger}. \label{eq:pilot95contamination} \end{align}\tag{10}\] Therefore, if we simply plug the jammer-contaminated linear channel estimate \(\hat{\mathbf{H}}\) from 10 into 7 as follows, \[\begin{align} \min_{\substack{\tilde{\mathbf{S}}_D \,\in\, \mathcal{S}^{U\times D},\\ \tilde{\mathbf{P}}\,\in\, \mathscr{G}_{B-I}(\mathbb{C}^B)}} \| \tilde{\mathbf{P}}(\mathbf{Y}\!_D - \hat{\mathbf{H}}\tilde{\mathbf{S}}_D) \|_F^2, \label{eq:jmd95problem2} \end{align}\tag{11}\] then the “true” jammer-nulling projection \(\mathbf{P}= \mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}\) also removes the jammer contamination of the channel estimate, since \(\mathbf{P}\hat{\mathbf{H}}=\mathbf{P}\mathbf{H}\). This implies that JMD can naturally incorporate jammer-resilient channel estimation, without requiring extra countermeasures to protect channel estimation from jammer interference.

Figure 1: The bound p_e (and its approximation \tilde{p}_e) on the probability that a single-antenna jammer eclipses vs. the number of jammed symbols \|\mathbf{w}\|_0, for different numbers U of UEs. The bound on the probability of eclipsing decreases exponentially in the number of jammed symbols.

To characterize the probability of successful data recovery for JMD in combination with linear channel estimation, we provide the following adapted (compared to ) definition of eclipsing:10

Definition 2 (Eclipsing with channel estimation). The jammer is eclipsed in a given frame if there exists a matrix \(\tilde{\mathbf{S}}_D\in\mathcal{S}^{U\times D}\setminus\!\{\mathbf{S}_D\}\), so that \(\mathbf{\Sigma}(\tilde{\mathbf{S}}_D)\triangleq[\mathbf{S}_D-\tilde{\mathbf{S}}_D;\) \(\mathbf{W}\!_D - \mathbf{W}_T\mathbf{S}_T^{\dagger}\tilde{\mathbf{S}}_D]\) is a matrix of rank less than or equal to \(I\).

With this modified definition of eclipsing, we obtain the following guarantee for JMD with linear channel estimation:

Theorem 3. If the thermal noise is zero (\(\mathbf{N}=\boldsymbol{0}\)), and if the jammer is not eclipsed, then the problem in 11 has the unique solution \(\{\hat{\mathbf{P}},\hat{\mathbf{S}}_D\} = \{\mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}, \mathbf{S}_D\}\).

As in , the question becomes: what is the probability that a jammer eclipses?11 The answer depends fundamentally on whether or not the jammer knows the pilots that are being used by the UEs. Let us first consider the case where the jammer knows the pilots (as well as the transmit constellation \(\mathcal{S}\)).

Theorem 4. If a single-antenna jammer knows at least one pilot sequence (i.e., one row of \(\mathbf{S}_T\)) as well as the constellation \(\mathcal{S}\), then it can jam all frame symbols (i.e., \(\|\mathbf{w}_T\|_0=T\) and \(\|\mathbf{w}_D\|_0=D\)) while eclipsing with probability \(1-4^{-D}\approx1\).

However, the situation is completely different if the (single-antenna) jammer does not know the pilots. This can be modeled by assuming that a (square) pilot matrix \(\mathbf{S}_T\) is drawn at random according to the power-normalized Haar measure (i.e., drawn uniformly from the set of unitary \(T\times T\) matrices and scaled to \(\|\mathbf{S}_T\|_F^2=U^2\)) [28] and not revealed to the jammer in advance:12

Theorem 5. If \(\mathbf{S}_T\in\mathbb{C}^{T\times U}\) is square (\(T=U\)) and drawn according to the power-normalized Haar measure without being revealed to the jammer, then the probability that a single-antenna jammer eclipses is upper bounded by \[\begin{align} p_e(\mathbf{w}_D) = \begin{cases} 1 &\text{if } \mathbf{w}_D = \mathbf{0} \\ 1-\left(1-\frac{2^{\|\mathbf{w}_D\|_0}-1}{4^{\|\mathbf{w}_D\|_0-1}}\right)^U & \text{else,} \end{cases} \label{eq:ind95chest95eclipsing} \end{align}\qquad{(2)}\] if the jammer does not jam the pilot phase (i.e., if \(\mathbf{w}_T=\mathbf{0}\)), and otherwise (i.e., if \(\mathbf{w}_T\neq\mathbf{0}\)), it is upper bounded by \[\begin{align} p_{e,\mathbf{w}_T\neq\mathbf{0}} = 1-\left(1-\frac{2^D-1}{4^D-1}\right)^U. \label{eq:jammer95jams95pilots} \end{align}\qquad{(3)}\]

Note that the bound in ?? , where the jammer is silent during the pilot phase, coincides with the bound from . The bound in ?? , where the jammer is active during the pilot phase, satisfies \(p_{e,\mathbf{w}_T\neq\mathbf{0}}\leq p_e(\mathbf{w}_D)\) with equality if and only if \(\mathbf{w}_D\) has full support, in which case \(p_{e,\mathbf{w}_T\neq\mathbf{0}}= p_e(\mathbf{w}_D) \approx 4U\cdot2^{-D}\).

It is evident that the guarantees for successful JMD jammer mitigation are fundamentally dependent on whether or not the jammer knows the pilots used for channel estimation. This finding is reminiscent of a more general information-theoretic result which shows that reliable communication in the presence of jamming is possible if the legitimate transmitter and the receiver share a common secret, but not otherwise[29]. Intuitively, the receiver needs a lever that enables it to distinguish between legitimate and illegitimate transmit signals. This lever could be knowledge of the legitimate UEs’ channel matrix \(\mathbf{H}\) (as in ). Or, if ground-truth channel knowledge is not available, it could be a mechanism which guarantees that the jammer cannot replicate a legitimate pilot sequence during the channel estimation phase (as in ).

4.2 Joint Channel Estimation↩︎

Besides linear channel estimation, there is a second possible approach for channel estimation in combination with JMD, namely to estimate the channel jointly with the data and the jammer-mitigating projection. Specifically, the idea is to use a pilot and a data phase as in 4 and 5 , and to then solve \[\begin{align} \min_{\substack{ \tilde{\mathbf{P}}\,\in\, \mathscr{G}_{B-I}(\mathbb{C}^B),\\ \tilde{\mathbf{H}}\,\in\, \mathbb{C}^{B\times U},\\ \tilde{\mathbf{S}}_D \,\in\, \mathcal{S}^{U\times D} }} \big\| \tilde{\mathbf{P}}([\mathbf{Y}_T, \mathbf{Y}\!_D] - \tilde{\mathbf{H}}[\mathbf{S}_T, \tilde{\mathbf{S}}_D]) \big\|_F^2. \label{eq:jmd95problem3} \end{align}\tag{12}\] Analogous to linear channel estimation (cf. ), this approach exploits the fact that the jammer’s interference during the channel estimation and the data phase is contained in the same subspace \(\textit{col}(\mathbf{J})\). Compared to linear channel estimation, joint channel estimation benefits from a joint channel estimation and data detection (JED) performance gain, at the price of having to solve a more complex optimization problem.

To characterize the probability of successful data recovery for JMD in combination with joint channel estimation, we can use the same definition of eclipsing () as in . Furthermore, it turns out that we also obtain the same success guarantee:

Theorem 6. If the noise is zero (\(\mathbf{N}=\boldsymbol{0}\)), and if the jammer is not eclipsed, then the problem in 12 has the unique solution \(\{\hat{\mathbf{P}},\hat{\mathbf{P}}\hat{\mathbf{H}},\hat{\mathbf{S}}_D\} = \{\mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}, \hat{\mathbf{P}}\mathbf{H}, \mathbf{S}_D\}\).13

Since the definition for eclipsing is the one from , the results of and specify the probability of eclipsing also for the case of joint channel estimation. Therefore, regardless of whether we estimate the channel matrix separately with a linear estimator (as in ) or jointly (as in this section), randomized pilots are necessary for meaningful success guarantees. However, we emphasize that these guarantees do not imply that joint channel estimation has the same performance as linear channel estimation under non-zero noise, i.e., for \(\mathbf{N}\neq\mathbf{0}\).

5 Algorithms↩︎

In the previous two sections, we have introduced the JMD paradigm (see ) and outlined two alternatives for channel estimation (linear and joint; see ). In this section, we provide concrete algorithms, i.e., we propose algorithms for approximately solving the optimization problems in 11 and 12 . The presented algorithms are by no means the only possible algorithms for these optimization problems. The development of other JMD algorithms is left as future work.

5.1 The SANDMAN Algorithm↩︎

We start by providing SANDMAN, an algorithm for JMD in conjunction with linear channel estimation as expressed through the optimization problem in 11 . Since solving 11 exactly is difficult, we solve it approximately. A key difficulty is that, due to the discreteness of \(\mathcal{S}\), the problem in 11 is NP-hard even when fixing \(\tilde{\mathbf{P}}\) and solving only for \(\tilde{\mathbf{S}}_D\) [30]. We thus relax the constraint set \(\mathcal{S}\) to its convex hull \(\mathcal{C}\triangleq \textit{conv}(\mathcal{S})\).14

To promote symbol estimates at, or near, the corners of \(\mathcal{C}\), we add a concave regularizer \(-\|\tilde{\mathbf{S}}_D\|_F^2\) weighted by \(\alpha>0\) to the objective in 11. We colloquially refer to the resulting constraint and regularizer as a box prior. The use of this box prior is motivated by our assumption (cf. ) that the transmit constellation is QPSK, so that the corners of \(\mathcal{C}\) correspond to the constellation symbols \(\mathcal{S}\). If larger constellations such as 16-QAM or 64-QAM are used, other signal priors are more effective [15]. The relaxed problem is \[\begin{align} \min_{\substack{\tilde{\mathbf{S}}_D \,\in\, \mathcal{C}^{U\times D},\\ \tilde{\mathbf{P}}\,\in\, \mathscr{G}_{B-I}(\mathbb{C}^B)}} \big\| \tilde{\mathbf{P}}(\mathbf{Y}\!_D - \hat{\mathbf{H}}\tilde{\mathbf{S}}_D) \big\|_F^2 - \alpha\|\tilde{\mathbf{S}}_D\|_F^2. \label{eq:sandman95problem} \end{align}\tag{13}\] Due to the nonconvex constraint set \(\mathscr{G}_{B-I}(\mathbb{C}^B)\), the relaxed problem is still non-convex. But the following result holds:

Theorem 7. When \(\tilde{\mathbf{P}}\) is fixed and \(\alpha\leq \lambda_{\min}(\hat{\mathbf{H}}^{\textrm{H}}\tilde{\mathbf{P}}\hat{\mathbf{H}})\), then the objective in 13 is convex in \(\tilde{\mathbf{S}}_D\). Vice versa, when \(\tilde{\mathbf{S}}_D\) is fixed, then the objective in 13 is minimized with respect to \(\tilde{\mathbf{P}}\) by \(\mathbf{I}_B - \mathbf{U}_I\mathbf{U}_I^{\text{H}}\), where \(\mathbf{U}_I\in\mathbb{C}^{B\times I}\) consists of the \(I\) principal left-singular vectors of \(\mathbf{Y}\!_D - \hat{\mathbf{H}}\tilde{\mathbf{S}}_D\) (i.e., the left-singular vectors corresponding to the \(I\) largest singular values).

