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

Channel Estimation in MIMO Systems with
One-bit Spatial Sigma-delta ADCs

R.S. Prasobh Sankar, , and Sundeep Prabhakar Chepuri The authors are with the Department of Electrical Communication Engineering, Indian Institute of Science, Bangalore, India. Email:{rsprasobh,spchepuri}@iisc.ac.in. This work was supported in part by Nokia Faculty Research Award (NSN Oy, Espoo, Finland) and MHRD, India. The conference precursor of this paper appeared in the 46th IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP), June 2021 [1].
Abstract

This paper focuses on channel estimation in single-user and multi-user MIMO systems with multi-antenna base stations equipped with 1-bit spatial sigma-delta analog-to-digital converters (ADCs). A careful selection of the quantization voltage level and phase shift used in the feedback loop of 1-bit sigma-delta ADCs is critical to improve its effective resolution. We first develop a quantization noise model for 1-bit spatial sigma-delta ADCs. Using the developed noise model, we then present a two-step channel estimation algorithm to estimate a multipath channel parameterized by the gains, angles of arrival (AoAs), and angles of departure (AoDs). Specifically, in the first step, the AoAs and path gains are estimated using uplink pilots, which excite all the angles uniformly. Next, in the second step, the AoDs are estimated by progressively refining uplink beams through a recursive bisection procedure. For this algorithm, we propose a technique to select the quantization voltage level and phase shift. Through numerical simulations, we demonstrate that with the proposed parametric channel estimation algorithm, MIMO systems with 1-bit spatial sigma-delta ADCs perform significantly better than those with regular 1-bit ADCs and are on par with MIMO systems with high-resolution ADCs.

Index Terms:
Angular channel model, channel estimation, mmWave MIMO, 1-bit quantization, quantization noise modeling, spatial sigma-delta ADC.

I Introduction

Millimeter wave (mmWave) multiple-input multiple-output (MIMO) systems have become very popular for sensing and wireless communications beyond 5G [2, 3, 4]. While the abundant spectrum available at the mmWave frequency bands enables higher cellular data rates and precise positioning, links at mmWave frequencies are very sensitive to blockages and have significantly higher path loss. These issues are alleviated by beamforming with very large antenna arrays, typically packed in small areas. MIMO systems operating at mmWave frequencies, commonly referred to as massive MIMO systems, are either single-user MIMO (SU-MIMO) systems with multi-antenna user equipment (UE) and a base station (BS) having a large antenna array or multi-user MIMO (MU-MIMO) systems with many single antenna UEs communicating with a BS having a large array.

High-resolution analog-to-digital converters (ADCs) and digital-to-analog converters (DACs) for every antenna in the array significantly increase the radio frequency (RF) complexity and power consumption of massive MIMO systems. Low-resolution quantizers (e.g., 1-bit) are thus preferred albeit their deteriorated performance [5, 6, 7]. Sigma-delta (ΣΔ\Sigma\Delta) quantization is a popular technique frequently used to increase the effective resolution of low-resolution quantizers [8]. In a 1-bit ΣΔ\Sigma\Delta quantizer, the time-domain signal is first oversampled at a rate significantly higher than the Nyquist rate. Then the difference between the input and the 1-bit quantized output, i.e., the quantization noise, is fed back in time by adding it to the input at the next time instance. This operation leads to noise shaping with the quantization noise pushed to higher temporal frequencies. This means that the effective quantization noise is negligible for a low-pass signal, and it would be as if the signal were quantized by a high-resolution quantizer. This classical architecture to increase the effective resolution of time-domain signals by using a simple 1-bit quantizer with feedback has been recently adapted to the spatial domain [9, 10] and is receiving steady attention for multi-antenna communications [11, 12, 13, 14, 15].

In a 1-bit spatial ΣΔ\Sigma\Delta quantizer, oversampling and feedback are performed in the spatial domain, i.e., across antennas. To perform spatial oversampling, the antenna elements of an array are placed less than half wavelength apart. The quantization noise of each antenna is fed back along with the input of the next antenna. Analogous to its temporal counterpart, a spatial ΣΔ\Sigma\Delta quantizer pushes and shapes the quantization noise to higher spatial frequencies away from the array broadside. In other words, the quantization noise at lower spatial frequencies in a spatial ΣΔ\Sigma\Delta quantizer is reduced as if it was arising from a higher-resolution quantizer [9, 10]. By introducing phase shifts to the quantization noise before feedback allows angle steering so that the quantization noise (respectively, the effective resolution) will be lower (respectively, higher) for signals arriving around the steering angle [12, 14]. Therefore, with angle steering, it is possible to obtain a higher effective resolution for signals of interest in a spatial sector of certain width centered around any desired angle. In addition, a careful selection of the quantization voltage level assigned to 1-bit quantized signals significantly improves the inference performance when working with spatial ΣΔ\Sigma\Delta quantizers.

In essence, 1-bit spatial ΣΔ\Sigma\Delta quantization is an attractive architecture for massive MIMO systems, requiring only 1-bit quantizers per antenna element with feedback across the elements. However, feedback and 1-bit quantization make the channel estimation required for beamforming and symbol detection very challenging. This work focuses on channel estimation in MIMO systems with 1-bit spatial ΣΔ\Sigma\Delta quantizers.

I-A Related prior works

For rich scattering environments with a large number of multipath components, the MIMO channel matrix does not have any apparent structure. Such unstructured channel models are useful for channels at sub-6 GHz frequency bands [14, 6]. In contrast, at mmWave frequencies, due to the extreme path loss, the mmWave channel matrix is sparse in the angular domain and can be parameterized with the angles of departure (AoD), angles of arrival (AoA), and the complex gain of each path. Such angular channel models are commonly used at mmWave frequencies, e.g., at 28 GHz [4, 3].

Channel estimation with unstructured models in MIMO systems with 1-bit or few-bit quantizers is typically performed by first linearizing the non-linear quantizer using the so-called Bussgang decomposition [16, 17, 6] followed by computing a linear minimum mean squared error (LMMSE) estimate of the MIMO channel matrix [6, 18]. Bussgang decomposition based techniques [6, 17] of linearizing low-resolution quantizers have also been extended to spatial ΣΔ\Sigma\Delta quantizers for MU-MIMO channel estimation with unstructured models [13, 14].

For channel estimation with angular models, Bussgang decomposition based methods are not useful as the channel correlation matrix required for computing the Bussgang decomposition, LMMSE estimate, or setting the quantization voltage level in spatial ΣΔ\Sigma\Delta quantizers is not available. This is because knowing the channel correlation matrix for angular channel models amounts to knowing the unknown parameters, namely, AoAs and AoDs that characterize the channel. Thus, for channel estimation in MIMO systems with 1-bit ADCs and angular models, techniques based on optimization to recover the missing amplitudes [19], sparse recovery [20], and deep learning [21] have been proposed.

To summarize, existing works on channel estimation in MIMO systems with 1-bit spatial ΣΔ\Sigma\Delta quantizers focus on unstructured models [13, 14], and they cannot be directly extended to angular channel models. Therefore, in this work, we focus on channel estimation with angular channel models in MIMO systems having 1-bit spatial ΣΔ\Sigma\Delta quantizers.

I-B Contributions and main results

This paper is an extension of the precursor [1], wherein we presented a parametric channel estimation technique for SU-MIMO systems with a single line-of-sight (LoS) path. In this work, we extend [1] in several aspects to estimate multipath channels in SU-MIMO and MU-MIMO systems by leveraging angle steering in spatial ΣΔ\Sigma\Delta quantizers and describe methods to choose the quantization voltage level. The major contributions and results are summarized as follows.

  • Quantization noise model: For channel estimation with angular models, as discussed before, Bussgang decomposition based linearization techniques cannot be used. Therefore, we derive a model for the quantization noise in 1-bit spatial ΣΔ\Sigma\Delta quantizers based on the deterministic input-output relation in 1-bit temporal ΣΔ\Sigma\Delta quantizers. Specifically, we derive a closed-form expression for the correlation matrix of the approximation error due to linearization of the 1-bit spatial ΣΔ\Sigma\Delta quantizer. To do so, we use one of the main results of the paper that for most of the antenna elements in a large array, the quantization noise is uncorrelated with the corresponding input.

  • Channel estimation and quantization voltage selection: Leveraging the proposed quantization noise model, we develop algorithms to estimate multipath channels admitting angular models for SU-MIMO and MU-MIMO systems. We use uplink pilots to perform channel estimation at the BS equipped with a 1-bit spatial ΣΔ\Sigma\Delta quantizer. For SU-MIMO and MU-MIMO systems, we present techniques to choose the quantizer voltage level, which, when not chosen carefully, leads to significant performance degradation due to the extreme quantization.

    For channel estimation in SU-MIMO systems, i.e., to estimate the AoAs at the BS, AoDs of the paths from the UE, and path gains, we propose a two-step channel estimation algorithm, which is computationally efficient and has low overhead. In Step 1, the multi-antenna UE omnidirectionally transmits pilot symbols to the BS, which estimates the AoAs using a Bartlett beamformer and the complex path gains using a weighted least squares estimator. Since the AoDs are not known in Step 1, the proposed omnidirectional transmission ensures that sufficient power reaches the BS via all the paths. In addition, it allows us to choose a suitable quantization voltage level, which is essential for path gain estimation. Next, in Step 2, to estimate the AoDs, the UE transmits precoded pilot symbols using a sequence of adaptively chosen beamformers from a codebook hierarchically. To reduce the overhead, we assume a 1-bit feedback link between the BS and UE. We also provide a method to choose a quantization voltage level in Step 2. We show that with angle steering and the proposed voltage level, the beampatterns (from the designed codebook) as seen at the output of the 1-bit spatial ΣΔ\Sigma\Delta quantizer is comparable to that at the input (i.e., without quantization), and thereby resulting in channel estimates that are on par with that of unquantized MIMO systems.

    We then specialize the SU-MIMO channel estimation algorithm for MU-MIMO systems to estimate the AoAs and path gains at the BS. Specifically, we reformulate the MU-MIMO channel estimation problem using orthogonal pilots and separately estimate the single-input multiple-output (SIMO) channels between each single antenna UE and multi-antenna BS with a 1-bit spatial ΣΔ\Sigma\Delta quantizer.

  • Performance: Through numerical simulations, performance of the proposed channel estimation algorithms, in terms of normalized mean squared error (NMSE), are found to be significantly better than that of the regular (non-ΣΔ\Sigma\Delta) 1-bit channel estimation algorithms and are comparable with MIMO systems having infinite resolution quantizers for most of the SNRs and for paths with angles not far from the array broadside. Performance of algorithms with 1-bit spatial ΣΔ\Sigma\Delta quantizers is limited for paths with angles away from the array broadside because of the quantization noise shaping. The proposed channel estimation algorithm also performs better than the state-of-the-art unstructured channel estimation algorithm for MIMO systems with 1-bit ΣΔ\Sigma\Delta quantizers [14], where we use as input to the unstructured channel estimation algorithm a realistic approximation of the channel correlation matrix.

Although we focus on channel estimation with angular models, the developed quantization noise model is useful for estimating unstructured channels using classical estimation techniques (e.g., using least squares) whenever unquantized channel correlation information is not available a priori.

I-C Organization and notation

The remainder of the paper is organized as follows. In Sections II and III, we describe 1-bit spatial ΣΔ\Sigma\Delta quantizers and model the quantization noise in 1-bit spatial ΣΔ\Sigma\Delta quantizers, respectively. In Section IV, we present the SU-MIMO and MU-MIMO system models, which we use for channel estimation. In Sections V and VI, we propose channel estimation algorithms for SU-MIMO and MU-MIMO systems, respectively. In Section VII, we discuss results from numerical experiments and conclude the paper in Section VIII.

Throughout the paper, we use lowercase letters to denote scalars and boldface lowercase (respectively, uppercase) to denote vectors (respectively, matrices). We use ()(\cdot)^{*}, ()T(\cdot)^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}, and ()H(\cdot)^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}} to denote complex conjugation, transpose, and Hermitian (i.e., complex conjugate transpose) operations, respectively. [𝐱]n[{{\mathbf{x}}}]_{n} or xnx_{n} denotes the nn-th entry of the vector 𝐱{{\mathbf{x}}}. 𝐀𝐁{\mathbf{A}}\odot{\mathbf{B}} denotes the Khatri-Rao (or columnwise Kronecker) product of matrices 𝐀{\mathbf{A}} and 𝐁{\mathbf{B}}. Since we restrict ourselves to a 1-bit spatial ΣΔ\Sigma\Delta quantizer only at the BS, henceforth, we simply refer to it as a 1-bit spatial ΣΔ\Sigma\Delta ADC.

Software to reproduce the results in this paper is available at https://ece.iisc.ac.in/~spchepuri/sw/SpatialSigmaDelta.zip.

II One-bit spatial sigma-delta ADC

In this section, we describe the architecture of a multi-channel first-order 1-bit spatial ΣΔ\Sigma\Delta ADC with angle steering. Let us denote the input and output of an NrN_{\rm r} channel 1-bit spatial ΣΔ\Sigma\Delta ADC, at time tt, as 𝐱(t)=[x1(t),x2(t),,xNr(t)]TNr{{\mathbf{x}}}(t)=[x_{1}(t),x_{2}(t),\dots,x_{N_{\rm r}}(t)]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}\in{\mathbb{C}}^{N_{\rm r}} and 𝐲(t)=[y1(t),y2(t),,yNr(t)]TNr{{\mathbf{y}}}(t)=[y_{1}(t),y_{2}(t),\dots,y_{N_{\rm r}}(t)]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}\in{\mathbb{C}}^{N_{\rm r}}, respectively.

Refer to caption
Figure 1: The 1-bit spatial ΣΔ\Sigma\Delta ADC architecture. 𝒬b\mathcal{Q}_{b} is the 1-bit quantizer with level bb and c\mathcal{L}_{c} is the amplitude limiter with level cc.

