You are on page 1of 16

This is a postprint version of the following published document:

Ostman, J., Lancho, A., Durisi, G. & Sanguinetti, L.


(2021). URLLC With Massive MIMO: Analysis and
Design at Finite Blocklength. IEEE Transactions on
Wireless Communications, 20(10), 6387-6401.

DOI: 10.1109/twc.2021.3073741

© 2021 IEEE. Personal use of this material is permitted. Permission


from IEEE must be obtained for all other uses, in any current or
future media, including reprinting/republishing this material
for advertising or promotional purposes, creating new collective
works, for resale or redistribution to servers or lists, or reuse
of any copyrighted component of this work in other works.
1

URLLC with Massive MIMO:


Analysis and Design at Finite Blocklength
Johan Östman, Student Member, IEEE, Alejandro Lancho, Member, IEEE, Giuseppe Durisi, Senior
Member, IEEE, and Luca Sanguinetti, Senior Member, IEEE

Abstract—The fast adoption of Massive MIMO for high- packet over independently fading frequency resources. Thus,
throughput communications was enabled by many research the spatial diversity offered by multiple antennas becomes crit-
contributions mostly relying on infinite-blocklength information- ical to achieve the desired reliability. The latest instantiation of
theoretic bounds. This makes it hard to assess the suitability of
Massive MIMO for ultra-reliable low-latency communications multiple antenna technologies is the so-called Massive MIMO
(URLLC) operating with short-blocklength codes. This paper (multiple-input multiple-output), which refers to a wireless
provides a rigorous framework for the characterization and network where base stations (BS), equipped with a very large
numerical evaluation (using the saddlepoint approximation) of number M of antennas, serve a multitude of UEs via linear
the error probability achievable in the uplink and downlink spatial signal processing [4]. Thanks to the intense research
of Massive MIMO at finite blocklength. The framework en-
compasses imperfect channel state information, pilot contami- performed since its inception in 2010, the advantages of
nation, spatially correlated channels, and arbitrary linear spatial Massive MIMO in terms of spectral efficiency [5], [6], energy
processing. In line with previous results based on infinite- efficiency [7], and power control [8] are well understood, and
blocklength bounds, we prove that, with minimum mean-square its key ingredients have made it into the 5G standard [9].
error (MMSE) processing and spatially correlated channels, the However, all these results have mainly been established in the
error probability at finite blocklength goes to zero as the number
M of antennas grows to infinity, even under pilot contamination. ergodic regime, where the propagation channel evolves ac-
However, numerical results for a practical URLLC network setup cording to a block-fading model, and each codeword spans an
involving a base station with M = 100 antennas, show that a increasingly large number of independent fading realizations
target error probability of 10−5 can be achieved with MMSE as the codeword length goes to infinity (infinite-blocklength
processing, uniformly over each cell, only if orthogonal pilot regime). Since these assumptions are highly questionable in
sequences are assigned to all the users in the network. Maximum
ratio processing does not suffice. URLLC scenarios [10], it remains unclear whether the design
guidelines that have been obtained so far for Massive MIMO
Index Terms—Massive MIMO, ultra-reliable low-latency com-
(see [11], [12] for a detailed review on the topic) apply to
munications, finite blocklength information theory, saddlepoint
approximation, outage probability, pilot contamination, MR and URLLC deployments.
MMSE processing, asymptotic analysis.
A. Prior Art
I. I NTRODUCTION Unlike the vast majority of literature on Massive MIMO,
which focuses on the aforementioned ergodic regime, the
Among the new use cases that will be supported by next authors in [13], [14] assume that the fading channel stays
generation wireless systems [2], some of the most challenging constant during the transmission of a codeword (the so-called
ones fall into the category of ultra-reliable low-latency com- quasi-static fading scenario) and use outage capacity [15]
munications (URLLC). For example, in URLLC for factory as asymptotic performance metric. Although the quasi-static
automation [3], small payloads on the order of 100 bits must fading scenario is relevant for URLLC, the infinite blocklength
be delivered within hundreds of microseconds and with a assumption may yield incorrect estimates of the error proba-
reliability no smaller than 99.999%. To achieve such a high bility. The use of outage capacity in the context of URLLC is
reliability, it is crucial to exploit diversity. Unfortunately, often justified by the results reported in [16], where it is proved
the stringent latency requirements prevent the exploitation of that short channel codes operate close to the outage capacity
diversity in time. Furthermore, the use of frequency diversity for quasi-static fading channels. More specifically, the authors
is problematic, especially in the uplink where current standard- of [16] proved that the difference between the outage capacity
ization rules do not allow user equipments (UEs) to spread a and the maximum coding rate, achievable at finite blocklength
Parts of this paper have been presented at the Asilomar Conf. Signals, Syst., over quasi-static fading channels, goes to zero much faster than
Comput., Pacific Grove, CA, USA, Dec. 2019 [1], and will be presented at the difference between the capacity and the maximum coding
the IEEE Int. Conf. Commun. (ICC), Montreal, Canada, Jun. 2021. rate achievable over additive white Gaussian noise (AWGN)
Johan Östman, Alejandro Lancho, and Giuseppe Durisi are with the
Department of Electrical Engineering, Chalmers University of Technology, channels. The intuition is that the dominant sources of errors
Gothenburg 41296, Sweden (e-mail: {johanos,lanchoa,durisi}@chalmers.se). in quasi-static fading channels are deep-fade events, which
Luca Sanguinetti is with the Dipartimento di Ingegneria dell’Informazione, cannot be alleviated through the use of channel codes, since
University of Pisa, 56122 Pisa, Italy (e-mail: luca.sanguinetti@unipi.it).
The work of Johan Östman, Alejandro Lancho and Giuseppe Durisi was channel coding provides protection only against additive noise.
partly supported by the Swedish Research Council under grant 2016-03293, The application of this result to Massive MIMO is prob-
and by the Wallenberg AI, Autonomous Systems, and Software Program. Luca lematic since, as M grows, we start observing channel hard-
Sanguinetti was in part supported by the Italian Ministry of Education and
Research (MIUR) in the framework of the CrossLab project (Departments of ening and the underlying effective channel (after precod-
Excellence). ing/combining) becomes more similar to an AWGN channel.
2

As a consequence, finite-blocklength effects become more network, with imperfect channel state information, pilot con-
pronounced, since additive noise turns into the dominating tamination, and spatially correlated channels. Both minimum
impairment. Another unsatisfactory feature of the outage- mean-square error (MMSE) and maximum ratio (MR) process-
capacity framework is its inability to account for the channel ing are considered. We remark that the application of the RCUs
state information (CSI) acquisition overhead, caused by the bound and saddlepoint approximation to characterize the error
transmission of pilot sequences. Indeed, quasi-static fading probability in this scenario is novel. Furthermore, differently
channels can be learnt perfectly at the receiver in the asymp- from [25], the proposed saddlepoint approximation involves
totic limit of large blocklength with no rate penalty: it is quantities that can be characterized in closed form. Hence,
enough to let the number of pilot symbols grow sublinearly it can be evaluated efficiently. We prove that the average
with the blocklength. The attempts made so far to include error probability at finite blocklength with MMSE tends to
channel-estimation overhead in the outage setup [13], [14] are zero as M → ∞, whereas it converges to a positive number
not convincing from a theoretical perspective. A theoretically when MR is used. These results are similar in flavor to those
satisfying framework must include the use of a mismatch about Massive MIMO ergodic rates in the infinite-blocklength
receiver that treats the channel estimate, obtained using a regime (see, e.g., [6] and [26]).
fixed number of pilot symbols, as perfect. One difficulty is Through numerical experiments, we estimate the error prob-
that a fundamental result commonly used in the ergodic case ability achievable for finite values of M and quantify the
to bound the mutual information, by treating the channel impact of spatial correlation and pilot contamination. Inspired
estimation error as noise (see, e.g., [17, Lemma B.0.1]), by [27], we use the network availability as performance metric,
does not apply to the outage case. This is because, in the which we define as the fraction of UE placements for which
outage setup, the fading channel stays constant over the entire the per-link error probability, averaged over the small-scale
codeword, and one is interested in computing an outage event fading and the additive noise, is below a given target. In the
over fading realizations. This means that both the channel and asymptotic outage setting, this quantity is obtained by charac-
its estimate must be treated as deterministic quantities when terizing the metadistribution of the signal-to-interference ratio
computing bounds on the instantaneous spectral efficiency. (SIR) [27]. At finite blocklength, the network availability turns
The limitation of both ergodic and outage setups can be out to be related to the metadistribution of the so called
overcome by performing a nonasymptotic analysis of the generalized information density [23, Eq. (3)].
error probability based on the finite-blocklength information- The numerical experiments show that, for finite values of
theoretic bounds introduced in [18] and extended to fading M , it is important to take into account spatial correlation to
channels in [16], [19], [20]. This approach has been pursued obtain realistic estimates of the error probability. Furthermore,
recently in [21], [22]. However, the analysis in these papers pilot contamination turns out to have a strong impact on perfor-
relies on the so called normal approximation [18, Eq. (291)], mance. Consider for example a network with four 75 m×75 m
whose tightness for the range of error probabilities of interest cells, K = 10 UEs, M = 100 BS antennas. Furthermore,
in URLLC is questionable. Also, the use of the normal assume a transmit power of 10 dBm in UL and DL, an error
approximation for the case of imperfect CSI in both [21], [22] probability target of 10−5 and a fixed frame of 300 symbols,
is not convincing, since the approximation does not depend which accommodates pilots and data transmission in UL and
on the instantaneous channel estimation error, but only on its DL. Assume also that in each data transmission phase, 160
variance. This is not compatible with a scenario in which the information bits need to be conveyed with an error probability
channel stays constant over the duration of each codeword. target of 10−5 . For this scenario, a network availability above
90% can be achieved with MMSE processing in UL and DL
B. Contributions only if pilot contamination is avoided by allocating as many
To verify if the design guidelines developed for Massive pilot symbols as the total number of UEs in the network.
MIMO in the context of non-delay limited, large-throughput, In contrast, when all cells use the same pilot sequences, a
communication links apply also to the URLLC setup, we network availability just above 50% is achieved despite the
present a rigorous nonasymptotic characterization of the error fact that the shorter duration of the pilot sequences allows for
probability achievable in Massive MIMO. Specifically, we a larger number of channel uses in the data phase. With MR
provide a firm upper bound on the error probability, which processing, the network availability remains below 50% for
is obtained by adapting the random-coding union bound with both UL and DL, even when pilot contamination is avoided.
parameter s (RCUs) introduced in [23] to the case of Massive These numerical results suggest the following guidelines for
MIMO communications. The resulting bound applies to Gaus- the design of Massive MIMO for URLLC applications: i) Pilot
sian codebooks, and holds for any linear processing scheme contamination must be avoided; ii) In line with [26], MMSE
and any pilot-based channel estimation scheme. Since the should be chosen in place of the simpler MR.
bound is in terms of integrals that are not known in closed form
and need to be evaluated numerically, which is impractical
when the targeted error probability is low, we also present C. Paper Outline and Notation
an accurate and easy-to-compute approximation, based on the In Section II, we present the finite-blocklength framework
saddlepoint method [24, Ch. XVI]. that will be used to analyze and design Massive MIMO
We then use the bound to evaluate the error probability networks. In Section III, the finite-blocklength framework
in the uplink (UL) and downlink (DL) of a Massive MIMO is used to analyze the impact on the error probability of
3