This theorem suggests using an alternating minimization strategy, since solving 13 for either \(\tilde{\mathbf{S}}_D\) or \(\tilde{\mathbf{P}}\) is straightforward as long as the other quantity is fixed:

5.1.1 Solving for \(\tilde{\mathbf{S}}_D\)↩︎

To solve 13 with respect to \(\tilde{\mathbf{S}}_D\) (for fixed \(\tilde{\mathbf{P}}\)), we use forward-backward splitting (FBS) [31]. FBS is a method for solving optimization problems of the form \[\begin{align} \min_{\tilde{\mathbf{s}}}~ f(\tilde{\mathbf{s}}) + g(\tilde{\mathbf{s}}), \label{eq:fbs95problem} \end{align}\tag{14}\] where \(f\) is convex and differentiable, and \(g\) is convex but not necessarily differentiable. FBS solves such problems iteratively by computing \[\begin{align} \tilde{\mathbf{s}}^{(t+1)} = \text{prox}_g\big(\tilde{\mathbf{s}}^{(t)} - \tau^{(t)}\nabla f(\tilde{\mathbf{s}}^{(t)}); \tau^{(t)}\big), \label{eq:fbs95iteration} \end{align}\tag{15}\] where \(\tau^{(t)}\) is the stepsize at iteration \(t\), \(\nabla f(\tilde{\mathbf{s}})\) is the gradient of \(f\) in \(\tilde{\mathbf{s}}\), and \(\text{prox}_g\) is the proximal operator of \(g\), defined as [32] \[\begin{align} \text{prox}_g(\tilde{\mathbf{s}}; \tau) = \arg\!\min_{\tilde{\mathbf{x}}~~~~} \tau g(\tilde{\mathbf{x}}) + \frac{1}{2} \|\tilde{\mathbf{s}}- \tilde{\mathbf{x}}\|_2^2. \end{align}\] FBS solves convex problems exactly (for a sufficient number of iterations with suitable stepsizes \(\tau^{(t)}\)), but it is also effective for approximately solving non-convex problems [31]. To solve the problem in 13 , we define the functions \(f\) and \(g\) as \[\begin{align} f(\tilde{\mathbf{S}}_D) &= \big\| \tilde{\mathbf{P}}(\mathbf{Y}\!_D - \hat{\mathbf{H}}\tilde{\mathbf{S}}_D) \big\|_F^2, \\ g(\tilde{\mathbf{S}}_D) &= - \alpha\big\|\tilde{\mathbf{S}}_D\big\|_F^2 + \chi_\mathcal{C}(\tilde{\mathbf{S}}_D), \end{align}\] where \(\chi_\mathcal{C}\) acts entrywise on \(\tilde{\mathbf{S}}_D\) as the indicator function of \(\mathcal{C}\), \[\begin{align} \chi_\mathcal{C}(\tilde{s}) = \begin{cases} 0 &: \tilde{s} \in \mathcal{C}\\ \infty &: \tilde{s} \notin \mathcal{C}. \end{cases} \end{align}\] The gradient of \(f\) in \(\tilde{\mathbf{S}}_D\) is given as \[\begin{align} \nabla f(\tilde{\mathbf{S}}_D) = -2\,\hat{\mathbf{H}}^{\text{H}}\tilde{\mathbf{P}}(\mathbf{Y}\!_D - \hat{\mathbf{H}}\tilde{\mathbf{S}}_D). \end{align}\] The proximal operator of \(g\) acts entrywise on \(\tilde{\mathbf{S}}_D\) and is given as \(\text{prox}_g(\tilde{s}; \tau) = \text{clip}(\tilde{s}/(1-\tau\alpha); \sqrt{\frac{1}{2}})\) when \(\alpha\tau<1\) (where \(\text{clip}(z;a)\) clips the real and imaginary part of \(z\in\mathbb{C}\) to the interval \([-a,a]\)), and otherwise as \(\arg\!\min_{\tilde{x} \in \{\pm\sqrt{\frac{1}{2}} \pm i\sqrt{\frac{1}{2}} \} } |\tilde{s} - \tilde{x}|^2\).

Figure 2: Approximate SVD [33]

5.1.2 Solving for \(\tilde{\mathbf{P}}\)↩︎

According to , we can solve 13 with respect to \(\tilde{\mathbf{P}}\) (for fixed \(\tilde{\mathbf{S}}_D\)) by calculating the \(I\) principal left-singular vectors \([\mathbf{u}_1,\dots,\mathbf{u}_I]=\mathbf{U}_I\) of \(\mathbf{Y}\!_D - \hat{\mathbf{H}}\tilde{\mathbf{S}}_D\). Instead of performing an exact but computationally expensive singular value decomposition (SVD), we approximate \(\mathbf{U}_I\) with the power method from [33]. We perform a single power iteration per dimension. The resulting procedure is outlined in .

The SANDMAN algorithm alternates between descent steps in \(\tilde{\mathbf{S}}_D\) and approximate computations of \(\tilde{\mathbf{P}}\) for a total of \(t_\text{max}\) iterations. We choose \(\alpha=2.5\), and the stepsizes \(\tau^{(t)}\) are selected using the Barzilai-Borwein criterion [34] detailed in [35]. SANDMAN is summarized in and has a computational complexity of \(O(t_\text{max}UDB)\), i.e., its complexity is linear in the number of UEs \(U\), the number of data samples \(D\), and the number of BS antennas \(B\).

One key feature of SANDMAN is that, for single antenna jammers for which \((\tilde{\mathbf{J}}^{(t)})^{\dagger}=(\tilde{\mathbf{j}}^{(t)})^{\dagger} = (\mathbf{j}^{(t)})^{\text{H}}/\|\mathbf{j}^{(t)}\|^2\), it requires no matrix inversion. This makes SANDMAN particularly attractive for hardware implementation. In fact, the first (and so far the only) jammer-mitigating MU-MIMO receiver ASIC [22] is based on SANDMAN, mainly thanks to the fact that no matrix inversion is required.

Figure 3: SANDMAN

5.2 The MAED Algorithm↩︎

MAED is an algorithm for JMD in conjunction with joint channel estimation as expressed through the optimization problem in 12 . A version of MAED for single-antenna jammers has originally been proposed in [15].15 Estimating the channel jointly enables MAED to achieve the detection gains associated with joint channel estimation and data detection (JED) [36][40]. Other than estimating the channel jointly as well, however, MAED works similarly to SANDMAN.

To derive the MAED algorithm, we start by noting that the objective in 12 is quadratic in \(\tilde{\mathbf{H}}\). The optimal \(\tilde{\mathbf{H}}\) as a function of \(\tilde{\mathbf{P}}\) and \(\tilde{\mathbf{S}}\triangleq[\mathbf{S}_T,\tilde{\mathbf{S}}_D]\) is therefore equal to \[\begin{align} \tilde{\mathbf{H}}= \tilde{\mathbf{P}}\mathbf{Y}\tilde{\mathbf{S}}^{\dagger}. \end{align}\] By plugging this expression back into 12 , we obtain an optimization problem which only depends on \(\tilde{\mathbf{P}}\) and \(\tilde{\mathbf{S}}\). Furthermore, as in 13 , we also relax the constraint set \(\mathcal{S}\) to its convex hull \(\mathcal{C}\) and add a concave regularizer \(-\alpha\|\tilde{\mathbf{S}}_D\|_F^2\) to promote symbol estimates near the constellation points. The resulting optimization problem is therefore \[\begin{align} \min_{\substack{ \tilde{\mathbf{P}}\,\in\, \mathscr{G}_{B-I}(\mathbb{C}^B),\\ \tilde{\mathbf{S}}=[\mathbf{S}_T,\tilde{\mathbf{S}}_D]:\, \tilde{\mathbf{S}}_D \,\in\, \mathcal{C}^{U\times D} }} \big\| \tilde{\mathbf{P}}\mathbf{Y}(\mathbf{I}_K - \tilde{\mathbf{S}}^{\dagger}\tilde{\mathbf{S}}) \big\|_F^2 - \alpha\|\tilde{\mathbf{S}}_D\|_F^2. \end{align}\] As in SANDMAN, we now use an alternating minimization strategy where we alternate between FBS-based descent steps in \(\tilde{\mathbf{S}}\) and approximate optimization steps in \(\tilde{\mathbf{P}}\). The gradient required for FBS is \[\begin{align} \nabla f(\tilde{\mathbf{S}}) = - (\mathbf{Y}\tilde{\mathbf{S}}^{\dagger})^{\text{H}}\tilde{\mathbf{P}}\mathbf{Y}(\mathbf{I}_K-\tilde{\mathbf{S}}^{\dagger}\tilde{\mathbf{S}}). \end{align}\] The proximal operator \(\text{prox}_g\) maps the first \(T\) columns of \(\tilde{\mathbf{S}}\) to \(\mathbf{S}_T\) and acts entrywise on the remaining columns as \(\text{prox}_g(\tilde{s}; \tau) = \text{clip}(\tilde{s}/(1-\tau\alpha); \sqrt{\frac{1}{2}})\) when \(\alpha\tau<1\) (where \(\text{clip}(z;a)\) clips the real and imaginary part of \(z\in\mathbb{C}\) to the interval \([-a,a]\)), and otherwise as \(\arg\!\min_{\tilde{x} \in \{\pm\sqrt{\frac{1}{2}} \pm i\sqrt{\frac{1}{2}} \} } |\tilde{s} - \tilde{x}|^2\).

As in SANDMAN, we choose \(\alpha=2.5\) and compute the stepsizes using the Barzilai-Borwein criterion of [35]. The resulting algorithm is summarized in and has computational complexity \(O(t_\text{max}UKB)\). While this is the same asymptotic complexity order as for SANDMAN, MAED requires the computation of a pseudoinverse (of \(\tilde{\mathbf{S}}\in\mathbb{C}^{U\times K}\)) in every algorithm iteration, which makes it less attractive for hardware implementation.

Figure 4: MAED

6 Evaluation↩︎

6.1 Simulation Setup↩︎

We evaluate our JMD-type methods SANDMAN and MAED using simulations. The channel vectors are generated with QuaDRiGa [41]. We simulate a MU-MIMO system with \(B=32\) BS antennas and \(U=16\) single-antenna UEs at a carrier frequency of 2 GHz using the 3GPP 38.901 urban macrocellular (UMa) channel model [42]. All antennas are omnidirectional. The BS antennas are arranged as a uniform linear array (ULA) and spaced at half wavelength. The UEs are uniformly distributed at distances between \(10\) m and \(250\) m in a \(120^\circ\) angular sector in front of the BS, and with a minimum angular separation of \(1^\circ\) between any two UEs. We assume \(\pm3\) dB per-UE power control. The specific jammer model varies between the different experiments, see below. In general, we consider \(J\geq1\) jammers placed randomly in the same area as the UEs, with a minimum angular separation of \(1^\circ\) between any two jammers as well as between any jammer and any UE. Every jammer is equipped with \(\frac{I}{J}\geq1\) antennas arranged as a ULA with half-wavelength spacing that faces towards the BS.

