Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

4592 IEEE TRANSACTIONS ON COMMUNICATIONS, VOL. 64, NO.

11, NOVEMBER 2016

Analyzing Uplink SINR and Rate in Massive


MIMO Systems Using Stochastic Geometry
Tianyang Bai, Student Member, IEEE, and Robert W. Heath, Jr., Fellow, IEEE

Abstract— This paper proposes a stochastic geometry frame- (in the limit of the number of base station antennas) is limited
work to analyze the signal-to-noise-and-interference ratio (SINR) by pilot contamination [2]. In this paper, we propose to study
and rate performance in a large-scale uplink massive multiple- the uplink signal-to-noise-and-interference ratio (SINR) and
input and multiple-output (MIMO) network. Based on the model,
expressions are derived for spatial average SINR distributions rate performance in a massive MIMO network with maximum
over user and base station distributions with maximum ratio ratio combining (MRC) and zero-forcing (ZF) receivers, for a
combining (MRC) and zero-forcing (ZF) receivers. We show random base station topology.
that, using massive MIMO, the uplink SINR in certain urban
marcocell scenarios is limited by interference. In the interference- A. Related Work
limited regime, the results reveal that for MRC receivers,
a superlinear (polynomial) scaling law between the number of The performance with MRC and ZF beamforming, in
base station antennas and scheduled users per cell preserves terms of SINR, spectrum efficiency, and energy effi-
the uplink signal-to-interference ratio (SIR) distribution, while ciency, was examined though in a simple network topology
a linear scaling applies to ZF receivers. ZF receivers are shown in [2] and [6]–[10], where the SINR and rate expressions
to outperform MRC receivers in the SIR coverage, and the were conditioned on specific user locations, or equivalently
performance gap is quantified in terms of the difference in
the number of antennas to achieve the same SIR distribution. the received power for each user. The conclusions drawn from
Numerical results verify the analysis. It is found that the optimal the conditional expressions, however, need not apply to the
compensation fraction in fractional power control to optimize system-level performance taking average over different users’
rate is generally different for MRC and ZF receivers. Besides, distributions. For example, the linear scaling between the num-
simulations show that the scaling results derived from the ber of users and antennas examined in [8] need not maintain
proposed framework apply to the networks, where base stations
are distributed according to a hexagonal lattice. the uplink signal-to-interference ratio (SIR) distribution, as
Index Terms— Massive MIMO, SINR, achievable rate, perfor-
will be shown in our analysis. This motivates the investiga-
mance evaluation, stochastic geometry. tion of the spatial average performance over different base
station and user distributions in large-scale massive MIMO
I. I NTRODUCTION networks [11]–[14]. While the spatial average performance
can be simulated using prior results conditioning on fixed
M ASSIVE multiple-input and multiple-output (MIMO) is
an approach to increase the area spectrum efficiency in
5G cellular systems [2]–[5]. By deploying large-scale antenna
base station and user realizations, an analytical framework
approximating the key features of the system is desirable, as
arrays, base stations can use multi-user MIMO to serve a large it helps reveal key dependencies of parameters in the system.
number of users and provide high cell throughput [2]–[5]. Stochastic geometry provides a powerful tool to analyze
In this paper, we focus on the defacto massive MIMO systems system-level performance in a large-scale network with ran-
operated below 6 GHz, where pilot-aided channel estimation domly distributed base stations and users. Assuming a single
is performed in the uplink, and pilots are reused across cells antenna at each base station, the spatial average downlink
to reduce the training overhead [2]–[5]. Prior work showed SINR and rate distributions were derived for a network with
that when the number of base station antennas grows large, Poisson point process (PPP) distributed base stations, which
a high throughput is achieved through simple signal process- were shown a reasonable fit with simulations using real base
ing, and that the asymptotic performance of massive MIMO station data [15]. The stochastic geometry framework in [15]
was further extended to analyze the performance of MIMO
Manuscript received October 8, 2015; revised March 8, 2016 and networks: the downlink SINR and rate of multi-user MIMO
May 31, 2016; accepted July 2, 2016. Date of publication July 13, 2016;
date of current version November 15, 2016. This work was supported cellular system were analyzed, e.g. in [16]–[19] assuming
by the National Science Foundation through the Division of Computing perfect channel state information (CSI), and in [20] with
and Communication Foundations under Grant Nos. 1218338, 1319556, and quantized CSI from limited feedback. For uplink analysis,
1514257. A conference version of this paper was presented at the 2015 IEEE
Global Conference on Communications [1]. The associate editor coordinating prior work [21]–[23] showed that the uplink and downlink
the review of this paper and approving it for publication was M. Di Renzo. SINR follows different distributions, due to the difference
T. Bai was with The University of Texas at Austin, Austin, TX 78701 USA. in network topology. In [23], a stochastic geometry uplink
He is now with Qualcomm Flarion Technologies, Inc., Bridgewater, NJ 08807
USA (e-mail: tybai@utexas.edu). model was proposed to incorporate the pairwise correlations
R. W. Heath is with The University of Texas at Austin, Austin, TX 78701 in the user locations, where the SINR distributions derived
USA (e-mail: rheath@utexas.edu). based on the analytical model were shown a good fit with the
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org. simulations. The prior results in [15]–[23], however, do not
Digital Object Identifier 10.1109/TCOMM.2016.2591007 directly apply to analyze uplink massive MIMO networks, as
0090-6778 © 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
BAI AND HEATH: ANALYZING UPLINK SINR AND RATE IN MASSIVE MIMO SYSTEMS 4593