pilot contamination, spatial correlation, and of the number A receiver operating according to (2) is known as mis-
of BS antennas, by focusing on a single-cell network with matched scaled nearest-neighbor (SNN) decoder [17].
two UEs. The analysis is extended to a general multicell Note that it coincides with the optimal maximum like-
multiuser setting in Section IV. Some conclusions are drawn lihood decoder if and only if gb = g.
in Section V. We are interested in deriving an upper bound on the error
Lower-case bold letters are used for vectors and upper-case probability  = P[b
q 6= q] achieved by the SNN decoding rule
bold letters are used for matrices. The circularly-symmetric (2). To do so, we follow a standard practice in information
Gaussian distribution is denoted by CN (0, σ 2 ), where σ 2 theory and use a random-coding approach [28]. Specifically,
denotes the variance. We use E[·] to indicate the expectation we consider a Gaussian random code ensemble, where the
operator, and P[·] for the probability of a set. The natural log- elements of each codeword are drawn independently from
arithm is denoted by log(·), and Q(·) stands for the Gaussian a CN (0, ρ) distribution.1 Here, ρ can be thought of as the
Q-function. The Frobenius and spectral norms of a matrix X average transmit power. We consider the cases where the
are denoted by kXkF and kXk2 , respectively. The operators channel gain g in (1) can be modelled as a deterministic or a
(·)T , (·)∗ , and (·)H denote transpose, complex conjugate, and random variable. In the literature, this latter case is commonly
d
Hermitian transpose, respectively. Finally, we use = to denote referred to as quasi-static fading setting [30, p. 2631].
equality in distribution while, for two random sequences an , Theorem 1: Assume that g ∈ C and gb ∈ C in (1) are
bn , we write an  bn to indicate that limn→∞ (an − bn ) = 0 deterministic. There exists a coding scheme with m = 2b
almost surely. codewords of length n operating according to the mismatched
SNN decoding rule (2), whose error probability  is upper-
D. Reproducible Research bounded by2
The Matlab code used to obtain the simulation results is q 6= q]
 = P[b
available at: https://github.com/infotheorychalmers/URLLC " n #
Massive MIMO.
X
≤P ıs (q[k], v[k]) + log(u) ≤ log(m − 1) (3)
k=1
II. A F INITE -B LOCKLENGTH U PPER -B OUND ON THE for all s > 0. Here, u is a random variable that is uniformly
E RROR P ROBABILITY distributed over the interval [0, 1] and ıs (q[k], v[k]) is the
In this section, we present a finite-blocklength upper bound generalized information density, given by
on the error probability and describe an efficient method for its
2
numerical evaluation, based on the saddlepoint approximation ıs (q[k], v[k]) = −s |v[k] − gbq[k]|
[24, Ch. XVI]. We start by considering the simple case in s|v[k]|2
g |2 . (4)

which the received signal is the superposition of a scaled + 2
+ log 1 + sρ|b
1 + sρ|bg|
version of the desired signal and additive Gaussian noise. This
simple channel model constitutes the building block for the Assume now that g ∈ C and gb ∈ C in (1) are random variables
analysis of the error probability achievable in the Massive drawn according to an arbitrary joint distribution. Then, for all
MIMO networks considered in Sections III and IV. s > 0, the error probability  is upper-bounded by
q 6= q]
 = P[b
A. Upper Bound for Deterministic and Random Channels " " n ##
X m − 1
Consider a discrete AWGN channel given by ≤ Eg,bg P ıs (q[k], v[k]) ≤ log g, gb (5)
u
k=1
v[k] = gq[k] + z[k], k = 1, . . . , n (1)
where the average is taken over the joint distribution of g and
where q[k] ∈ C and v[k] ∈ C are the input and output over gb. If g ∈ C is a random variable and gb ∈ C is deterministic,3
channel use k, respectively, and n is the codeword length. the average in (5) is only taken over the distribution of g.
Furthermore, g ∈ C is the channel gain, which is assumed to Proof: The proof for the case of g and gb being determinis-
remain constant during transmission of the n-length codeword. tic, which is given in Appendix A for completeness, follows by
The additive noise variables {z[k] ∈ C; k = 1, . . . , n}, particularizing the RCUs bound introduced in [23, Th. 1] to the
are independent and identically distributed (i.i.d.), CN (0, σ 2 ), considered setup. The upper bound for random g and gb readily
random variables. In what follows, we assume that: follows by taking an expectation over the joint distribution of
1) The receiver does not know the channel gain g but has g and gb.
an estimate gb of g that is treated as perfect. Coarsely speaking, Theorem 1 shows that the error prob-
2) To determine the transmitted codeword ability in the finite-blocklength regime can be characterized
q = [q[1], . . . , q[n]]T , the receiver seeks the codeword 1 Note that this ensemble is not optimal at finite blocklength, not even if
e from the codebook C that, once scaled by gb, is the
q b = g. However, it is commonly used to obtain tractable expressions and
g
closest to the received vector v = [v[1], . . . , v[n]]T ∈ Cn insights into the performance of communication systems [11], [12], [29]. Our
in Euclidean distance. Mathematically, the estimated analysis can be extended to other ensembles—see, e.g., [20].
2 Note that the probability in (3) is computed with respect to the channel
codeword q b is obtained as inputs {q[k]}n k=1 , the additive noise {z[k]}k=1 , and the random variable u.
n
3 This case will turn out important to analyze the DL of Massive MIMO
ek2 .
b = arg minkv − gbq
q (2)
e∈C
q
networks.
4

in terms of the probability that the empirical average of the main idea of the saddlepoint method is to perform an ex-
generalized information density ıs is smaller than the chosen ponential tilting [24, Ch. XVI.7] on the random variables
rate R = (log m)/n. In contrast, in the infinite-blocklength {ıs (q[k], v[k]), k = 1, . . . , n}, which moves their mean close
regime, the error (outage) probability, is given by the prob- to the desired rate R. This guarantees that a subsequent use
ability that the so-called generalized mutual information [17, of the normal approximation yields small errors.
Sec. III] Is = E[ıs (q[1], v[1])] is below the chosen rate. If g The saddlepoint method has been applied to obtain accurate
is known at the receiver, i.e., gb = g, it follows immediately approximations of the RCUs in, e.g., [31] and [25]. In the
from the decoding rule (2) that  → 0 when the SNR grows following, we particularize these expressions to the setup con-
unboundedly, i.e., ρ/σ 2 → ∞. The following lemma shows sidered in Theorem 1 and refer to [25], [31] for further details
that this is also true for the upper bounds (3) and (5). and proofs. While to obtain (7), it is sufficient to check that
Lemma 1: If g = gb, then the third central moment of ıs (q[k], v[k]) is bounded (which
" n # is indeed the case in our setup), the existence of a saddlepoint
X m−1 approximation requires the more stringent condition that the
lim P ıs (q[k], v[k]) ≤ log = 0. (6)
ρ/σ 2 →∞ u third derivative of the moment-generating function (MGF) of
k=1
−ıs (q[k], v[k]) exists in a neighborhood of zero. Specifically,
Proof: This result is easily established by setting v[k] = we require that there exist two values ζ < 0 < ζ such that
gq[k] and gb = g in (4) and by noting that one can make (4)
arbitrarily large by choosing s sufficiently large. d3 h −ζıs (q[k],v[k]) i
sup e < ∞. (9)
We anticipate that Lemma 1 will be important for the ζ<ζ<ζ dζ
3 E
characterization of the error probability of Massive MIMO in
the asymptotic limit of large antenna arrays, i.e., M → ∞. As shown in Appendix B, this condition is verified in our
The upper bounds in (3) and (5) involve the evaluation of setup. Specifically, we have that
a tail probability, which is not known in closed form and p
(βB − βA )2 + 4βA βB (1 − ν) + βA − βB
needs to be evaluated numerically. Furthermore, they can be ζ=− (10)
tightened by performing an optimization over the parameter 2βA βB (1 − ν)
p
s > 0, which also needs to be performed numerically. All this (βB − βA )2 + 4βA βB (1 − ν) − βA + βB
ζ= (11)
is computational demanding, especially when one targets the 2βA βB (1 − ν)
low error probabilities required in URLLC applications. In the where
next section, we discuss how this problem can be alleviated
by using a saddlepoint approximation. βA = s(ρ|g − gb|2 + σ 2 ) (12)
s
ρ|g|2 + σ 2 (13)

βB =
1 + sρ|bg |2
B. Saddlepoint Approximation 2
s2 ρ|g|2 + σ 2 − g ∗ gbρ

One possible way to numerically approximate (3) and (5) ν = . (14)
is to perform a normal approximation on the probability βA βB (1 + sρ|b g |2 )
term based on the Berry-Esseen central limit theorem [24, The saddlepoint approximation that will be provided in The-
Ch. XVI.5]. This leads to the following expansion: orem 2 below depends on the cumulant-generating function
" n # (CGF) of −ıs (q[k], v[k])
X m−1 h i
P ıs (q[k], v[k]) ≤ log κ(ζ) = log E e−ζıs (q[k],v[k]) (15)
u
k=1
nIs − log(m − 1) and on its first derivative κ0 (ζ) and second derivative κ00 (ζ).
   
