Open access peer-reviewed chapter

New Results on Single User Massive MIMO

Written By

Kasturi Vasudevan, Surendra Kota, Lov Kumar and Himanshu Bhusan Mishra

Reviewed: 06 July 2023 Published: 01 August 2023

DOI: 10.5772/intechopen.112469

From the Edited Volume

MIMO Communications - Fundamental Theory, Propagation Channels, and Antenna Systems

Edited by Ahmed A. Kishk and Xiaoming Chen

Chapter metrics overview

99 Chapter Downloads

View Full Metrics

Abstract

Achieving high bit rates is the main goal of wireless technologies like 5G and beyond. This translates to obtaining high spectral efficiencies using large number of antennas at the transmitter and receiver (single user massive multiple input multiple output or SU-MMIMO). It is possible to have a large number of antennas in the mobile handset at mm-wave frequencies in the range 30–300 GHz due to the small antenna size. In this work, we investigate the bit-error-rate (BER) performance of SU-MMIMO in two scenarios (a) using serially concatenated turbo code (SCTC) in uncorrelated channel and (b) parallel concatenated turbo code (PCTC) in correlated channel. Computer simulation results indicate that the BER is quite insensitive to re-transmissions and wide variations in the number of transmit and receive antennas. Moreover, we have obtained a BER of 10−5 at an average signal-to-interference plus noise ratio (SINR) per bit of just 1.25 dB with 512 transmit and receive antennas (512 × 512 SU-MMIMO system) with a spectral efficiency of 256 bits/transmission or 256 bits/sec/Hz in an uncorrelated channel. Similar BER results have been obtained for SU-MMIMO using PCTC in correlated channel. A semi-analytic approach to estimating the BER of a turbo code has been derived.

Keywords

  • single user massive multiple input multiple output (SU-MMIMO)
  • Rayleigh fading
  • serially concatenated turbo code (SCTC)
  • parallel concatenated turbo code (PCTC)
  • spectral efficiency (SE)
  • signal-to-interference plus noise ratio (SINR) per bit
  • spatial multiplexing
  • bit-error-rate (BER)

1. Introduction

As wireless technologies evolve beyond 5G [1, 2, 3], there is a growing need to attain peak data rates of about gigabits per second per user, which is required for high definition video, remote surgery, autonomous vehicles, gaming and so on, while at the same time consuming minimum transmit power. This can only be achieved by using multiple antennas at the transmitter and receiver [4, 5, 6, 7, 8], small constellations like quadrature shift keying (QPSK) and powerful error correcting codes like turbo or low density parity check (LDPC) codes. Having a large number of antennas in the mobile handset is feasible in mm-wave frequencies [9, 10, 11, 12] (30–300 GHz) due to the small antenna size. The main concern about mm wave communications has been its rather high attenuation in outdoor environments with rain and snow [13]. Therefore, at least in the initial stages, mm wave could be deployed indoors. The second issue relates to the poor penetration characteristics of mm wave through walls, doors, windows and other materials. This points towards to usage of mm wave [9] in a single room, say a big auditorium or underground parking and so on. Reconfigurable intelligent surface (RIS) [11, 12, 13, 14, 15, 16, 17] could be used to boost the propagation of mm waves, both indoors and outdoors. Most of the massive MIMO systems discussed in the literature are multi-user (MU) [18, 19, 20, 21, 22, 23, 24, 25, 26], that is, the base station has a large number of antennas and the mobile handset has only a single antenna (Nt=1). A large number of users are served simultaneously by the base station. A comparison between MU-MMIMO and SU-MMIMO is given in Table 1 [27, 28].

MU-MIMOSU-MMIMO
Beamforming possible in downlinkBeamforming possible in uplink & downlink
Spatial multiplexing not possibleSpatial multiplexing possible in uplink & downlink
Low spectral efficiency per userHigh spectral efficiency per user
High directivity in downlink in beamforming modeHigh directivity in uplink & downlink in beamforming mode

Table 1.

Comparison of MU-MMIMO and SU-MMIMO.

The base station in MU-MMIMO uses beamforming to improve the signal-to-noise ratio at the mobile handset. On the other hand, SU-MMIMO uses spatial multiplexing to improve the spectral efficiency in the downlink and uplink. The comparison between beamforming and spatial multiplexing is given in Table 2 [27, 28]. The total transmit power of SU-MMIMO using uncoded QPSK versus MU-MMIMO using M-ary QAM is shown in Table 3. The minimum Euclidean distance between symbols of all constellations is taken to be 2. The peak-to-average power ratio (PAPR) for SU-MMIMO using QPSK is compared with MU-MMIMO using M-ary QAM in Table 4 [27]. Of course in the case of frequency selective fading channels, OFDM needs to be used, which would result in PAPR greater than 0 dB even for QPSK signaling. It is clear from Tables 14 that technologies that use SU-MMIMO have a lot to gain.

BeamformingSpatial multiplexing
High directivityLittle or no directivity
Line-of-sight communication requiredRich scattering channel required
Low spectral efficiency per user since the same signal is transmitted from each antenna elementHigh spectral efficiency per user since different signals are transmitted from each antenna element
Spectral efficiency can be improved by increasing the constellation size resulting in high PAPRQPSK constellations with PAPR 0 dB can be used
Difficult to turbo/LDPC code large constellationsEasy to turbo/LDPC code QPSK
Large BER at average SINR per bit close to 0 dBSmall BER at average SINR per bit close to 0 dB

Table 2.

Comparison of beamforming and spatial multiplexing.

Spectral efficiency (bits/sec/Hz)QPSKM-ary QAM
Transmit antennas NtTotal average transmit powerM-ary QAMNt = 1 average transmit power
42416-QAM10
63664-QAM42
848256-QAM170
105101024-QAM682

Table 3.

SU-MMIMO using QPSK vs. MU-MMIMO using M-ary.

Spectral efficiency (bits/sec/Hz)QPSKM-ary Nt = 1
Transmit antennas NtPAPR (dB)MPAPR (dB)
42016-QAM2.5
63064-QAM3.7
840256-QAM4.23
10501024-QAM4.5

Table 4.