(i) they did not take account for the effects of pilot contam- 3) Comparison of MRC and ZF Receivers: We use the
ination, which becomes a limiting factor with large numbers scaling law results to quantify the performance gap between
of antennas [2]; (ii) the analysis in [15]–[20] was intended for ZF and MRC receivers, in terms of the difference in the
downlink performance, which follows different distributions number of antennas to provide the same SIR distribution. The
from the uplink network; and (iii) the results in [16]–[19] results show that ZF receivers provides better SIR coverage
were intended for MIMO networks with a few antennas, where than MRC receivers; the performance gap increases with the
the computational complexity for the analytical expressions number of scheduled users in a cell, and is reduced with a
grows with the number of antennas, and hinders their direct larger compensation fraction in the fractional power control,
applications to the massive MIMO scenarios. as it mitigates the near-far effect from intra-cell interference.
Stochastic geometry was first applied to study the asymp- Numerical results indicate different compensation fractions
totic SINR and rate (in the limit of base station antennas) in fractional power control to maximize the per-user-rate for
in a massive MIMO networks in [13] and [14], where the MRC and ZF receivers: the optimal fraction is around 0.5 for
asymptotic SINR was shown to be approached with impracti- MRC receivers, and 0.2 for ZF receivers.
cally large number of antennas, e.g. 104 antennas. To analyze
the performance with finite antennas, related work in [24] C. Organization and Notation
applied stochastic geometry to compute the uplink interference
This paper is organized as follows. We present the system
in a massive MIMO network. A linear scaling between the
model for the network topology and propagation channel
numbers of base station antennas and scheduled users was
in Section II. We analyze the performance of MRC receivers
found to maintain the mean interference, which need not
in Section III, and that of ZF receivers in Section IV-A,
preserve the SIR distribution. Our prior work in [1] derived the
followed by a performance comparison between two receivers
uplink SIR distribution for MRC receivers as a function of the
in Section IV-B. We present numerical results to verify the
number of antennas, assuming no power control. In this paper,
analysis in Section V, and conclude the paper in Section VI.
we extend the results by incorporating thermal noise into the
We use the following notation throughout this paper. Bold
analysis, assuming a general fractional power control for MRC
lower-case letters x are used to denote vectors, and bold upper-
receivers, and investigating the performance of ZF receivers.
case letters X are used to denote matrices. We use X[:, k] to
B. Contributions denote the k-th row of matrix X, X∗ as the Hermitian transpose
of X, and X† as the pseudo-inverse of X. We use E to denote
1) Stochastic Geometry Framework for Uplink Massive
expectation, and P to denote probability.
MIMO Networks: We propose a stochastic geometry frame-
work to derive the uplink SINR and rate distributions in a
large-scale cellular network using multi-user MIMO. To model II. S YSTEM M ODEL
the uplink topology, we propose an exclusion ball model In this section, we introduce the system model for an uplink
based on prior work [23], which simplifies the computation. massive MIMO cellular network. We focus on the networks
Channel estimation error due to pilot contamination is also operated in the sub-6 GHz band; the proposed model can be
considered in the system model. The proposed framework extended for massive MIMO at millimeter wave (mmWave)
also incorporates the fractional power control by compensating frequencies by incorporating key differences in propagation
for a fraction of the path loss as in the long term evolution and hardware constraints [26]. Each base station is assumed
(LTE) systems [25]. Based on the framework, we derive to have M antennas. In each time-frequency resource block, a
analytical expressions for the uplink SINR distribution for base station can simultaneously schedule K users in its cell.
both MRC and ZF receivers in the massive MIMO regime. Let X  be the location of the -th base station, Y(k) be the
Unlike prior work analyzing asymptotic performance with location of the k-th scheduled user in the cell of -th base
(k) (k)
infinity antennas [13], [14], the SINR coverage is examined station, and h the channel vector from X  to Y .
as a function of the number of base station antennas and We consider a cellular network with perfect synchroniza-
scheduled users per cell. tion, and assume the following pilot-aided channel estima-
2) Scaling Laws to Maintain SIR Distributions: We apply tion in the uplink. In the uplink channel training stage,
the SINR results to investigate the interference-limited case, the scheduled users Y(k) send their assigned pilots tk , and
as numerical results show that the impact of noise becomes base stations X  estimate the channels by correlating the
minor in the urban macro-cell scenario. We derive scaling laws corresponding pilots and using an minimum mean square error
between the number of base station antennas and scheduled (MMSE) estimator; in the uplink data transmission, the base
users per cell to maintain the uplink SIR distributions. Unlike stations will apply either MRC or ZF receivers, based on the
the linear scaling law examined in prior work [8], [24], we find channel estimates. Further, we assume the pilots {tk }1≤k≤K
that a super-linear scaling is generally required for MRC are orthogonal and fully reused in the network. Note that the
receivers to maintain the uplink SIR distributions, due to the system model assumption applies to general uplink multi-user
near-far effect from intra-cell interference. For ZF receivers, MIMO networks with pilot-aided channel estimation in the
we show that a linear scaling law still holds, as the intra- uplink, including but not limited to the time-division duplex
cell interference is negligible. Simulations verify our analysis, (TDD) massive MIMO [2].
and indicate that the scaling laws derived from the stochastic Now, we introduce the channel model assumptions. The
geometry framework also apply to the hexagonal model. channel is assumed to be constant during one resource block,