Quantization noise from each antenna is fed back along with the input of the next antenna to realize the ΣΔ\Sigma\Delta architecture in space. Due to the presence of the feedback, the quantization noise in a spatial ΣΔ\Sigma\Delta ADC may become unbounded, overloading the output [22, 12]. To prevent overloading, the input signal, xx\in{\mathbb{C}}, is clipped using c(){\mathcal{L}}_{c}(\cdot) with clipping level c>0c>0 as

c[x]=sign((x)){|(x)|}c+ȷsign((x)){|(x)|}c,{\mathcal{L}}_{c}[x]={\rm sign}(\Re(x))\left\{|\Re(x)|\right\}_{c}+\jmath\,{\rm sign}(\Im(x))\left\{|\Im(x)|\right\}_{c},

where the operator {x}c\left\{x\right\}_{c} is defined as {x}c=max{x,c}\left\{x\right\}_{c}=\max\{x,c\} and ȷ=1\jmath=\sqrt{-1}. The clipped signal is then quantized using a 1-bit quantizer 𝒬b[]{\mathcal{Q}}_{b}[\cdot] with quantization voltage level bb as

𝒬b[c[x]]=bsign((c[x]))+ȷbsign((c[x])).{\mathcal{Q}}_{b}[{\mathcal{L}}_{c}[x]]=b\,{\rm sign}(\Re({\mathcal{L}}_{c}[x]))+\jmath\,b\,{\rm sign}(\Im({\mathcal{L}}_{c}[x])).

Let rn(t)r_{n}(t) denote the input to channel nn of the ADC at time tt. Then the corresponding output yn(t)y_{n}(t) is given by

yn(t)=𝒬b[rn(t)]=rn(t)+en(t),y_{n}(t)={\mathcal{Q}}_{b}[r_{n}(t)]=r_{n}(t)+e_{n}(t), (1)

where en(t)=yn(t)rn(t)e_{n}(t)=y_{n}(t)-r_{n}(t) is the quantization error. The quantization error at channel nn is phase shifted by φ\varphi and added to the input of channel n+1n+1 as

rn(t)=c[xn(t)]eȷφen1(t),r_{n}(t)={\mathcal{L}}_{c}[x_{n}(t)]-e^{\jmath{\varphi}}e_{n-1}(t), (2)

where φ=2πdsinψ\varphi=2\pi d\sin{\psi} with ψ\psi being the steering angle and dd being the inter-element spacing in the array in wavelengths. Since there is no feedback to the first channel, e0(t)= 0e_{0}(t)=\leavevmode\nobreak\ 0. From (1) and (2), we have the recursion

yn(t)\displaystyle y_{n}(t) =𝒬b[k=1neȷ(nk)φc[xk(t)]\displaystyle={\mathcal{Q}}_{b}\left[\sum_{k=1}^{n}e^{\jmath(n-k){\varphi}}{\mathcal{L}}_{c}[x_{k}(t)]\right.
k=1n1eȷ(nk)φyk(t)].\displaystyle\hskip 85.35826pt\left.-\sum_{k=1}^{n-1}e^{\jmath(n-k){\varphi}}y_{k}(t)\right]. (3)

We are interested in estimating parameters underlying the input signal 𝐱(t){{\mathbf{x}}}(t) from the output of the 1-bit spatial ΣΔ\Sigma\Delta ADC 𝐲(t){{\mathbf{y}}}(t). This is a challenging problem because of the cascade of two non-linearities in (II). Moreover, the information loss introduced by the 1-bit spatial ΣΔ\Sigma\Delta ADC is mainly determined by the clipping level cc and the quantization level bb. For a given quantization level bb, we can prevent the system from overloading by choosing the clipping voltage cc as [12]

c=b(2|cos(φ)||sin(φ)|).c=b(2-|{\rm cos}(\varphi)|-|{\rm sin}(\varphi)|). (4)

A small value of cc leads to severe loss in the input information due to clipping, which cannot be compensated later by any choice of the quantization level bb. Hence, it is necessary to choose a sufficiently large cc to avoid information loss due to clipping. In this work, we propose to choose cc such that Pr(c[𝐱(t)]𝐱(t)2>δ)ϵ{\rm Pr}\left(\|{\mathcal{L}}_{c}[{{\mathbf{x}}}(t)]-{{\mathbf{x}}}(t)\|_{2}>\delta\right)\leq\epsilon for appropriately selected constants δ,ϵ>0\delta,\epsilon>0. See Section V-C for more details on the selection of cc. With such a choice of cc, the output of the clipper can be approximated as c[xn(t)]xn(t),n=1,2,,Nr{\mathcal{L}}_{c}[x_{n}(t)]\approx x_{n}(t),\>n=1,2,\ldots,N_{\rm r}. Then the 1-bit spatial ΣΔ\Sigma\Delta recursion in (II) simplifies to

𝐲(t)=𝒬b[𝐔𝐱(t)𝐕𝐲(t)],{{\mathbf{y}}}(t)={\mathcal{Q}}_{b}[{\mathbf{U}}{{\mathbf{x}}}(t)-{\mathbf{V}}{{\mathbf{y}}}(t)], (5)

where the Nr×NrN_{\rm r}\times N_{\rm r} lower triangular matrices 𝐔{\mathbf{U}} and 𝐕{\mathbf{V}} are defined as

𝐔=[1eȷφ1eȷ(Nr1)φeȷφ1]and𝐕=𝐔𝐈{\mathbf{U}}=\begin{bmatrix}1&&&&\\ e^{\jmath{\varphi}}&1&&&\\ \vdots&\ddots&\ddots&&\\ e^{\jmath(N_{\rm r}-1){\varphi}}&\cdots&e^{\jmath{\varphi}}&1\end{bmatrix}\quad\text{and}\quad{\mathbf{V}}={\mathbf{U}}-{\mathbf{I}}

with 𝐈{\mathbf{I}} being the Nr×NrN_{\rm r}\times N_{\rm r} identity matrix. The architecture of the first-order 1-bit spatial ΣΔ\Sigma\Delta ADC is shown in Fig. 1.

III Quantization noise modeling

Quantization is a non-linear and irreversible operation that makes its statistical analysis complicated. To simplify the analysis of a quantizer, the usual approach is to linearize the quantizer and account for the error due to linearization through additive noise, which is often assumed to be uniformly distributed and uncorrelated with the input of the quantizer [23, 22, 10, 11, 12]. However, this assumption is reasonable only for a multi-level quantizer with many levels and when the input has a sufficiently large dynamic range. This classical approach of modeling the approximation error due to linearization as additive uniform noise is not suitable for 1-bit quantizers. Therefore, in what follows, we develop a noise model for 1-bit spatial ΣΔ\Sigma\Delta ADCs.

To begin with, we first express the input-output relation of a spatial ΣΔ\Sigma\Delta ADC in terms of the floor[]{\rm floor}[\cdot] function by drawing inspiration from [22], where a similar expression for the deterministic error in a temporal 1-bit ΣΔ\Sigma\Delta quantizer was developed. Next, we propose to linearize the floor[]{\rm floor}[\cdot] function by interpreting it as a multi-level quantizer to compute the second-order statistics of the quantization noise that is useful for decorrelating observations when solving parametric estimation and detection problems involving 1-bit spatial ΣΔ\Sigma\Delta ADCs.

The error en(t)e_{n}(t) in (1) due to 1-bit spatial ΣΔ\Sigma\Delta quantization admits a closed-form expression as given in the next Lemma.

Lemma 1.

For a 1-bit spatial ΣΔ\Sigma\Delta ADC with ψ=0\psi=0, the quantization error as a function of the input is given by

(en(t))=b2b12(n1)+12bk=1n(xk(t))\displaystyle\Re({e_{n}(t)})=b-2b\left\langle\frac{1}{2}(n-1)+\frac{1}{2b}\sum\limits_{k=1}^{n}\Re(x_{k}(t))\right\rangle (6a)
(en(t))=b2b12(n1)+12bk=1n(xk(t))\displaystyle\Im({e_{n}(t)})=b-2b\left\langle\frac{1}{2}(n-1)+\frac{1}{2b}\sum\limits_{k=1}^{n}\Im(x_{k}(t))\right\rangle (6b)

for n=1,,Nrn=1,\ldots,N_{\rm{r}}. Here, \langle\cdot\rangle is the fractional part function.

The above expressions are derived in the appendix. Let us collect the output of all the channels of the 1-bit spatial ΣΔ\Sigma\Delta ADC in a vector 𝐲(t){{\mathbf{y}}}(t) and rewrite (5) as

𝐲(t)=𝐔𝐱(t)𝐕𝐲(t)+𝐞(t).{{\mathbf{y}}}(t)={\mathbf{U}}{{\mathbf{x}}}(t)-{\mathbf{V}}{{\mathbf{y}}}(t)+{{\mathbf{e}}}(t). (7)

From Lemma 1, we have

𝐞(t)=2bμ𝟏2b𝝂12b𝐔𝐱(t),{{\mathbf{e}}}(t)=2b\mu\boldsymbol{1}-2b\left\langle{\boldsymbol{\nu}}-\frac{1}{2b}{\mathbf{U}}{{\mathbf{x}}}(t)\right\rangle, (8)

where 𝝂=μ𝐕𝟏Nr{\boldsymbol{\nu}}=\mu{\mathbf{V}}\boldsymbol{1}\in{\mathbb{C}}^{N_{\rm r}} with μ=0.5+ȷ0.5\mu=0.5+\jmath 0.5, 𝟏\boldsymbol{1} is the length-NrN_{\rm r} column vector with all ones, and the fractional part function is applied elementwise. Using the fact that x=xfloor[x],x\langle x\rangle=x-{\rm floor}[x],\>\forall\>x\in\mathbb{R}, and from (7), the deterministic relation between the input and output of a 1-bit spatial ΣΔ\Sigma\Delta ADC is given by

0.5b1𝐔𝐲(t)+𝝂μ𝟏=floor[0.5b1𝐔𝐱(t)+𝝂],0.5b^{-1}{\mathbf{U}}{{\mathbf{y}}}(t)+{\boldsymbol{\nu}}-\mu\boldsymbol{1}={\rm floor}\left[0.5b^{-1}{\mathbf{U}}{{\mathbf{x}}}(t)+{\boldsymbol{\nu}}\right], (9)

where we have transformed the non-linearity due to the 1-bit quantization 𝒬b[]{\mathcal{Q}}_{b}[\cdot] to the non-linear floor[]{\rm floor[\cdot]} function.

Refer to caption
Figure 2: Quantization noise in a 128 channel 1-bit ΣΔ\Sigma\Delta ADCIn Fig. 2, the correlation and noise power spectrum are computed by averaging over 10410^{4} independent experiments in which we use the same simulation setting as in Fig. 4 in Section VII-A with a path impinging on the BS array from the angular sector [30,30][-30^{\circ},30^{\circ}] and with a signal-to-noise ratio of 0 dB.. (a) Correlation between the input and quantization noise of a regular one-bit and a one-bit spatial ΣΔ\Sigma\Delta ADC for different steering angles. (b) Quantization noise shaping.

Now we leverage the fact that the input-output relation of a floor[]{\rm floor}[\cdot] function is similar to that of a multi-level quantizer with a step size of one. For the spatial ΣΔ\Sigma\Delta ADC channels (corresponding to antenna elements) with larger indices that are away from the 11st channel, the dynamic range of the real and imaginary parts of [0.5b1𝐔𝐱(t)+𝝂]n[0.5b^{-1}{\mathbf{U}}{{\mathbf{x}}}(t)+{\boldsymbol{\nu}}]_{n} is large when compared to [0,1)[0,1). This means that for antenna elements with indices away from the 11st element, the floor[]{\rm floor}[\cdot] function acts like a multi-level quantizer with an input having a large dynamic range. Therefore, it is now reasonable to model the approximation error due to the linearization of the floor[]{\rm floor}[\cdot] function as additive noise, 𝐰(t)Nr{{\mathbf{w}}}(t)\in\leavevmode\nobreak\ \mathbb{C}^{N_{\rm r}}, which is uniformly distributed and uncorrelated with the input for channels having larger indices, i.e., we have

floor[0.5b1𝐔𝐱(t)+𝝂]=0.5b1𝐔𝐱(t)+𝝂+𝐰(t){\rm floor}\left[0.5b^{-1}{\mathbf{U}}{{\mathbf{x}}}(t)+{\boldsymbol{\nu}}\right]=0.5b^{-1}{\mathbf{U}}{{\mathbf{x}}}(t)+{\boldsymbol{\nu}}+{{\mathbf{w}}}(t)

with the real and imaginary parts of [𝐰(t)]n[0,1)[{{\mathbf{w}}}(t)]_{n}\in[0,1). Substituting in (9) yields

𝐲(t)=𝐱(t)+𝐪(t),\mathbf{y}(t)=\mathbf{x}(t)+{{\mathbf{q}}}(t), (10)

where 𝐪(t)=2b𝐔1(𝐰(t)+μ𝟏){{\mathbf{q}}}(t)=2b{\mathbf{U}}^{-1}({{{\mathbf{w}}}}(t)+\mu\boldsymbol{1}), which is an affine transformation of 𝐰(t){{\mathbf{w}}}(t), is also uniformly distributed and [𝐪(t)]n[{{\mathbf{q}}}(t)]_{n} is uncorrelated with the input [𝐱(t)]n[{{\mathbf{x}}}(t)]_{n} for larger channel (or antenna) indices nn. In massive MIMO systems with large number of antennas, there are many antennas for which xn(t)x_{n}(t) and [𝐪(t)]n[{{\mathbf{q}}}(t)]_{n} are uncorrelated. The correlation of [𝐪(t)]n=yn(t)xn(t){[{{\mathbf{q}}}(t)]}_{n}=y_{n}(t)-x_{n}(t) with xn(t)x_{n}(t), i.e., 𝔼[xn(ynxn)]\mathbb{E}[x_{n}(y_{n}-x_{n})^{*}] for an antenna array having Nr=128N_{\rm r}=128 elements with an inter-element spacing of one eighth the signal wavelength is illustrated in Fig. 2(a).