PAPR of SU-MMIMO using QPSK vs. MU-MMIMO using M-ary.

Moreover, since all transmit antennas use the same carrier frequency, there is no increase in bandwidth. SU-MMIMO with equal number of transmit and receive antennas is given in [29, 30]. The probability of erasure in MIMO-OFDM is presented in [31]. A practical SU-MMIMO receiver with estimated channel, carrier frequency offset and timing is described in [32, 33]. SU-MMIMO with unequal number of transmit and receive antennas and precoding is discussed in [34, 35] and the case without precoding in [36, 37]. All the earlier research on SU-MMIMO involved the use of a parallel concatenated turbo code (PCTC) and uncorrelated channel. In this work, we investigate the performance of SU-MMIMO using (a) serial concatenated turbo code (SCTC) in uncorrelated channel and (b) PCTC in correlated channel. Throughout this article we assume that the channel is known perfectly at the receiver. Perfect carrier and timing synchronization is also assumed.

This work is organized as follows. Section II discusses SU-MMIMO with SCTC in uncorrelated channel, the procedure for bit-error-rate (BER) estimation and computer simulation results. Section III deals with SU-MMIMO using PCTC in correlated channel with and without precoding along with computer simulation results. Section IV presents the conclusions and scope for future work.

Advertisement

2. SU-MMIMO with SCTC

2.1 System model

Consider the block diagram in Figure 1 [36, 38]. The input bits ai, 1iLd1 is passed through an outer rate-1/2 recursive systematic convolutional (RSC) encoder to obtain the coded bit stream bi, 1iLd, where

Figure 1.

SU-MMIMO with serially concatenated turbo code.

Ld=2Ld1.E1

Now bi is input to an interleaver to generate ci, 1iLd. Next ci is passed through an inner rate-1/2 RSC encoder and mapped to symbols Si, 1iLd, in a quadrature phase shift keyed (QPSK) constellation having symbol coordinates ±1±j, where j=1. Throughout this article we assume that bit “0” maps to +1 and bit “1” maps to 1. The set of Ld QPSK symbols constitute a “frame” and are transmitted using Nt antennas. We assume that

LdNt=an integerE2

so that all symbols in the frame are transmitted using Nt antennas. The set of QPSK symbols transmitted simultaneously using Nt antennas constitute a “block”. The generator matrix for both the inner and outer rate-1/2 RSC encoder is given by

GD=11+D21+D+D2.E3

Hence, both encoders have SE=4 states in the trellis. Assuming uncorrelated Rayleigh flat fading, the received signal for the kth re-transmission (0kNrt1, k is an integer) is given by (2) of [36], which is repeated here for convenience

Rk=HkS+WkE4

where SCNt×1 whose elements are drawn from the QPSK constellation, HkCNr×Nt whose elements are mutually independent and CN02σH2 and WkCNr×1 is the additive white Gaussian noise (AWGN) vector whose elements are mutually independent and CN02σW2. Note that σH2,σW2 denote the variance per dimension (real part or imaginary part) and Nr is the number of receive antennas. We assume that Hk and Wk are independent across blocks and re-transmissions, hence (4) in [29] is valid with N replaced by Nt. Recall that (see also (16) of [36])

Ntot=Nt+Nr.E5

Following the procedure given in Section 4 of [36] we get (see (36) of [36])

Yi=FiSi+Uifor1iNt.E6

After concatenation over blocks, Yi in (6) for 1iLd is sent to the turbo decoder (see also the sentence after (25) in [29]). For the sake of consistency with earlier work [38], we re-index i as 0iLd1 and use the same index i for ai, bi, ci and Yi without any ambiguity. In the next subsection, we discuss the turbo decoding (BCJR) algorithm [39, 40] for the inner code.

2.2 BCJR for the inner code

Let Dn denote the set of states that diverge from state n in the trellis [38, 40]. Similarly, let Cn denote the set of states that converge to state n. Let αi,n denote the forward sum-of-products (SOP) at time i, 0iLd2, at state n, 0nSE1. Then the forward SOP can be recursively computed as follows (see also (30) of [38]):

αi+1,n=mCnαi,mγi,m,nPci,m,n;α0,n=1;αi+1,n=αi+1,nn=0SE1αi+1,nE7

where Pci,m,n denotes the a priori probability of the systematic bit corresponding to the transition from encoder state m to n, at time i (this is set to 0.5 at the beginning of the first iteration). The last equation in (7) is required to prevent numerical instabilities [40]. We have

γi,m,n=expYiSm,n22σU2E8

where Yi is given by (6), Sm,n is the QPSK symbol corresponding to the transition from encoder state m to n and σU2 is given by (38) of [36] which is repeated here for convenience:

EUi2=8σH4NrNt1+4σW2σH2NrNrt=ΔσU2.E9

Robust turbo decoding (see Section 4.2 of [41]) can be employed to compute γi,m,n in (8). Similarly, let βi,m denote the backward SOP at time i, 1iLd1, at state m, 0mSE1. Then the backward SOP can be recursively computed as (see also (33) of [38]):

βi,m=nDmβi+1,nγi,m,nPci,m,n;βLd,m=1;βi,m=βi,mm=0SE1βi,m.E10

Let ρ+n denote the state that is reached from encoder state n when the input symbol is +1. Similarly let ρn denote the state that can be reached from encoder state n when the input symbol is 1. Then for 0iLd1 we compute

Ci+=n=0SE1αi,nγi,n,ρ+nβi+1,ρ+n;Ci=n=0SE1αi,nγi,n,ρnβi+1,ρn.E11

Finally, the extrinsic information that is fed to the BCJR algorithm for the outer code is computed as, for 0iLd1, (see (36) of [38]):

Eci=+1=Ci+Ci++Ci;Eci=1=Ci/Ci++Ci.E12

Next, we describe the BCJR for the outer code.