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
4594 IEEE TRANSACTIONS ON COMMUNICATIONS, VOL. 64, NO. 11, NOVEMBER 2016

(k)
and fades independently from block to block. Moreover, we 1) The distances R from a user to their associated
apply a narrowband channel model, as frequency selectivity base stations are assumed to√be IID Rayleigh random
in fading can be minimized by techniques like orthogo- variables with a mean of 0.5 1/λb [21].
nal frequency-division multiplexing (OFDM) and frequency 2) The other-cell scheduled user process Nu(k) is modeled
domain equalization [27]. In this paper, we focus on the case of by a homogeneous PPP of density λb outside an exclu-
identically and independently distributed (IID) Rayleigh fading sion ball centered at the tagged base station X 0 with a
channels, and defer the incorporation of spatial correlations radius Re .
to future work. Therefore, we express the channel vector 3) The scheduled users processes using different pilots
(k) (k) (k  )
hn ∈ C M×1 as Nu and Nu are assumed to be independent
  for k = k  .
(k) (k) 1/2 (k)
hn = βn wn , (1) Note that in the exclusion ball model, we equivalently use a
step function λb (1 − I(r < Re )) to approximate the density
(k)
where βn is the large-scale path loss, wn (k)
∈ C M×1 is a function in (3), where I(·) is the√indicator function.
Gaussian vector with the distribution CN (0, I M ) for Rayleigh In this paper, we let Re = 1/ (πλb ) by matching the
fading. average number of the excluded points from a homogenous
The large-scale path loss gain βn(k)
is computed as PPP of density λb in the  ∞step function and in (3), i.e., by
letting λb π Re2 = 2πλb 0 e−λb πr r dr = 1. An alternative
2
 