The covariance matrix of the noise vector 𝐪(t){{\mathbf{q}}}(t) in (10) is given by 𝐑q=𝔼[𝐪(t)𝐪H(t)]=2b23𝐔1𝐔-H{\mathbf{R}}_{q}={\mathbb{E}}\left[{{\mathbf{q}}}(t){{\mathbf{q}}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t)\right]=\frac{2b^{2}}{3}{\mathbf{U}}^{-1}{\mathbf{U}}^{\raisebox{1.0pt}{-$\rm\scriptscriptstyle H$}}. When ψ=0\psi=0, the matrix 𝐔1{\mathbf{U}}^{-1} with ones on the main diagonal, 1-1 on the first sub-diagonal, and zeros elsewhere, is a spatial high-pass filter, which shapes the quantization noise to higher spatial frequencies. We illustrate the angular power spectrum of the quantization noise, i.e., 𝔼[|𝐚BSH(θ)(𝐲(t)𝐱(t))|2]\mathbb{E}[|{{\mathbf{a}}}_{\rm BS}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\theta)({{\mathbf{y}}}(t)-{{\mathbf{x}}}(t))|^{2}], for different values of the steering angle in Fig. 2(b), where 𝐚BS(θ)=[1,eȷ2πdsin(θ),,eȷ(Nr1)2πdsin(θ)]T{{\mathbf{a}}}_{\rm BS}(\theta)=[1,e^{-\jmath 2\pi d{\rm sin}(\theta)},\ldots,e^{-\jmath(N_{\rm r}-1)2\pi d{\rm sin}(\theta)}]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}} is the steering vector of the uniform linear array (ULA) having NrN_{\rm r} elements at the BS with an inter-element spacing dd wavelengths. We can see that the quantization noise is very small for the directions around the steering angle.

Before we end this section, we make the following two remarks. Although in the development of the noise model, the spatial ΣΔ\Sigma\Delta ADC is steered to the array broadside with ψ=0\psi=0 for ease of exposition, introducing a phase shift in the feedback does not alter the uncorrelatedness between the input and the quantization noise as can be seen in Fig. 2(a). Next, for the floor[]{\rm floor}[\cdot] function in (9) to behave as a multi-level quantizer, the quantization voltage level bb plays an important role. When bb is very large, the dynamic range of the real and imaginary parts of the argument of the floor[]{\rm floor}[\cdot] function, i.e., 0.5b1𝐔𝐱(t)+𝝂0.5b^{-1}\mathbf{U}\mathbf{x}(t)+{\boldsymbol{\nu}}, can be very small for which the uniform noise or uncorrelated noise assumption might fail. Hence, carefully selecting bb (as discussed in Section V-C) becomes crucial.

Next, we use the developed spatial ΣΔ\Sigma\Delta signal and noise models to estimate MIMO channels with angular models.

IV Angular channel model

In this paper, we consider the two commonly encountered SU-MIMO and MU-MIMO settings in MIMO communications. In the SU-MIMO setting, a single UE with NtN_{\rm t} antennas communicates with a multi-antenna BS, whereas in the MU-MIMO setting, KK single antenna UEs communicate with a multi-antenna BS. In both these settings, the BS has a ULA with NrN_{\rm r} antennas and it receives and processes uplink training pilots.

Let us collect the uplink training pilots in 𝐒=[𝐬(1),𝐬(2),,𝐬(T)]{\mathbf{S}}=[{{\mathbf{s}}}(1),{{\mathbf{s}}}(2),\ldots,{{\mathbf{s}}}(T)], where TT is the pilot length. Without loss of generality, let us assume that the columns of the pilot matrix have unit norm. Let PP denote the total transmit power at the UE and 𝐇{\mathbf{H}} denote the MIMO channel matrix. The signal received at the BS prior to quantization, denoted as 𝐗=[𝐱(1),𝐱(2),,𝐱(T)]Nr×T{\mathbf{X}}=[{{\mathbf{x}}}(1),{{\mathbf{x}}}(2),\ldots,{{\mathbf{x}}}(T)]\in{\mathbb{C}}^{N_{\rm r}\times T}, can be expressed as

𝐗=P𝐇𝐒+𝐙,{\mathbf{X}}=\sqrt{P}{\mathbf{H}}{\mathbf{S}}+{\mathbf{Z}}, (11)

where PP is also the uplink SNR of the system and 𝐙Nr×T{\mathbf{Z}}\in{\mathbb{C}}^{N_{\rm r}\times T} is the additive white Gaussian receiver noise matrix with entries [𝐙]ij𝒞𝒩(0,1)[{\mathbf{Z}}]_{ij}\sim{\mathcal{C}}{\mathcal{N}}(0,1). The received signal is then quantized using an NrN_{\rm r}-channel 1-bit spatial ΣΔ\Sigma\Delta ADC to obtain 𝐘=[𝐲(1),𝐲(2),,𝐲(T)]Nr×T{\mathbf{Y}}=[{{\mathbf{y}}}(1),{{\mathbf{y}}}(2),\ldots,{{\mathbf{y}}}(T)]\in{\mathbb{C}}^{N_{\rm r}\times T}, which is given by

𝐘=𝒬b[𝐔𝐗𝐕𝐘]=𝒬b[P𝐔𝐇𝐒𝐕𝐘+𝐔𝐙].\displaystyle{\mathbf{Y}}={\mathcal{Q}}_{b}[{\mathbf{U}}{\mathbf{X}}-{\mathbf{V}}{\mathbf{Y}}]={\mathcal{Q}}_{b}[\sqrt{P}{\mathbf{U}}{\mathbf{H}}{\mathbf{S}}-{\mathbf{V}}{\mathbf{Y}}+{\mathbf{U}}{\mathbf{Z}}].

Next, we parameterize 𝐇{\mathbf{H}} by assuming a narrowband spatially sparse model for MU-MIMO and SU-MIMO systems.

IV-A SU-MIMO channel model

Suppose there are LL scatterers that result in LL paths between the UE and BS. Let us denote the AoDs of the LL paths from the UE by ϕ=[ϕ1,,ϕL]T\boldsymbol{\phi}=[\phi_{1},\ldots,\phi_{L}]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}, the AoAs of the LL paths at the BS by 𝜽=[θ1,,θL]T\boldsymbol{\theta}=[\theta_{1},\ldots,\theta_{L}]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}, and the complex path gains by 𝜶=[α1,,αL]T\boldsymbol{\alpha}=[\alpha_{1},\ldots,\alpha_{L}]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}. The MIMO channel matrix is then expressed in terms of these parameters as

𝐇\displaystyle{\mathbf{H}} =1Lk=1Lαk𝐚BS(θk)𝐚UEH(ϕk)\displaystyle=\frac{1}{\sqrt{L}}\sum_{k=1}^{L}\alpha_{k}{{\mathbf{a}}}_{\rm BS}(\theta_{k}){{\mathbf{a}}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\phi_{k})
=1L𝐀BS(𝜽)diag(𝜶)𝐀UEH(ϕ),\displaystyle=\frac{1}{\sqrt{L}}{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}){\rm diag}(\boldsymbol{\alpha}){\mathbf{A}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\boldsymbol{\phi}), (12)

where 𝐀BS(𝜽)=[𝐚BS(θ1),,𝐚BS(θL)]Nr×L{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta})=[{{\mathbf{a}}}_{\rm BS}(\theta_{1}),\ldots,{{\mathbf{a}}}_{\rm BS}(\theta_{L})]\in{\mathbb{C}}^{N_{\rm r}\times L} is the array manifold of the ULA at the BS and 𝐀UE(ϕ)=[𝐚UE(ϕ1),,𝐚UE(ϕL)]Nt×L{\mathbf{A}}_{\rm UE}(\boldsymbol{\phi})=[{{\mathbf{a}}}_{\rm UE}(\phi_{1}),\ldots,{{\mathbf{a}}}_{\rm UE}(\phi_{L})]\in{\mathbb{C}}^{N_{\rm t}\times L} is the array manifold at the UE. The columns of 𝐀UE(ϕ){\mathbf{A}}_{\rm UE}(\boldsymbol{\phi}) contain the array response vector of the critically spaced ULA at the UE with NtN_{\rm t} elements, and is given by 𝐚UE(ϕ)=[1,eȷπsin(ϕ),,eȷ(Nt1)πsin(ϕ)]T{{\mathbf{a}}}_{\rm UE}(\phi)=[1,e^{-\jmath\pi{\rm sin}(\phi)},\ldots,e^{-\jmath(N_{\rm t}-1)\pi{\rm sin}(\phi)}]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}. Similarly, the columns of 𝐀BS(𝜽){\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}) contain the array response vector of the oversampled ULA at the BS with an inter-element spacing dd wavelengths, and is given by 𝐚BS(θ)=[1,eȷ2πdsin(θ),,eȷ(Nr1)2πdsin(θ)]T{{\mathbf{a}}}_{\rm BS}(\theta)=[1,e^{-\jmath 2\pi d{\rm sin}(\theta)},\ldots,e^{-\jmath(N_{\rm r}-1)2\pi d{\rm sin}(\theta)}]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}.

IV-B MU-MIMO channel model

Suppose that there are KK single antenna users. The SIMO channel 𝐡kNr{{\mathbf{h}}}_{k}\in{\mathbb{C}}^{N_{\rm r}} between the kkth UE and the BS with LkL_{k} paths is given by

𝐡k=1Lkj=1Lkαk,j𝐚BS(θk,j)=1Lk𝐀BS(𝜽k)𝜶k,{{\mathbf{h}}}_{k}=\frac{1}{\sqrt{L_{k}}}\sum_{j=1}^{L_{k}}\alpha_{k,j}{{\mathbf{a}}}_{\rm BS}(\theta_{k,j})=\frac{1}{\sqrt{L_{k}}}{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}_{k})\boldsymbol{\alpha}_{k}, (13)

where 𝐀BS(𝜽k)=[𝐚BS(θk,1),,𝐚BS(θk,Lk)]Nr×Lk{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}_{k})=[{{\mathbf{a}}}_{\rm BS}(\theta_{k,1}),\ldots,{{\mathbf{a}}}_{\rm BS}(\theta_{k,L_{k}})]\in{\mathbb{C}}^{N_{\rm r}\times L_{k}} denotes the array manifold at the BS, 𝜽k=[θk,1,,θk,Lk]T\boldsymbol{\theta}_{k}=[\theta_{k,1},\ldots,\theta_{k,L_{k}}]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}} collects AoAs of paths from kkth UE at the BS, and 𝜶k=[αk,1,,αk,Lk]TLk\boldsymbol{\alpha}_{k}=[\alpha_{k,1},\ldots,\alpha_{k,L_{k}}]^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}\in{\mathbb{C}}^{L_{k}} denotes the corresponding complex path gains. Then the overall channel matrix for the MU-MIMO system is given by

𝐇=[𝐡1𝐡2𝐡K]Nr×K.{\mathbf{H}}=\begin{bmatrix}{{\mathbf{h}}}_{1}&{{\mathbf{h}}}_{2}&\ldots&{{\mathbf{h}}}_{K}\end{bmatrix}\in{\mathbb{C}}^{N_{\rm r}\times K}.

Our aim is to estimate the MIMO channels (12) and (13) by estimating the underlying angles (namely, the AoAs and AODs) and path gains from the output of the 1-bit spatial ΣΔ\Sigma\Delta ADC at the BS given the uplink pilots 𝐒{\mathbf{S}}, the SNR PP, and the number of paths.

V SU-MIMO channel estimation

In this section, we present the proposed algorithm for channel estimation with angular models in SU-MIMO systems having 1-bit spatial ΣΔ\Sigma\Delta ADCs at the BS. Specifically, we propose a two-step algorithm to estimate the channel parameters (𝜽,ϕ,𝜶)(\boldsymbol{\theta},\boldsymbol{\phi},\boldsymbol{\alpha}) from uplink pilots. In the first step, we estimate the AoAs (i.e., 𝜽\boldsymbol{\theta}) and the path gains (i.e., 𝜶\boldsymbol{\alpha}) using precoded uplink pilots, which excite all the angles uniformly. Next, in the second step, to estimate the AoDs, i.e., ϕ\boldsymbol{\phi}, we select precoders from a codebook using a recursive bisection procedure that leverages 1-bit feedback between the BS and UE to reduce the number of channel uses and thus the channel estimation overhead.

As we discuss later, it is crucial to carefully select the clipping voltage levels to benefit from the advantages of the 1-bit spatial ΣΔ\Sigma\Delta ADCs. The proposed two-step channel estimation procedure is developed keeping in mind the dependence of the unknown parameters on voltage level selection.

V-A Step 1: AoA and path gain estimation

V-A1 AoA estimation

To estimate the AoAs at the BS, the UE transmits precoded pilot symbols 𝐬(t)=𝐩1(t){{\mathbf{s}}}(t)={{\mathbf{p}}}_{1}(t) for t=1,,T1t=1,\ldots,T_{1}, where 𝐩1(t)Nt{{\mathbf{p}}}_{1}(t)\in{\mathbb{C}}^{N_{\rm t}} is the precoder with 𝐩1(t)2=1\|{{\mathbf{p}}}_{1}(t)\|_{2}=1 so that the total transmit power is PP. As we do not yet know the AoDs, instead of selecting 𝐩1(t){{\mathbf{p}}}_{1}(t) to focus energy in specific directions, we select it such that all the departure angles are excited uniformly as

𝐩1H(t)𝐚UE(ϕ~d)=1,d=1,2,,D,{{\mathbf{p}}}_{1}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t){{\mathbf{a}}}_{\rm UE}(\tilde{\phi}_{d})=1,\,d=1,2,\ldots,D, (14)

where 𝒟={ϕ~1,ϕ~2,,ϕ~D}{\mathcal{D}}=\{\tilde{\phi}_{1},\tilde{\phi}_{2},\ldots,\tilde{\phi}_{D}\} denotes the set of DD candidate AoDs. An obvious choice of 𝐩1(t){{\mathbf{p}}}_{1}(t) that satisfies (14) is

𝐩1(t)=𝐩1=[100]T,t=1,,T1.{{\mathbf{p}}}_{1}(t)={{\mathbf{p}}}_{1}=\begin{bmatrix}1&0&\ldots&0\end{bmatrix}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}},\,\,t=1,\ldots,T_{1}. (15)

This means that we turn off all the antennas at the UE except the first one to perform an omnidirectional transmission.