To be able to compare our methods with training-period-based methods, we now consider communication frames of length \(L=K+R\) instead of \(K\). That is, we replace 3 with \[\begin{align} \mathbf{Y}= \mathbf{H}\mathbf{X}+ \mathbf{J}\mathbf{W}+ \mathbf{N}\in \mathbb{C}^{B\times L}, \end{align}\] where the columns of \(\mathbf{X}\) consist of \(R\) (for “redundancy”) evenly distributed all-zero columns used as a jammer training period as well as the columns of \(\mathbf{S}=[\mathbf{S}_T,\mathbf{S}_D]\).16 The submatrices of \(\mathbf{Y}\) consisting of the columns of the jammer training period, the pilot phase, and the data phase are denoted \(\mathbf{Y}_J\in\mathbb{C}^{B\times R}\), \(\mathbf{Y}_T\in\mathbb{C}^{B\times T}\), and \(\mathbf{Y}_D\in\mathbb{C}^{B\times D}\), respectively. Our JMD methods will use \(R=0\) since they do not require a jammer training period. We assume a frame duration of \(L=100\) channel uses.

We consider QPSK transmission, and we use a square pilot matrix (\(T=U=16\)). The pilots are selected as rows of a \(U\times U\) Haar matrix (normalized to unit symbol energy) and not revaled to the jammer. The signal-to-noise ratio (SNR) is \[\begin{align} \textit{SNR} \triangleq\frac{\mathop{\mathrm{\mathbb{E}}}_{\mathbf{s}_k}\mathopen{}\left[\|\mathbf{H}\mathbf{s}_k\|_s^2\right]}{\mathop{\mathrm{\mathbb{E}}}_{\mathbf{n}_k}\mathopen{}\left[\|\mathbf{n}_k\|_2^2\right]} = \frac{\|\mathbf{H}\|_F^2}{BN_0}. \end{align}\] Furthermore, we characterize the receive power of the jammer interference relative to the receive power of the average UE as \[\begin{align} \rho \triangleq\frac{\frac{1}{L}\|\mathbf{J}\mathbf{W}\|_F^2}{\frac{1}{U}\mathop{\mathrm{\mathbb{E}}}_{\mathbf{s}_k}\mathopen{}\left[\|\mathbf{H}\mathbf{s}_k\|_2^2\right]} = \frac{\frac{1}{L}\|\mathbf{J}\mathbf{W}\|_F^2}{\frac{1}{U}\|\mathbf{H}\|_F^2}, \end{align}\] where we deterministically scale the interference \(\mathbf{J}\mathbf{W}\) to some specified \(\rho\). As performance metrics, we consider the uncoded bit error rate (BER) and a metric called modulation error ratio (MER) between the data symbols \(\mathbf{S}_D\) and their estimate \(\hat{\mathbf{S}}_D\), \[\begin{align} \textit{MER}\triangleq \sqrt{\frac{\mathbb{E}\big[\|\hat{\mathbf{S}}_D - \mathbf{S}_D\|_F^2\big]}{\mathbb{E}\big[\|\mathbf{S}_D\|_F^2\big]}}. \end{align}\] We use the MER as a surrogate for error vector magnitude (EVM), which the 3GPP 5G NR technical specification requires to be below 17.5% [43] for QPSK transmission.

6.2 Higher Data Rates↩︎

The first advantage of JMD compared to training-period based mitigation schemes is increased data rates since no symbol time slots are reserved for estimating the jammer’s channel.

Jammer Model↩︎

The rate advantage due to the absence of a training period is shown for the following jammer model:

Single-antenna barrage jammer↩︎

One single-antenna jammer transmits i.i.d. circularly-symmetric complex Gaussian noise during the entire frame. The jammer power is \(\rho=30\) dB.

Baselines↩︎

The following receivers are used as baselines:

POS-JED↩︎

This receiver uses \(0\leq R \leq L-T\) channel uses in each frame for estimating the jammer subspace as the \(I\) principal left singular vectors of \(\mathbf{Y}_J\). This reduces the number of channel uses for data transmission from \(D=K-T\) to \(D=K-T-R\). The jammer is mitigated by first projecting \(\mathbf{Y}_T\) and \(\mathbf{Y}_D\) onto the orthogonal complement of the estimated subspace, followed by performing FBS-based joint channel estimation and data detection analogous to MAED.

G-POS-JED↩︎

This receiver serves as an upper bound to the achievable performance of MAED. It works analogous to MAED but is furnished with ground-truth knowledge of the jammer channel matrix \(\mathbf{J}\). It therefore uses \(R=0\) and sets \(\tilde{\mathbf{P}}^{(t)}\) on line 7 of to the optimal projector \(\mathbf{P}= \mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}\).

POS-BOX↩︎

Like POS-JED, this receiver uses \(0\leq R \leq L-T\) channel uses of each frame for estimating the jammer subspace as the \(I\) principal left singular vectors of \(\mathbf{Y}_J\). The jammer is mitigated by first projecting \(\mathbf{Y}_T\) and \(\mathbf{Y}_D\) onto the orthogonal complement of the estimated subspace, followed by performing LS channel estimation and FBS-based data detection with a box prior analogous to SANDMAN.

G-POS-BOX↩︎

This receiver serves as an upper bound to the achievable performance of SANDMAN. It works analogous to SANDMAN but is furnished with ground-truth knowledge of the jammer’s channel matrix \(\mathbf{J}\). It therefore uses \(R=0\) and fixes \(\tilde{\mathbf{P}}^{(t)}\) on line 7 of to the optimal projector \(\mathbf{P}= \mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}\).

All of the above algorithms run \(t_\text{max}=30\) iterations.

Results↩︎

In order to demonstrate the rate increase due to the absence of a jammer training period, we consider the tradeoff between the lowest SNR for which a receiver satisfies \(\text{MER}\leq 17.5\%\) and the ratio \[\begin{align} r = \frac{L-T-R}{L-T}, \end{align}\] which is the number of channel uses \(L-T-R\) per communication frame that are available for data transmission, normalized with respect to the maximum number of samples \(L-T\) available without a jammer estimation phase. The ratio \(r\) translates directly to the achieved data rate (measured in terms of bits/s). shows the results:

Since SANDMAN and MAED as well as G-POS-BOX and G-POS-JED have no jammer training period (\(R=0\)), they achieve \(r=1\). In contrast, the detection accuracy of POS-BOX and POS-JED increases with the redundancy \(R\) (a longer training period yields a better estimate of the jammer channel, which enables more precise nulling), at the expense of \(r\). Our results demonstrate that SANDMAN and MAED, which can utilize the receive data of the entire frame to estimate the jammer subspace, achieve virtually the same performance as their genie-assisted counterparts G-POS-BOX and G-POS-JED. In contrast, POS-BOX and POS-JED can only use a subset of the receive samples to estimate the jammer interference, and so perform significantly worse even when dedicating a significant fraction of the frame to jammer training. In fact, POS-BOX does not reliably outperform SANDMAN even when using almost the entire frame for jammer training (so that no almost no channel uses remain for data transmission, \(r\to0\)). The performance of POS-JED even deteriorates when \(r\to0\) (since joint channel estimation and data detection degenerates when there are very few data symbols), and it is always at least \(0.18\) dB away from the performance of MAED.

Figure 5: Trade-off between the relative achievable rate r and the lowest SNR for which the different receivers satisfy the criterion \text{MER}\leq17.5\% when mitigating a single-antenna barrage jammer .
Figure 6: Uncoded bit error-rate (BER) vs. SNR performance of different receivers when mitigating different kinds of jammers, including smart, distributed, and dynamic multi-antenna jammers.

6.3 Mitigating Smart, Distributed, and Multi-Antenna Jammers↩︎

In this experiment, we analyze the ability of JMD to mitigate smart jammers that might not jam during any training phase but only at specific instants, or spatially distributed single-antenna jammers, or multi-antenna jammers that use dynamic beamforming to change their interference subspace.

Jammer Models↩︎

Besides the single-antenna barrage jammer (), we consider the following types of jammers:

Single-antenna data jammer↩︎

A single-antenna jammer that does not jam the training period and pilot phase, and that transmits i.i.d. complex Gaussian noise with \(\rho=30\) dB in the data phase. Note that this jammer is smart (i.e., protocol-aware).

Single-antenna pilot jammer↩︎

A single-antenna jammer that does not jam the training period and data phase, and that transmits i.i.d. complex Gaussian noise with \(\rho=30\) dB during the pilot phase. Note that this jammer is smart, too.

Single-antenna sparse jammer↩︎

A non-smart single-antenna jammer that only jams for a (randomly selected) single symbol time per communication frame, with \(\rho=30\) dB.

Distributed single-antenna barrage jammers↩︎

Four distributed and statistically independent single-antenna jammers transmit i.i.d. complex Gaussian noise with \(\rho=30\) dB.

Slowly varying multi-antenna jammer↩︎

A non-smart four-antenna jammer where the \(\tilde{\mathbf{w}}_k\) (see 2 ) are white Gaussian random vectors with \(\rho=30\) dB. The matrices \(\mathbf{A}_k\) are constructed as follows: Only the leftmost column \(\mathbf{a}_{k,1}\) of \(\mathbf{A}_k\) is nonzero. For randomly selected instants \(k_1<\dots<k_M, M=5\), the vector \(\mathbf{a}_{k_m,1}\) is fixed to randomly drawn vectors \(\{\mathbf{a}^{(m)}\}_{m=1}^M\). For \(k_m\!<\!k\!<\!k_{m+1}\), \(\mathbf{a}_{k,1}\) interpolates between \(\mathbf{a}^{(m)}\) and \(\mathbf{a}^{(m+1)}\).

Abruptly varying multi-antenna jammer↩︎

A non-smart four-antenna jammer that transmits only with a random subset of one or two of its antennas at each instant \(k\). To this end, a randomly selected subset of the rows of the beamforming matrix \(\mathbf{A}_k\) (see 2 ) is populated with i.i.d. \(\mathcal{C}\mathcal{N}(0,1)\) entries, and the other rows are set to zero. The beamforming matrix \(\mathbf{A}_{k+1}\) is equal to \(\mathbf{A}_k\) with probability \(0.95\) and otherwise is redrawn (including the support set of the rows of \(\mathbf{A}_{k+1}\)). The vectors \(\tilde{\mathbf{w}}_k\) are white Gaussian random vectors with \(\rho=30\,\)dB.