1
=Q √ +o √ (7)
nVs n In our setup, these quantities can be computed in closed form
for all ζ ∈ (ζ, ζ) and are given by (see Appendix B)
where Is = E[ıs (q[1], v[1])] is the so-called generalized
mutual information [17, Sec. III], g |2

κ(ζ) = − ζ log 1 + sρ|b
− log 1 + (βB − βA ) ζ − βA βB (1 − ν)ζ 2

Vs = E |ıs (q[1], v[1]) − Is |2 (8)
 

(16)
is the variance of the information density, typically
√ referred to 0 2

as channel dispersion [18, Sec. IV], √ and o(1/ n) accounts κ (ζ) = − log 1 + sρ|b g|
for terms that decay faster than 1/ n as n → ∞. The (βB − βA ) − 2βA βB (1 − ν)ζ
− (17)
so-called
√ normal approximation obtained by neglecting the
1 + (βB − βA ) ζ − βA βB (1 − ν)ζ 2
o(1/ n) term in (7) is accurate only when R = (log m)/n
2
(βB − βA ) − 2βA βB (1 − ν)ζ

is close to Is [25]. Unfortunately, this is typically not the κ00 (ζ) =
1 + (βB − βA ) ζ − βA βB (1 − ν)ζ 2
case in URLLC since one needs to operate at rates much 2βA βB (1 − ν)
lower than Is to obtain the required low error probabilities + . (18)
1 + (βB − βA ) ζ − βA βB (1 − ν)ζ 2
at SNR values of practical interest (see, e.g., [25, Fig. 3]).
A more accurate approximation, that holds for all values Note that −κ(ζ) coincides with the so-called Gallager’s E0
of R, can be obtained using the saddlepoint method. The function for the mismatched case [23, Eq. (22)]. As a conse-
5

quence, we have that Is = −κ0 (0). Furthermore, the so-called error-exponent term, i.e., the exponential term in (20), (22),
critical rate Rscr (see [28, Eq. (5.6.30)]) is given by and (24), which governs the exponential decay of the error
probability as a function of the blocklength, and then to use
Rscr = −κ0 (1). (19) the Berry-Esseen central-limit theorem to characterize only
We are now ready to present the saddlepoint expansion of the the pre-exponential factor, i.e., the factor that multiplies the
RCUs bound (3). exponential term. It is also worth highlighting that since all
Theorem 2: Let m = enR for some R > 0, and let ζ ∈ (ζ, ζ) quantities in (20), (22), and (24) are known is closed form,
be the solution to the equation R = −κ0 (ζ).4 If ζ ∈ [0, 1], then the evaluation of the saddlepoint approximation, for a given ζ
Rscr ≤ R ≤ Is and and its corresponding rate R = −κ0 (ζ), entails a complexity
" n # similar to that of the normal approximation (7).
X enR − 1 Note that both the saddlepoint approximation and the
P ıs (q[k], v[k]) ≤ log
u normal approximation can be tightened by performing an
k=1
   optimization over s, which may be time consuming. One
1
=e n[κ(ζ)+ζR]
Ψn,ζ (ζ) + Ψn,ζ (1 − ζ) + o √ (20) way to avoid this step is to choose an s that is optimal
n in some asymptotic regime. One can for example set s so
where as to maximize the generalized mutual information Is . The
u2 00
 p  corresponding value for s can be obtained in closed form [17,
Ψn,ζ (u) , en 2 κ (ζ) Q u nκ00 (ζ) (21) Eq. (64)].
√ √
and o(1/ n) comprises terms that vanish faster than 1/ n
and are uniform in ζ. C. Outage Probability and Normal Approximation
If ζ > 1, then R < Rscr and Equipped with the bound (5) and with an efficient method
" n # for the numerical evaluation of the probability term within (5),
enR − 1
we can now evaluate the error probability achievable for short
X
P ıs (q[k], v[k]) ≤ log
u blocklengths and investigate whether the outage probability is
k=1
an accurate performance metric in Massive MIMO systems for
  
= en[κ(1)+R] Ψ e n (0, −1) + O √1
e n (1, 1) + Ψ (22)
n URLLC applications. For the sake of simplicity, we consider
a single-UE multiantenna system in which the BS has a large
where number M of antennas. We denote by h ∈ CM the channel
κ00 (1)
between the UE and the BS array and assume that it can be
h i
e n (a1 , a2 ) = ena1 Rscr −R+ 2
Ψ
! modelled as uncorrelated Rayleigh fading h ∼ CN (0M , βIM )
p n(Rcr − R) where β is the large-scale fading gain [11, Sec. 1.3.2]. If
×Q a1 nκ00 (1) + a2 p s (23)
nκ00 (1) perfect CSI is available at the receiver and MR combining is
√ √ used for detection, the UL channel input-output relation can
and O(1/ n) comprises terms that are of order 1/ n and be expressed as
are uniform in ζ. If ζ < 0, then R > Is and
hH hH 0
v[k] = hq[k] + z [k], k = 1, . . . , n (25)
" n #
X enR − 1 khk khk
P ıs (q[k], v[k]) ≤ log
u
k=1
 where z0 [k] ∼ CN (0M , σ 2 IM ) is the thermal noise over
= 1 − en[κ(ζ)+ζR] Ψn,ζ (−ζ) − Ψn,ζ (1 − ζ) the antenna array over channel use k. Note that (25) can
hH
be mapped into (1) by setting g = khk h = khk and
 
1 H
h
z[k] = khk z0 [k] ∼ CN (0, σ 2 ). Since h is perfectly known at
+o √ . (24)
n the receiver, we have that gb = g = khk. In the limit n → ∞, it
Proof: The proof follows along steps similar to [31, App. can be shown that the probability term in (5), once optimized
E] and to [25, App. I], and it is thus omitted because of space over the parameter s, is equal to 1 if log(1 + ρ|g|2 /σ 2 ) < R
limitations. and 0 otherwise. This means that the bound in (5) converges
We√will refer to the approximations obtained by ignoring the to the outage probability

o(1/ n) terms and the O(1/ n) terms in (20), (22), and (24)  
ρg 2
 
as saddlepoint approximations. Note that the exponential term P log 1 + 2 < R . (26)
σ
on the right-hand side of (20) and (22) corresponds to the Gal-
lager error exponent for the mismatch decoding scenario [32]. Here, the probability is evaluated with respect to the random
This means that the saddlepoint approximation provides an variable g = khk.
estimate of the subexponential factor, thereby allowing one In Fig. 1, we depict the outage probability in (26) as a
to obtain accurate approximations of error probability values function of the number of BS antennas M . Comparisons
for which the error exponent is inaccurate. In a nutshell, the are made with the upper bound in (5), evaluated by means
key idea of the saddlepoint method is to isolate the Gallager of both Monte-Carlo integration (exact) and the saddlepoint
approximation in Theorem 2. We also depict the normal ap-
4 The existence of such a solution for all rates R ≥ 0 follows from (17). proximation obtained by averaging (7) over g. In the evaluation
6

100
when M > 20. Both approximations are not accurate at the
outage capacity low error probabilities of interest in URLLC. The saddlepoint
normal approximation
10−1 normal approx. M → ∞ approximation is instead very accurate for all M values.
In Fig. 1b we report the error probability with no power
Error probability, ǫ

RCUs bound, saddlepoint


RCUs bound, exact scaling so that the average received SNR increases as M
10−2
increases. Specifically, we consider a fixed transmit power
10−3 ρ = −24 dBm. Hence, for M = 320 the average received
SNR in Fig. 1b equals 1 dB, which coincides with the average
10−4 received SNR in Fig. 1a. With no power scaling, the outage
probability (26) is an accurate approximation for the RCUs
10−5 0
10 101 102 103 bound (5) only for very large values of the error probability,
Number of antennas, M whereas the accuracy of the normal approximation (7) is
(a) Fixed average received SNR = 1 dB. acceptable for  within the range [10−3 , 1]. The saddlepoint
0
approximation again is on top of the RCUs bound for all values
10
of error probability considered in the figure.
10−1 Based on the above results, we conclude that outage prob-
ability and the normal approximation do not always provide
Error probability, ǫ

10−2
accurate estimates of the error probability achievable in large-
10−3
antenna systems with short-packet communications over quasi-
10−4
outage capacity static channels. The accuracy of these approximations becomes
10−5 normal approximation even more questionable in the presence of imperfect CSI. This
normal approx. M → ∞
RCUs bound, saddlepoint
problem can be avoided altogether by using the nonasymptotic
10−6
RCUs bound, exact bound (5) in Theorem 1, which can be efficiently evaluated
10−7 by means of the saddlepoint approximation in Theorem 2. In
100 150 200 250 300 350 400
Number of antennas, M the next two sections, we will show how the simple input-
output relation (1) can be used as building block for the
(b) Fixed transmit power ρ = −24 dBm.
analysis of practical Massive MIMO networks with imperfect
Fig. 1: Average error probability in the UL of a single-UE multiantenna system CSI, pilot contamination, spatial correlation among antennas,
when g b = g = khk with h ∼ CN (0M , βIM ), n = 100, and R = 0.6
bits per channel use. The UE is assumed to be at a distance from the BS that and both inter-cell and intra-cell interference. Theorem 1 and
results in β/σ 2 = 1. Theorem 2 will then be used to efficiently evaluate the average
error probability also in these more realistic scenarios.