From (10), the symbols received at the BS in Step 1 can be compactly expressed as

𝐘1\displaystyle{\mathbf{Y}}_{1} =PL𝐀BS(𝜽)diag(𝜶)𝐀UEH(ϕ)𝐒+𝐍\displaystyle=\sqrt{\frac{P}{L}}{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}){\rm diag}(\boldsymbol{\alpha}){\mathbf{A}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\boldsymbol{\phi}){\mathbf{S}}+{\mathbf{N}}
=PL𝐀BS(𝜽)diag(𝜶)𝐄+𝐍,\displaystyle=\sqrt{\frac{P}{L}}{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}){\rm diag}(\boldsymbol{\alpha}){\mathbf{E}}+{\mathbf{N}}, (16)

where 𝐒=𝐩1𝟏TNt×T1{\mathbf{S}}={{\mathbf{p}}}_{1}{\bf 1}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}\in{\mathbb{C}}^{N_{\rm t}\times T_{1}} is the transmitted pilot matrix and the effective noise matrix 𝐍=[𝐧(1),,𝐧(T1)]Nr×T1{\mathbf{N}}=[{{\mathbf{n}}}(1),\cdots,{{\mathbf{n}}}(T_{1})]\in{\mathbb{C}}^{N_{\rm r}\times T_{1}} is defined as 𝐍=𝐙+𝐐{\mathbf{N}}={\mathbf{Z}}+{\mathbf{Q}} is the sum of the receiver noise 𝐙{\mathbf{Z}} and the quantization noise 𝐐=[𝐪(1),,𝐪(T1)]{\mathbf{Q}}=[{{\mathbf{q}}}(1),\cdots,{{\mathbf{q}}}(T_{1})]. The L×T1L\times T_{1} matrix 𝐄=𝐀UEH(ϕ)𝐒=𝐀UEH(ϕ)𝐩1𝟏TL×T1{\mathbf{E}}{=}{\mathbf{A}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\boldsymbol{\phi}){\mathbf{S}}={\mathbf{A}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\boldsymbol{\phi}){{\mathbf{p}}}_{1}{\bf 1}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}\in{\mathbb{C}}^{L\times T_{1}} has all one entries. Thus, the precoder in (15) makes 𝐘1{\mathbf{Y}}_{1} independent of the AoDs.

The AoAs can now be estimated from 𝐘1{\mathbf{Y}}_{1} using standard direction-finding techniques like Bartlett beamforming, minimum variance distortionless response (MVDR) beamforming [24], sparse recovery, or using a maximum likelihood based estimator. Since a massive MIMO BS typically has a large number (of the order of hundreds) of antennas, we may use a computationally less intensive method, such as the Bartlett beamforming for AoA estimation. Specifically, the AoA estimates, denoted by {θ^k}k=1L\{\hat{\theta}_{k}\}_{k=1}^{L}, are the locations of the LL local maxima of the Bartlett spatial spectrum

𝒥(θ)=𝐚BSH(θ)𝐑y1𝐚BS(θ),\mathcal{J}(\theta)={{\mathbf{a}}}_{\rm BS}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\theta){\mathbf{R}}_{y1}{{\mathbf{a}}}_{\rm BS}(\theta), (17)

where 𝐑y1=1T1𝐘1𝐘1H{\mathbf{R}}_{y1}=\frac{1}{T_{1}}{\mathbf{Y}}_{1}{\mathbf{Y}}_{1}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}} is the sample covariance matrix.

V-A2 Path gain estimation

Next, to estimate the path gains, we form the estimated array manifold at the BS as 𝐀^BS=[𝐚BS(θ^1),𝐚BS(θ^2),,𝐚BS(θ^L)]\hat{{\mathbf{A}}}_{\rm BS}=[{{\mathbf{a}}}_{\rm BS}(\hat{\theta}_{1}),{{\mathbf{a}}}_{\rm BS}(\hat{\theta}_{2}),\cdots,{{\mathbf{a}}}_{\rm BS}(\hat{\theta}_{L})] and approximate (16) as

𝐘^1=PL𝐀^BSdiag(𝜶)𝐄+𝐍.\hat{{\mathbf{Y}}}_{1}=\sqrt{\frac{P}{L}}\hat{{\mathbf{A}}}_{\rm BS}{\rm diag}(\boldsymbol{\alpha}){\mathbf{E}}+{\mathbf{N}}. (18)

Recall from the quantization noise modeling in Section III that the covariance matrix of the quantization noise is given by 𝐑q=2b23𝐔1𝐔-H{\mathbf{R}}_{q}=\frac{2b^{2}}{3}{\mathbf{U}}^{-1}{\mathbf{U}}^{\raisebox{1.0pt}{-$\rm\scriptscriptstyle H$}} and that the quantization noise is uncorrelated with the input. Thus the covariance matrix of the effective noise term in (16) is

𝐑n=𝔼[𝐧(t)𝐧H(t)]=𝐈+2b23𝐔1𝐔-H.{\mathbf{R}}_{n}=\mathbb{E}[{{\mathbf{n}}}(t){{\mathbf{n}}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t)]=\mathbf{I}+\frac{2b^{2}}{3}\mathbf{U}^{-1}\mathbf{U}^{\raisebox{1.0pt}{-$\rm\scriptscriptstyle H$}}. (19)

We now prewhiten the observations 𝐘^1\hat{{\mathbf{Y}}}_{1} to obtain

𝐘1=𝐑n1/2𝐘^1=PL𝐑n1/2𝐀^BSdiag(𝜶)𝐄+𝐍,{{\mathbf{Y}}}_{1}^{\prime}={\mathbf{R}}_{n}^{-1/2}\hat{{\mathbf{Y}}}_{1}=\sqrt{\frac{P}{L}}{\mathbf{R}}_{n}^{-1/2}\hat{{\mathbf{A}}}_{\rm BS}{\rm diag}(\boldsymbol{\alpha}){\mathbf{E}}+{\mathbf{N}}^{\prime},

where 𝐑n1/2{\mathbf{R}}_{n}^{-1/2} is the prewhitening matrix, which can be obtained using an eigenvalue decomposition of 𝐑n{\mathbf{R}}_{n} and 𝐍=𝐑n1/2𝐍{\mathbf{N}}^{\prime}={\mathbf{R}}_{n}^{-1/2}{\mathbf{N}} is the whitened noise term. Using the property that vec(𝐀diag(𝐛)𝐂)=(𝐂T𝐀)𝐛{\rm vec}({\mathbf{A}}{\rm{diag}({{\mathbf{b}}})}{\mathbf{C}})=({\mathbf{C}}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}\odot{\mathbf{A}}){{\mathbf{b}}}, we have vec(𝐘1)=𝚿𝜶+vec(𝐍){\rm vec}({{\mathbf{Y}}}_{1}^{\prime})=\boldsymbol{\Psi}\boldsymbol{\alpha}+{\rm vec}({\mathbf{N}}^{\prime}) with 𝚿=(𝐄TPL𝐑n1/2𝐀^BS)T1Nr×L\boldsymbol{\Psi}=({\mathbf{E}}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}\odot\sqrt{\frac{P}{L}}{\mathbf{R}}_{n}^{-1/2}\hat{{\mathbf{A}}}_{\rm BS})\in{\mathbb{C}}^{T_{1}N_{\rm r}\times L} and T1NrLT_{1}N_{\rm r}\gg L as the BS usually has large number of antennas and the mmWave MIMO channel is sparse in the angular domain. The path gains can then be estimated using least squares as

𝜶^=(𝚿H𝚿)1𝚿Hvec(𝐘1).\hat{\boldsymbol{\alpha}}=(\boldsymbol{\Psi}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}\boldsymbol{\Psi})^{-1}\boldsymbol{\Psi}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{\rm vec}({{\mathbf{Y}}}_{1}^{\prime}). (20)

V-B Step 2: AoD estimation

Now what remains is to estimate the AoDs. To do so, we propose a recursive bisection procedure, which divides the spatial sector into two subsectors at each stage and measures the power of the signal received in the direction corresponding to the estimated AoA. The method selects the subsector with the largest received power as the new sector to be used in the next bisection stage. This procedure is continued till the desired subsector resolution is obtained and is repeated for each of the estimated AoAs corresponding to LL paths.

Let us denote the precoder that we use in Step 2 by 𝐩{{\mathbf{p}}} with 𝐩=1\|{{\mathbf{p}}}\|=1, as before. Let us also define the inner product between 𝐚UE(ϕl){{\mathbf{a}}}_{\rm UE}(\phi_{l}) and 𝐩{{\mathbf{p}}} as ρl(𝐩)=𝐚UEH(ϕl)𝐩\rho_{l}({{\mathbf{p}}})={{\mathbf{a}}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\phi_{l}){{\mathbf{p}}}, where we have 0|ρl(𝐩)|Nt0\leq|\rho_{l}({{\mathbf{p}}})|\leq\sqrt{N_{\rm t}} and when 𝐩=1Nt𝐚UE(ϕl){{\mathbf{p}}}=\frac{1}{\sqrt{N_{\rm t}}}{{\mathbf{a}}}_{\rm UE}(\phi_{l}), the upper bound is achieved with equality. Therefore, by choosing a precoder 𝐩{{\mathbf{p}}} that yields the maximum ρl(𝐩)\rho_{l}({{\mathbf{p}}}), we can indirectly refine the AoD sector of the llth path to compute the AoD.

Suppose the UE transmits T2T_{2} precoded symbols 𝐒=𝐩𝟏TNt×T2{\mathbf{S}}={{\mathbf{p}}}{\bf 1}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}\in\mathbb{C}^{N_{\rm t}\times T_{2}}, where the subscript “2” denotes Step 2 of the algorithm. Using the combiner 𝐜l=1Nr𝐚BS(θl){{\mathbf{c}}}_{l}=\frac{1}{\sqrt{N_{\rm r}}}{{\mathbf{a}}}_{\rm BS}({\theta}_{l}), the component of the received uplink pilot arriving along the llth path at the BS can be expressed as

𝐜lH𝐘\displaystyle{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{\mathbf{Y}} =PLk=1Lαk𝐜lH𝐚BS(θk)𝐚UEH(ϕk)𝐩𝟏T+𝐜lH𝐍\displaystyle=\sqrt{\frac{P}{L}}\sum_{k=1}^{L}\alpha_{k}{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{{\mathbf{a}}}_{\rm BS}(\theta_{k}){{\mathbf{a}}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\phi_{k}){{\mathbf{p}}}\boldsymbol{1}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}+{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{\mathbf{N}}
=PLαl𝐜lH𝐚BS(θl)ρl(𝐩)𝟏T+𝐜lH𝐍\displaystyle=\sqrt{\frac{P}{L}}\alpha_{l}{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{{\mathbf{a}}}_{\rm BS}(\theta_{l})\rho_{l}({{\mathbf{p}}})\boldsymbol{1}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}+{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{\mathbf{N}}
+PLk=1,klLαk𝐜lH𝐚BS(θk)ρk(𝐩)𝟏T.\displaystyle\quad\quad\quad+\sqrt{\frac{P}{L}}\sum_{k=1,k\neq l}^{L}\alpha_{k}{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{{\mathbf{a}}}_{\rm BS}(\theta_{k})\rho_{k}({{\mathbf{p}}})\boldsymbol{1}^{\raisebox{1.2pt}{$\rm\scriptstyle T$}}. (21)

Since we have a large array at the BS and assuming that the AoAs are sufficiently separated, we can approximate 𝐜lH𝐚BS(θk)=1Nr𝐚BSH(θl)𝐚BS(θk)0{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{{\mathbf{a}}}_{\rm BS}(\theta_{k})=\frac{1}{\sqrt{N_{\rm r}}}{{\mathbf{a}}}_{\rm BS}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}({\theta}_{l}){{\mathbf{a}}}_{\rm BS}(\theta_{k})\approx 0 for lkl\neq k. Using this approximation and multiplying both sides of (21) with 𝟏{\bf 1}, we have

𝐜lH𝐘𝟏=PNrLT2αlρl(𝐩)+𝐜lH𝐍𝟏,{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{\mathbf{Y}}{\bf 1}=\sqrt{\frac{PN_{\rm r}}{L}}T_{2}\alpha_{l}\rho_{l}({{\mathbf{p}}})+{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{\mathbf{N}}{\bf 1}, (22)

which allows us to compute the energy of the T2T_{2} received symbols arriving from the llth path as

El(𝐩)=|1T2𝐜lH𝐘𝟏|2.E_{l}({{\mathbf{p}}})=\left|\frac{1}{T_{2}}{{\mathbf{c}}}_{l}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{\mathbf{Y}}{\bf 1}\right|^{2}. (23)

In practice, we compute El(𝐩)E_{l}({{{\mathbf{p}}}}) using the estimate θ^l\hat{\theta}_{l} from Step 1 to form the combiner as 𝐜l=1Nr𝐚BS(θ^l){{{\mathbf{c}}}}_{l}=\frac{1}{\sqrt{N_{\rm r}}}{{\mathbf{a}}}_{\rm BS}(\hat{\theta}_{l}).