Baselines↩︎

We compare SANDMAN and MAED with the baselines from . The training period-based methods POS-BOX and POS-JED use a redundancy of \(R=4\) and \(D=80\) data transmission symbols per frame (yielding a relative rate of \(r=80/84\approx0.95\)) while the other methods use \(R=0\) and \(D=84\) data transmission symbols (yielding \(r=1\)). All iterative algorithms run for \(t_\text{max}=30\) iterations when facing a single-antenna jammer, and for \(t_\text{max}=50\) iterations when facing distributed or multi-antenna jammers. The algorithms are provided with knowledge of the number \(I\) of jammer antennas. Other than that, the configurations of the algorithms do not depend on the type of jammer being faced. For context, we also include the following unmitigated receiver as an additional baseline:

Unmitigated↩︎

This receiver does not mitigate the jammer. It performs LS channel estimation and (jammer-oblivious) LMMSE data detection.

Results↩︎

depicts the results. The performance of the G-POS-BOX and G-POS-JED baselines depends on the jammer type only via the number of jammer antennas.17 In contrast, the performance of the unmitigated receiver suffers to varying degrees—but always significantly—under the different types of jammers: the least harmful jammers for an unmitigated receiver are the pilot jammer and the sparse jammer (with a BER of around \(1\%\) at high SNR), and the most harmful are the data jammer as well as distributed and multi-antenna jammers , , (with a BER of \(20\%\) - \(50\%\) even at high SNR). As expected, the training period-based baselines POS-BOX and POS-JED mitigate the barrage jammers and successfully, but fail against all smart or dynamic jammers.18

In contrast, the JMD-type methods SANDMAN and MAED are able to mitigate all jammers more or less successfully:

  • They achieve virtually the same performance as their genie-assisted counterparts G-POS-BOX and G-POS-JED against the barrage jammers and as well as against the data jammer and the pilot jammer . This indicates that SANDMAN and MAED null these jammers essentially perfectly. Note also that SANDMAN and MAED outperform their training period-based counterparts POS-BOX and POS-JED even against the barrage jammers and (as expected based on ).

  • Also against the smart jammers and , SANDMAN and MAED achieve virtually the same performance as their genie-assisted counterparts. In contrast, their training period-based counterparts fail completely against such smart jammers that do not jam during the training period.

  • Even against the sparse jammer , which might be expected to be their Achilles heel (based on the discussion in and ), SANDMAN and MAED perform outperform the unmitigated baseline as well as their training-period based counterparts. They manage to achieve a BER of below \(1\%\) at high SNR.

  • Against the dynamic multi-antenna jammers and , SANDMAN and MAED achieve BERs significantly below \(1\%\) at high SNR. In contrast, the BERs of their training period-based counterparts POS-BOX and POS-JED remain significantly above \(1\%\) even against such non-smart dynamic jammers. Note, however, that also SANDMAN and MAED eventualy hit an error floor due to difficulty of disentangling a high-dimensional and time-varying interference subspace from the signal subspace. The error floor of SANDMAN is lower than that of MAED because the optimization problem approximated by SANDMAN is less complex, and thus less prone to misconvergence, than the one approximated by MAED.

In summary, SANDMAN and MAED outperform their training period-based counterparts POS-BOX and POS-JED for every considered jammer type (sometimes decivisely). They also consistently outperform the unmitigated baseline. This demonstrates the suitability of JMD-based methods for mitigating smart and/or dynamic single-antenna jammers, distributed jammers, or multi-antenna jammers.

7 Discussion and Conclusion↩︎

We have proposed joint jammer mitigation and data detection (JMD), a novel paradigm for mitigating jammers in MIMO systems that does not use a dedicated jammer training period. As a result, JMD is able to mitigate smart jammers regardless of (i) when they are active and (ii) how they vary their multi-antenna transmit beamforming. We have provided theoretical success guarantees for the case of smart single-antenna jammers, and we have proposed two JMD-type algorithms (SANDMAN and MAED) whose efficacy against single- and multi-antenna jammers has been demonstrated through extensive simulations.

At the moment, JMD still exhibits certain drawbacks: (i) we only provide success guarantees for single-antenna jammers, (ii) JMD requires the number of jammer antennas (or the rank of jamming interference) to be known at the receiver in advance, and (iii) the JMD-type algorithms SANDMAN and MAED exhibit an error floor against sparse jammers (which are prone to eclipsing, see and ) and against multi-antenna jammers (for which solving the respective optimization problems is challenging). All of these issues can be remedied by combining JMD with a recently developed technique in which the transmit signals are transformed using a linear time-domain transform. Preliminary results are shown in [45] and will be detailed more fully in future work.

Proofs↩︎

7.1 Proof of↩︎

We start by noting that the Frobenius norm in 7 is nonnegative and evaluates to zero if and only if its argument is the all-zero matrix. That \(\{\hat{\mathbf{P}},\hat{\mathbf{S}}_D\} = \{\mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}, \mathbf{S}_D\}\) is a minimizer follows then directly from \[\begin{align} \!\!\hat{\mathbf{P}}(\mathbf{Y}\!_D - \mathbf{H}\hat{\mathbf{S}}_D) &= (\mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}) (\mathbf{H}\mathbf{S}_D + \mathbf{J}\mathbf{W}\!_D - \mathbf{H}\mathbf{S}_D)\!\! \\ &= (\mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger})\mathbf{J}\mathbf{W}\!_D \\ &= \mathbf{0}, \end{align}\] since \((\mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger})\mathbf{J}= \mathbf{0}\). It remains to show that \(\{\hat{\mathbf{P}},\hat{\mathbf{S}}_D\}\) is the only minimizer of 7 . For this, we rewrite the argument of the Frobenius norm in 7 as \[\begin{align} \tilde{\mathbf{P}}(\mathbf{H}\mathbf{S}_D + \mathbf{J}\mathbf{W}\!_D - \mathbf{H}\tilde{\mathbf{S}}_D) &= \tilde{\mathbf{P}}\begin{bmatrix}\mathbf{H},\mathbf{J}\end{bmatrix} \begin{bmatrix} \mathbf{S}_D - \tilde{\mathbf{S}}_D\\ \mathbf{W}\!_D \end{bmatrix}. \label{eq:data95objective95rewritten} \end{align}\tag{16}\] Since \(\tilde{\mathbf{P}}\in\mathscr{G}_{B-I}(\mathbb{C}^B)\), it can null a matrix only if that matrix has rank less than or equal to \(I\). But by assumption, the matrix \([\mathbf{H},\mathbf{J}]\) has full column rank \(U+I\). So its product with \([\mathbf{S}_D - \tilde{\mathbf{S}}_D; \mathbf{W}\!_D]\) can give a matrix whose rank does not exceed \(I\) only if the rank of \([\mathbf{S}_D - \tilde{\mathbf{S}}_D; \mathbf{W}\!_D]\) itself does not exceed \(I\). By our assumption that the jammer is not eclipsed, this can be the case only if \(\tilde{\mathbf{S}}_D = \mathbf{S}_D\), in which case 16 simplifies to \(\tilde{\mathbf{P}}\mathbf{J}\mathbf{W}\!_D\). Since the jammer is not eclipsed, the rank of \(\mathbf{W}\!_D\) itself is equal to \(I\). Hence \(\textit{col}(\mathbf{J}\mathbf{W}\!_D)=\textit{col}(\mathbf{J})\), so that \(\tilde{\mathbf{P}}\mathbf{J}\mathbf{W}\!_D=\mathbf{0}\) if and only if \(\tilde{\mathbf{P}}=\mathbf{I}_B - \mathbf{J}\mathbf{J}^{\dagger}\). \(\blacksquare\)

7.2 Proof of↩︎

Since signals from the pilot phase play no role for this result, it will be convenient to simply call \(\mathbf{w}_D\) the jammer transmit signal (instead of “the transmit signal during the data phase”).

We start by defining the coset \[\begin{align} \mathfrak{E}(\mathbf{S}_D) \triangleq \Big\{ \mathbf{E}(\tilde{\mathbf{S}}_D; \mathbf{S}_D)=\mathbf{S}_D - \tilde{\mathbf{S}}_D : \tilde{\mathbf{S}}_D \in \mathcal{S}^{U\times D}\!\setminus\!\{\mathbf{S}_D\} \Big\}. \end{align}\] Note that, by definition, the coset \(\mathfrak{E}(\mathbf{S}_D)\) does not include the all-zero matrix. We can now rewrite as follows:

Definition 3 (Eclipsing with perfect CSI). A single-antenna jammer is eclipsed in a given frame if there exists a matrix \(\mathbf{E}(\tilde{\mathbf{S}}_D; \mathbf{S}_D)\in\mathfrak{E}(\mathbf{S}_D)\) such that \([\mathbf{E}(\tilde{\mathbf{S}}_D; \mathbf{S}_D); \mathbf{w}_D^{\text{T}}]\) is a matrix of rank \(1\).

Thus, the jammer eclipses if (i) \(\mathfrak{E}(\mathbf{S}_D)\) includes a matrix \(\mathbf{E}(\tilde{\mathbf{S}}_D; \mathbf{S}_D)\) whose rows are all collinear and (ii) \(\mathbf{w}_D^{\text{T}}\) is collinear with these rows. Without loss of generality, for all \(\mathbf{E}(\tilde{\mathbf{S}}_D; \mathbf{S}_D)\in\mathfrak{E}(\mathbf{S}_D)\), denote the \(u\)th row of \(\mathbf{E}(\tilde{\mathbf{S}}_D; \mathbf{S}_D)\) by \(\mathbf{e}_{(u)}(\tilde{\mathbf{s}}_{(u)};\mathbf{s}_{(u)})^{\text{T}}\), where \(\tilde{\mathbf{s}}_{(u)}^{\text{T}}\) and \(\mathbf{s}_{(u)}^{\text{T}}\) denote the corresponding rows of \(\tilde{\mathbf{S}}_D\) and \(\mathbf{S}_D\), respectively. We then define the cosets \[\begin{align} \mathfrak{e}_{(u)}(\mathbf{s}_{(u)}) \!\triangleq \! \big\{\mathbf{e}_{(u)}(\tilde{\mathbf{s}}_{(u)};\mathbf{s}_{(u)})\!=\tilde{\mathbf{s}}_{(u)}\!-\mathbf{s}_{(u)} : \tilde{\mathbf{s}}_{(u)}\!\!\in\!\mathcal{S}^D\!\setminus\!\{\mathbf{s}_{(u)}\} \!\big\} \end{align}\] for \(u=1,\dots,U\). Note that, by definition, \(\mathfrak{e}_{(u)}(\mathbf{s}_{(u)})\) does not include the all-zero vector. We have the following lemmas:

Lemma 8. A single-antenna jammer with transmit signal \(\mathbf{w}_D\) eclipses if and only if, for some \(u\in[1:U]\), there exists an \(\mathbf{e}\in\mathfrak{e}_{(u)}(\mathbf{s}_{(u)})\) that is collinear with \(\mathbf{w}_D\).