of (5), we set gb = g and optimize over the parameter s by III. A T WO -UE S INGLE -C ELL M ASSIVE MIMO
means of a bisection search.5 We 2assume
 2 that σ =2−94 dBm6
2
S CENARIO
and set β = σ so that E g /σ = βM/σ = M .
2
We consider a single-cell network where the BS is equipped
Furthermore, we consider a codeword length n = 100 and
with M antennas and serves K = 2 single-antenna UEs.
a rate of R = 60/100 = 0.6 bits per channel use.
We denote by hi ∈ CM the channel vector between the BS
In Fig. 1a, we illustrate the error probability for a transmit and UE i for i = 1, 2. We use a correlated Rayleigh fading
power ρ that decreases as 1/M . Specifically, we set ρ = ρe/M model where hi ∼ CN (0M , Ri ) remains constant for the
with ρe = 1 dB. Since g 2 /M → β as M → ∞ and we assume duration of a codeword transmission. The normalized trace
β = σ 2 , it thus follows that the instantaneous SNR ρg 2 /σ 2 βi = tr(Ri )/M determines the average large-scale fading
converges to the deterministic value ρe as M → ∞. This means between UE i and the BS, while the eigenstructure of Ri
that, as M → ∞, the normal approximation for i.i.d. Gaussian describes its spatial channel correlation [11, Sec. 2.2]. We
inputs given in (7) for a fixed g converges to a deterministic assume that R1 and R2 are known at the BS; see, e.g.,
quantity. Specifically, in the limit M → ∞, we have that [26], [33] for a description of practical estimation methods.
Is = log(1 + ρeβ/σ 2 ) (achieved for s = 1/σ 2 ) and Vs = This setup is sufficient to demonstrate the usefulness of the
2eρβ/(eρβ + σ 2 ) [29, Eq. (2.55)]. The resulting approximation framework developed in Section II for the analysis and design
is of interest because it does not require any Monte-Carlo of Massive MIMO networks. A more general setup will be
averaging over the realizations of the fading channel. From considered in Section IV.
Fig. 1a, we see that the outage probability (26) approximates
well the exact RCUs bound (5) only when M is small, i.e.,
M < 5, whereas the normal approximation loses accuracy A. Uplink pilot transmission
We consider the standard time-division duplex (TDD) Mas-
5 In all numerical simulations presented throughout the paper, we will sive MIMO protocol, where the UL and DL transmissions are
always evaluate the error probability bound in (5) using the saddlepoint assigned n channel uses in total, divided in np channel uses for
approximation in Theorem 2, and optimize it over the parameter s via a UL pilots, nul channel uses for UL data, and ndl = n−np −nul
bisection search.
6 With the distance-dependent pathloss model that will be introduced in (39), channel uses for DL data. We assume that the np -length pilot
this corresponds to a distance of 36.4 m. sequence φi ∈ Cnp with φHi φi = np is used by UE i for
7

channel estimation. The elementsp of φi are scaled by the Note that (31) has the same form as (1) with v[k] = y1ul [k],
square-root of the pilot power ρul and transmitted over np q[k] = xul1 [k], g = u1 h1 , and z[k] = u1 h2 x2 [k] + u1 z [k].
H H ul H ul

channel uses. When the UEs transmit their pilot sequences, Furthermore, given {h1 , u1 , h2 }, the random variables {z[k] :
the received pilot signal Ypilot ∈ CM ×np is k = 1, . . . , nul } are conditionally i.i.d. and z[k] ∼ CN (0, σ 2 )
p p with σ 2 = ρul |uH 1 h2 | + ku1 k σul .
2 2 2
Ypilot = ρul h1 φH1 + ρul h2 φH2 + Zpilot (27) We assume that the BS treats the acquired (noisy) channel
where Zpilot ∈ CM ×np is the additive noise with i.i.d. estimate h b 1 as perfect. This implies that, to recover the
elements distributed as CN (0, σul
2
). Assuming that R1 and transmitted codeword, which we assume to be drawn from
R2 are known at the BS, the MMSE estimate of hi is [11, a codebook C ul , it performs mismatched SNN decoding with
Sec. 3.2] gb = uH1 h1 . Specifically, the estimated codeword x 1 is ob-
bul
b
q tained as
b i = ρul np Ri Q−1 Ypilot φi (28)

h i bul
x ul Hb
xul
1 = arg minky1 − (u1 h1 )e 1 k
2
(32)
eul
for i = 1, 2 with 1 ∈C
x ul

Qi = ρul R1 φH1 φi + ρul R2 φH2 φi + σul


2
IM . (29) with y1ul = [y1ul [1], . . . , y1ul [nul ]]T and xeul
1 =
[e ul
e1 [nul ]] . It thus follows that (3) provides
x1 [1], . . . , xul T

The MMSE estimate h b i and the estimation error h ei = a bound on the conditional error probability for UE 1 given
hi − hi are independent random vectors, distributed as h
b bi ∼ g and gb. To obtain the average error probability, we need
CN (0, Φi ) and hi ∼ CN (0, Ri − Φi ), respectively, with
e to take an expectation over g = uH 1 h1 , g 1 h1 , and
b = uH b
Φi = ρul np Ri Q−1 i Ri . σ = ρ |u1 h2 | + ku1 k σul , which results in
2 ul H 2 2 2

It follows from (29) that if the two UEs use orthogonal pilot
ul
sequences, i.e., φH 1 φ2 = 0, they do not interfere, whereas they
1" "n ##
ul
interfere if they use the same pilot sequence, i.e. φ1 = φ2 . X
ul ul m − 1 2
≤E P ıs (y1 [k], x1 [k]) ≤ log g, gb, σ .
This interference is known as pilot contamination and has u
k=1
two main consequences in the channel estimation process [11, (33)
Sec. 3.2.2]. The first is a reduced estimation quality; the second
is that the estimates h b 1 and h b 2 become correlated. To see The saddlepoint approximation in Theorem 2 can be applied
this, observe that if φ1 = φ2 then Ypilot φ1 = Ypilot φ2 verbatim to efficiently compute the conditional probability
and Q1 = Q2 = Q with Q = ρul np R1 + ρul np R2 + in (33). The average error probability for UE 2 can be
2
σul IM . Hence, h b 2 can be written as h b 2 = R2 (R1 )−1 h b1 evaluated similarly.
provided that R1 is invertible. This implies that the two The combining vector u1 is selected at the BS based on
estimates arei correlated with cross-correlation matrix given the channel estimates hb 1 and h b 2 . The simplest choice is to
use MR combining: uMR = b 1 /M . A more computationally
h
by E h b H = Υ12 = ρul np R1 Q−1 R2 . This holds even
b1h 1 h
2
intensive choice is MMSE combining:
though the underlying channels h1 and h2 are statistically
!−1
independent, which implies that E[h1 hH2 ] = 0M . Observe that X2
MMSE
if there is no spatial correlation, i.e., Ri = βi IM , i = 1, 2, u 1 = h
bih H
b +Z
i h
b1 (34)
then the channel estimates are identical up to a scaling factor, i=1
i.e., they are linearly dependent. We will return to the issue of P2 2
σul
where Z = Φi + I .
ρul M
pilot contamination in Section III-E. i=1

C. Downlink data transmission


B. Uplink data transmission
Assume that, to transmit to UE i with i = 1, 2, the BS uses
During UL data transmission, the received complex base- the precoding vector wi ∈ CM , which determines the spatial
band signal rul [k] ∈ CM over an arbitrary channel use k, directivity  of the transmission and satisfies the normalization
where k = 1, . . . , nul , is given by E kwi k2 = 1. During DL data transmission, the received


rul [k] = h1 xul ul ul


(30) signal y1dl [k] ∈ C at UE 1 over channel use k, where k =
1 [k] + h2 x2 [k] + z [k]
1, . . . , ndl , is
7
where xul
i [k] ∼ CN (0, ρ ) is the information bearing signal
ul
y1dl [k] = hH dl H dl dl
1 w1 x1 [k] + h1 w2 x2 [k] + z1 [k] (35)
transmitted by UE i with ρ being the average UL transmit
ul

power and zul [k] ∼ CN (0, σul2


IM ) is the independent additive where xdl
i [k]∼ CN (0, ρ ) is the data signal intended for
dl

noise. The BS detects the signal xul


1 [k] by using the combining UE i and z1dl [k] ∼ CN (0, σdl 2
) is the receiver noise at
vector u1 ∈ CM , to obtain UE 1. Again, we can put (35) in the same form as (1)
by setting v[k] = y1dl [k], q[k] = xdl 1 [k], g = h1 w1 and
H
y1ul [k] = uH ul
1 r [k] z[k] = h1 w2 x2 [k] + z1 [k]. Note that, given {h1 , w1 , w2 },
H dl dl
= u1 h1 xul
H H ul H ul
1 [k] + u1 h2 x2 [k] + u1 z [k]. (31) the random variables {z[k] : k = 1, . . . , ndl } are conditional
7 As detailed in Section II, we will evaluate the error probability for a
i.i.d. and z[k] ∼ CN (0, σ 2 ) with σ 2 = ρdl |hH
1 w2 | + σdl .
2 2

Gaussian random code ensemble, where the elements of each codeword are Since no pilots are transmitted in the DL, the UE does not
drawn independently from a CN (0, ρul ) distribution. know the precoded channel g = hH 1 w1 in (35). Instead, we
8

assume that the UE has access its expected value E hH


 
1 w1 100
and uses this quantity to performmismatched SNN decoding. uncorrelated MR
Specifically, we have that gb = E hH
1 w1 and

10−1

Error probability, ǫ
bdl
x dl
1 = arg minky1 − g
bxedl
1 k
2
(36)
edl
10−2
1 ∈C
x dl

with y1dl = [y1dl [1], . . . , y1dl [ndl ]]T and x edl


1 = 10−3
[edl
e1 [ndl ]] . Obviously, channel hardening is
x1 [1], . . . , x dl T
MMSE
critical for this choice to result in good performance [11, 10−4 M = 100
Sec. 2.5.1]. Since gb = E hH 1 w1 is deterministic, the error
M = 200

probability at UE 1 in the DL can be evaluated as follows: 10−5


−20 −10 0 10 20 30 40 50 60 70 80
Angle of UE 2, ϕ2
" "n ##
dl
dl
X
dl dl m − 1
2
1 ≤ E P ıs (y1 [k], x1 [k]) ≤ log g, σ .