(k) (k) −α explanation for our choice of Re is to let the size of the
βn = C Rn , (2)
exclusion ball π Re2 equal the average cell size 1/λb [29].
where C is a constant determined by the carrier frequency and In Section V, we will show in simulations that the proposed
reference distance, α > 2 is the path loss exponent. exclusion ball model, when applied together with subsequent
Next, we introduce the network topology assumptions based approximations in (23) and (26), causes minor errors in the
on stochastic geometry. We assume the base stations are SINR distributions. The approximate process generated by
distributed as a PPP with a density λb . A user is assumed to the exclusion ball approximation alone, however, need not be
be associated with the base station that provides the minimum representative of the exact scheduled user process, in terms of
path loss signal. The users are assumed to be distributed as other properties, e.g. the null probability.
an independent PPP with a sufficiently high density. In a Fractional power control, as used in the LTE systems [25],
resource block, each base station randomly schedules K users is assumed in both the uplink training and uplink data stages:
that are independently and uniformly distributed in its Voronoi the user Y(k) transmits with power
cell [23], [28]. Without loss of generality, a typical scheduled  
(1) (k) (k) −
user Y0 is fixed at the origin, and its serving base station X 0 P = Pt β , (4)
is denoted as the tagged base station in this paper. We will
(k)
investigate the SINR and rate performance at this typical user. where β is the path loss in the corresponding signal link,
Now we focus on modeling the distribution of scheduled  ∈ [0, 1] is the fraction of the path loss compensation, and
user process in a resource block. For 1 ≤ k ≤ K , the k-th Pt is the open loop transmit power with no power control.
scheduled user Y(k) in each cell is assigned with the same pilot We omit the constraint on the maximum uplink transmit
(k)
tk . Let Nu be the point process formed by the locations of the power for simplicity, which will increase the average transmit
k-th scheduled users Y(k) from all the cells. Note that the power, and reduce the impact of noise. As shown in Fig. 2(b),
however, the maximum transmit power constraint can have a
scheduled user process Nu(k) is non-stationary (also non-PPP),
secondary impact on the SINR distributions in certain massive
as their locations are correlated with the base station process,
(k) MIMO networks, as due to the large antenna array gains,
and one scheduled user Y prohibits the presence of all
(k) mobile stations can transmit with a lower open loop power Pt
other users in Nu in cell X  [21]–[23]. Unfortunately, the to meet the receiver sensitivity requirement. The incorporation
correlations in the scheduled users’ locations make the exact of more complicated power control algorithms is deferred to
analysis intractable. In [23], the authors proposed an uplink future work. The noise power is denoted as σ 2 .
model to account for the pairwise correlations, where the In the uplink training stage, after correlating the received
other-cell scheduled users for base station X 0 in Nu(k) is training signal with the corresponding pilot, base station X 0
modelled as an inhomogeneous PPP with a density function of (1)
has an observation of the channel h00 as
 
λu (r ) = λb 1 − e−λb πr ,
2
(3)   
(1) (1) (1) (1) (1)
u00 = P0 h00 + P h0 + nt ,
where r is the distance to base station X 0 . To further simplify >0
the analysis, e.g., the computation in (23) and (26), we propose
where nt is the noise
 vector
 in the training stage following the
an exclusion ball approximation, as a first-order approximation
distribution CN 0, σK I M .
2
of the model in [23], on the distribution of the scheduled user
process Nu(k) as follows. We assume for  > 0, the large-scale path losses β0 (1)

Assumption 1: The following assumptions are made to are perfectly known to base station X 0 . Since the channels
(k) (1)
approximate the exact scheduled users’ process Nu . are assumed to be IID Rayleigh fading, the channel h00 is

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
BAI AND HEATH: ANALYZING UPLINK SINR AND RATE IN MASSIVE MIMO SYSTEMS 4595

estimated by an MMSE estimator as by the interference in typical urban macro-cell scenarios.