Proof. We start with the “only if” direction: The \(\mathfrak{e}_{(u)}(\mathbf{s}_{(u)})\), \(u=[1:U],\) contain the rows of the elements of \(\mathfrak{E}(\mathbf{S}_D)\) which are distinct from zero. Thus, if there exists no \(\mathbf{e}\in\mathfrak{e}_{(u)}(\mathbf{s}_{(u)})\) that is collinear with \(\mathbf{w}_D\)—for any \(u\in[1:U]\)—then for any matrix \(\mathbf{E}\) which contains at least one row from \(\bigcup_{u=1}^U \mathfrak{e}_{(u)}(\mathbf{s}_{(u)})\), the augmented matrix \([\mathbf{E}; \mathbf{w}_D^{\text{T}}]\) has at least least rank \(2\). Since all matrices \(\mathbf{E}\in\mathfrak{E}(\mathbf{S}_D)\) contain at least one such row, it follows that the jammer does not eclipse.

For the “if” direction, we may assume that there exists a \(u\in[1:U]\) and an \(\mathbf{e}\in\mathfrak{e}_{(u)}(\mathbf{s})\) such that \(\mathbf{w}_D\) is collinear with \(\mathbf{e}\). We now construct a matrix \(\tilde{\mathbf{S}}_D\in\mathcal{S}^{U\times D}\) as follows: In all rows except the \(u\)th one, \(\tilde{\mathbf{S}}_D\) equals \(\mathbf{S}_D\); in the \(u\)th one, \(\tilde{\mathbf{S}}_D\) equals the corresponding row of \(\mathbf{S}_D\) minus \(\mathbf{e}^{\text{T}}\). By being constructed this way, \(\tilde{\mathbf{S}}_D\) has the properties that (i) \(\mathbf{S}_D-\tilde{\mathbf{S}}_D\in\mathfrak{E}(\mathbf{S}_D)\), and (ii) \([\mathbf{S}_D-\tilde{\mathbf{S}}_D;\mathbf{w}_D^{\text{T}}]=[\mathbf{0}^{\text{T}};\dots;\mathbf{0}^{\text{T}};\mathbf{e}^{\text{T}};\mathbf{0}^{\text{T}};\dots;\mathbf{0}^{\text{T}};\mathbf{w}_D^{\text{T}}]\) has rank one. Thus, the jammer is eclipsed. ◻

Lemma 9.

Let \(\mathbf{w}_D\) be the transmit signal of a single-antenna jammer, let \(\mathbf{s}\sim\textrm{Unif}[\mathcal{S}^D]\) be a random vector, and define \[\begin{align} \mathfrak{e}(\mathbf{s}) \triangleq \big\{\tilde{\mathbf{s}}-\mathbf{s}: \tilde{\mathbf{s}}\in\mathcal{S}^D\setminus\{\mathbf{s}\} \big\}. \label{eq:ecoset} \end{align}\qquad{(4)}\] Let furthermore \(u\in[1:U]\) be fixed. Then the following two events have identical probabilities:

  1. There is a \(\mathbf{e}\!\in\!\mathfrak{e}_{(u)}(\mathbf{s}_{(u)})\) such that \(\mathbf{w}_D\) is collinear with \(\mathbf{e}\).

  2. There is a \(\mathbf{e}\!\in\!\mathfrak{e}(\mathbf{s})\) such that \(\mathbf{w}_D\) is collinear with \(\mathbf{e}\).

Proof. Since \(\mathbf{s}_{(u)}\sim\textrm{Unif}[\mathcal{S}^D]\) for any \(u\in[1:U]\) (which follows since the transmit symbols are assumed to be i.i.d. uniform, cf. ) , the random sets \(\mathfrak{e}_{(u)}(\mathbf{s}_{(u)})\) and \(\mathfrak{e}(\mathbf{s})\) have the same distribution. The result follows immediately. ◻

and give rise to the following corollary:

Corollary 10. Let \(\mathbf{w}_D\) be the transmit signal of a single-antenna jammer. If \(\mathbf{s}\sim\textrm{Unif}[\mathcal{S}^D]\) and \(\mathfrak{e}(\mathbf{s})\) is defined as in ?? , and if \[\begin{align} \!\!\! q(\mathbf{w}_D) \!\triangleq\! \mathop{\mathrm{\mathbb{P}}}(\exists \mathbf{e}\in\mathfrak{e}(\mathbf{s}) \text{ such that } \mathbf{w}_D \text{ is collinear with } \mathbf{e}),\!\! \label{eq:define95q} \end{align}\qquad{(5)}\] then the jammer eclipses with probability \(1-(1-q(\mathbf{w}_D))^U\).

Proof. By , the jammer eclipses if and only if there exists a \(\mathbf{e}\in\mathfrak{e}_{(u)}(\mathbf{s}_{(u)})\) that is collinear with \(\mathbf{w}_D\), for some \(u\in[1:U]\). Eclipsing corresponds therefore to the union of these \(U\) events. Since the \(U\) rows of \(\mathbf{S}_D\) are independent (cf. ), these events are independent, and by , they all have probability \(q(\mathbf{w}_D)\). The result follows immediately. ◻

To complete the proof of , it now only remains to show that the probability \(q(\mathbf{w}_D)\) as defined in ?? is at most \[\begin{align} \frac{2^{\|\mathbf{w}_D\|_0}-1}{4^{\|\mathbf{w}_D\|_0-1}}. \end{align}\] For this, we define the concept of an eclipsing-optimal jammer.

Definition 4. An eclipsing-optimal jammer* is a jammer which, for a given zero-norm \(\|\mathbf{w}_D\|_0\), transmits a signal with maximal probability of eclipsing, i.e., a signal from the set \[\begin{align} \mathop{\mathrm{arg\;max}}_{\mathbf{w}'\in\mathbb{C}^D:\|\mathbf{w}'\|_0=\|\mathbf{w}_D\|_0} q(\mathbf{w}'). \end{align}\]*

Lemma 11. To be eclipsing-optimal, a single-antenna jammer has to transmit a signal \(\mathbf{w}_D\in\hat{\mathcal{W}}_\alpha^D\) for some \(\alpha\in\mathbb{C}\), where \(\mathcal{W}_\alpha\triangleq \{0, \alpha , \alpha i, -\alpha , -\alpha i\}\).

Proof. By , for \(\mathbf{s}\sim\text{Unif}[\mathcal{S}^D]\), the probability of eclipsing is increasing in the probability that there exists an \(\mathbf{e}\in\mathfrak{e}(\mathbf{s})\) such that \(\mathbf{w}\) is collinear with \(\mathbf{e}\). The entries of any element \(\mathbf{e}\in\mathfrak{e}(\mathbf{s})\) take value in the set \(\Delta\mathcal{S}\triangleq \{s-\tilde{s} : s,\tilde{s} \in \mathcal{S}\}\) illustrated in . We have \[\begin{align} \!\!\!\!\Delta\mathcal{S}=& \hat{\mathcal{W}}_{\sqrt2} \nonumber \\ &\! \cup \big\{\sqrt2(1\!+\!i), \sqrt2(1\!-\!i), \sqrt2(-1\!+\!i), \sqrt2(-1\!-\!i)\big\}.\!\!\! \end{align}\] So the \(\mathbf{w}_D\in\mathbb{C}^D\) that lead to nonzero probability of eclipsing are exactly those contained in \(\alpha\Delta\mathcal{S}^D\triangleq\{\alpha\Delta\mathbf{s}:\Delta\mathbf{s}\in\Delta\mathcal{S}^D\}\) for some \(\alpha\in\mathbb{C}\). However, not all \(\mathbf{e}\in\Delta\mathcal{S}^D\) are contained in \(\mathfrak{e}(\mathbf{s})\) with equal probability, and so not all \(\mathbf{w}_D\in\alpha\Delta\mathcal{S}^D\) lead to the same probability of eclipsing. Define \[\begin{align} \mathfrak{e}_k(s_k) \triangleq \{e_k : \mathbf{e}\in\mathfrak{e}(\mathbf{s}) \} \label{eq:coset95entries} \end{align}\tag{17}\] to be the \(k\)th “entry” of \(\mathfrak{e}(\mathbf{s})\), which depends on \(\mathbf{s}\) only through \(s_k\).Vice versa, \(\mathfrak{e}(\mathbf{s})=\{\mathbf{e}:e_k\in\mathfrak{e}_k(s_k),k=1,\dots,D\}\). For all \(\mathbf{e}\in\Delta\mathcal{S}^D\setminus\{\boldsymbol{0}\}\), we have \[\begin{align} \Pr(\mathbf{e}\in\mathfrak{e}(\mathbf{s})) = \prod_{k=1}^D\Pr(e_k\in\mathfrak{e}_k(s_k)). \label{eq:e95prod} \end{align}\tag{18}\] The reason why not all \(\mathbf{w}_D\in\alpha\Delta\mathcal{S}^D\) have the same probability of eclipsing is that, for given \(k\), \(\mathfrak{e}_k(s_k)\) does not contain all elements of \(\Delta\mathcal{S}\) with equal probability, cf. :

First, we have \(P(0\in\mathfrak{e}_k(s_k))=1\). That is, the origin is contained in \(\mathfrak{e}_k(s_k)\) with probability one (i.e, for all possible realizations of \(s_k\)), as it can be “reached” (by subtracting some \(\tilde{s}_k\in\mathcal{S}\)) from any symbol \(s_k\in\mathcal{S}\), by setting \(\tilde{s}_k = s_k\).19 Note that by transmitting a zero, the jammer does not increase \(\|\mathbf{w}\|_0\).

The four points \(\sqrt2, i\sqrt2, -\sqrt2, -i\sqrt2\) on the coordinate axes are each contained in \(\mathfrak{e}_k(s_k)\) with probability \(\frac{1}{2}\), as they can be reached from \(s_k\) if \(s_k\) takes on either of the two constellation values adjacent to that value of the coordinate semi-axis.

Finally, the four corner points \(\sqrt2, i\sqrt2, -\sqrt2, -i\sqrt2\) are each contained in \(\mathfrak{e}_k(s_k)\) with probability \(\frac{1}{4}\), as they can be reached from \(s_k\) only if \(s_k\) takes on the constellation value in the corresponding quadrant.

So, to be optimal for some \(\|\mathbf{w}\|_0\geq0\), a jammer should restrict its transmit alphabet to the zero symbol and (a scaled version of) the four points \(\sqrt2, i\sqrt2, -\sqrt2, -i\sqrt2\). That is, the jammer should restrict its alphabet to \(\hat{\mathcal{W}}_\alpha\) for some \(\alpha\in\mathbb{C}\). ◻

Figure 7: Illustration of \mathcal{S} and \Delta\mathcal{S}. Values in \Delta\mathcal{S} which are contained in \mathfrak{e}_k(s_k) with probability 1 are circumscribed with a drawn red line; values that are contained with probability \frac{1}{2} are circumscribed with a dashed red line; values that are contained with probability \frac{1}{4} are circumscribed with a dotted red line.