Refer to caption
Figure 3: (a) First two stages of the precoders that partition the spatial sector into two and four subsectors. Corresponding beampatterns as seen at the BS (b) without any quantizer and (c) with 1-bit ΣΔ\Sigma\Delta ADC, but without any angle steering or voltage level selection. The UE has Nt=32N_{\rm t}=32 critically-spaced antennas and the BS has Nr=128N_{\rm r}=128 antennas spaced one eighth wavelengths apart.

We can estimate the AoDs by finding precoders, 𝐩{{\mathbf{p}}}, that maximize El(𝐩)E_{l}({{{\mathbf{p}}}}) for each path l=1,,Ll=1,\ldots,L, where recall that the structure of the precoder vector is constrained to 𝐩=1Nt𝐚UE(ϕ){{\mathbf{p}}}=\frac{1}{\sqrt{N_{\rm t}}}{{\mathbf{a}}}_{\rm UE}(\phi). Here, we assume that the number of paths, LL, is known. Performing this maximization by exhaustively searching over all the departure angles in a predefined grid 𝒟{{\mathcal{D}}} results in excessive training overhead. Therefore, we present a recursive bisection procedure, where for each path ll, we start with a wide beam and progressively refine it to maximize El(𝐩)E_{l}({{{\mathbf{p}}}}).

V-B1 Design of the precoder codebook

Before we describe the proposed recursive biscection procedure, we first provide the design of a codebook that contains precoders required for each stage of the procedure. Let us discretize the departure angular domain into DD grid points 𝒟={ϕ~1,ϕ~2,,ϕ~D}{\mathcal{D}}=\{\tilde{\phi}_{1},\tilde{\phi}_{2},\cdots,\tilde{\phi}_{D}\} by uniformly sampling the direction cosine space so that ϕ~d=sin1(1+2D1(d1))\tilde{\phi}_{d}={\rm sin}^{-1}\left(-1+\frac{2}{D-1}(d-1)\right) for d=1,2,,Dd=1,2,\ldots,D. Let us define the index set s,i{\mathcal{I}}_{s,i} as

s,i={D(i1)2s+1,D(i1)2s+2,,Di2s}{\mathcal{I}}_{s,i}=\left\{\frac{D(i-1)}{2^{s}}+1,\frac{D(i-1)}{2^{s}}+2,\ldots,\frac{Di}{2^{s}}\right\}

for i=1,2,,2si=1,2,\ldots,2^{s} with |s,i|=D2s|{\mathcal{I}}_{s,i}|=\frac{D}{2^{s}}. Next, we the partition the set 𝒟{\mathcal{D}} into 2s2^{s} partitions for the ssth stage of the recursive procedure with a total number of Ns=log2(D)N_{\rm s}={\rm log}_{2}(D) stages. In other words, at Stage ss, we have 2s2^{s} spatial sectors with the iith sector formed from angles ϕ~is,i\tilde{\phi}_{i}\in{\mathcal{I}}_{s,i} and in the last stage, i.e., at Stage NsN_{\rm s}, we have sectors with DD angular grid points in 𝒟\mathcal{D}. For example, in the first stage, we have two spatial sectors {ϕ~1,ϕ~2,,ϕ~D2}\{\tilde{\phi}_{1},\tilde{\phi}_{2},\cdots,\tilde{\phi}_{\frac{D}{2}}\} and {ϕ~D2+1,ϕ~D2+2,,ϕ~D}\{\tilde{\phi}_{\frac{D}{2}+1},\tilde{\phi}_{\frac{D}{2}+2},\cdots,\tilde{\phi}_{D}\}.

Let 𝐩s,i{{\mathbf{p}}}_{s,i} denote the iith precoder for the ssth stage. The precoders are designed to focus on a desired angular sector as

𝐚UEH(ϕ~j)𝐩s,i={1,js,i,0,otherwise{{\mathbf{a}}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\tilde{\phi}_{j}){{\mathbf{p}}}_{s,i}=\begin{cases}1,&\quad j\in{\mathcal{I}}_{s,i},\\ 0,&\quad\text{otherwise}\end{cases} (24)

for s=1,2,,Nss=1,2,\ldots,N_{\rm s} and i=1,2,,2si=1,2,\ldots,2^{s}. Defining the dictionary matrix 𝐃=[𝐚UE(ϕ~1),,𝐚UE(ϕ~D)]HD×Nt{\mathbf{D}}=[{{\mathbf{a}}}_{\rm UE}(\tilde{\phi}_{1}),\ldots,{{\mathbf{a}}}_{\rm UE}(\tilde{\phi}_{D})]^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}\in{\mathbb{C}}^{D\times N_{\rm t}} with DNtD\gg N_{\rm t}, and the precoder matrix 𝐏s=[𝐩s,1,𝐩s,2,,𝐩s,2s]Nt×2s{\mathbf{P}}_{s}=[{{\mathbf{p}}}_{s,1},{{\mathbf{p}}}_{s,2},\ldots,{{\mathbf{p}}}_{s,2^{s}}]\leavevmode\nobreak\ \in\leavevmode\nobreak\ \mathbb{C}^{N_{\rm t}\times 2^{s}}, we can compactly express (24) as the linear system

𝐃𝐏s=𝚿s,{\mathbf{D}}{\mathbf{P}}_{s}=\mbox{$\mbox{\boldmath$\Psi$}$}_{s}, (25)

where the desired beampattern matrix 𝚿s=[𝟙s,1,,𝟙s,2s]{1,0}D×2s\mbox{$\mbox{\boldmath$\Psi$}$}_{s}=[\mathbbm{1}_{s,1},\cdots,\mathbbm{1}_{s,2^{s}}]\in\{1,0\}^{D\times 2^{s}} with 𝟙s,i{0,1}D\mathbbm{1}_{s,i}\in\{0,1\}^{D} being the indicator vector with entries equal to one at locations indexed by the set s,i{\mathcal{I}}_{s,i}. Then the precoders for the ssth stage can be computed using least squares as

𝐏s=(𝐃H𝐃)1𝐃H𝚿s.{\mathbf{P}}_{s}=({\mathbf{D}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{\mathbf{D}})^{-1}{\mathbf{D}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}\mbox{$\mbox{\boldmath$\Psi$}$}_{s}. (26)

To obtain unit-norm precoding vectors, we normalize each column of 𝐏s{\mathbf{P}}_{s} to unity. We repeat this procedure for s=1,2,,Nss=1,2,\ldots,N_{\rm s} to compute precoders for all the stages. This completes the design of the precoder codebook.

In Fig. 3(a), we show the beampattern |𝐚UEH(ϕ)𝐩|2|{{\mathbf{a}}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\phi){{\mathbf{p}}}|^{2} for ϕ[0,2π]\phi\in[0,2\pi] corresponding to Stage 1 with 𝐩{𝐩1,1,𝐩1,2}{{\mathbf{p}}}\in\{{{\mathbf{p}}}_{1,1},{{\mathbf{p}}}_{1,2}\} and Stage 2 with 𝐩{𝐩2,1,𝐩2,2,𝐩2,3,𝐩2,4}{{\mathbf{p}}}\in\{{{\mathbf{p}}}_{2,1},{{\mathbf{p}}}_{2,2},{{\mathbf{p}}}_{2,3},{{\mathbf{p}}}_{2,4}\} of the codebook. In Figs. 3(b) and 3(c), we show the beampatterns corresponding to the first two stages as seen by the BS without any quantizer, i.e., |𝐜H𝐱|2\left|{{\mathbf{c}}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{{\mathbf{x}}}\right|^{2} and by the BS with 1-bit ΣΔ\Sigma\Delta ADC, i.e., |𝐜H𝐲|2\left|{{\mathbf{c}}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}{{\mathbf{y}}}\right|^{2}, respectively, where 𝐱=Pα𝐚BS(θ)𝐚UEH(ϕ)𝐩+𝐳{{\mathbf{x}}}=\sqrt{P}\alpha{{\mathbf{a}}}_{\rm BS}(\theta){{\mathbf{a}}}_{\rm UE}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(\phi){{\mathbf{p}}}+{{\mathbf{z}}} as in (11), 𝐲=𝐱+𝐪{{\mathbf{y}}}={{\mathbf{x}}}+{{\mathbf{q}}} as in (10), and we vary ϕ\phi and 𝐩{{\mathbf{p}}} as before. Here, we use θ=30\theta=30^{\circ}, 𝐜=1Nr𝐚BS(30){{\mathbf{c}}}=\frac{1}{\sqrt{N_{\rm r}}}{{\mathbf{a}}}_{\rm BS}(30^{\circ}), |α|=1|\alpha|=1, and SNR of 10 dB. We see that the received beampattern in Fig. 3(c) is significantly distorted as the 1-bit ΣΔ\Sigma\Delta ADC is not steered to the θ\theta, i.e., ψθ\psi\neq\theta, and more importantly, because the quantization voltage level bb is set to an arbitrary level. While we can steer the 1-bit ΣΔ\Sigma\Delta ADC based on the AoA estimate from Step 1 as ψ=θ^\psi=\hat{\theta}, we emphasize that a procedure to select an appropriate voltage level bb is crucial. Before describing the procedure to select bb, we next present the recursive bisection procedure to estimate the AoDs.

V-B2 Recursive bisection procedure

We estimate the AoDs of the llth path by maximizing El(𝐩)E_{l}({{{\mathbf{p}}}}) in (23) by recursively bisecting the spatial sector and selecting a subsector that yields the highest received energy. The BS informs via a 1-bit error-free feedback link the selected subsector to the UE, which further bisects the selected subsector to transmit pilots for the next stage. The procedure is continued for NsN_{\rm s} stages, where at the last stage, we select one of the angular grid points in 𝒟\mathcal{D} as the estimated AoD of the llth path. We repeat this procedure for all the LL paths.

To estimate the AoD of the llth path, we proceed as follows. Recall from (24) that 𝐩s,i{{\mathbf{p}}}_{s,i} has a unit response in the sector formed by departure angles ϕ~is,i\tilde{\phi}_{i}\in\mathcal{I}_{s,i}. In the first stage, the UE transmits pilots using precoders 𝐩1,1{{\mathbf{p}}}_{1,1} and 𝐩1,2{{\mathbf{p}}}_{1,2}. The BS then computes the energy of the received symbols using (23) and selects the subsector (or the precoder) that yeilds max{El(𝐩1,1),El(𝐩1,2)}\max\,\{E_{l}({{\mathbf{p}}}_{1,1}),E_{l}({{\mathbf{p}}}_{1,2})\}. If El(𝐩1,1)>El(𝐩1,2)E_{l}({{\mathbf{p}}}_{1,1})>E_{l}({{\mathbf{p}}}_{1,2}), the BS sends a feedback of 0 to the UE through an error-free 1-bit feedback link indicating that the sector ϕ~i1,1\tilde{\phi}_{i}\in\mathcal{I}_{1,1} is selected. Similarly, it sends a 1 indicating that the sector ϕ~i1,2\tilde{\phi}_{i}\in\mathcal{I}_{1,2} is selected. In the second stage, the UE then bisects the selected subsector and transmits pilots using the precoders {𝐩2,1,𝐩2,2}\{{{\mathbf{p}}}_{2,1},{{\mathbf{p}}}_{2,2}\} (respectively, {𝐩2,3,𝐩2,4}\{{{\mathbf{p}}}_{2,3},{{\mathbf{p}}}_{2,4}\}) if the received feedback from the BS is 0 (respectively, 1). More generally, at Stage s, suppose the UE transmits pilots using precoders {𝐩s,m,𝐩s,m+1}\{{{\mathbf{p}}}_{s,m},{{\mathbf{p}}}_{s,m+1}\} and El(𝐩s,m)>El(𝐩s,m+1)E_{l}({{\mathbf{p}}}_{s,m})>E_{l}({{\mathbf{p}}}_{s,m+1}). The BS transmits a 1 to the UE, which then selects the precoders {𝐩s+1,2m1,𝐩s+1,2m}\{{{\mathbf{p}}}_{s+1,2m-1},{{\mathbf{p}}}_{s+1,2m}\} corresponding to a narrower subsector for the next stage. We continue this procedure for NsN_{\rm s} stages, where the partition selected in the final stage corresponds to the index of the estimated AoD. The same procedure is repeated for all the LL paths.

V-C Selection of clipping and quantization voltage levels

In Fig. 3(c), we have seen that the choice of bb and cc play a crucial role in determining the channel estimation and beamforming performance of MIMO systems with 1-bit spatial ΣΔ\Sigma\Delta ADCs. As discussed in Section II, we choose the clipping level cc based on the standard deviation of input that allows us to place a bound on the clipping error using the Chebyshev inequality Pr(c[𝐱(t)]𝐱(t)2>δ)ϵ{\rm Pr}\left(\|{\mathcal{L}}_{c}[{{\mathbf{x}}}(t)]-{{\mathbf{x}}}(t)\|_{2}>\delta\right)\leq\epsilon for constants δ,ϵ>0\delta,\epsilon>0. Therefore, the choice of cc depends on the statistics of the unquantized signal 𝐱(t){{\mathbf{x}}}(t) received at the BS, leading to different choices of voltage levels for Step 1 and Step 2 of the proposed channel estimation algorithm.