u (a) Uplink transmission.
k=1
100
(37)
uncorrelated
MR
Similarly to (33), the saddlepoint approximation in The- Error probability, ǫ
10−1
orem 2 can be used to evaluate the conditional probability
in (37) efficiently. 10−2
Similar to the UL, the upper bound (37) holds for any
10−3
precoder vector that is selected on the basis of the chan-
MMSE
nel estimates available at the BS. Different precoders yield
10−4
different tradeoffs between the error probability achievable M = 100
M = 200
at the UEs. A common heuristic comes from UL-DL dual- 10−5
ity [11, Sec. 4.3.2], which suggests to choose the precoding −20 −10 0 10 20 30 40 50 60 70 80
vectors wip as the following function of the combining vectors: Angle of UE 2, ϕ2
wi = ui / E[kui k2 ]. By selecting ui as one of the uplink (b) Downlink transmission.
combining schemes described earlier, the corresponding pre- Fig. 2: Average error probability  for UE 1 versus the nominal angle of UE
coding scheme is obtained; that is, ui = uMR i yields MR 2 when φ1 = φ2 . Here, ρul = ρdl = 10 dBm, ∆ = 25◦ , ϕ1 = 30◦ , b =
precoding and ui = uMMSEi yields MMSE precoding. 160, np = 2, and n = 300. The curves are obtained using the saddlepoint
approximation; the circles indicate the values of the RCUs bound, computed
directly via (5).
D. Numerical Analysis
In this section, we use the finite blocklength bound in
Theorem 1 to study the impact of imperfect CSI, pilot con- transmit powers are equal and given by ρul = ρdl = 10 mW.
tamination, and spatial correlation in both UL and DL. We We assume a total of n = 300 channel uses, out of which np
assume that the K = 2 UEs are within a square area of channel uses are allocated for pilot transmission and nul =
75 m × 75 m, with the BS at the center of the square. The ndl = (n − np )/2 channel uses are assigned to the UL and
BS is equipped with a horizontal uniform linear array (ULA) DL data transmissions, respectively. In each data-transmission
with antenna elements separated by half a wavelength. The phase, b = 160 information bits are to be conveyed. These
antennas and the UEs are located in the same horizontal plane, parameters are in agreement with the stringent low-latency
thus the azimuth angle is sufficient to determine the directivity. setups described in [3, App. A.2.3.1].
We assume that the scatterers are uniformly distributed in the Fig. 2 shows the UL and DL error probability  of UE
angular interval [ϕi − ∆, ϕi + ∆], where ϕi is the nominal 1 with MR and MMSE combining, when the two UEs use
angle-of-arrival (AoA) of UE i and ∆ is the angular spread. the same pilot sequence (i.e., pilot contamination is present)
Hence, the (m1 , m2 )th element of Ri is equal to [11, Sec. 2.6] and M = 100 or 200. The uncorrelated Rayleigh-fading case
Z ∆ where Ri = βi IM , i = 1, 2, is also reported as reference. The
βi nominal angle of UE 1 is fixed at ϕ1 = 30◦ while the angle
[Ri ]m1 ,m2 = ejπ(m1 −m2 ) sin(ϕi +ϕ̄) dϕ̄. (38)
2∆ −∆ of UE 2 varies from −20◦ to 80◦ . We let d1 = d2 = 36.4 m,
We assume ∆ = 25◦ and let the large-scale fading coefficient, which leads to β1 = β2 = −94 dB. Fig. 2 reveals that a low
measured in dB, be error probability can be achieved if the UEs are well-separated
  in the angle domain, even when the channel estimates are
di
βi = −35.3 − 37.6 log10 (39) affected by pilot contamination. MMSE combining/precoding
1m achieves a much lower error probability for a given angle
where di is the distance between the BS and UE i. The separation. These results are in agreement with the findings
communication takes place over a 20 MHz bandwidth with reported in the asymptotic regime of large packet size in [6],
a total receiver noise power of σul 2
= σdl2
= −94 dBm [26].
(consisting of thermal noise and a noise figure of 7 dB in the Fig. 2 shows that the error probability with MR combining
receiver hardware) at both the BS and UEs. The UL and DL in the UL is worse than that of MR precoding in the DL. This
9

phenomenon can be clarified by comparing the input-otput 1


relations in (31) and (35) for the case of perfect CSI at both MMSE
BS and UEs. Specifically, when the desired signal experiences 0.9

Network availability, η
a deep fade, the magnitude of the UL interference is unaffected
whereas the DL interference becomes small. This results in a 0.8
larger error probability in the UL compared to the DL. The
same argument holds also for the case of imperfect CSI with 0.7
MR
and without pilot contamination. Note that this phenomenon
does not occur when MMSE combining/precoding is used. 0.6 φ1 = φ2
On the contrary, with MMSE combining/precoding the DL φH
1 φ2 = 0
0.5 −5
performs slightly worse than the UL because DL decoding 10 10−4 10−3 10−2 10−1 100
relies on channel hardening. Target error probability, ǫtarget
Assume now that the 2 UEs are positioned independently
(a) Uplink transmission.
and uniformly at random within the square area of 75 m×75 m,
with a minimum distance from the BS of 5 m. Fig. 3 shows the 1

UL and DL network availability η with both MR and MMSE


0.9
when M = 100. We define η as
Network availability, η
MMSE
MR
η = P[ ≤ target ] (40) 0.8

MR, CSI@UE
and represents the probability that the target error probability 0.7
target is achieved on a link between a randomly positioned
UE and its corresponding BS, in the presence of randomly 0.6 φ1 = φ2
positioned interfering UEs (in this case, just one). Note that φH
1 φ2 = 0

the error probability  is averaged with respect to the small- 0.5 −5


10 10−4 10−3 10−2 10−1 100
scale fading and the additive noise, given the UEs location, Target error probability, ǫtarget
whereas the network availability is computed with respect to
the random UEs locations. We consider both the scenario in (b) Downlink transmission.
which the UEs use orthogonal pilot sequences, i.e., φH1 φ2 = 0, Fig. 3: Network availability η with and without pilot contamination with M =
and the one in which φ1 = φ2 . 100, ρul = ρdl = 10 dBm, np = 2, b = 160, n = 300, and ∆ = 25◦ .
The results of Fig. 3 show that pilot contamination reduces
significantly the network availability irrespective of the pro-
cessing scheme. MR performs better in the DL than in the UL M → ∞ and K = 2.8 Specifically, we prove that, in the
when orthogonal pilot sequences are used. This is in agreement presence of spatial correlation, the error probability vanishes
with what stated when discussing Fig. 2. However, in the case as M → ∞, provided that MMSE combining/precoding is
of pilot contamination, the UL achieves better performance used. To this end, we will proceed similarly as in [6] and
than the DL when the UE relies on channel hardening (and make the following two assumptions.
slightly worse performance than the DL when the UE has Assumption 1: For i = 1, 2, lim inf M M 1
tr(Ri ) > 0 and
access to perfect CSI). Note that this does not contradict what lim supM kRi k2 < ∞.
stated after Fig. 2. Indeed, due to the random UE placements, Assumption 2: For (λ1 , λ2 ) ∈ R2 and i = 1, 2,
the correlation matrix may have low rank. This affects channel
hardening and, consequently, results in a deterioration of the 1
lim inf inf kλ1 R1 + λ2 R2 k2F > 0. (41)
DL performance. For MMSE processing, the UL is always M {(λ1 ,λ2 ):λi =1} M
superior to the DL because the DL relies on channel hardening.
The first condition in Assumption 1 implies that the array
gathers an amount of signal energy that is proportional to
E. Asymptotic Analysis as M → ∞ M . The second condition implies that the increased signal
It is well known that, for spatially uncorrelated Rayleigh energy is spread over many spatial dimensions, i.e., the rank
fading channels, the interference caused by pilot contamination of Ri must be proportional to M . These two conditions are
limits the spectral efficiency of Massive MIMO in the large- commonly invoked in the asymptotic analysis of Massive
blocklength ergodic setup as M → ∞ and the number of UEs MIMO [34]. Assumption 2 requires R1 and R2 to be asymp-
K is fixed, for both MR and MMSE combining/precoding [4], totically linearly independent [26].
[34]. However, it was recently shown in [6] that Massive In Theorem 3 below, we establish that, with MR combining,
MIMO with MMSE combining/precoding is not asymptoti- the probability of error vanishes as M → ∞ if the two UEs
cally limited by pilot contamination when the spatial corre- transmit orthogonal pilot sequences. However, it converges to
lation exhibited by practically relevant channels is taken into a positive constant if they share the same pilot sequence.
consideration.
We show next that a similar conclusion holds for the 8 We consider the case K = 2 for simplicity, although a similar result can
average error probability in the finite-blockength regime when be obtained for arbitrary K using the same approach.
10

100
power, are at the same distance from the BS, and use the same
pilot sequence. Furthermore, we assume that their nominal
10−1 MR limit angles are ϕ1 = 30◦ and ϕ2 = 40◦ . Note that the angle
MR between the two UEs is small. Hence, we expect pilot contam-
Error probability, ǫ

10−2 ination to have a significant impact on the error probability.


As in Fig. 2, we assume that σul 2
= σdl 2
= −94 dBm and
10−3 that the UEs are located 36.4 m away from the BS so that
MMSE β1 = β2 = −94 dB. In Fig. 4a, we illustrate the average
10−4 φ1 = φ2 error probability as a function of M with MR and MMSE.
φH
1 φ2 = 0
We see that, in the presence of pilot contamination, the error
10−5 0
10 101 102 103 104 probability with MR converges to a nonzero constant as M
Number of antennas, M grows, in accordance with Theorem 3. In contrast, the error
(a) Uplink transmission. probability with MMSE goes to 0 as M → ∞, in accordance
0
with Theorem 4. However, a comparison with the orthogonal-
10
pilot case reveals that, for fixed M , pilot contamination has
MR limit a significant impact on the error probability of MMSE. As
10−1
MR shown in Fig. 4b, similar conclusions can be drawn for the
Error probability, ǫ

10−2 DL.