Lemma 12. For an optimal jammer, the probability \(q\) as defined in ?? is equal to \[\begin{align} q(\mathbf{w}_D) = \begin{cases} 1 &\text{if } \mathbf{w}_D=\mathbf{0} \\ \frac{2^{\|\mathbf{w}_D\|_0}-1}{4^{\|\mathbf{w}_D\|_0-1}} &\text{else}. \end{cases} \end{align}\]

Proof. Without loss of generality, an optimal jammer uses the transmit alphabet \(\hat{\mathcal{W}}_{\sqrt{2}}=\{0,\sqrt{2},i\sqrt{2},-\sqrt{2},-i\sqrt{2}\}\). First, if the jammer only transmits zeros, \(\mathbf{w}_D=\mathbf{0}\), then \(\mathbf{w}_D\) is always collinear with any \(\mathbf{e}\in\mathfrak{e}(\mathbf{s})\), and hence \(q(\mathbf{w}_D)=1\). In contrast, if \(\mathbf{w}_D\neq\mathbf{0}\), then we can ignore the zeros in the jammer’s transmit signal for our analysis: Let \(\mathbf{w}\in\mathbb{C}^L\) be a jammer transmit signal (of arbitrary length \(L\)) without zeros, and if \(\mathbf{w}'\in\mathbb{C}^{L'}\) with \(L'>L\) is a zero-padded version of \(\mathbf{w}\), with \(L'-L\) zeros inserted at arbitrary indices \(\mathcal{I}\subset[1:L']\). If and only if \(\mathbf{w}\) is collinear with some \(\mathbf{e}\in\mathfrak{e}(\mathbf{s})\), then \(\mathbf{w}'\) is collinear with \(\mathbf{e}'\), where \(\mathbf{e}'\in\mathbb{C}^{L'}\) is the zero-padded version of \(\mathbf{e}\) (with zeros inserted at indices \(\mathcal{I}\)) and is contained in \(\mathfrak{e}(\mathbf{s}')\) for any \(\mathbf{s}'\in\mathcal{S}^{L'}\) that satisfies \(\mathbf{s}'_{[\mathcal{I}]}=\mathbf{s}\). So we have both \(q(\mathbf{w})=q(\mathbf{w}')\) and \(\|\mathbf{w}\|_0=\|\mathbf{w}'\|_0\). It follows that we can ignore the presence of zero-valued entries in the jammer’s transmit signal.

Furthermore, note that the \(\mathfrak{e}_k(s_k)\) as defined in 17 are independent and identically distributed for all \(k\) (since the \(s_k\) are i.i.d.). Therefore, and since the \(s_k\) are uniformly distributed, we can assume without loss of generality that all non-zero symbols that the jammer transmits are equal to \(\sqrt2\).

If the jammer only transmits one non-zero symbol, \(w_1=\sqrt{2}\), then for any realization of \(s_1\), we have eclipsing: An \(\mathbf{e}\in\mathfrak{e}(\mathbf{s})\) that is collinear with \(\mathbf{w}\) can always be found by setting \(\tilde{s}_1\in\mathcal{S}\setminus\{s_1\}\) and \(\tilde{s}_k = s_k\) for any \(k\neq1\).

If the jammer transmits two or more nonzero symbols \(w_\ell=\sqrt2\), \(\ell=1,\dots,L\), \(L\geq2\), then we only have eclipsing if all \(s_\ell, \ell=1,\dots,L\) lie on the same half-plane of the constellation. More precisely: An \(\mathbf{e}\in\mathfrak{e}(\mathbf{s})\) that is collinear with \([w_1,\dots,w_{L}]\) exists if and only if all \(s_\ell, \ell=1,\dots,L\) lie on the same half-plane of the constellation. To get the probability of this event, we can simply count the allowable combinations, since all realizations of \(\mathbf{s}\) are equally likely: There are four different half-planes (left, right, top, bottom), and in each half-plane, there are \(2^{L}=2^{\|\mathbf{w}_D\|_0}\) different sequences. But in counting like this, we count each of the four constant sequences \(e^{i\frac{\pi}{4}}\mathbf{1}, e^{i\frac{3\pi}{4}}\mathbf{1}, e^{i\frac{5\pi}{4}}\mathbf{1}, e^{i\frac{7\pi}{4}}\mathbf{1}\) twice (e.g., \(e^{i\frac{\pi}{4}}\mathbf{1}\) is counted both for the right and the top half-plane), so we need to subtract these four double-counted realizations. This gives us \(4\cdot2^{\|\mathbf{w}_D\|_0}-4\) realizations of \(\mathbf{s}\) that lead to eclipsing, out of \(4^{\|\mathbf{w}_D\|_0}\) sequences in total. So the probability \(q(\mathbf{w}_D)\) is \[\begin{align} q(\mathbf{w}_D) = \frac{4\cdot2^{\|\mathbf{w}\|_0}-4}{4^{\|\mathbf{w}\|_0}} = \frac{2^{\|\mathbf{w}_D\|_0}-1}{4^{\|\mathbf{w}_D\|_0-1}}. \end{align}\] ◻

now follows from and , where the approximation for large \(\|\mathbf{w}\|_0\) follows by approximating \(q(\mathbf{w}_D)\) as \(4\cdot2^{-\|\mathbf{w}\|_0}\) and then using a first-order Taylor approximation around \(q(\mathbf{w}_D)=0\). \(\blacksquare\)

7.3 Proof of↩︎

Using the theorem’s assumptions and 10 , we rewrite the argument of the optimization objective in 11 as \[\begin{align} & \tilde{\mathbf{P}}(\mathbf{Y}\!_D - \hat{\mathbf{H}}\tilde{\mathbf{S}}_D) \\ &= \tilde{\mathbf{P}}(\mathbf{H}\mathbf{S}_D + \mathbf{J}\mathbf{W}\!_D - (\mathbf{H}+ \mathbf{J}\mathbf{W}_T\mathbf{S}_T^{\dagger}) \tilde{\mathbf{S}}_D) \\ &= \tilde{\mathbf{P}}\begin{bmatrix}\mathbf{H},\mathbf{J}\end{bmatrix} \begin{bmatrix} \mathbf{S}_D - \tilde{\mathbf{S}}_D\\ \mathbf{W}\!_D - \mathbf{W}_T\mathbf{S}_T^{\dagger}\tilde{\mathbf{S}}_D \end{bmatrix} \end{align}\] From here on, the proof is very similar to the one of (considering the modified notion of eclipsing as defined in ), and so is omitted. \(\blacksquare\)

7.4 Proof of↩︎

Assume without loss of generality that the jammer knows the pilot sequence of the first UE, i.e., the first row of \(\mathbf{S}_T\), which we denote \(\mathbf{s}_{T,1}^{\text{T}}\). Then the jammer can transmit \(\mathbf{w}_T=\mathbf{s}_{T,1}\) and \(\mathbf{w}_D\in\mathcal{S}^D\). We therefore have \(\mathbf{w}_T^{\text{T}}\mathbf{S}_T^{\dagger}=[1,0,\dots,0]\) and so \[\begin{align} \mathbf{\Sigma}&= [\mathbf{S}_D-\tilde{\mathbf{S}}_D;\mathbf{w}_D^{\text{T}} - \mathbf{w}_T^{\text{T}}\mathbf{S}_T^{\dagger}\tilde{\mathbf{S}}_D] \\ &= [\mathbf{S}_D-\tilde{\mathbf{S}}_D;\mathbf{w}_D^{\text{T}} - \tilde{\mathbf{s}}_{D,1}^{\text{T}}], \end{align}\] where \(\tilde{\mathbf{s}}_{D,1}^{\text{T}}\) is the first row of \(\tilde{\mathbf{S}}_D\). In that case, consider for \(\tilde{\mathbf{S}}_D\) the matrix which is equal to \(\mathbf{S}_D\) on all rows except the first row, where it is equal to \(\mathbf{w}_D^{\text{T}}\). If \(\mathbf{w}_D^{\text{T}}\neq\mathbf{s}_{D,1}^{\text{T}}\), then \(\tilde{\mathbf{S}}_D\neq\mathbf{S}_D\) and \(\mathbf{\Sigma}\) is a matrix of rank one, meaning that the jammer eclipses.

Since \(\mathbf{s}_{D,1}^{\text{T}}\) is drawn uniform at random from \(\mathcal{S}\), the probability that \(\mathbf{w}_D^{\text{T}}\neq\mathbf{s}_{D,1}^{\text{T}}\) is equal to \(1-4^{-D}\). \(\blacksquare\)

7.5 Proof of↩︎

If the jammer does not jam during the pilot phase, \(\mathbf{w}_T=\mathbf{0}\), then eclipsing with channel estimation coincides with eclipsing with perfect CSI (cf. ), and the result follows from .

If the jammer does jam during the pilot phase, \(\mathbf{w}_T\neq\mathbf{0}\), then \(\mathbf{w}_T\) is independent of \(\mathbf{S}_T\) (since the jammer does not know \(\mathbf{S}_T\)). We now focus on the last row of \(\mathbf{\Sigma}\), which is \(\mathbf{w}_D^{\text{T}}-\mathbf{w}_T^{\text{T}}\mathbf{S}_T^{\dagger}\tilde{\mathbf{S}}_D\). Since \(\mathbf{S}_T\) is Haar distributed (and therefore unitary) up to a scale-factor, \(\mathbf{S}_T^{\dagger}=\frac{1}{T}\mathbf{S}_T^{\text{H}}\) is Haar distributed up to a scale-factor, too. Hence, \(\mathbf{x}^{\text{T}}\triangleq\mathbf{w}_T^{\text{T}}\mathbf{S}_T^{\dagger}\) is uniformly distributed over the complex \(U\)-dimensional sphere of radius \(\|\mathbf{w}_T\|_2/T\) [28]. We can therefore also write \(\mathbf{x}= \frac{\|\mathbf{w}_T\|_2}{T \|\mathbf{z}\|_2}\mathbf{z}\), where \(\mathbf{z}\sim\mathcal{C}\mathcal{N}(\mathbf{0},\mathbf{I}_U)\), and rewrite the last row of \(\mathbf{\Sigma}\) as \(\mathbf{w}_D^{\text{T}} - \frac{\|\mathbf{w}_T\|_2}{T \|\mathbf{z}\|_2}\mathbf{z}^{\text{T}}\tilde{\mathbf{S}}_D\). Here, \(\mathbf{z}^{\text{T}}\tilde{\mathbf{S}}_D\) is a (transposed) complex Gaussian vector with covariance matrix \(\tilde{\mathbf{S}}_D^{\text{H}}\tilde{\mathbf{S}}_D\), so that its entries are complex circularly-symmetric scalar Gaussians with variance \(U\) (corresponding to the energy of the columns of \(\tilde{\mathbf{S}}_D\)). Hence, the last row \(\mathbf{w}_D^{\text{T}}-\mathbf{w}_T^{\text{T}}\mathbf{S}_T^{\dagger}\tilde{\mathbf{S}}_D\) has full support \(D\) with probability one. Since this row does not depend on \(\mathbf{S}_D\), we can simply refer to , with the full-support (with probability one) vector \(\mathbf{w}_D^{\text{T}}-\mathbf{w}_T^{\text{T}}\mathbf{S}_T^{\dagger}\tilde{\mathbf{S}}_D\) in lieu of the potentially sparse vector \(\mathbf{w}_D^{\text{T}}\), and the result follows. \(\blacksquare\)