2.3 BCJR for the outer code

Let αi,n denote the forward SOP at time i, 0iLd12, at state n, 0nSE1. Then the forward SOP is recursively computed as follows:

αi+1,n=mCnαi,mγsys,i,m,nγpar,i,m,nPαi,m,n;α0,n=1;αi+1,n=αi+1,nn=0SE1αi+1,nE13

where Pai,m,n denotes the a priori probability of the systematic bit corresponding to the transition from state m to state n, at time i. In the absence of any other information, we assume ai,m,n=0.5 [42]. We also have for 0iLd11 (similar to (38) of [38])

γsys,i,m,n=Ecπ2i=+1ifH1Ecπ2i=1ifH2;γpar,i,m,n=Ecπ2i+1=+1ifH3Ecπ2i+1=1ifH4E14

where π· denotes the interleaver map and

H1:systematicbitfrom statemtonis+1; H2:systematicbitfrom statemtonis1

H3:paritybitfrom statemtonis+1;H4:paritybitfrom statemtonis1.E15

Observe that in (14) and (15) it is assumed that after the parallel-to-serial conversion in Figure 1, b2i corresponds to the systematic (data) bits and b2i+1 corresponds to the parity bits for 0iLd11. Similarly, let βi,m denote the backward SOP at time i, 1iLd11, at state m, 0mSE1. Then the backward SOP can be recursively computed as:

βi,m=nDmβi+1,nγsys,i,m,nγpar,i,m,nPai,m,n;βLd1,m=1;βi,m=βi,mm=0SE1βi,m.E16

Next, for 0iLd11 we compute

B2i+=n=0SE1αi,nγpar,i,n,ρ+nβi+1,ρ+n;B2i=n=0SE1αi,nγpar,i,n,ρnβi+1,ρn.E17

Let μ+n, and μn denote the states that are reached from state n when the parity bit is +1 and 1respectively. Similarly for 0iLd11 compute

B2i+1+=n=0SE1αi,nγsys,i,n,μ+nβi+1,μ+n;B2i+1=n=0SE1αi,nγsys,i,n,μnβi+1,μn.E18

The extrinsic information that is sent to the inner decoder for 0iLd1 is computed as

Ebi=+1=Bi+Bi++Bi;Ebi=1=Bi/Bi++BiE19

where Bi+,Bi are given by (17)or (18) depending on whether i is even or odd respectively. Note that Pci,m,n for 0iLd1 in (7) and (10) is equal to

Pci,m,n=Ebπ1i=+1ifH1Ebπ1i=1ifH2E20

where π1· denotes the inverse interleaver map. Note that ci,m,n are the systematic (data) bits for the inner encoder.

After the convergence of the BCJR algorithm in the last iteration, the final a posteriori probabilities of ai for 0iLd11 is given by

Pai=+1=Eb2i=+1Ecπ2i=+1;Pai=1=Eb2i=1Ecπ2i=1E21

where Eci=±1 and Ebi=±1 are given by (12) and (19) respectively. Finally note that for 0iLd11

ai=b2i=cπ2i.E22

In the next section we present the estimation of the bit-error-rate (BER) of the SCTC.

2.4 Estimation of BER

The estimation of BER of SCTC is based on the following propositions:

Proposition1.The extrinsic information as computed in(12)and(19)lies in the range01(0 and 1 included). The extrinsic information in the range01, 0 and 1 excluded, is Gaussian distributed [43] for each frame.

This is illustrated in Figure 2 for different values of the frame length Ld1, over many frames (F). We find that for large values of Ld1, the histogram better approximates the Gaussian characteristic. It may be noted that the extrinsic information at the output of one decoder is equal to the a priori probabilities for the other decoder.

Figure 2.

Normalized histogram for Ntot = 1024, Nt = 512, Nrt = 2 (a) Ld1 = 1024, SNRav, b = 1.25 dB, F = 105 frames (b) Ld1 = 50,176, SNRav, b = 0.3 dB, F = 2000 frames (c) expanded view of (around) r3,i = 0 and (d) Ld1 = 50,176, SNRav, b = 0.5 dB, F = 2000 frames.

Proposition 2.After convergence of the BCJR algorithm in the final iteration, the extrinsic information at a decoder output has the same mean and variance as that of the a priori probability at its input.

Proposition 3.The mean and variance of the Gaussian distribution may vary from frame to frame.

This is illustrated in Figure 3 over two frames, that is, F=2.

Pe=12erfcA2σ2.E23

Figure 3.

Normalized histogram over two frames (F = 2) for Ntot = 1024, Nt = 512, Nrt = 2 (a) Ld1 = 1024, SNRav, b = 1.25 dB and (d) Ld1 = 50,176, SNRav, b = 0.5 dB.

Based on Propositions 1 & 2 and (22), after convergence of the BCJR algorithm, we can write for 0iLd11

Eb2i=+1=1σ2πer1,iA2/2σ2;Ecπ2i=+1=1σ2πer2,iA22σ2E24

where it is assumed that bit “0” maps to A and bit “1” maps to A and

r1,i=±A+w1,i;r2,i=±A+w2,iE25

where w1,i,w2,i denote real-valued samples of zero-mean additive white Gaussian noise (AWGN) with variance σ2. Similarly we have

Eb2i=1=1σ2πer1,i+A2/2σ2;Ecπ2i=1=1σ2πer2,i+A22σ2.E26

Clearly

lnEb2i=+1Eb2i=1=2Aσ2r1,i;lnEcπ2i=+1Ecπ2i=1=2Aσ2r2,i.E27

From (21) and (26) we have for 0iLd11

lnPai=+1Pai=1=2Aσ2r1,i+r2,i=Δ2Aσ2r3,i.E28

Consider the average

Y=2Aσ2Ld2i=0Ld21air3,i=4A2σ2+ZE29

where

Z=2Aσ2Ld2i=0Ld21aiw1,i+w2,i;.Ld2<Ld1E30

Note that the average in (28) is done over less than Ld1 terms to avoid situations like

Pai=±1=1or0.E31

In fact, only those time instants i have been considered in the summation of (28) for which

Pai=±1>e500.E32

Now

EZ=0;EZ2=4A2σ4Ld22i=0Ld212σ2=8A2σ2Ld2E33

where we have used the fact that w1,i,w2,i are independent. Now, we know that the probability of error for the BPSK signal in (27), that is equal to [40].

r3,i=r1,i+r2,i=±2A+w1,i+w2,iE34

Therefore from (28), (32) and (34) we have

Pfe12erfcY4E35

where Pfe denotes the probability of bit error for frame “f” and

EZ20forLd21.E36

Observe that it is necessary to take the absolute value of Y in (35) since there is a possibility that it can be negative. The average probability of bit error over F frames is given by

Pe=1Ff=0F1Pfe.E37

In the next section we present computer simulation results for SU-MMIMO using SCTC in uncorrelated channel.