10−3 IV. M ASSIVE MIMO N ETWORK


MMSE
10−4 φ1 = φ2
We will now extend the analysis in Section III to a Massive
φH
1 φ2 = 0
MIMO network with L cells, each comprising a BS with M
10−5 0 antennas and K UEs. We denote by hjli ∼ CN (0M , Rjli ) the
10 101 102 103 104
channel between UE i in cell l and the BS in cell j. The
Number of antennas, M
np -length pilot sequence of UE i in cell j is denoted by the
(b) Downlink transmission. vector φji ∈ Cnp and satisfies kφji k2 = np . We assume that
Fig. 4: Average error probability  of UE 1 versus number of antennas M the K UEs in a cell use mutually orthogonal pilot sequences
with and without pilot contamination. Here, ρul = ρdl = 10 dBm, np = 2, and these pilot sequences are reused in a fraction 1/f of the L
b = 160, n = 300, ∆ = 25◦ , ϕ1 = 30◦ , and ϕ2 = 40◦ .
cells with np = Kf . The channel vectors are estimated using
the MMSE estimator given in [11, Sec. 3.2].
Theorem 3: Let c > 0 be a positive real-valued scalar.
If MR combining is used with uMR
1 = M h1 , then under
1 b
A. Uplink
Assumption 1, The data signal from UE i0 in cell l over an arbitrary
lim ul
1 = 0, if φ1 φ2 = 0,
H
(42) time instant k is denoted by xul li0 [k] ∼ CN (0, ρ ), with ρ
ul ul
M →∞ being the transmit power. To detect xji [k], BS j selects the
ul
lim ul
1 = c, if φ1 = φ2 . (43) combining vector uji ∈ CM , which is multiplied with the
M →∞
received signal rul
j [k] to obtain
Proof: See Appendix C.
Next, we show that, if MMSE combining is used, the error gq[k] z[k]
probability vanishes as M → ∞ even in the presence of pilot
z }| { z }| {
K
contamination. j ul
X
ul
yji [k] = uHji rul H
j [k] = uji hji xji [k] + uHji hjji0 xul
ji0 [k]
Theorem 4: If MMSE combining is used with uMMSE1 given i0 =1,i0 6=i
by (34), then under Assumption 1 and Assumption 2, the | {z } | {z }
Desired signal Intra-cell interference
average error probability ul
1 goes to zero as M → ∞, both
when φH1 φ2 = 0 and when φ1 = φ2 . z
z[k]
}| {
z[k]
z }| {
Proof: The proof is given in Appendix D. It makes use L
X K
X
of the asymptotic analysis presented in [6, App. B] to show + uHji hjli0 xul H ul
li0 [k] + uji zj [k] (44)
that y1ul  xul
1 as M → ∞, even in the presence of pilot l=1,l6=j i0 =1
contamination. Once this is proved, the result follows by
| {z } | {z }
Inter-cell interference Noise
applying Lemma 1 from Section II.
Note that Theorem 3 and Theorem 4 can be extended to for k = 1, . . . , nul . We note that (44) can be put in the same
the DL with a similar methodology. Details are omitted due form as (1) if we set v[k] = yji ul
[k], q[k] = xul ji [k], g =
PK
to space limitations.
H j
uji hji , gb = uji hji , and z[k] = i0 =1,i0 6=i uHji hjji0 xul
H bj
ji0 [k] +
PL PK j
To validate the asymptotic analysis provided by Theorems 3 i0 =1 uji hli0 xli0 [k] + uji zj [k]. Given all chan-
H ul H ul
l=1,l6=j
and 4 and to quantify the impact of pilot contamination for nels and combining vectors, the random variables {z[k] :
values of M of practical interest, we numerically evaluate the k = 1, . . . , nul } are conditionallyPi.i.d. and z[k] ∼
K j 2
UL error probability when the 2 UEs transmit at the same CN (0, σ 2 ) with σ 2 = σul 2
kuji k2 + ρul i0 =1,i0 6=i |uH
ji hji0 | +
11

PL PK j 2
ρul l=1,l6=j i0 =1 |uH ji hli0 | . An upper bound on the error 1
ji then follows by applying (3) in Theorem 1 and
probability ul 0.9 MMSE
then by averaging over g, gb and σ 2 . This bound holds for 0.8

Network availability, η
any choice of vji . In the numerical results, we will consider 0.7
multicell MMSE and MR combining. 0.6
pilot contamination no pilot contamination
0.5
0.4
B. Downlink 0.3
The BS in cell j transmits the DL signal xdl j [k] = 0.2
PK
w 0 x
ji0 =1 ji ji0
dl
[k] where x dl
ji0 [k] ∼ CN (0, ρdl
) is the DL 0.1 MR
data signal intended for UE i0 in cell j over the time index 0
10 20 30 40 50 100 250
k, assigned to a precoding vector wji0 ∈ CM that satisfies Pilot sequence length, np (log scale)
kwji0 k2 = 1 so that ρdl represents the transmit power. The
(a) Uplink transmission.
received signal yji dl
[k] ∈ C for k = 1, . . . , ndl at UE i in cell j
1
is given by
0.9 MMSE, CSI@UE
gq[k] z[k] Network availability, η 0.8
z }| { z }| { MMSE
K 0.7
X
dl
yji [k] = (hjji )H wji xdl
ji [k] + (hjji )H wji0 xdl
ji0 [k]
0.6
pilot contamination no pilot contamination
i0 =1,i0 6=i
0.5
0.4
MR
| {z } | {z }
Desired signal Intra-cell interference 0.3
z[k] z[k] 0.2
z }| { z }| {
L K 0.1
X X
+ (hlji )H wli0 xdl dl
(45) 0
li0 [k] + zji [k] 10 20 30 40 50 100 250
l=1,l6=j i0 =1 Pilot sequence length, np (log scale)
| {z } | {z }
Inter-cell interference Noise (b) Downlink transmission.
where zji dl
[k] ∼ CN (0, σdl 2
) is the receiver noise. The de- Fig. 5: Network availability for target = 10−5 . Here, L = 4, K = 10,
sired signal to UE i in cell j propagates over the pre- ∆ = 25◦ , the cell size is 75 × 75 m, ρul = ρdl = 10 dBm, M = 100,
b = 160, and n = 300.
coded channel gji = (hjji )H wji . The UE does not know
gji and relies on channel hardening h to approximate
i it with
j H
its mean value E[gji ] = E (hji ) wji . As in the UL,
Section III-D, we assume n = 300, nul = ndl = (n − np )/2,
we note that (45) can be put in the same form as (1) if
j H b = 160 and ρul = ρdl = 10 dBm.
we seth v[k] = iyji dl
[k], q[k] = xdl ji [k], g = (hji ) wji ,
K In Fig. 5, we plot the network availability (40) for a fixed
gb = E (hjji )H wji , and z[k] = i0 =1,i6=i (hjji )H wji0 xdl
P
ji0 [k] + target = 10−5 (which is in agreement with the URLLC
PL PK
l=1,l6=j i0 =1 (hji ) wli0 xli0 [k] + zji [k]. Given all chan-
l H dl dl
requirements) versus the number of pilot symbols np = f K,
nels and precoding vectors, the random variables {z[k] : where we recall that f is the pilot reuse factor. The results
k = 1, . . . , ndl } are conditionally PK i.i.d. and z[k] ∼ presented in Fig. 3 suggest that pilot contamination should
CN (0, σ 2 ) with σ 2 = σdl 2
+ ρdl i0 =1,i6=i0 |(hjji )H wji0 |2 + be avoided and that MMSE should be preferred to MR. The
PL PK
ρdl l=1,l6=j i0 =1 |(hlji )H wli0 |2 . An upper bound on the er- results presented in Fig. 5 confirm these design guidelines.
ror probability dl With multicell MMSE, a network availability above 90% can
ji then follows by applying (3) in Theorem 1
and then by averaging over g and σ 2 . As for the UL, the be achieved in UL and DL by setting a pilot reuse factor
above results hold for any choice of wji . In the numerical f = 4 such that np = f K = 40. This is the minimum
simulations, we will consider both multicell MMSE and MR value of np that results in no pilot contamination in a network
precoding. with L = 4 cells. Increasing np further has a deleterious
effect on the network availability, especially in the DL. Indeed,
the corresponding reduction in the number of channel uses
C. Numerical Analysis ndl = (300 − np )/2 available for data transmission in the DL
The simulation setup consists of L = 4 square cells, overcomes the benefits of a more accurate CSI. As already
each of size 75 m × 75 m, containing K = 10 UEs each, discussed, the difference in performance between UL and DL
independently and uniformly distributed within the cell, at a with multicell MMSE processing is due to the assumption
distance of at least 5 m from the BS. As in Section III-D, that the UE has no CSI and performs mismatched decoding
we consider a horizontal ULA with M = 100 antennas and by relying on channel hardening. Indeed, when the UEs
half-wavelength spacing. The correlation matrix and large- are provided with perfect CSI (black curve), the network
scale fading coefficient associated with each UE follow the availability achievable in UL and DL is the same. If needed,
models given in (38) and (39), respectively. Furthermore, we additional network-availability gains can be achieved by, e.g.,
employ a wrap-around topology as in [11, Sec. 4.1.3]. As in increasing the number of BS antennas, by reducing the number
12

of UEs that are served simultaneously, or by using scheduling the union bound. We next apply the Chernoff bound to f (q, v)
to avoid serving at the same time UEs that are difficult to and obtain that
separate spatially via linear precoding. For example, in the Eq̄ exp −skv − gbq̄k2
 
scenario considered in Fig. 5b, a network availability above f (q, v) ≤ (47)
exp(−skv − gbqk2 )
98% can be achieved by halving the number of scheduled
UEs. Finally, note that the network availability achievable with for s > 0. Substituting (47) into (46), we conclude that
MR is below 50% even when pilot contamination is avoided. " (
This implies that in practical scenarios, MR is too sensitive  ≤ E min 1, exp log(m − 1)
to interference to achieve the low error probability targets
required in URLLC. Eq̄ exp −skv − gbq̄k2
  !)#
+ log (48)
exp(−skv − gbqk2 )
V. C ONCLUSIONS " (
We presented guidelines on the design of Massive MIMO = E min 1, exp log(m − 1)
systems supporting the transmission of short information pack-
ets under the high reliability targets demanded in URLLC. n  !)#
exp −s|v[k] − gbq[k]|2
Specifically, we showed that, for a BS equipped with up
X
− log . (49)
to 100 antennas, it is imperative to avoid pilot contamina- Eq̄[k] [exp(−s|v[k] − gbq̄[k]|2 )]
k=1
tion and to use MMSE spatial processing in place of the Let now the generalized information density be defined as
computationally less intensive MR spatial processing. Our
exp −s|v[k] − gbq[k]|2

guidelines were based on a firm nonasymptotic bound on
ıs (q[k], v[k]) = log . (50)
the error probability, which is based on recent results in Eq̄[k] [exp(−s|v[k] − gbq̄[k]|2 )]
finite-blocklength information theory, and applies to a real-
Using (50), we can rewrite (49) as
istic Massive MIMO network, with imperfect channel state
information, pilot contamination, spatially correlated channels,
" (
arbitrary linear spatial processing, and randomly positioned  ≤ E min 1, exp log(m − 1)
UEs. We provided an accurate approximation for this bound, !)#
n
based on the saddlepoint method, which makes its evaluation X
computationally efficient for the low error probabilities tar- − ıs (q[k], v[k]) . (51)
k=1
geted in URLLC. Finally, we showed that analyses based on
performance metrics such as outage probability and normal The desired bound (3) follows by observing that, for every
approximation, although appealing because of the simplicity positive random variable w, we have that E[min{1, w}] =
of the underlying mathematical formulas, may result in a P[w ≥ u] where u is uniformly distributed on [0, 1].
significant underestimation of the error probability, which is To conclude the proof, it remains to show that the general-
clearly undesirable when designing URLLC links. Results ized information density defined in (50) can be expressed as
relied on the assumption that the channel covariance matrix in (4). Since q̄[k] ∼ CN (0, ρ), it follows that, for a given v[k]
is perfectly known to the receiver. If no such knowledge is 
√ !2 
2
available, the BS can perform instead least-square channel d |b
g | ρ |v[k]| 2
|v[k] − gbq̄[k]|2 = √ + u1 + u22 
estimation followed by regularized zero forcing, at the cost