7.6 Proof of↩︎

This theorem was already proved for the special case of a single-antenna jammer in [15]. The extension to multi-antenna jammers is straightforward using arguments as in the proofs of and and so is omitted. \(\blacksquare\)

7.7 Proof of↩︎

We first prove convexity in \(\tilde{\mathbf{S}}_D\) for fixed \(\tilde{\mathbf{P}}\): For fixed \(\tilde{\mathbf{P}}\), the objective in 13 can be rewritten as \[\begin{align} \!\!\! \hat{f}(\tilde{\mathbf{S}}_D) \triangleq& \mathop{\mathrm{\mathbb{T} r}}\big(\mathbf{Y}\!_D^{\text{H}}\tilde{\mathbf{P}}\mathbf{Y}\!_D \nonumber \\ & + \tilde{\mathbf{S}}_D^{\text{H}}[\hat{\mathbf{H}}^{\text{H}}\tilde{\mathbf{P}}\hat{\mathbf{H}}-\alpha \mathbf{I}_U]\tilde{\mathbf{S}}_D - 2\Re\{\mathbf{Y}\!_D^{\text{H}}\tilde{\mathbf{P}}\hat{\mathbf{H}}\tilde{\mathbf{S}}_D\}\big).\!\! \label{eq:traceform} \end{align}\tag{19}\] \(\hat{f}\) is a real-valued function of the complex matrix \(\tilde{\mathbf{S}}_D\). Following [46], we represent the dependence on this complex input matrix by using both \(\tilde{\mathbf{S}}_D\) and \(\tilde{\mathbf{S}}_D^\ast\), giving rise to the four different Hessian matrices \(\mathcal{H}_{\tilde{\mathbf{S}}_D, \tilde{\mathbf{S}}_D^\ast} \hat{f}, \mathcal{H}_{\tilde{\mathbf{S}}_D^\ast, \tilde{\mathbf{S}}_D^\ast} \hat{f}, \mathcal{H}_{\tilde{\mathbf{S}}_D, \tilde{\mathbf{S}}_D} \hat{f}\), and \(\mathcal{H}_{\tilde{\mathbf{S}}_D^\ast, \tilde{\mathbf{S}}_D} \hat{f}\) of \(\hat{f}\). The objective is convex if the matrix \[\begin{align} \mathcal{H}\hat{f} = \begin{bmatrix} \mathcal{H}_{\tilde{\mathbf{S}}_D, \tilde{\mathbf{S}}_D^\ast} \hat{f} & \mathcal{H}_{\tilde{\mathbf{S}}_D^\ast, \tilde{\mathbf{S}}_D^\ast} \hat{f} \\ \mathcal{H}_{\tilde{\mathbf{S}}_D, \tilde{\mathbf{S}}_D} \hat{f}& \mathcal{H}_{\tilde{\mathbf{S}}_D^\ast, \tilde{\mathbf{S}}_D} \hat{f} \end{bmatrix} \end{align}\] is positive semidefinite. Recognizing that the second-order derivatives of the constant and affine terms in 19 are zero, and following [46], we obtain \[\begin{align} \mathcal{H}\hat{f} = \begin{bmatrix} [\hat{\mathbf{H}}^{\text{H}}\tilde{\mathbf{P}}\hat{\mathbf{H}}\!-\!\alpha \mathbf{I}_U]^{\text{T}}\mathop{\mathrm{\otimes}}\mathbf{I}_U & \mathbf{0} \\ \mathbf{0} & [\hat{\mathbf{H}}^{\text{H}}\tilde{\mathbf{P}}\hat{\mathbf{H}}\!-\!\alpha \mathbf{I}_U] \mathop{\mathrm{\otimes}}\mathbf{I}_U \end{bmatrix}. \!\! \end{align}\] \(\mathcal{H}\hat{f}\) is positive semidefinite if all its eigenvalues are non-negative. The eigenvalues of \(\mathcal{H}\hat{f}\) are simply the eigenvalues of \(\hat{\mathbf{H}}^{\text{H}}\tilde{\mathbf{P}}\hat{\mathbf{H}}\!-\!\alpha \mathbf{I}_U\), and they are all non-negative as long as \(\alpha\leq \lambda_\text{min}\), where \(\lambda_\text{min}\) is the smallest eigenvalue of \(\hat{\mathbf{H}}^{\text{H}}\tilde{\mathbf{P}}\hat{\mathbf{H}}\).

We now turn to the minimization with respect to \(\tilde{\mathbf{P}}\), and define \(\mathbf{E}\triangleq \mathbf{Y}\!_D - \hat{\mathbf{H}}\tilde{\mathbf{S}}_D\). Note that any \(\tilde{\mathbf{P}}\in \mathscr{G}_{B-I}(\mathbb{C}^B)\) can be written as \(\tilde{\mathbf{P}}= \mathbf{I}_B - \mathbf{Q}\mathbf{Q}^{\text{H}}\), where \(\mathbf{Q}\in\mathbb{C}^{B\times I}\) consists of \(I\) orthonormal columns. The second term of 13 does not depend on \(\tilde{\mathbf{P}}\), and since the squaring of the norm does not affect the minimizing argument (which is what we are interested in), it suffices to consider \[\begin{align} \min_{\tilde{\mathbf{P}}\,\in\, \mathscr{G}_{B-I}(\mathbb{C}^B)} \| \tilde{\mathbf{P}}\mathbf{E}\|_F &= \min_{\mathbf{Q}} \| \mathbf{E}- \mathbf{Q}\mathbf{Q}^{\text{H}}\mathbf{E}\|_F \tag{20} \\ &\geq \min_{\tilde{\mathbf{E}}\,:\,\mathop{\mathrm{rank}}{\tilde{\mathbf{E}}}\leq I} \|\mathbf{E}- \tilde{\mathbf{E}}\|_F. \tag{21} \end{align}\] Here, 21 follows since the matrix \(\mathbf{Q}\mathbf{Q}^{\text{H}}\mathbf{E}\) has at most rank \(I\) (since \(\mathbf{Q}\) has dimensions \(B\times I\)), so the optimization range in 20 is a subset of the optimization range in 21 . By the Eckart-Young-Mirksy theorem, 21 is minimized for \(\tilde{\mathbf{E}}= \mathbf{U}_I \boldsymbol{\Sigma}_I \mathbf{V}_I^{\text{H}}\), where \(\boldsymbol{\Sigma}_I = \text{diag}(\sigma_1, ..., \sigma_I)\) is the diagonal matrix whose diagonal entries are the \(I\) largest singular values of \(\mathbf{E}\), and \(\mathbf{U}_I\) and \(\mathbf{V}_I\) consist of the corresponding left- and right-singular vectors, respectively [47]. But if we let \(\mathbf{E}=\mathbf{U}\boldsymbol{\Sigma}\mathbf{V}^{\text{H}}\) be the singular-value decomposition of \(\mathbf{E}\) and choose \(\mathbf{Q}=\mathbf{U}_I\), then \(\mathbf{Q}\mathbf{Q}^{\text{H}}\mathbf{E}= \mathbf{U}_I \boldsymbol{\Sigma}_I \mathbf{V}_I^{\text{H}}\), meaning that 21 holds with equality and that 20 is minimized for \(\mathbf{P}=\mathbf{I}_B - \mathbf{U}_I \mathbf{U}_I^{\text{H}}\). \(\blacksquare\)