V-C1 For estimating AoAs and path gains in Step 1

From (16) and (10), the unquantized signal received at the iith antenna of the BS in Step 1 of the proposed channel estimation technique is

[𝐱1(t)]i=PLk=1Lαkeȷ2πd(i1)sin(θk)+[𝐳1(t)]i,[{{\mathbf{x}}}_{1}(t)]_{i}=\sqrt{\frac{P}{L}}\sum_{k=1}^{L}\alpha_{k}e^{-\jmath 2\pi d(i-1){\rm sin}(\theta_{k})}+[{{\mathbf{z}}}_{1}(t)]_{i},

where αk\alpha_{k} and [𝐳1(t)]i[{{\mathbf{z}}}_{1}(t)]_{i} follow a complex Gaussian distribution with zero mean and unit variance. Since the complex path gains and the additive noise are mutually independent, [𝐱1(t)]i[{{\mathbf{x}}}_{1}(t)]_{i} follows a complex Gaussian distribution with zero mean and variance P+1P+1. In other words, ([𝐱1(t)]i)𝒩(0,P+12)\Re([{{\mathbf{x}}}_{1}(t)]_{i})\sim{\mathcal{N}}(0,\frac{P+1}{2}) and ([𝐱1(t)]i)𝒩(0,P+12)\Im([{{\mathbf{x}}}_{1}(t)]_{i})\sim{\mathcal{N}}(0,\frac{P+1}{2}). Let us recall that the probability that a Gaussian random variable takes values away from the mean by more than thrice the standard deviation is less than 1%1\%. Hence, to ensure

Pr(|([𝐱1(t)]i)|>c)\displaystyle{\rm Pr}(|\Re([{{\mathbf{x}}}_{1}(t)]_{i})|>c)
=Pr(|([𝐱1(t)]i)c[([𝐱1(t)]i]|>0)0.01,\displaystyle\quad={\rm Pr}(|\Re([{{\mathbf{x}}}_{1}(t)]_{i})-{\mathcal{L}}_{c}[\Re([{{\mathbf{x}}}_{1}(t)]_{i}]|>0)\leq 0.01,

in Step 1, we choose the clipping voltage level

c=3P+12.c=3\sqrt{\frac{P+1}{2}}. (27)

The clipping voltage level for the imaginary part is computed similarly. The quantization voltage level bb is then selected using the overload condition (4).

V-C2 For estimating AoDs in Step 2

In the second step, from (21) and (10), the unquantized signal received at the iith antenna of the BS related to the uplink pilot transmission with the precoder 𝐩{{\mathbf{p}}} is

[𝐱2(t)]i=PLk=1Lαkρk(𝐩)eȷ2πd(i1)sin(θk)+[𝐳2(t)]i,[{{\mathbf{x}}}_{2}(t)]_{i}=\sqrt{\frac{P}{L}}\sum_{k=1}^{L}\alpha_{k}\rho_{k}({{\mathbf{p}}})e^{-\jmath 2\pi d(i-1){\rm sin}(\theta_{k})}+[{{\mathbf{z}}}_{2}(t)]_{i},

where [𝐱2(t)]i[{{\mathbf{x}}}_{2}(t)]_{i} follows a complex Gaussian distribution with zero mean and variance PLk=1L|ρk(𝐩)|2+1\frac{P}{L}\sum_{k=1}^{L}|\rho_{k}({{\mathbf{p}}})|^{2}+1. Since ρk(𝐩)\rho_{k}({{\mathbf{p}}}) is bounded from above by Nt\sqrt{N_{\rm t}}, the worst-case variance of [𝐱2(t)]i[{{\mathbf{x}}}_{2}(t)]_{i} is PNt+1PN_{\rm t}+1. Therefore, we choose the clipping level as

c=3PNt+12,c=3\sqrt{\frac{PN_{\rm t}+1}{2}}, (28)

in Step 2, to ensure that the worst-case clipping probability of [𝐱2(t)]i[{{\mathbf{x}}}_{2}(t)]_{i}, i=1,,Nri=1,\ldots,N_{\rm r} is less than 1%\%.

Since the clipping voltage level that we select in Step 1 is different from Step 2, a joint estimator (e.g., sparse recovery based method as in [3]) to jointly estimate the channel parameters is not straightforward when dealing with observations from 1-bit spatial ΣΔ\Sigma\Delta ADCs.

V-D Computations and number of pilot transmissions

In Step 1, we transmit T1T_{1} pilots. In Step 2, we transmit each beam T2T_{2} times with 2 beams in each stage. Since there are Ns=log2DN_{\rm s}=\log_{2}D stages, and LL paths, the total pilot transmission overhead in the second stage is 2LT2Ns2LT_{2}N_{\rm s}.

For a search grid of size AA, computing the Bartlett beamforming spectrum in (17) costs about order ANrT1AN_{\rm r}T_{1} flops. The least squares estimator to compute the path gains incurs about order L2NrT1L^{2}N_{\rm r}T_{1} flops. In Step 2, for each beam and path, we compute the received power, which costs about order LT2NsNrLT_{2}N_{\rm s}N_{\rm r} flops. In contrast, a scheme that exhaustively searches over all possible AoA and AoD combinations with DT2DT_{2} pilot transmissions incurs a computational complexity of about order ADNrT2ADN_{\rm r}T_{2} flops, which is higher than Nr(AT1+LT2Ns)N_{\rm r}(AT_{1}+LT_{2}N_{\rm s}) flops incurred by the proposed method as typically DLNsD\gg LN_{\rm s}. The computational complexity of the proposed channel estimation algorithm scales linearly with the number of receive antennas and log-linearly with the search grid size, thus making it well-suited for massive MIMO systems.

VI MU-MIMO channel estimation

In this section, we specialize the proposed channel estimation algorithm to the MU-MIMO setting with the angular channel model described in Section IV-B. We estimate the channel by estimating the AoAs and the path gains using Step 1 of the algorithm developed in the previous section. Since we assume that the UEs have a single antenna in the MU-MIMO setup, there is no AoD estimation step.

Let 𝐒(t)K×K{\mathbf{S}}(t)\in{\mathbb{C}}^{K\times K} denote the orthogonal pilot matrix transmitted at time instance tt from the KK UEs such that 𝐒(t)𝐒H(t)=𝐒H(t)𝐒(t)=𝐈{\mathbf{S}}(t){\mathbf{S}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t)={\mathbf{S}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t){\mathbf{S}}(t)={\mathbf{I}}. From (11) and (10), the received signal at the BS with 1-bit spatial ΣΔ\Sigma\Delta ADC is

𝐘(t)=P𝐇𝐒(t)+𝐍(t),t=1,2,,T,{\mathbf{Y}}(t)=\sqrt{P}{\mathbf{H}}{\mathbf{S}}(t)+{\mathbf{N}}(t),\,t=1,2,\cdots,T,

where TT is the total number of channel uses, 𝐍(t){\mathbf{N}}(t) is the sum of additive white Gaussian noise and quantization noise, as defined before. Using the known pilots, we preprocess the received signal to separate the signal components from the UEs to obtain 𝐘𝐒H(t)=P𝐇+𝐍𝐒H(t),{\mathbf{Y}}{\mathbf{S}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t)=\sqrt{P}{\mathbf{H}}+{\mathbf{N}}{\mathbf{S}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t), where the kkth column of 𝐘𝐒H(t){\mathbf{Y}}{\mathbf{S}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t), denoted by 𝐲k,tNr{{{\mathbf{y}}}}_{k,t}\in{\mathbb{C}}^{N_{\rm r}}, is related to the SIMO channel between the kkth UE and the BS at time instance tt, and is given by [cf. (13)]

𝐲k,t\displaystyle{{{\mathbf{y}}}}_{k,t} =PLk𝐀BS(𝜽k)𝜶k+𝐧k,t\displaystyle=\sqrt{\frac{P}{L_{k}}}{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}_{k})\boldsymbol{\alpha}_{k}+{{{\mathbf{n}}}}_{k,t}
=PLk𝐀BS(𝜽k)diag(𝜶k)𝟏+𝐧k,t.\displaystyle=\sqrt{\frac{P}{L_{k}}}{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}_{k}){\rm diag}(\boldsymbol{\alpha}_{k}){\bf 1}+{{{\mathbf{n}}}}_{k,t}.

Here, 𝐧k,t{{{\mathbf{n}}}}_{k,t} denotes the kkth column of 𝐍𝐒H(t){\mathbf{N}}{\mathbf{S}}^{\raisebox{1.0pt}{$\rm\scriptscriptstyle H$}}(t). Suppose 𝐘k=[𝐲k,1,,𝐲k,T]{\mathbf{Y}}_{k}=[{{{\mathbf{y}}}}_{k,1},\cdots,{{{\mathbf{y}}}}_{k,T}] and 𝐍k=[𝐧k,1,,𝐧k,T]{\mathbf{N}}_{k}=[{{{\mathbf{n}}}}_{k,1},\cdots,{{{\mathbf{n}}}}_{k,T}]. Then we have 𝐘k=PLk𝐀BS(𝜽k)diag(𝜶k)𝐄+𝐍k,{\mathbf{Y}}_{k}=\sqrt{\frac{P}{L_{k}}}{\mathbf{A}}_{\rm BS}(\boldsymbol{\theta}_{k}){\rm diag}(\boldsymbol{\alpha}_{k}){{\mathbf{E}}}+{{\mathbf{N}}}_{k}, which readily resembles the signal model in (16). Therefore, we can use Step 1 of the SU-MIMO channel estimation algorithm developed in the previous section to estimate the channel parameters (𝜽k,𝜶k)(\boldsymbol{\theta}_{k},\boldsymbol{\alpha}_{k}) for each user. Also, we choose the same voltage level c=3P+12c=3\sqrt{\frac{P+1}{2}} as the one used in Step 1 of SU-MIMO channel estimation.

VII Numerical simulations

In this section, we present results from a number of numerical simulations to demonstrate the efficacy of the developed quantization noise model, voltage level selection, and channel estimation algorithms in mmWave MIMO systems with 1-bit spatial ΣΔ\Sigma\Delta ADCs. We compare different algorithms in terms of normalized mean square error (NMSE) and angle estimation error. We define NMSE of a path gain estimate 𝜶^\hat{\boldsymbol{\alpha}} and a channel estimate 𝐇^\hat{{\mathbf{H}}} as

NMSE(𝜶^)=𝔼[𝜶^𝜶22]𝔼[𝜶22]{\rm NMSE}(\hat{\boldsymbol{\alpha}})=\frac{{\mathbb{E}}\left[\|\hat{\boldsymbol{\alpha}}-\boldsymbol{\alpha}\|_{2}^{2}\right]}{{\mathbb{E}}\left[\|\boldsymbol{\alpha}\|_{2}^{2}\right]}

and

NMSE(𝐇^)=𝔼[𝐇^𝐇F2]𝔼[𝐇F2],{\rm NMSE}(\hat{{\mathbf{H}}})=\frac{{\mathbb{E}}\left[\|\hat{{\mathbf{H}}}-{\mathbf{H}}\|_{F}^{2}\right]}{{\mathbb{E}}\left[\|{\mathbf{H}}\|_{F}^{2}\right]},

respectively. Let 𝒜={θ~1,,θ~A}{\mathcal{A}}=\{\tilde{\theta}_{1},\ldots,\tilde{\theta}_{A}\} denote the AoA search grid of size AA used in (17). Let us denote the index set of AoAs corresponding to the angles in 𝒜{\mathcal{A}} as

𝔸(𝜽)={j:[𝜽]l=θ~j,1lL,1jA}.\mathbb{A}({\boldsymbol{\theta}})=\{j\,:\,[{\boldsymbol{\theta}}]_{l}=\tilde{\theta}_{j},1\leq l\leq L,1\leq j\leq A\}.

Let us also denote the index set of AoDs corresponding to the angles in AoD grid 𝒟={ϕ~1,ϕ~2,,ϕ~D}{\mathcal{D}}=\{\tilde{\phi}_{1},\tilde{\phi}_{2},\cdots,\tilde{\phi}_{D}\} of size DD as

𝔻(ϕ)={j:[ϕ]l=ϕ~j,1lL,1jD}.\mathbb{D}({\boldsymbol{\phi}})=\{j\,:\,[{\boldsymbol{\phi}}]_{l}=\tilde{\phi}_{j},1\leq l\leq L,1\leq j\leq D\}.

We then define the AoA and AoD estimation errors as

Eθ=Pr(𝔸(𝜽^)𝔸(𝜽))andEϕ=Pr(𝔻(ϕ^)𝔻(ϕ)),E_{\theta}={\rm Pr}\left(\mathbb{A}(\hat{\boldsymbol{\theta}})\neq\mathbb{A}({\boldsymbol{\theta}})\right)\quad\text{and}\quad E_{\phi}={\rm Pr}\left(\mathbb{D}(\hat{\boldsymbol{\phi}})\neq\mathbb{D}({\boldsymbol{\phi}})\right),

where 𝜽^\hat{\boldsymbol{\theta}} and ϕ^\hat{\boldsymbol{\phi}} are the estimated AoAs and AoDs, respectively.

VII-A The SU-MIMO setting

We consider a SU-MIMO setup with the UE having Nt= 32N_{\rm t}\leavevmode\nobreak\ =\leavevmode\nobreak\ 32 antennas, the BS having Nr=128N_{\rm r}=128 antennas that are spaced d=1/8d=1/8 wavelengths apart. For Bartlett beamforming, we use the search grid 𝒜={90,89,,89,90}{{\mathcal{A}}}=\{-90^{\circ},-89^{\circ},\cdots,89^{\circ},90^{\circ}\} with A=181A=181 points. The AoD grid 𝒟{\mathcal{D}} is obtained by uniformly sampling the direction cosine space in the interval 1-1 to 11 with D=128D=128 points as described in Section V-B1.

Refer to caption
Figure 4: SU-MIMO performance with L=1L=1. (a) AoA. (b) Path gain. (c) AoD.
Refer to caption
Figure 5: Impact of voltage level selection on Step 1 with L=1L=1. (a) Bartlett spectrum with true AoA at 1010^{\circ}. (b) AoA estimation error. (c) Path gain estimation error.
Refer to caption
Figure 6: (a) Beam patterns of the Stage 2 of the proposed codebook as seen at the receiver for different voltage levels with and without angle steering. (b) Impact of angle steering on AoD estimation. (c) Impact of voltage level on AoD estimation.