2 |b
g| ρ
of a performance loss recently quantified in [35].
d g |2 ρ
|b
= θ (52)
A PPENDIX A - P ROOF OF T HEOREM 1 2
Let q = [q[1], . . . , q[n]] ∼ CN (0, ρIn ) be the transmitted where u1 and u2 are independent N (0, 1) random vari-
codeword and v = [v[1], . . . , v[n]] be the corresponding ables and θ follows a noncentral chi-squared distribution
channel output obtained via the input-output relation (1). with 2 degrees of freedom and noncentrality parameter λ =
Finally, let q̄ = [q̄[1], . . . , q̄[n]] be a vector of i.i.d. CN (0, ρ) g |2 ). The MGF of θ is given by
2|v[k]|2 /(ρ|b
random variables, independent of both q and v. Intuitively, q̄
 
λζ
exp 1
stands for any codeword different from the transmitted one. E eζθ =
  1−2ζ
, ζ< . (53)
A simple generalization of the random coding union bound (1 − 2ζ) 2
in [18, Th. 16] to the mismatched SNN decoder (2) results in
Using (53) in (50) with ζ = −s|b
g |2 ρ/2, we conclude that (50)
the following bound
coincides with (4).
 ≤ E[min{1, (m − 1)f (q, v)}] (46)
A PPENDIX B - P ROOF OF (10) AND (11)
where f (q, v) = Pr{kv − gbq̄k2 ≤ kv − gbqk2 |q, v}. The
bound (46) is obtained by observing that, when the mis- In this appendix, we prove that (9) holds for every ζ ∈ [ζ, ζ],
matched SNN decoder (2) is used, an error occurs if, after where ζ and ζ are given in (10) and (11), respectively. Let
being scaled by gb, the codeword q̄ is closer in Euclidean q ∼ CN (0, ρ) and v = gq + z where z ∼ CN (0, σ 2 ), so
distance to v than q, and then by using a tightened version of that v ∼ CN (0, σv2 ) with σv2 = ρ|g|2 + σ 2 Furthermore, set
13

A = s|v − gbq|2 and B = γ|v|2 with γ = s/(1 + sρ|b g |2 ). We sσx2 and βB = γσv2 . Recall also that E[A] = βA , Var(A) =
9
can then rewrite the information density (4) as 2
βA , E[B] = βB , Var(B) = βB2
. Hence, we conclude that
g |2 (54) sγ(|E[v ∗ x]|2 + σv2 σx2 ) − βA βB sγ|E[v ∗ x]|2

ıs (q, v) = B − A + log 1 + sρ|b
ν= = . (61)
βA βB βA βB
It then follows that A and B are dependent exponentially-
distributed random variables with rate parameter 1/βA defined To obtain (14), we use that γ = s/(1 + sρ|b
g |2 ) and that
∗ 2 ∗
in (12) and 1/βB defined in (13), respectively. This implies E[v x] = σv − g gbρ.
that the random variable ıs (q, v) involves the difference be-
tween two dependent exponentially-distributed random vari- A PPENDIX C - P ROOF OF T HEOREM 3
ables. Let ∆ = B −A. The probability density function (PDF)
of ∆ is [36, Cor. 8] Substituting uMR = 1 h
1
b 1 into (31), we obtain
M

1 1 bH 1 bH 1 b H ul
f∆ (δ) = p y1ul [k] = h h1 xul
1 [k]+ h h2 xul
2 [k]+ h z [k]. (62)
(βB − βA )2
+ 4βA βB (1 − ν) M 1 M 1 M 1
Under Assumption 1 and using [6, Lem. 3], we have that, in
p !
|δ| (βB − βA )2 + 4βA βB (1 − ν)
× exp − the limit M → ∞,10
2βA βB (1 − ν)
1 b H (a) 1 b H b 1 1 b H ul (b)
δ(βB − βA ) h h1  h h1  tr(Φ1 ) and h z [k]  0.
 
× exp (55) M 1 M 1 M M 1
2βA βB (1 − ν) (63)
where ν = Cov(A, B) / Var(A) Var(B) is the correlation
p
b 1 and the pair (h
Here, (a) and (b) follow because h e 1 , zul [k])
coefficient between A and B. Using (55), we can express the are independent. Similarly, we have that
MGF of −ıs (q, v) as follows:
1 bH
1
Z ∞ h h2  0, if φ1 6= φ2 with φH1 φ2 = 0, (64)
M 1
h i
−ζıs (q,v)
E e = exp(−ζδ) f∆ (δ)dδ
g |2 )ζ −∞
(1 + sρ|b 1 bH 1 b H b (c) 1
h1 h2  h h2  tr(Υ12 ), if φ1 = φ2 (65)
(1 + sρ|bg |2 )−ζ M M 1 M
= (56)
1 + (βB − βA )ζ − βA βB (1 − ν)ζ 2 where (c) follows from the fact that h b 1 and h b 2 are correlated
where the last step holds for all ζ ∈ (ζ, ζ), with ζ and ζ given under pilot contamination. Using (63), (64), and (65) in (62),
in (10) and (11), respectively. The desired result in (9) follows we conclude that
because the right-hand side of (56) is infinitely differentiable. 1
y1ul [k]  tr(Φ1 )xul 1 [k], if φ1 6= φ2 with φ1 φ2 = 0,
H

To conclude the proof, we need to show that ν in (55) is M


given by (14). By definition, Cov(A, B) = E[AB]−E[A] E[B] (66)
where tr(Φ 1 )x ul
1 [k] + tr(Υ 12 )x ul
2 [k]
y1ul [k]  , if φ1 = φ2 . (67)
E[AB] = sγ E |v − gbq|2 |v|2 . (57)
 
M
To compute this correlation, it turns out convenient to set x = This implies that, in the absence of pilot contamination, the
v − gbq and to express x as the MMSE estimate of v given x input-output relation becomes that of a deterministic noiseless
plus the uncorrelated estimation error e: channel as M → ∞, while it converges to that of an AWGN
channel with transmit power limM →∞ [ M 1
tr(Φ1 )]2 and noise
x = αv + e. (58) variance limM →∞ [ M tr(Υ12 )] when the two UEs use the
1 2

same pilot sequence. Note also that g  gb  M 1


tr(Φ1 ) where
Here, α is the MMSE coefficient, given by α = E[v ∗ x] /σv2 , g and gb are defined in (33). The desired result then follows
and  CN (0, σe2) where2 σe = σx −|E[v x]| /σv , with σx =
 e 2∼
2 2 2 ∗ 2 2 2
from (6).
E |x| = |g − gb| ρ + σ . Note that since e is Gaussian and
uncorrelated with v, then e and v are independent. Using (58),
we can rewrite the expectation on the right-hand side of (57) A PPENDIX D - P ROOF OF T HEOREM 4
as follows: We only consider the case in which the two UEs use the
same pilot sequence, i.e., φ1 = φ2 . By applying the matrix
E |v − gbq|2 |v|2 = E |x|2 |v|2
   
inversion lemma (see, e.g., [6, Lem. 4]) we can rewrite (34)
= E |αv + e|2 |v|2
 
as
= |α|2 E |v|4 + E |v|2 E |e|2
     
2
!−1
X
= 2|α|2 σv4 + σv2 σe2 (59) uMMSE
1 = b b H
hi h + Z i h
b1 (68)
∗ 2 2 2
= |E[v x]| + σv σx . (60) i=1
−1
1 b bH
Here, in (59) we used that E |v| = 2σv4 . Furthermore, (60)
 4 = ul
h2 h2 + Z h
b1 (69)
1 + γ1
follows by the definition of α and of σe2 . Note now that βA =
10 Under Assumption 1, R Q−1 R has uniformly bounded spectral
k i
9 We drop the indices in q and v because immaterial for the proof. norm—a result that follows from [6, Lem. 4].
14