2.5 Simulation results

The simulation parameters are given in Table 5. We can make the following observations from Figures 46 [36]:

ParameterValue(s)
ModulationQPSK
Total Antennas (Ntot = Nt + Nr)1024322
Transmit antennas (Nt)400512712161
Frame length (Ld1)1200 50,4001024 50,1761001100810241001
Frames simulated (F)104. 105 for Ld1 range 1001 to 1200, 200, 2000 for Ld1 = 50,176, 50,400
Spectral eff. For Nrt = 1 (bits/sec/Hz)2002563.5680.5

Table 5.

Simulation parameters for results in Figures 46.

Figure 4.

Simulation results for Ntot = 1024.

Figure 5.

Simulation results for Ntot = 32.

Figure 6.

Simulation results for Ntot = 2, Nt = 1.

The theoretical prediction of BER closely matches with simulations.

  • For Ntot = 32, 1024, the BER is quite insensitive to wide variations in the total number of antennas Ntot, transmit antennas Nt and retransmissions Nrt.

  • For Ntot = 2, the BER improves significantly with increasing retransmissions.

In Figure 4(c) we observe that there is more than 1 dB improvement in SINR compared to Figures 46(a, b). However, large values of Ld1 may introduce more latency which is contrary to the requirements of 5G and beyond. In the next section we present SU-MMIMO using PCTC in correlated channel.

Advertisement

3. SU-MMIMO using PCTC in correlated channel

3.1 System model

The block diagram of the system is identical to Figure 2 in [36] and the received signal is given by (4). Note that in (4), the channel autocorrelation matrix is given by

RHH=12EHkHHk=NrINtE38

where the superscript “H” denotes Hermitian and INt denotes the Nt×Nt identity matrix. In this section, we investigate the situation where RHH is not an identity matrix, but is a valid autocorrelation matrix [40]. As mentioned in [36], the elements of Hk – given by Hk,i,j for the kth re-transmission, ith row, jth column of Hk – are zero-mean, complex Gaussian random variables with variance per dimension equal to σH2. The in-phase and quadrature components of Hk,i,j – denoted by Hk,i,j,I and Hk,i,j,Q respectively – are statistically independent. Moreover, we assume that the rows of Hk are statistically independent. Following the procedure in [36] for the case without precoding, we now find the expression for the average SINR per bit before and after averaging over re-transmissions (k). All symbols and notations have the usual meaning, as given in [36].

3.2 SINR analysis

The ith element of HkHRk is given by (25) of [36] which is repeated here for convenience

Yk,i=Fk,i,iSi+Ik,i+Vk,ifor1iNtE39

where

Vk,i=j=1NrHk,j,iWk,j;Ik,i=j=1jiNtFk,i,jSj;Fk,i,j=l=1NrHk,l,iHk,l,j.E40

We have

EFk,i,i2=El=1NrHk,l,i2m=1NrHk,m,i2=El=1NrHk,l,i,I2+Hk,l,i,Q2m=1NrHk,m,i,I2+Hk,m,i,Q2=4σH4NrNr+1E41

which is identical to (27) in [36] and we have used the following properties

  1. The in-phase and quadrature components of Hk,i,j are independent.

  2. The rows of Hk are independent.

  3. For zero-mean, real-valued Gaussian random variable X,. with variance equal to σX2,EX4=3σX4.

The interference power is

EIk,i2=Ej=1jiNtFk,i,jSjl=1liNtFk,i,lSl=j=1jiNtl=1liNtEFk,i,jFk,i,lESjSl=Pavj=1jiNtEFk,i,j2.E42

where we have used (9) in. Similarly the noise power is

EVk,i2=Ej=1NrHk,j,iWk,jm=1NrHk,m,iWk,m=j=1Nrm=1NrEHk,j,iHk,m,iEWk,mWk,j=j=1Nrm=1Nr2σH2δKjm2σW2jm=4NrσH2σW2E43

which is identical to (29) in [36] and we have used the following properties:

  1. Rows of Hk are independent.

  2. Sifting property of the Kronecker delta function.

  3. Noise and channel coefficients are independent.

Now in (42)

EFk,i,j2=El=1NrHk,l,iHk,l,jm=1NrHk,m,iHk,m,j=l=1NrEHk,l,iHk,l,jHk,l,iHk,l,j+m=1mlNrHk,m,iHk,m,j
=l=1NrEHk,l,i2Hk,l,j2+m=1mlNrHk,l,iHk,l,jHk,m,iHk,m,j.E44

Now the first summation in (44) is equal to

E1=EHk,l,i2Hk,l,j2=EHk,l,i,I2+Hk,l,i,Q2Hk,l,j,I2+Hk,l,j,Q2=4σH4+4RHH,ji2E45

where we have used the property that for real-valued, zero-mean Gaussian random variables Xi, 1i4 [44, 45]

EX1X2X3X4=C12C34+C13C24+C14C23E46

where

Cij=EXiXjfor1i,j4E47

and

RHH,ji=EHk,l,i,IHk,l,j,I=EHk,l,i,QHk,l,j,Q=12EHk,l,iHk,l,j=RHH,ijE48

is the real-valued autocorrelation of Hk,m,n and we have made the assumption that the in-phase and quadrature components of Hk,m,n are independent. The second summation in (44) can be written as

E2=m=1mlNrEHk,l,iHk,l,jHk,m,iHk,m,j=m=1mlNrEHk,l,iHk,l,jEHk,m,iHk,m,j=m=1mlNr4RHH,ji2=4Nr1RHH,ji2E49

where we have used the property that the rows of Hk are independent. Therefore (44) becomes