VII-A1 Angle steering and voltage level selection

We first begin with a discussion on the estimation performance of the channel parameters (𝜽,𝜶)(\boldsymbol{\theta},\boldsymbol{\alpha}) in Step 1 and ϕ\boldsymbol{\phi} in Step 2 of the proposed channel estimator for a single path channel with L=1L=1. This allows us to discuss the impact of voltage level selection and angle steering on the estimators. The AoA is drawn uniformly at random from the sector [x,x][-x^{\circ},x^{\circ}] for x=10,30,60x=10,30,60. The AoD is drawn randomly from the sector [75,75][-75^{\circ},75^{\circ}]. The path gain is assumed to be unit modulus. We use T1=10T_{1}=10 and T2=1T_{2}=1. Unless otherwise mentioned, we use 10610^{6} independent channel realizations to compute NMSE and angle errors. In this subsection, we compare the estimation performance of the proposed technique with 1-bit spatial ΣΔ\Sigma\Delta ADC, referred to as “ΣΔ\Sigma\Delta”, with an equivalent channel estimation method applied on unquantized data (wherein the BS is assumed to have a very high-resolution ADC). We label it as “Unquantized” in the plots. Since estimates from “Unquantized” will be better than the same scheme applied on quantized data for a comparable number of snapshots, we use this as a baseline to illustrate the loss due to 1-bit ΣΔ\Sigma\Delta ADC.

We illustrate the channel estimation performance for different SNRs in terms of EθE_{\theta} in Fig. 4(a), NMSE(𝜶^){\rm NMSE}(\hat{\boldsymbol{\alpha}}) in Fig. 4(b) and EϕE_{\phi} in Fig. 4(c). Here, “Unquantized, θ[10,10]\theta\in[-10^{\circ},10^{\circ}]” and “ΣΔ,θ[10,10]\Sigma\Delta,\theta\in[-10^{\circ},10^{\circ}]” indicate that the AoA is drawn uniformly at random from the sector [10,10][-10^{\circ},10^{\circ}] in each channel realization. We can see that “ΣΔ\Sigma\Delta” performs similar to that of “Unquantized” for the path with AoA arriving close to the array broadside, whereas the gap between “Unquantized” and “ΣΔ\Sigma\Delta” increases for the path arriving with angles away from the array broadside. This is mainly due to the quantization noise shaping at higher spatial frequencies. Performance degradation at higher SNRs is inevitable due to the larger choice of bb at high SNRs, which results in higher quantization noise [cf. (19)].

Next, in Fig. 5(a), Fig. 5(b), and Fig. 5(c), respectively, we illustrate the impact of clipping voltage level selection on 𝒥(θ){\mathcal{J}}(\theta), EθE_{\theta}, and EϕE_{\phi} by comparing the proposed clipping voltage level from Section V-C1 with a fixed, SNR independent arbitrary clipping voltage level c=1c=1. While we observe that the Bartlett beampatterns as well as the AoA estimation errors are not sensitive to the choice of clipping voltage levels, we, however, can observe from Fig. 5(c) that not selecting a correct clipping voltage level leads to severe performance degradation in path gain estimation. In other words, Fig. 5(c) demonstrates that a blind application of least squares without an appropriate selection of the clipping voltage level does not result in satisfactory performance.

Recall in Fig. 3(c), we discussed the impact of clipping voltage level selection and angle steering on the beampatterns in Step 2 as observed at the 1-bit spatial ΣΔ\Sigma\Delta ADC. We now extend that discussion in Fig. 6(a), where we focus on the second stage of the proposed codebook as observed at the receiver of an unquantized system and compare it to the beampattern as seen at the receiver with a 1-bit spatial ΣΔ\Sigma\Delta ADC. We consider a scenario where the path arrives at the BS with an AoA of θ=300\theta=30^{0} and an SNR of 1010 dB. We observe that the beampatterns are severely distorted whenever the steering angle, voltage level, or both are incorrectly selected. Nevertheless, with a careful selection of voltage levels and steering angle, the beampatterns at the receiver with a 1-bit spatial ΣΔ\Sigma\Delta ADC are comparable to that of the unquantized system.

In Fig. 6(b) and Fig. 6(c), we demonstrate, respectively, the impact of angle steering and clipping voltage level selection on AoD estimation, where the AoA is drawn uniformly at random from the sector [30,30][-30^{\circ},30^{\circ}] in each channel realization to compute EϕE_{\phi}. We can observe that failing to exploit angle steering leads to serious loss in performance. While clipping voltage selection was not crucial for AoA estimation, we observe that fixing the voltage level to an arbitrary value, such as c=1c=1, or varying it in an inappropriate manner, e.g., using the clipping voltage level designed for Step 1 in Step 2 (indicated as “ΣΔ\Sigma\Delta, cc from Step1”), do not provide satisfactory AoD estimation performance. In other words, a direct application of existing hierarchical codebook-based channel estimation methods, e.g., [3] without a careful selection of clipping voltage levels and phase shifts in the feedback loop will not provide reasonable performance for MIMO systems with 1-bit spatial ΣΔ\Sigma\Delta ADCs.

VII-A2 Multipath channel

We now consider a multipath SU-MIMO channel with L={2,3}L=\{2,3\}, which are typical at mmWave frequencies. We assume that the complex path gains αi\alpha_{i} follow a truncated Gaussian distribution with |(αi)|,|(αi)|τ|\Re(\alpha_{i})|,|\Im(\alpha_{i})|\geq\tau for i=1,,Li=1,\ldots,L. To ensure that all the LL paths are sufficiently strong, we choose τ=0.5\tau=0.5. The AoDs are drawn uniformly at random from the sector [75,75][-75^{\circ},75^{\circ}] with a minimum spacing of 0.10.1 in the direction cosine space and the AoAs are drawn uniformly at random from the sector [x,x][-x^{\circ},x^{\circ}] with a minimum spacing of 2020^{\circ}, where x={45,60}x^{\circ}=\{45^{\circ},60^{\circ}\}. We choose T1=10T_{1}=10 and T2=1T_{2}=1. Recall that the proposed method with 1-bit spatial ΣΔ\Sigma\Delta ADC, referred to as “ΣΔ\Sigma\Delta”, requires reception of T1T_{1} pilots in Step 1 and 2LT2Ns2LT_{2}N_{s} pilots in Step 2, where the reception at each step is with a different clipping voltage level. Hence, designing a scheme that utilizes all the available T1+2LT2NsT_{1}+2LT_{2}N_{s} measurements for channel estimation with 1-bit spatial ΣΔ\Sigma\Delta ADCs is not straightforward.

We compare the performance of the proposed method with amplitude retrieval based one-bit SU-MIMO channel estimation algorithm [19], referred to as “AR” in the plots. In “AR”, after the amplitudes are recovered from one-bit measurements, standard channel estimation algorithm can be used. Since “AR” does not involve any clipping voltage selection, here, we use all the T1+2LT2NsT_{1}+2LT_{2}N_{s} measurements to perform amplitude recovery as in [19], AoA and path gain estimation using Step 1 and AoD estimation using Step 2 as described in Section V. In addition to “Unquantized”, which serves as a benchmark for “ΣΔ\Sigma\Delta” as it uses T1T_{1} pilots in Step 1 and 2LT2Ns2LT_{2}N_{s} pilots in Step 2, we also report the performance of an unquantized system, labelled as “Unquantized full data”, that uses all the T1+2LT2NsT_{1}+2LT_{2}N_{s} pilots in both Step 1 and Step 2. Thus, the total number of pilot transmissions in all the methods that we compare with are the same. We have observed the runtime of the “AR” algorithm is significantly higher than the proposed method. Due to this reason, we use 500500 independent Monte-Carlo (MC) experiments to compute the NMSE of “AR”, whereas 4000040000 MC experiments are used to obtain plots for “Unquantized,” “Unquantized full data,” and “ΣΔ\Sigma\Delta” methods. NMSE of “AR” is, in general, much higher than that of the other methods, which suggests that a fewer number of MC runs is sufficient to obtain curves of comparable precision.

Refer to caption
Figure 7: SU-MIMO channel with (a) L=2L=2 and (b) L=3L=3.

In Fig. 7, we show the channel estimation NMSE, where θ[45,45]\theta\in[-45^{\circ},45^{\circ}] or θ[60,60]\theta\in[-60^{\circ},60^{\circ}] indicate the sector from which the AoAs are drawn in each channel realization. We can observe that the performance of “ΣΔ\Sigma\Delta” is comparable to that of “Unquantized” and is better than that of “AR” except at very low SNRs. At extremely low SNRs, we observe that the performance of “AR” is slightly better than that of “Unquantized” and “ΣΔ\Sigma\Delta” as the effective number of snapshots available to estimate the channel parameters is larger for “AR”. At high SNRs, on the other hand, higher quantization noise in 1-bit systems leads to a deteriorated performance of “AR”. Similarly, at high SNRs and for angles away from broadside, there is an inevitable gap between the NMSE of “ΣΔ\Sigma\Delta” and “Unquantized” due to the increase in the quantization noise. Furthermore, the NMSE of channel estimation for L=3L=3 is slightly larger than that of L=2L=2 due to the larger number of parameters to be estimated in the latter case. As expected, the benchmark scheme “Unquantized full data” has the lowest NMSE due to the absence of quantization and efficient use of all available snapshots to carry out channel estimation. In essence, the proposed method achieves performance comparable to that of “Unquantized” and significantly better than that of “AR” for most of the SNRs, making it an attractive choice for massive MIMO systems.

Refer to caption
Figure 8: The MU-MIMO setting. (a) LoS (i.e., Lk=1L_{k}=1) with K=8K=8 and Nr=128N_{\rm r}=128. (b) Multipath (i.e., Lk=3L_{k}=3) with K=8K=8 and Nr=128N_{\rm r}=128. (b) Multipath (i.e., Lk=3L_{k}=3) with K=8K=8 and Nr=256N_{\rm r}=256. (c) Multipath (i.e., Lk=3L_{k}=3) with K=8K=8 and Nr=128N_{\rm r}=128, where the true channel correlation is used to compute the Bussgang decomposition.

VII-B The MU-MIMO setting

For the MU-MIMO setting, we consider K=8K=8 users and a BS having Nr{128,256}N_{\rm r}\in\{128,256\} antennas that are d=1/8d=1/8 wavelengths apart. We use Lk{1,3}L_{k}\in\{1,3\} for k=1,,8k=1,\ldots,8, T=1T=1, and use 500500 independent channel realizations for computing NMSE. For each path, the AoAs are drawn uniformly at random from the sector [45,45][-45^{\circ},45^{\circ}] with a minimum spacing of 2020^{\circ}. We compare the channel estimation performance of the proposed method, referred to as “ΣΔ\Sigma\Delta proposed” with the following state-of-the-art techniques in MU-MIMO channel estimation: (A) Bussgang decomposition followed by computing LMMSE MU-MIMO channel estimation for 1-bit MIMO systems [6], referred to as “BLMMSE”, (B) MU-MIMO channel estimation with 1-bit spatial ΣΔ\Sigma\Delta ADC based on the Bussgang decomposition [14], labelled as “ΣΔ\Sigma\Delta Bussgang”, (C) an LMMSE MU-MIMO channel estimator that uses unquantized data, referred to as “Unquantized LMMSE”, and (D) the proposed channel estimation algorithm from Section IV-B applied to unquantized data, referred to as “Unquantized angular”. “Unquantized angular” serves as the benchmark scheme for “ΣΔ\Sigma\Delta proposed” and illustrates the loss due 1-bit spatial ΣΔ\Sigma\Delta quantization. We reemphasize that “Unquantized LMMSE”, “BLMMSE”, and “ΣΔ\Sigma\Delta Bussgang” require the channel correlation information to perform the Bussgang decomposition and for subsequent channel estimation, and that this amounts to knowing the angles that parameterize the angular channel model. However, for the sake of comparison, we generate an approximate channel correlation by assuming that each UE has 9191 paths, each separated by 1 degree in the range of [45,45][-45^{\circ},45^{\circ}] with 𝜶k𝒞𝒩(0,𝐈){\boldsymbol{\alpha}_{k}}\sim{\mathcal{C}}{\mathcal{N}}(0,{\mathbf{I}}) [cf. (13)].

The NMSE performance of different methods for (Lk,Nr)=(1,128)(L_{k},N_{\rm r})=(1,128), (Lk,Nr)=(3,128)(L_{k},N_{\rm r})=(3,128) and (Lk,Nr)=(3,256)(L_{k},N_{\rm r})=(3,256) are presented in Fig. 8(a), Fig. 8(b), and Fig. 8(c), respectively. It can be observed that the proposed scheme performs better in terms of NMSE than existing methods, namely, “BLMMSE” and “ΣΔ\Sigma\Delta Bussgang”. At low-to-moderate SNRs, the performance of “ΣΔ\Sigma\Delta Proposed” is comparable to that of “Unquantized angular”. This corroborates the developed theory that 1-bit spatial ΣΔ\Sigma\Delta ADCs have higher effective resolution, which can be leveraged for parametric estimation. At higher SNRs, the NMSE performance reduces for both “ΣΔ\Sigma\Delta Bussgang” and “ΣΔ\Sigma\Delta Proposed” due to the inevitable increase in quantization noise that is bound to occur at high SNRs. Nonetheless, even at high SNRs, the proposed method outperforms “ΣΔ\Sigma\Delta Bussgang” by a margin of about 8dB8\leavevmode\nobreak\ {\rm dB} (respectively, 12dB12\leavevmode\nobreak\ {\rm dB}) for Nr=128N_{\rm r}=128 (respectively, Nr=256N_{\rm r}=256).

As expected, the performance of the parametric channel estimation techniques (i.e., “Unquantized angular” and “ΣΔ\Sigma\Delta Proposed”) is better than their non-parametric counterparts (i.e., “Unquantized LMMSE” and “ΣΔ\Sigma\Delta Bussgang”) as the parametric techniques exploit the structure in the angular channel model. As expected, the channel estimation performance of “ΣΔ\Sigma\Delta Proposed” and “Unquantized angular” is better in Fig. 8(a) with Lk=1L_{k}=1 as compared to the multipath scenario in Fig. 8(b). Furthermore, in Fig. 8(a), we can see that “ΣΔ\Sigma\Delta Proposed” is about 812dB8-12\leavevmode\nobreak\ {\rm dB} better when compared to the state-of-the-art method “ΣΔ\Sigma\Delta Bussgang”. Also, when NrN_{\rm r} is doubled from 128128 in Fig. 8(b) to 256256 in Fig. 8(c), due to the improved resolution and quantization noise shaping of the larger antenna array in the latter, we see that the proposed technique significantly outperforms existing techniques.