 −1
where we have set γ1ul = h bH h
1
b2h bH + Z
2
b 1 . Substitut-
h Under Assumption 1 and using [6, Lem. 3], we have that, as
ing (69) into (31) we obtain, after multiplying and dividing M → ∞,
each term by M , 1 b H −1 ul 1 b H −1 ul
h Z z  0 and h Z z 0 (78)
1 1 bH b bH −1 M 1 M 2
y1ul = h h 2 h + Z h1 xul b 2 ) and zul are independent.
1
+
ul
γ1 M 1 2 1
where we have used that (h b1, h
M M
1 1 bH b bH −1 Therefore, we have that
+ γ1ul M 1
h h h
2 2 + Z h2 xul2 1 1 bH b bH −1
1
M + M ul h1 h2 h2 + Z zul  0. (79)
+ 1 M
1 γ
1 1 bH b bH −1 M M
+ h h 2 h + Z zul . (70)
1
+
ul
γ1 M 1 2
Combining all the above results, we conclude that y1ul  xul
1.
M M
This implies that, as M → ∞, the input-output relation (70)
We begin by considering the first term. Under Assumption 1 converges to that of a deterministic noiseless channel. The
and using [6, Lem. 3], we obtain desired result then follows from (6).
1 bH b bH −1 γ ul
h1 h2 h2 + Z h1  1 (71) R EFERENCES
M M
[1] J. Östman, A. Lancho, and G. Durisi, “Short-packet transmission over
since h1 = h b1 + he 1 with h
e 1 being independent from h
b 1 and
a bidirectional massive MIMO link,” in Proc. Asilomar Conf. Signals,
ul
γ
b 2 . We note that, under Assumptions 1 and 2, 1 converges
h Syst., Comput., Pacific Grove, CA, USA, Dec. 2019.
M
to a finite value as M → ∞ [6, App. B]. This ensures that [2] 3GPP, “NR; NR and NG-RAN Overall description; Stage-2,” 3rd Gener-
ation Partnership Project (3GPP), Technical Specification (TS), 9 2019,
γ1ul version 15.7.0.
M
xul ul
(72) [3] ——, “Service requirements for cyber-physical control applications in
γ1ul 1  x1 . vertical domains,” 3rd Generation Partnership Project (3GPP), Technical
1
M + M Specification (TS) 22.104, 12 2019, version 17.2.0.
By applying [6, Lem. 5] to the second term in (70), we obtain [4] T. L. Marzetta, “Noncooperative cellular wireless with unlimited num-
bers of base station antennas,” IEEE Trans. Wireless Commun., vol. 9,
1 1 bH b bH −1 no. 11, pp. 3590–3600, Nov. 2010.
h
γ1ul M 1
h h
2 2 + Z h2 [5] H. Q. Ngo, E. G. Larsson, and T. L. Marzetta, “Energy and spectral effi-
1
M + M ciency of very large multiuser MIMO systems,” IEEE Trans. Commun.,
vol. 61, no. 4, pp. 1436–1449, Apr. 2013.
1 1 b H −1 [6] E. Björnson, J. Hoydis, and L. Sanguinetti, “Massive MIMO has
= h Z h2 unlimited capacity,” IEEE Trans. Wireless Commun., vol. 17, no. 1, pp.
1
+
γ1ul M 1
M M 574–590, Jan. 2018.
[7] E. Björnson, L. Sanguinetti, J. Hoydis, and M. Debbah, “Optimal design
!
1 b H −1 b 1 b H −1
M h1 Z h2 M h2 Z h2 of energy-efficient multi-user MIMO systems: Is Massive MIMO the
− . (73)
1 1 b H −1 b answer?” IEEE Trans. Wireless Commun., vol. 14, no. 6, pp. 3059–3075,
M + M h2 Z h2
Jun. 2015.
Under Assumption 1 and using [6, Lem. 3], we have that11 , [8] H. V. Cheng, E. Björnson, and E. G. Larsson, “Optimal pilot and
payload power control in single-cell massive MIMO systems,” IEEE
as M → ∞, Trans. Signal Process., vol. 65, no. 9, pp. 2363–2378, May 2017.
1 b H −1 1 b H −1 b 1 [9] E. Björnson, L. Sanguinetti, H. Wymeersch, J. Hoydis, and T. L.
h1 Z h2  h1 Z h2  tr(Υ12 Z−1 ) , β12 (74) Marzetta, “Massive MIMO is a reality—what is next?: Five promis-
M M M ing research directions for antenna arrays,” Digital Signal Processing,
1 b H −1 b 1 vol. 94, pp. 3–20, Nov. 2019.
h2 Z h2  tr(Φ2 Z−1 ) , β22 . (75) [10] G. Durisi, T. Koch, and P. Popovski, “Towards massive, ultra-reliable,
M M
and low-latency wireless communication with short packets,” Proc.
Substituting (74) and (75) in (73) and using Assumption 2, we IEEE, vol. 104, no. 9, pp. 1711–1726, Sep. 2016.
conclude that [11] E. Björnson, J. Hoydis, and L. Sanguinetti, “Massive MIMO networks:
1 −1   Spectral, energy, and hardware efficiency,” Foundations and Trends® in
M

H b bH M β12 β22 Signal Processing, vol. 11, no. 3-4, pp. 154–655, Nov. 2017.
h
b h h
2 2 + Z h2  β12 − [12] T. L. Marzetta, E. G. Larsson, H. Yang, and H. Q. Ngo, Fundamentals
1 γ1ul 1 γ1ul β22
M + M of massive MIMO. London, U.K.: Cambridge Univ. Press, 2016.
=0 (76) [13] M. Karlsson, E. Björnsson, and E. G. Larsson, “Performance of in-band
transmission of system information in massive MIMO systems,” IEEE
γ ul Trans. Wireless Commun., vol. 17, no. 3, pp. 1700–1712, Mar. 2018.
since M1 converges to a finite value as M → ∞ [6, App. B].
[14] A. Bana, G. Xu, E. D. Carvalho, and P. Popovski, “Ultra reliable low
For the third term in (70), we have latency communications in massive multi-antenna systems,” in Proc.
1 1 bH b bH −1 Asilomar Conf. Signals, Syst., Comput., Pacific Grove, CA, USA, Oct.
ul
γ1 M
h1 h2 h2 + Z zul 2018, pp. 188–192.
1
+ [15] L. H. Ozarow, S. Shamai (Shitz), and A. D. Wyner, “Information
M M theoretic considerations for cellular mobile radio,” IEEE Trans. Veh.
1 1 b H −1 ul Technol., vol. 43, no. 2, pp. 359–378, May 1994.
= h Z z [16] W. Yang, G. Durisi, T. Koch, and Y. Polyanskiy, “Quasi-static multiple-
1
+
γ1ul M 1 antenna fading channels at finite blocklength,” IEEE Trans. Inf. Theory,
M M
1 b H −1 b H 1 b H −1 ul
! vol. 60, no. 7, pp. 4232–4265, Jul. 2014.
M h1 Z h2 M h2 Z z [17] A. Lapidoth and S. Shamai (Shitz), “Fading channels: How perfect need
− 1 1 b H −1 b
. (77) ‘perfect side information’ be?” IEEE Trans. Inf. Theory, vol. 48, no. 5,
M + M h2 Z h2 pp. 1118–1134, May 2002.
[18] Y. Polyanskiy, H. V. Poor, and S. Verdú, “Channel coding rate in the
11 Under Assumption 1, Q−1 R Z−1 R has uniformly bounded spectral finite blocklength regime,” IEEE Trans. Inf. Theory, vol. 56, no. 5, pp.
i k
norm, which can be proved using in [6, Lem. 4]. 2307–2359, May 2010.
15

[19] G. Durisi, T. Koch, J. Östman, Y. Polyanskiy, and W. Yang, “Short-


packet communications over multiple-antenna Rayleigh-fading chan-
nels,” IEEE Trans. Commun., vol. 64, no. 2, pp. 618–629, Feb. 2016.
[20] J. Östman, G. Durisi, E. G. Ström, M. C. Coskun, and G. Liva,
“Short packets over block-memoryless fading channels: Pilot-assisted
or noncoherent transmission?” IEEE Trans. Commun., vol. 67, no. 2,
pp. 1521–1536, Feb. 2019.
[21] J. Zeng, T. Lv, R. P. Liu, X. Su, Y. J. Guo, and N. C. Beaulieu, “Enabling
ultra-reliable and low-latency communications under shadow fading by
massive MU-MIMO,” IEEE Internet of Things J., vol. 7, no. 1, pp.
234–246, Jan. 2020.
[22] H. Ren, C. Pan, Y. Deng, M. Elkashlan, and A. Nallanathan, “Joint pilot
and payload power allocation for massive-MIMO-enabled URLLC IIoT
networks,” IEEE J. Sel. Areas Commun., vol. 38, no. 5, pp. 816–830,
May 2020.
[23] A. Martinez and A. Guillén i Fàbregas, “Saddlepoint approximation of
random–coding bounds,” in Proc. Inf. Theory Applicat. Workshop (ITA),
San Diego, CA, USA, Feb. 2011.
[24] W. Feller, An Introduction to Probability Theory and Its Applications,
2nd ed., New York, NY, USA, 1971, vol. II.
[25] A. Lancho, J. Östman, G. Durisi, T. Koch, and G. Vazquez-Vilar,
“Saddlepoint approximations for short-packet wireless communications,”
IEEE Trans. Wireless Commun., vol. 19, no. 7, pp. 4831–4846, Jul. 2020.
[26] L. Sanguinetti, E. Björnsson, and J. Hoydis, “Towards massive MIMO
2.0: Understanding spatial correlation, interference suppression, and
pilot contamination,” IEEE Trans. Commun., vol. 68, no. 1, pp. 232–257,
Jan. 2020.
[27] M. Haenggi, “The meta distribution of the SIR in Poisson bipolar and
cellular networks,” IEEE Trans. Wireless Commun., vol. 15, no. 4, pp.
2577–2589, Apr. 2016.
[28] R. G. Gallager, Information Theory and Reliable Communication. New
York, NY, U.S.A.: John Wiley & Sons, 1968.
[29] E. MolavianJazi, “A unified approach to Gaussian channels with finite
blocklength,” Ph.D. dissertation, University of Notre Dame, Notre Dame,
IN, 2014.
[30] E. Biglieri, J. G. Proakis, and S. Shamai (Shitz), “Fading channels:
Information-theoretic and communications aspects,” IEEE Trans. Inf.
Theory, vol. 44, no. 6, pp. 2619–2692, Oct. 1998.
[31] J. Scarlett, A. Martinez, and A. Guillén i Fàbregas, “Mismatched
decoding: Error exponents, second-order rates and saddlepoint approxi-
mations,” IEEE Trans. Inf. Theory, vol. 60, no. 5, pp. 2647–2666, May
2014.
[32] G. Kaplan and S. Shamai, “Information rates and error exponents of
compound channels with application to antipodal signaling in a fading
environment,” Arch. Elek. Über, vol. 47, no. 4, pp. 228–239, 1993.
[33] E. Björnson, L. Sanguinetti, and M. Debbah, “Massive MIMO with
imperfect channel covariance information,” in 2016 50th Asilomar
Conference on Signals, Systems and Computers, 2016, pp. 974–978.
[34] J. Hoydis, S. ten Brink, and M. Debbah, “Massive MIMO in the UL/DL
of cellular networks: How many antennas do we need?” IEEE J. Sel.
Areas Commun., vol. 31, no. 2, pp. 574–590, Feb. 2013.
[35] A. Lancho, J. Östman, G. Durisi, and L. Sanguinetti, “A finite-
blocklength analysis for URLLC with massive MIMO,” in Proc. IEEE
Int. Conf. Commun. (ICC), Montreal, Canada, Jun. 2021.
[36] H. Holm and M.-S. Alouini, “Sum and difference of two squared cor-
related Nakagami variates in connection with the McKay distribution,”
IEEE Trans. Commun., vol. 52, no. 8, pp. 1367–1376, Aug. 2004.

You might also like