EFk,i,j2=NrE1+E2=4NrσH4+RHH,ji2+Nr1RHH,ji2=4NrσH4+NrRHH,ji2.E50

The total power of interference plus noise is

EIk,i+Vk,i2=EIk,i2+EVk,i2=4PavNrj=1jiNtσH4+NrRHH,ji2+4NrσH2σW2E51

where we have made the assumption that noise and symbols are independent. The average SINR per bit for the ith transmit antenna is similar to (31) of [36] which is repeated here for convenience

SINRav,b,i=EFk,i,iSi2×2NrtEIk,i+Vk,i2for1iNtE52

into which (41) and (51) have to be substituted. The upper bound on the average SINR per bit for the ith transmit antenna is obtained by setting σW2=0 in (51), (52) and is given by, for 1iNt

SINRav,b,UB,i=σH41+Nr×2Nrtj=1jiNtσH4+NrRHH,ji2.E53

Observe that in contrast to (31) and (32) in [36], the average SINR per bit and its upper bound depend on the transmit antenna. Let us now compute the average SINR per bit after averaging over retransmissions. The received signal after averaging over retransmissions is given by (6) with (see also (20) of [36])

Fi=1Nrtk=0Nrt1Fk,i,i
U=1Nrtk=0Nrt1Ik,i+Vk,i=1Nrtk=0Nrt1Uk,isayE54

where Fk,i,i, Ik,i and Vk,i are given in (39). The power of the signal component of (6) is

ESi2Fi2=PavEFi2=PavNrt2Ek=0Nrt1Fk,i,il=0Nrt1Fl,i,i=PavNrt2k=0Nrt1l=0lkNrt1EFk,i,iEFl,i,i+EFk,i,i2E55

where we have used the fact that the channel is independent across retransmissions, therefore

EFk,i,iFl,i,i=EFk,i,iEFl,i,iforkl.E56

Now

EFk,i,i=El=1NrHk,l,i2=2NrσH2.E57

Substituting (41) and (57) in (55) we get

ESi2Fi2=4NrPavσH4Nrt1+NrNrt.E58

The power of the interference component in (6) and (54) is

EUi2=1Nrt2Ek=0Nrt1Ik,i+Vk,il=0Nrt1Il,i+Vl,i=1Nrt2k=0Nrt1l=0Nrt1EIk,iIl,i+EVk,iVl,iE59

where we have used the following properties from (40)

EIk,i=EVk,i=0;EIk,iVl,i=EVk,iIl,i=0forallk,lE60

since Sj and Wk,j are mutually independent with zero-mean. Now

EIk,iIl,i=Ej=1jiNtFk,i,jSjn=1niNtFl,i,nSn=j=1jiNtn=1niNtEFk,i,jFl,i,nESjSn=j=1jiNtn=1niNtEFk,i,jFl,i,nPavδKjn=Pavj=1jiNtEFk,i,jFl,i,jE61

where we have used the property that the symbols are uncorrelated and δK· is the Kronecker delta function [40]. When k=l, (61) is given by (42) and (50). When kl, (61) is given by

EIk,iIl,i=Pavj=1jiNtEFk,i,jEFl,i,j=Pavj=1jiNt4Nr2RHH,ji2E62

where we have used (40) and (48). Similarly, we have

EVk,iVl,i=4NrσH2σW2δKklE63

where we have used (43). Substituting (42), (50), (62) and (63) in (59) we get

EUi2=1Nrt24PavNrNrtj=1jiNtσH4+NrRHH,ji2+4PavNr2NrtNrt1j=1jiNtRHH,ji2+4NrNrtσH2σW2=1Nrt4PavNrj=1jiNtσH4+NrRHH,ji2+4PavNr2Nrt1j=1jiNtRHH,ji2]+4NrNrtσH2σW2.E64

The average SINR per bit for the ith transmit antenna, after averaging over retransmissions (also referred to as “combining” [36]) is given by

SINRav,b,C,i=2PavEFi2EUi2E65

into which (58) and (64) have to be substituted. The upper bound on the average SINR per bit after “combining” for the ith transmit antenna is given by

SINRav,b,C,UB,i=SINRav,b,C,iσW2=0.E66

The plots of the average SINR per bit for the ith transmit antenna before and after “combining” are shown in Figures 7 and 8 respectively for Ntot=1024 and Nrt=2. The channel correlation is given by

Figure 7.

Plot of SINRav,b,UB,i for Ntot = 1024, Nrt = 2. (a) Back view. (b) Sideview. (c) Front view.

Figure 8.

Plot of SINRav,b,C,UB,i for Ntot = 1024, Nrt = 2. (a) Back view. (b) Side view. (c) Front view.

RHH,ji=0.9jiσH2E67

in (48), which is obtained by passing samples of white Gaussian noise through a unit-energy, first-order infinite impulse response (IIR) lowpass filter with a=0.9 (see (30) of [46]).

We observe in Figures 7 and 8 that

The upper bound on the average SINR per bit decreases rapidly with increasing transmit antennas Nt and falls below 0 dB for Nt>5 (see Figures 7(b) and 8(b)). Since the spectral efficiency of the system is Nt/2Nrt bits/sec/Hz (see (33) of [36]), the system would be of no practical use, since the BER would be close to 0.5 for Nt>5.

The upper bound on the average SINR per bit after “combining” is less than that before “combining”. Therefore retransmissions are ineffective.

In view of the above observation, it becomes necessary to design a better receiver using precoding. This is presented in the next section.