In Fig. 8(d), we compare the performance of the proposed channel estimator when the true channel correlation is used to compute the Bussgang decomposition. In this setting, though not realizable in practice, “Unquantized LMMSE” is the optimal channel estimator. We can see that “Unquantized angular”, which is the proposed technique that works with unquantized data performs similar to that of the optimal estimator at low-to-moderate SNRs. Also, we can see that the performance of our method is comparable (in terms of channel estimation NMSE) to “ΣΔ\Sigma\Delta Bussgang” at moderate-to-high SNRs.

VIII Conclusions

In this paper, we have presented an algorithm for channel estimation with angular models in massive MIMO systems employing 1-bit spatial ΣΔ\Sigma\Delta ADC. We have developed a quantization noise model for 1-bit spatial ΣΔ\Sigma\Delta ADCs that is useful, in general, for large array processing applications. Although computing the complete quantization noise probability density function is difficult, the developed noise model allows us to compute its second-order statistics that can be used to prewhiten data when solving parametric estimation problems. When the quantization voltage levels and phase shifts in the feedback loop are carefully selected, the effective resolution of 1-bit spatial ΣΔ\Sigma\Delta ADCs can be improved and hence are comparable to unquantized systems in most operating regimes of interest. We have developed a two-step channel estimation procedure to estimate the AoAs, AoDs, and path gains that characterize the MIMO channel. The proposed algorithm allows us to select the phase shifts and quantization voltage levels, which depend on the unknown channel parameters. Through numerical simulations, we have demonstrated that with the proposed channel estimation algorithm, MIMO systems with 1-bit spatial ΣΔ\Sigma\Delta ADCs perform significantly better than MIMO systems with regular 1-bit quantization and are often on par with that of unquantized MIMO systems for low-to-moderate SNRs. [Proof of Lemma 1] The expressions in (6) are obtained by adapting the derivation in [22] to 1-bit spatial ΣΔ\Sigma\Delta ADCs and are derived here for self containment.

From (1) and (2) with ψ=0\psi=0, we have the recursion

yn(t)=xn(t)+en(t)en1(t).y_{n}(t)=x_{n}(t)+e_{n}(t)-e_{n-1}(t). (29)

Let us define the normalized quantization error as

εn(t)=0.5b1en(t)+μ,\varepsilon_{n}(t)=0.5b^{-1}e_{n}(t)+\mu, (30)

where μ=0.5+ȷ0.5\mu=0.5+\jmath 0.5 and 0(εn(t)),(εn(t))10\leq\Re(\varepsilon_{n}(t)),\Im(\varepsilon_{n}(t))\leq 1 as the quantization error en(t)e_{n}(t) is bounded when the voltage levels are chosen as in (4).

Using (1), we can express εn(t)\varepsilon_{n}(t) and rn(t)r_{n}(t) in terms of the input and output of the quantizer as

(εn(t))=12b((𝒬b[rn(t)])(rn(t)))+0.5\displaystyle\Re(\varepsilon_{n}(t))=\frac{1}{2b}\left(\Re({\mathcal{Q}}_{b}[r_{n}(t)])-\Re(r_{n}(t))\right)+0.5 (31)

with

(rn(t))=(xn(t))2b(εn1(t))+b.\Re(r_{n}(t))=\Re(x_{n}(t))-2b\Re(\varepsilon_{n-1}(t))+b.

When (rn(t))>0\Re(r_{n}(t))>0, or equivalently, when (εn1(t))0.5b1(xn(t))<0.5\Re(\varepsilon_{n-1}(t))-0.5b^{-1}\Re(x_{n}(t))<0.5, we have 𝒬b[rn(t)]=b{\mathcal{Q}}_{b}[r_{n}(t)]=b. This allows us to constrain (31) as

0(εn(t))=(εn1(t))0.5b1(xn(t))+0.5<1.0\leq\Re(\varepsilon_{n}(t))=\Re(\varepsilon_{n-1}(t))-0.5b^{-1}\Re(x_{n}(t))+0.5<1.

Thus, (εn(t))=(εn1(t))0.5b1(xn(t))+0.5\Re(\varepsilon_{n}(t))=\langle\Re(\varepsilon_{n-1}(t))-0.5b^{-1}\Re(x_{n}(t))+0.5\rangle as the fractional part function x=x\langle x\rangle=x for 0x<10\leq x<1 . Similarly, when (rn(t))<0\Re(r_{n}(t))<0, or equivalently, when (εn1(t))0.5b1(xn(t))>0.5\Re(\varepsilon_{n-1}(t))-0.5b^{-1}\Re(x_{n}(t))>0.5, we have 𝒬b[rn(t)]=b{\mathcal{Q}}_{b}[r_{n}(t)]=-b,

(εn(t))=(εn1(t))0.5b1(xn(t))0.5,\Re(\varepsilon_{n}(t))=\Re(\varepsilon_{n-1}(t))-0.5b^{-1}\Re(x_{n}(t))-0.5,

and

1(εn1(t))0.5b1(xn(t))+0.5<2,1\leq\Re(\varepsilon_{n-1}(t))-0.5b^{-1}\Re(x_{n}(t))+0.5<2, (32)

where the lower and upper bounds are due to the fact that 0(εn1(t))10\leq\Re(\varepsilon_{n-1}(t))\leq 1 and due to clipping b(xn(t))b-b\leq\Re(x_{n}(t))\leq b. For x[1,2)x\in[1,2), we have x=x1.\langle x\rangle=x-1. Therefore, when (rn(t))<0\Re(r_{n}(t))<0, from (32), we again have

(εn(t))=(εn1(t))0.5b1(xn(t))+0.5.\Re(\varepsilon_{n}(t))=\left<\Re(\varepsilon_{n-1}(t))-0.5b^{-1}\Re(x_{n}(t))+0.5\right>. (33)

Recursively substituting for (εi1(t))\Re(\varepsilon_{i-1}(t)), (εi2(t))\Re(\varepsilon_{i-2}(t)), and so on in (33), and using the fact that x+y=x+y,x,y\langle\langle x\rangle+y\rangle=\langle x+y\rangle,\leavevmode\nobreak\ \forall x,y\in{\mathbb{R}}, we obtain

(εn(t))\displaystyle\Re(\varepsilon_{n}(t)) =0.5(n+1)0.5b1k=1n(xk(t)).\displaystyle=\left\langle 0.5(n+1)-0.5b^{-1}\sum_{k=1}^{n}\Re(x_{k}(t))\right\rangle.

Using the above expression in (30) yields

(en(t))=2b0.5(n+1)0.5b1k=1n(xn(t))b.\Re(e_{n}(t))=2b\left\langle 0.5(n+1)-0.5b^{-1}\sum_{k=1}^{n}\Re(x_{n}(t))\right\rangle-b.

Using the fact that x+(n1)=x+1\langle-x+(n-1)\rangle=-\langle x\rangle+1, n,x\forall n\in\mathbb{Z},x\in\mathbb{R}, the above expression can be equivalently expressed as (6a). The expression for the imaginary part in (6b) can be derived along the same lines.

References

  • [1] R. S. P. Sankar and S. P. Chepuri, “Millimeter wave MIMO channel estimation with 1-bit spatial sigma-delta analog-to-digital converters,” in Proc. of the IEEE International Conference on Acoustics, Speech, and Signal Processing (ICASSP), Toronto, Canada, Jun. 2021.
  • [2] T. S. Rappaport, S. Sun, R. Mayzus, H. Zhao, Y. Azar, K. Wang, G. N. Wong, J. K. Schulz, M. Samimi, and F. Gutierrez, “Millimeter wave mobile communications for 5G cellular: It will work!” IEEE Access, vol. 1, pp. 335–349, May 2013.
  • [3] A. Alkhateeb, O. E. Ayach, G. Leus, and R. W. Heath, “Channel estimation and hybrid precoding for millimeter wave cellular systems,” IEEE J. Sel. Topics Signal Process., vol. 8, no. 5, pp. 831–846, Oct. 2014.
  • [4] A. Ghosh, T. A. Thomas, M. C. Cudak, R. Ratasuk, P. Moorut, F. W. Vook, T. S. Rappaport, G. R. MacCartney, S. Sun, and S. Nie, “Millimeter-wave enhanced local area systems: A high-data-rate approach for future wireless networks,” IEEE J. Sel. Areas Commun., vol. 32, no. 6, pp. 1152–1163, June 2014.
  • [5] J. Mo and R. W. Heath, “Capacity analysis of one-bit quantized MIMO systems with transmitter channel state information,” IEEE Trans. Signal Process., vol. 63, no. 20, pp. 5498–5512, Jul. 2015.
  • [6] Y. Li, C. Tao, G. Seco-Granados, A. Mezghani, A. L. Swindlehurst, and L. Liu, “Channel estimation and performance analysis of one-bit massive MIMO systems,” IEEE Trans. Signal Process., vol. 65, no. 15, pp. 4075–4089, Aug. 2017.
  • [7] K. Roth, H. Pirzadeh, A. L. Swindlehurst, and J. A. Nossek, “A comparison of hybrid beamforming and digital beamforming with low-resolution ADCs for multiple users and imperfect CSI,” IEEE J. Sel. Topics Signal Process., vol. 12, no. 3, pp. 484–498, Mar 2018.
  • [8] P. M. Aziz, H. V. Sorensen, and J. Van Der Spiegel, “An overview of sigma-delta converters: How a 1-bit ADC achieves more than 16-bit resolution,” IEEE Signal Process. Mag., vol. 13, no. 1, pp. 61–84, Jan. 1996.
  • [9] R. M. Corey and A. C. Singer, “Spatial sigma-delta signal acquisition for wideband beamforming arrays,” in Proc. Int. ITG Workshop Smart Antennas, Munich, Germany, Mar. 2016.
  • [10] D. Barać and E. Lindqvist, “Spatial sigma-delta modulation in a massive MIMO cellular system,” Master’s thesis, Chalmers University of Technology, Sweden, June 2016.
  • [11] V. Venkateswaran and A.-J. van der Veen, “Multichannel ΣΔ\Sigma\Delta ADCs with integrated feedback beamformers to cancel interfering communication signals,” IEEE Trans. Signal Process., vol. 59, no. 5, pp. 2211–2222, May 2011.
  • [12] M. Shao, W. Ma, Q. Li, and A. L. Swindlehurst, “One-bit sigma-delta MIMO precoding,” IEEE J. Sel. Topics Signal Process., vol. 13, no. 5, pp. 1046–1061, Sept. 2019.
  • [13] S. Rao, A. L. Swindlehurst, and H. Pirzadeh, “Massive MIMO channel estimation with 1-bit spatial sigma-delta ADCs,” in Proc. of the IEEE Int. Conf. on Acoustics, Speech and Signal Process. (ICASSP), Brighton, UK, May 2019.
  • [14] S. Rao, G. Seco-Granados, H. Pirzadeh, J. A. Nossek, and A. L. Swindlehurst, “Massive MIMO channel estimation with low-resolution spatial sigma-delta ADCs,” IEEE Access, vol. 9, pp. 109 320–109 334, Jul. 2021.
  • [15] H. Pirzadeh, G. Seco-Granados, S. Rao, and A. L. Swindlehurst, “Spectral efficiency of one-bit sigma-delta massive MIMO systems,” IEEE J. Sel. Areas Commun., vol. 38, no. 9, pp. 2215–2226, Sept. 2020.
  • [16] J. J. Bussgang, “Crosscorrelation functions of amplitude-distorted Gaussian signals,” Technical report, Research Laboratory of Electronics, Massachusetts Institute of Technology, Mar. 1952. [Online]. Available: http://hdl.handle.net/1721.1/4847
  • [17] Ö. T. Demir and E. Björnson, “The Bussgang decomposition of non-linear systems: Basic theory and MIMO extensions,” arXiv preprint arXiv:2005.01597, May 2020.
  • [18] S. Jacobsson, G. Durisi, M. Coldrey, U. Gustavsson, and C. Studer, “Throughput analysis of massive MIMO uplink with low-resolution ADCs,” IEEE Trans. Wireless Commun., vol. 16, no. 6, pp. 4038–4051, June 2017.
  • [19] C. Qian, X. Fu, and N. D. Sidiropoulos, “Amplitude retrieval for channel estimation of MIMO systems with one-bit ADCs,” IEEE Signal Process. Lett., vol. 26, no. 11, pp. 1698–1702, Nov. 2019.
  • [20] J. Mo, P. Schniter, N. G. Prelcic, and R. W. Heath, “Channel estimation in millimeter wave MIMO systems with one-bit quantization,” in Proc. of the Asilomar Conference on Signals, Systems and Computers, Pacific Grove, USA, Nov. 2014.
  • [21] Y. Zhang, M. Alrabeiah, and A. Alkhateeb, “Deep learning for massive MIMO with 1-bit ADCs: When more antennas need fewer pilots,” IEEE Wireless Commun. Lett., vol. 9, no. 8, pp. 1273–1277, Apr. 2020.
  • [22] R. M. Gray, W. Chou, and P. W. Wong, “Quantization noise in single-loop sigma-delta modulation with sinusoidal inputs,” IEEE Trans. Commun., vol. 37, no. 9, pp. 956–968, Sept. 1989.
  • [23] R. M. Gray, “Oversampled sigma-delta modulation,” IEEE Trans. Commun., vol. 35, no. 5, pp. 481–489, May 1987.
  • [24] H. L. Van Trees, Optimum Array Processing: Part IV of Detection, Estimation, and Modulation Theory.   USA: John Wiley & Sons, Ltd, 2002.