References↩︎

[1]
G. Marti and C. Studer, “Joint jammer mitigation and data detection for smart, distributed, and multi-antenna jammers,” in Proc. IEEE Int. Conf. Commun. (ICC), May 2023, pp. 1382–1387.
[2]
5G Threat Model Working Panel, “Potential threat vectors to 5G infrastructure,” CISA, NSA, and DNI.
[3]
T. Harrison, K. Johnson, M. Young, N. Wood, and A. Goessler, “Space threat assessment 2022,” Center Strategic Int. Studies (CSIS), Apr. 2022.
[4]
“The unmanned future is at risk,” Indian Defense Review, Aug. 2021. [Online]. Available: http://www.indiandefencereview.com/spotlights/the-unmanned-future-is-at-risk/.
[5]
“Satellite-navigation systems such as GPS are at risk of jamming,” The Economist, May 2021. [Online]. Available: https://www.economist.com/science-and-technology/2021/05/06/satellite-navigation-systems-such-as-gps-are-at-risk-of-jamming.
[6]
H. Pirayesh and H. Zeng, “Jamming attacks and anti-jamming strategies in wireless networks: A comprehensive survey,” IEEE Commun. Surveys Tuts., vol. 9, no. 2, pp. 767–809, Mar. 2022.
[7]
W. Shen, P. Ning, X. He, H. Dai, and Y. Liu, MCR decoding: A MIMO approach for defending against wireless jamming attacks,” in Proc. IEEE Conf. Commun. Netw. Security (CNS), Oct. 2014, pp. 133–138.
[8]
G. Marti, O. Castañeda, and C. Studer, “Jammer mitigation via beam-slicing for low-resolution mmWave massive MU-MIMO,” IEEE Open J. Circuits Syst., vol. 2, pp. 820–832, Dec. 2021.
[9]
Q. Yan, H. Zeng, T. Jiang, M. Li, W. Lou, and Y. T. Hou, “Jamming resilient communication using MIMO interference cancellation,” IEEE Trans. Inf. Forensics Security, vol. 11, no. 7, pp. 1486–1499, Jul. 2016.
[10]
T. T. Do, E. Björnsson, E. G. Larsson, and S. M. Razavizadeh, “Jamming-resistant receivers for the massive MIMO uplink,” IEEE Trans. Inf. Forensics Security, vol. 13, no. 1, pp. 210–223, Jan. 2018.
[11]
H. Akhlaghpasand, E. Björnsson, and S. M. Razavizadeh, “Jamming suppression in massive MIMO systems,” IEEE Trans. Circuits Syst. II, vol. 68, no. 1, pp. 182–186, Jan. 2020.
[12]
T. T. Nguyen and K.-K. Nguyen, “Anti-jamming in cell free mMIMO systems,” in Proc. IEEE Global Commun. Conf. (GLOBECOM).IEEE, Dec. 2021, pp. 1–6.
[13]
Y. Léost, M. Abdi, R. Richter, and M. Jeschke, “Interference rejection combining in LTE networks,” Bell Labs Tech. J., vol. 17, no. 1, pp. 25–50, Jun. 2012.
[14]
X. Jiang, X. Liu, R. Chen, Y. Wang, F. Shu, and J. Wang, “Efficient receive beamformers for secure spatial modulation against a malicious full-duplex attacker with eavesdropping ability,” IEEE Trans. Veh. Technol., vol. 70, no. 2, pp. 1962–1966, 2021.
[15]
G. Marti, T. Kölle, and C. Studer, “Mitigating smart jammers in multi-user MIMO,” IEEE Trans. Signal Process., vol. 71, pp. 756–771, 2023.
[16]
L. M. Hoang, J. A. Zhang, D. N. Nguyen, X. Huang, A. Kekirigoda, and K.-P. Hui, “Suppression of multiple spatially correlated jammers,” IEEE Trans. Veh. Technol., vol. 70, no. 10, pp. 10 489–10 500, Sep. 2021.
[17]
H. Zeng, C. Cao, H. Li, and Q. Yan, “Enabling jamming-resistant communications in wireless MIMO networks,” in Proc. IEEE Conf. Commun. Netw. Security (CNS), Oct. 2017, pp. 1–9.
[18]
L. Yang, X. Jiang, F. Shu, W. Zhang, and J. Wang, “Estimation of covariance matrix of interference for secure spatial modulation against a malicious full-duplex attacker,” IEEE Trans. Veh. Technol., vol. 71, no. 8, pp. 9050–9054, Aug. 2022.
[19]
L. M. Hoang, D. Nguyen, J. A. Zhang, and D. T. Hoang, “Multiple correlated jammers suppression: A deep dueling Q-learning approach,” in Proc. IEEE Wireless Commun. Netw. Conf. (WCNC), Apr. 2022, pp. 998–1003.
[20]
Z. Gülgün, E. Björnson, and E. G. Larsson, “Is massive MIMO robust against distributed jammers?” IEEE Trans. Commun., vol. 69, no. 1, pp. 457–469, Jan. 2021.
[21]
J. Vinogradova, E. Björnsson, and E. G. Larsson, “Detection and mitigation of jamming attacks in massive MIMO systems using random matrix theory,” in Proc. IEEE Int. Workshop Signal Process. Advances Wireless Commun. (SPAWC), Jul. 2016.
[22]
F. Bucheli, O. Castañeda, G. Marti, and C. Studer, “A jammer-mitigating 267Mb/s 3.78mm\(^{\text{2}}\) 583mW 32\(\times\)8 multi-user MIMO receiver in 22FDX,” in IEEE Symp. VLSI Technol. Circ., Jun. 2024.
[23]
G. Marti and C. Studer, “Single-antenna jammers in MIMO-OFDM can resemble multi-antenna jammers,” IEEE Commun. Lett., vol. 27, no. 11, pp. 3103–3107, Nov. 2023.
[24]
T. C. Clancy, “Efficient OFDM denial: Pilot jamming and pilot nulling,” in Proc. IEEE Int. Conf. Commun. (ICC), Jun. 2011, pp. 1–5.
[25]
S. Sodagari and T. C. Clancy, “Efficient jamming attacks on MIMO channels,” in IEEE Int. Conf. Commun. (ICC), Jun. 2012, pp. 852–856.
[26]
C. Shahriar, S. Sodagari, and T. C. Clancy, “Performance of pilot jamming on mimo channels with imperfect synchronization,” in Proc. IEEE Int. Conf. Commun. (ICC), Jun. 2012, pp. 898–902.
[27]
T. T. Nguyen and K.-K. Nguyen, “Pilot-partitioning protocol and anti-jamming methods in distributed massive MIMO systems,” IEEE Trans. on Cogn. Commun. Netw., 2023.
[28]
E. S. Meckes, The random matrix theory of the classical compact groups.Cambridge University Press, 2019, vol. 218.
[29]
A. Lapidoth and P. Narayan, “Reliable communication under channel un-certainty,” IEEE Trans. Inf. Theory, vol. 44, no. 6, pp. 2148–2177, 1998.
[30]
M. Grötschel, L. Lovász, and A. Schrijver, Geometric algorithms and combinatorial optimization.Springer, 2012.
[31]
T. Goldstein, C. Studer, and R. G. Baraniuk, “A field guide to forward-backward splitting with a FASTA implementation,” Feb. 2016. [Online]. Available: https://arxiv.org/abs/1411.3406.
[32]
N. Parikh and S. Boyd, “Proximal algorithms,” Found. Trends Optim., vol. 1, no. 3, pp. 127–239, Jan. 2014.
[33]
E. Liberty, Algorithms in Data Mining, Lecture 7: Singular Value Decomposition,” Lecture Notes, Tel Aviv University, Fall 2013.
[34]
J. Barzilai and J. M. Borwein, “Two-point step size gradient methods,” IMA J. Numer. Anal., vol. 8, no. 1, pp. 141–148, 1988.
[35]
B. Zhou, L. Gao, and Y.-H. Dai, “Gradient methods with adaptive step-sizes,” Comp. Optimization Applicat., vol. 35, no. 1, pp. 69–86, Mar. 2006.
[36]
H. Vikalo, B. Hassibi, and P. Stoica, “Efficient joint maximum-likelihood channel estimation and signal detection,” IEEE Trans. Wireless Commun., vol. 5, no. 7, pp. 1838–1845, Jul. 2006.
[37]
W. Xu, M. Stojnic, and B. Hassibi, “On exact maximum-likelihood detection for non-coherent MIMO wireless systems: a branch-estimate-bound optimization framework,” in Proc. IEEE Int. Symp. Inf. Theory (ISIT), Jul. 2008, pp. 2017–2021.
[38]
E. Kofidis, C. Chatzichristos, and A. L. de Almeida, “Joint channel estimation/data detection in MIMO-FBMC/OQAM systems–a tensor-based approach,” in Proc. Eur. Signal Process. Conf. (EUSIPCO), Aug. 2017, pp. 420–424.
[39]
B. B. Yilmaz and A. T. Erdogan, “Channel estimation for massive MIMO: A semiblind algorithm exploiting QAM structure,” in Proc. Asilomar Conf. Signals, Syst., Comput., Nov. 2019, pp. 2077–2081.
[40]
H. He, C.-K. Wen, S. Jin, and G. Y. Li, “Model-driven deep learning for MIMO detection,” IEEE Trans. Signal Process., vol. 68, pp. 1702–1715, Feb. 2020.
[41]
S. Jaeckel, L. Raschkowski, K. Börner, and L. Thiele, QuaDRiGa: A 3-D multi-cell channel model with time evolution for enabling virtual field trials,” IEEE Trans. Antennas Propag., vol. 62, no. 6, pp. 3242–3256, Jun. 2014.
[42]
3GPP, 3GPP TR 38.901,” Mar. 2022, version 17.0.0.
[43]
3GPP, 5G; NR; base station (BS) radio transmission and reception,” Mar. 2021, TS 38.104 version 17.1.0 Rel. 17.
[44]
G. Marti, A. Stutz-Tirri, and C. Studer, “Fundamental limits for jammer-resilient communication in finite-resolution MIMO,” in Proc. Asilomar Conf. Signals, Syst., Comput., Oct. 2024, pp. 1–8.
[45]
G. Marti and C. Studer, “Universal MIMO jammer mitigation via secret temporal subspace embeddings,” in Proc. Asilomar Conf. Signals, Syst., Comput., Oct. 2023, pp. 309–316.
[46]
A. Hjørungnes, Complex-Valued Matrix Derivatives: With Applications in Signal Processing and Communications.Cambridge Univ. Press, 2011.
[47]
C. Eckart and G. Young, “The approximation of one matrix by another of lower rank,” Psychometrika, vol. 1, no. 3, pp. 211–218, 1936.

  1. A short version of this paper has been presented at IEEE ICC 2023 [1].↩︎

  2. The work of GM and CS was supported in part by an ETH Research Grant.↩︎

  3. GM and CS are with the Department of Information Technology and Electrical Engineering, ETH Zurich, Switzerland; e-mail: gimarti@ethz.ch and studer@ethz.ch↩︎

  4. Another hard-to-mitigate threat is posed by multiple single-antenna distributed jammers, which also cause high-rank interference [20], [21].↩︎

  5. In particular, this insight has enabled the first application-specific integrated circuit (ASIC) implementation of an algorithm that mitigates smart jammers in multi-user MIMO [22].↩︎

  6. Our methods are also translatable to other MIMO contexts. For example, an extension to MIMO-OFDM in frequency-selective channels is possible but not straightforward, see [23].↩︎

  7. The model in 1 can also represent distributed single- or multi-antenna jammers with a total of \(I\) antennas. We consider this case in .↩︎

  8. For the moment, we neglect estimation of the UE channel matrix \(\mathbf{H}\) and assume perfect channel knowledge. Channel estimation is considered in . In this section, signals from the pilot phase 4 are therefore ignored.↩︎

  9. Also, it is evident from our proofs that the combinatorics involved in the analysis are quite involved already for single-antenna jammers, and we expect them to significantly increase in difficulty for multi-antenna jammers.↩︎

  10. This definition of eclipsing coincides with [15] for \(I\!=\!1\) and so can be seen as a generalization of the concept of eclipsing to multi-antenna jammers.↩︎

  11. As in , we can characterize this probability analytically for the single-antenna jammer case, and we refer to the simulations of for showing the efficacy against multi-antenna jammers empirically.↩︎

  12. Note that the following result is much stronger than [15], which assumes that both \(\mathbf{w}_D\) and \(\mathbf{w}_T\) are distinct from zero, and which gives a nontrivial bound only if the number of UEs satisfies \(U>3\).↩︎

  13. Because of the projection \(\tilde{\mathbf{P}}\) in 12 , the optimal value \(\hat{\mathbf{H}}\) of \(\tilde{\mathbf{H}}\) itself is not uniquely determined—only its composition with the projection \(\hat{\mathbf{P}}\) is.↩︎

  14. This relaxation, which allows us to use gradient-based methods for optimization, is only temporary—ultimately, entrywise rounding has to be used for converting a penultimate estimate \(\tilde{\mathbf{S}}_D\in\mathcal{C}^{U\times D}\) to one that is contained in \(\mathcal{S}^{U\times D}\). For simplicity, our algorithms omit this detail.↩︎

  15. Apart from the number of jammer antennas, the MAED version provided in [15] differs slightly from the one provided in this paper. The conference version of this paper [1] has therefore referred to the newer version of MAED as “MAED 2.0.” For simplicity, however, we have decided to return to the original name “MAED,” since these minor differences between the algorithms do not seem to justify a new name.↩︎

  16. Spreading the training period over the frame is the most sensible strategy against dynamic multi-antenna jammers as considered in , so this gives the strongest possible baseline.↩︎

  17. Since the jammer is perfectly nulled using ground-truth knowledge, there is no residual jamming interference. However, each jammer antenna entails the loss of one degree of freedom after nulling, see [44].↩︎

  18. POS-BOX and POS-JED often perform even worse than the unmitigated method (see, e.g., the pilot jammer ). This is due to the fact that the nonlinear data detectors of POS-BOX and POS-JED try to fit the receive signal with a signal model that is accurate only if the jammer subspace is correctly estimated, and which otherwise can fail catastrophically. In contrast, the linear detection of the unmitigated baseline is less susceptible to model mismatches.↩︎

  19. This is also the reason why 18 does not hold for \(\mathbf{e}=\boldsymbol{0}\): the right-hand-side is equal to \(1\), but the left-hand-side is equal to \(0\), cf. ?? .↩︎