3.3 Precoding

Similar to (4) consider the modified received signal given by

Rk=HkBS+WkE68

where

B=100a1,110aNt1,Nt1aNt1,11T=ΔATE69

where ·T denotes transpose. In (69), A is an Nt×Nt lower triangular matrix with diagonal elements equal to unity and ai,j denotes the jth coefficient of the optimum ith-order forward prediction filter [40] and B is the precoding matrix. Let

Yk=BHHkHRk=BHHkHHkBS+BHHkHWk.E70

Define

Zk=HkB=Zk,1,1Zk,1,NtZk,Nr,1Zk,Nr,Nt.E71

Now [40]

12EZkHZk=NrσZ,12000σZ,22000σZ,Nt2=ΔRZZE72

is an Nt×Nt diagonal matrix and σZ,i2 denotes the variance per dimension of the optimum i1th-order forward prediction filter. Note that [40]

σZ,12=σH2;σZ,i2σZ,j2fori<j.E73

Let

Vk=ZkHWk=Vk,1Vk,NtTE74

which is an Nt×1 vector. Now

EVk,iVk,m=Ej=1NrZk,j,iWk,jl=1NrZk,l,mWk,l=j=1Nrl=1NrEZk,l,mZk,j,iEWk,jWk,l
=j=1Nrl=1Nr2σZ,i2δKimδKjl×2σW2δKjl=4NrσZ,i2σW2δKimE75

where we have used (72). Let

Fk=ZkHZkE76

which is an Nt×Nt matrix. Substituting (76) in (70) we get

Yk=FkS+Vk.E77

Similar to (39), the ith element of Yk in (77) is given by

yk,i=Fk,i,iSi+Ik,i+Vk,ifor1iNtE78

where

Vk,i=j=1NrZk,j,iWk,j;Ik,i=j=1jiNtFk,i,jSj;Fk,i,jl=1NrZk,l,iZk,l,j.E79

Note that from (72) and (76) we have

EFk,i,i=2NrσZ,i2E80

Now

EFk,i,i2=El=1NrZk,l,i2m=1NrZk,m,i2=l=1NrZk,l,i4++m=1mlNrEZk,l,i2EZk,m,i2=4NrNr+1σZ,i2.E81

Similarly

EIk,i2=Ej=1jiNtFk,i,jSjl=1liNtFk,i,lSl=Pavj=1jiNtEFk,i,j2.E82

Now

EFk,i,j2=El=1NrZk,l,iZk,l,jm=1NrZk,m,iZk,m,j=l=1Nrm=1Nr4σZ,i2σZ,j2δKlm=4NrσZ,i2σZ,j2E83

where we have used (72). Substituting (83) in (82) we get

EIk,i2=4PavNrσZ,i2j=1jiNtσZ,j2.E84

Note that

EIk,i+Vk,i2=EIk,i2+EVk,i2.E85

The average SINR per bit for the ith transmit antenna is given by (52) and is equal to

SINRav,b,i=EFk,i,iSi2×2NrtEIk,i+Vk,i2=PavNr+1σZ,i2×2NrtPavj=1jiNtσZ,j2+σW2E86

where we have used (75), (81) and (84). The upper bound on the average SINR per bit for the ith transmit antenna is obtained by setting σW2=0 in (86) and is equal to

SINRav,b,UB,i=Nr+1σZ,i2×2Nrtj=1jiNtσZ,j2E87

which is illustrated in Figure 9 for Ntot=1024 and Nrt=2. The value of the upper bound on the average SINR per bit for Nt=i=50 is 18.6 dB. The channel correlation is given by (67). Note that a first-order prediction filter completely decorrelates the channel with [40]

Figure 9.

Plot of SINRav,b,UB,i for Ntot = 1024, Nrt = 2 after precoding. (a) Back view. (b) Sideview. (c) Front view.

ai,1=0.9for1iNt1;ai,j=0for2iNt1,2ji.E88

We also have [40]

σZ,i2=σZ,22=10.92σZ,12=0.19σZ,12fori>2.E89

Therefore we see in Figure 9 that the first transmit antenna i=1 has a high SINRav,b,UB,i due to low interference power from remaining transmit antennas, whereas for i1 the SINRav,b,UB,i is low due to high interference power from the first transmit antenna (i=1). The received signal after “combining” is given by (6) and (54). Note that from (54) and (79)

EFi2=1Nrt2Ek=0Nrt1Fk,i,il=0Nrt1Fl,i,i=1Nrt2k=0Nrt1EFk,i,i2++l=0lkNrt1EFk,i,iFl,i,i=4NrσZ,i2Nrt2k=0Nrt1Nr+1+Nrt1Nr=4NrσZ,i2Nrt21+NrNrtE90

where we have used (56), (80) and (81). Similarly from (54), (75), (84) and (85) we have

EUi2=1Nrt2Ek=0Nrt1Uk,il=0Nrt1Ul,i=1Nrt2k=0Nrt1l=0Nrt1EUk,iUl,i=1Nrt2k=0Nrt1l=0Nrt1EUk,i2δKkl
=1NrtEUk,i2=1NrtEIk,i2+EVk,i2=4NrσZ,i2NrtPavj=1jiNtσZ,j2+σW2.E91

Substituting (90) and (91) in (65) we have, after simplification, for 1iNt

SINRav,b,C,i=2PavEFi2EUi2=NrNrt+1σZ,i2×2PavPavj=1jiNtσZ,j2+σW2.E92

The upper bound on the average SINR per bit for the ith transmit antenna is obtained by substituting (92) in (66) and is equal to

SINRav,b,C,UB,i=NrNrt+1σZ,i2×2j=1jiNtσZ,j2SINRav,b,UB,iE93

for 1iNt, Nr1. This is illustrated in Figure 10 for Ntot=1024 and Nrt=2. We again observe that the first transmit antenna (i=1) has a high upper bound on the average SINR per bit, after “combining”, compared to the remaining transmit antennas. The value of the upper bound on the average SINR per bit after “combining” for Nt=i=50, Ntot=1024 is 18.6 dB. After concatenation, Yi for 0iLd1, in (6) and (54) is given to the turbo decoder [29, 40]. Let (see (26) of [29]):

Figure 10.

Plot of SINRav,b,C,UB,i for Ntot = 1024, Nrt = 2 after precoding. (a) Back view. (b) Side view. (c) Front view.

γ1,i,m,n=expYiFiSm,n22σU,i2;γ2,i,m,n=expYi1Fi1Sm,n22σU,i2E94
Y1=Y1YLd11;Y2=YLd1YLd1.E95

Thenwhere

i1=i+Ld1for0iLd11.E96

The rest of the turbo decoding algorithm is similar to that discussed in [29, 40] will not be repeated here. In the next subsection we present the computer simulation results for correlated channel with precoding and PCTC.

3.4 Simulation results

The channel correlation is given by (67). The BER results for Ntot=1024 with precoding are depicted in Figure 11. Incidentally, the value of the upper bound on the average SINR per bit before and after “combining” for Nt=i=512, Ntot=1024 is 6 dB. The BER results for Ntot=32 with precoding are depicted in Figure 12. Note that since the average SINR per bit depends on the transmit antenna, the minimum average SINR per bit is indicated along the x-axis of Figures 11 and 12. We also observe from Figures 11(a,b) and 12 that there is a large difference between theory and simulations. This is probably because, the average SINR per bit is not identical for all transmit antennas. In particular, we observe from Figures 9 and 10 that the first transmit antenna has a large average SINR per bit compared to the remaining antennas. However, in Figure 11(c,d) there is a close match between theory and simulations. This could be attributed to having a large number of blocks in a frame, as given by (2), resulting in better statistical properties. Even though the number of blocks is large in 12, the number of transmit antennas is small, resulting in inferior statistical properties. In order to improve the accuracy of the BER estimate for Ntot=32, we propose to transmit “dummy data” from the first transmit antenna and “actual data” from the remaining antennas. The BER results shown in Figure 13 indicates a good match between theory and practice. However, comparison of Figures 11 and 14 demonstrates that “dummy data” is ineffective for large number of transmit antennas.

Figure 11.

Simulation results with precoding for Ntot = 1024.

Figure 12.

Simulation results with precoding for Ntot = 32.

Figure 13.

Simulation results with precoding and dummy data for Ntot = 32.

Figure 14.

Simulation results with precoding and dummy data for Ntot = 1024.

Advertisement

4. Conclusions and future work

This article presents the advantages of single-user massive multiple input multiple output (SU-MMIMO) over multi-user (MU) MMIMO systems. The bit-error-rate (BER) performance of SU-MMIMO using serially concatenated turbo codes (SCTC) over uncorrelated channel is presented. A semi-analytic approach to estimating the BER of a turbo code is derived. A detailed signal-to-interference-plus-noise ratio analysis for SU-MMIMO over correlated channel is presented. The BER performance of SU-MMIMO with parallel concatenated turbo code (PCTC) over correlated channel is studied. Future work could involve estimating the MMIMO channel, since the present work assumes perfect knowledge of the channel.

References

  1. 1. Chowdhury MZ, Shahjalal M, Ahmed S, Jang YM. 6G wireless communication systems: Applications, requirements, technologies, challenges, and research directions. IEEE Open Journal of the Communications Society. 2020;1:957-975
  2. 2. Pereira de Figueiredo FA. An overview of massive MIMO for 5G and 6G. IEEE Latin America Transactions. 2022;20(6):931-940
  3. 3. Masaracchia A, Sharma V, Canberk B, Dobre OA, Duong TQ. Digital twin for 6G: Taxonomy, research challenges, and the road ahead. IEEE Open Journal of the Communications Society. 2022;3:2137-2150
  4. 4. Kshetrimayum RS, Mishra M, Aïssa S, Koul SK, Sharawi MS. Diversity order and measure of MIMO antennas in single-user, multiuser, and massive MIMO wireless communications. IEEE Antennas and Wireless Propagation Letters. 2023;22(1):19-23
  5. 5. Gkonis PK. A survey on machine learning techniques for massive MIMO configurations: Application areas, performance limitations and future challenges. IEEE Access. 2023;11:67-88
  6. 6. Zakavi MJ, Nezamalhosseini SA, Chen LR. Multiuser massive MIMO-OFDM for visible light communication systems. IEEE Access. 2023;11:2259-2273
  7. 7. Ning B, Tian Z, Mei W, Chen Z, Han C, Li S, et al. Beamforming technologies for ultra-massive MIMO in terahertz communications. IEEE Open Journal of the Communications Society. 2023;4:614-658
  8. 8. Molodtsov V, Bychkov R, Osinsky A, Yarotsky D, Ivanov A. Beamspace selection in multi-user massive MIMO. IEEE Access. 2023;11:18761-18771
  9. 9. Elayan H, Amin O, Shihada B, Shubair RM, Alouini MS. Terahertz band: The last piece of RF spectrum puzzle for communication systems. IEEE Open Journal of the Communications Society. 2020;1:1-32
  10. 10. Izadinasab K, Shaban AW, Damen O. Detection for hybrid beamforming millimeter wave massive MIMO systems. IEEE Communications Letters. 2021;25(4):1168-1172
  11. 11. Kebede T, Wondie Y, Steinbrunn J, Kassa HB, Kornegay KT. Precoding and beamforming techniques in mm wave massive MIMO: Performance assessment. IEEE Access. 2022;10:16365-16387
  12. 12. Wei C, Yang Z, Dang J, Li P, Wang H, Yu X. Accurate wideband channel estimation for Thz massive MIMO systems. IEEE Communications Letters. 2023;27(1):293-297
  13. 13. Zhang P, Li J, Wang H, You X. Millimeter-Wave Space-Time Propagation Characteristics in Urban Macrocell Scenarios. In: ICC 2019 - 2019 IEEE International Conference on Communications (ICC), Shanghai, China. 2019. pp. 1-6. DOI: 10.1109/ICC.2019.8761087
  14. 14. Nadeem Q-U-A, Alwazani H, Kammoun A, Chaaban A, Debbah M, Alouini M-S. Intelligent reflecting surface assisted multi-user MISO communication: Channel estimation and beamforming design. IEEE Open Journal of the Communications Society. 2020;1:661-680
  15. 15. Liu Y, Liu X, Mu X, Hou T, Xu J, Di Renzo M, et al. Reconfigurable intelligent surfaces: Principles and opportunities. IEEE Communications Surveys & Tutorials. 2021;23(3):1546-1577
  16. 16. Björnson E, Wymeersch H, Matthiesen B, Popovski P, Sanguinetti L, de Carvalho E. Reconfigurable intelligent surfaces: A signal processing perspective with wireless applications. IEEE Signal Processing Magazine. 2022;39(2):135-158
  17. 17. Zhang S, Zhang L, Lu Y, Ding T. Research on the propagation characteristics of millimeter wave signals in complicated enclosed spaces. In: Proceedings of the 5th China Aeronautical Science and Technology Conference. Singapore: Springer Singapore; 2022. pp. 1015-1021
  18. 18. Lu L, Li GY, Swindlehurst AL, Ashikhmin A, Zhang R. An overview of massive MIMO: Benefits and challenges. IEEE Journal of Selected Topics in Signal Processing. 2014;8(5):742-758
  19. 19. Wu M, Yin B, Wang G, Dick C, Cavallaro JR, Studer C. Large-scale MIMO detection for 3GPP LTE: Algorithms and FPGA implementations. IEEE Journal of Selected Topics in Signal Processing. 2014;8(5):916-929
  20. 20. Ma J, Ping L. Data-aided channel estimation in large antenna systems. IEEE Transactions on Signal Processing. 2014;62(12):3111-3124
  21. 21. Ciuonzo D, Rossi PS, Dey S. Massive MIMO channel aware decision fusion. IEEE Transactions on Signal Processing. 2015;63(3):604-619
  22. 22. Wang S, Li Y, Wang J. Multiuser detection in massive spatial modulation MIMO with low-resolution ADCs. IEEE Transactions on Wireless Communications. 2015;14(4):2156-2168
  23. 23. Peng Y, Li Y, Wang P. An enhanced channel estimation method for millimeter wave systems with massive antenna arrays. IEEE Communications Letters. 2015;19(9):1592-1595
  24. 24. Qin X, Yan Z, He G. A near-optimal detection scheme based on joint steepest descent and Jacobi method for uplink massive MIMO systems. IEEE Communications Letters. 2016;20(2):276-279
  25. 25. Choi J, Mo J, Heath RW. Near maximum-likelihood detector and channel estimator for uplink multiuser massive MIMO systems with one-bit ADCs. IEEE Transactions on Communications. 2016;64(5):2005-2018
  26. 26. Khwandah SA, Cosmas JP, Lazaridis PI, Zaharis ZD, Chochliouros IP. Massive MIMO systems for 5G communications. Wireless Personal Communications. 2021;120(3):2101-2115
  27. 27. Vasudevan K, Kota S, Pathak GK, Reddy APK, Kumar L. Turbo coded single user massive MIMO. In: 1st Massive MIMO Workshop. IEEE Future Networks; 2021
  28. 28. Sun H, Ng C, Huo Y, Hu RQ, Wang N, Chen C-M, et al. Massive MIMO. In: International Network Generations Roadmap 2022 Edition. IEEE Future Networks; 2022
  29. 29. Vasudevan K, Madhu K, Singh S. Data detection in single user massive MIMO using Re-transmissions. The Open Signal Processing Journal. 2019;6:15-26
  30. 30. Vasudevan K, Madhu K, and Singh S. Scilab Code for Data Detection in Single User Massive MIMO using Re-transmissions. 2019. Available from: https://www.codeocean.com/
  31. 31. Vasudevan K, Reddy APK, Pathak GK, Singh S. On the probability of erasure for MIMO OFDM. Semiconductor Science and Information Devices. 2020;2(1):1-5
  32. 32. Vasudevan K, Singh S, Reddy APK. Coherent receiver for turbo coded single-user massive MIMO-OFDM with retransmissions. In: Mohammady S, editor. Multiplexing. London: IntechOpen; 2019. pp. 1-21
  33. 33. Vasudevan K, Singh S, and Reddy APK. Scilab Code for Coherent Receiver for Turbo Coded Single-User Massive MIMO-OFDM with Retransmissions. 2019. Available from: https://www.codeocean.com/
  34. 34. Vasudevan K, Pathak GK, Reddy APK. Turbo coded single user massive MIMO with precoding. In: Proc. of the 1st IFSA Winter Conference on Automation, Robotics & Communications for Industry 4.0 (ARCI’ 2021). Chamonix-Mont-Blanc: IFSA; 2021. pp. 6-11
  35. 35. Vasudevan K, Pathak GK, and Reddy APK. Scilab Code for Turbo Coded Single User Massive MIMO with Precoding. 2021. Available from: https://www.codeocean.com/
  36. 36. Vasudevan K, Reddy APK, Pathak GK, Albreem M. Turbo coded single user massive MIMO. Sensors & Transducers Journal. 2021;252(5):65-75
  37. 37. Vasudevan K, Reddy APK, Pathak GK, and Albreem MAM. Scilab code for turbo coded single user massive MIMO. 2021. Available from: https://www.codeocean.com/
  38. 38. Vasudevan K. Turbo equalization of serially concatenated turbo codes using a predictive DFE-based receiver. Signal, Image and Video Processing. 2007;1(3):239-252
  39. 39. Bahl L, Cocke J, Jelinek F, Raviv J. Optimal decoding of linear codes for minimizing symbol error rate. IEEE Transactions on Information Theory. 1974;20(2):284-287
  40. 40. Vasudevan K. Digital Communications and Signal Processing, Second Edition (CDROM Included). Hyderabad: Universities Press (India); 2010. Available from: https://www.universitiespress.com
  41. 41. Vasudevan K. Coherent detection of turbo-coded OFDM signals transmitted through frequency selective Rayleigh fading channels with receiver diversity and increased throughput. Wireless Personal Communications. 2015;82(3):1623-1642. DOI: 10.1007/s11277-015-2303-8
  42. 42. Tüchler M, Koetter R, Singer AC. Turbo equalization: Principles and new results. IEEE Transactions on Communications. 2002;50(5):754-767
  43. 43. ten Brink S. Convergence behaviour of iteratively decoded parallel concatenated codes. IEEE Transactions on Communications. 2001;49(10):1727-1737
  44. 44. Papoulis A. Probability, Random Variables and Stochastic Processes. 3rd ed. New York: McGraw-Hill; 1991
  45. 45. Vasudevan K. Analog Communications: Problems & Solutions. Springer: Ane Books; 2018
  46. 46. Vasudevan K. Detection of signals in correlated interference using a predictive VA. Signal Processing Journal, Elsevier Science. 2004;84(12):2271-2286

Written By

Kasturi Vasudevan, Surendra Kota, Lov Kumar and Himanshu Bhusan Mishra

Reviewed: 06 July 2023 Published: 01 August 2023