 We derive a scaling law between the number of users and
(1)
P0(1) β00
(1)
(1) antennas that maintains the uplink SIR distribution for the
h̄00 =  (1) (1) u00 , (5) typical user. Finally, we present a method to compute the per-
σ2
 P β0 + K user achievable rate and cell throughput, based on the SINR
(1) (1) distribution.
where h̄00 is the estimation of h00 . Due to the orthogonality
(1)
principle, the channel vector h00 can be decomposed as
A. SINR Coverage Analysis
(1) (1) (1)
h00 = h̄00 + ĥ00 , (6) Now we investigate the uplink SINR coverage based on
(1) the system model. With MRC receivers, we assume that base
where (k)
 ĥ00 is the estimation error
following the distribution station X 0 applies the combining vector g00 as a scaled version
(1) P0(1) β00
(1)
(1) (1)
CN 0, β00 1 −  (1) (1) σ 2 I . of the channel estimate h̄00 to decode the signal from Y00 :
P β0 +
(k)
 K
(k)  (1) (1) σ 2
Let s be the uplink data symbol for user Y with (1) P β + K (1) (1)

(k)

(k)
g00 =   0 h̄00 = u00 . (9)
E |s |2 = P . In uplink data transmission, base station (1) (1)
P0 β00
(k)
X  is assumed to use the combiner vector g to decode s(k) Note the scaling on the combining vector is intended to sim-
(k) (k) plify expressions, and will not change the SINR distribution.
from Y , based on the channel estimate h̄ .
Then, at base
station X 0 , the decoded symbol ŝ0(1) for the typical user X 0(1) is Then, using the combining vector in (9), the SINR expression
can be simplified in (10), as shown  at the bottom of this
(1) (1)∗ (1) (1)
= g00 h̄00 s0 (k)  (k) − (k)
β0 + Kσ Pt , and (k)
ŝ0 2
 page, where 1 = >0 β 2 =
(1)∗ (1) (1)
+ g00 ĥ00 s0 +
(1)∗ (k) (k) (1)∗
g00 h0 s + g00 nu , (7)   −2  2
(k) (k)
(,k) =(0,1) >0 β β0 . The derivation to obtain (10) is
  (k) (k)
unknown at base station given in Appendix A. Note that 1 and 2 correspond
to the sum of certain interference terms from other-cell users.
where nu ∈ C M×1 is the thermal noise vector in the uplink Next, we denote the exact SINR distribution for (10) (using
data transmission. Treating the unknown terms at base station the exact scheduled user distribution defined in Section II
X 0 as uncorrelated additive noise, the uplink SINR for the without the exclusion ball approximation) as P(SINR > T ).
typical user Y0(1) is given in (8), as shown at the bottom of Due to pilot contamination, the combining vector g00 (k)
is
this page, where the expectation operators are taken over the correlated with certain interference channel vectors as shown
(1) (k)
channel estimation error ĥ00 and small-scale fading w0 in the in (9). As a result, the denominator in (10) contains cross-
interference links. Note that the SINR given in (8) is a random products of the path losses from different interferers. More-
variable, due to the randomness in the large scale path losses. over, different cross-product terms in the denominator of (10)
We will investigate the SINR distributions for MRC and ZF can be correlated, as they may contains common path loss
receivers in the following sections. terms, which renders the exact derivation of P(SINR > T )
The proposed system model represents a simple multi-user intractable. Therefore, we compute an approximate SINR
MIMO systems in which the SINR expression can be analyzed distribution P̄(SINR > T ), which we argue in Section V is a
using stochastic geometry. In the following sections, we will good match for P(SINR > T ), in Theorem 1.
study the uplink SINR and rate distributions for MRC and ZF Theorem 1 (MRC SINR): In the proposed massive MIMO
receivers, when the number of base station antennas is large. networks, an approximate uplink SINR distribution with MRC
receivers can be computed as
III. P ERFORMANCE A NALYSIS FOR MRC R ECEIVERS N 
N
In this section, we derive an approximate uplink SINR P̄(SINR > T ) = (−1)n+1
n
distribution for MRC receivers, where the approximation n=1
 ∞ α
becomes tight in the massive MIMO regime, e.g. with M ≥ 64 α(1−) −T ηC t 2 (1−)
e−t −T ηC1 t 2 C3 (t)dt,
antennas. Then, we focus on the interference limited case, as 0
numerical results show that the uplink SINR is dominated (11)

P0(1) |g00
(1)∗ (1) 2
h̄00 |
SINR = (1) (1)∗ (1)
 (k) (1)∗ (k) (1)
, (8)
P0 Eĥ(1) |g00 ĥ00 |2 + P Ew(k) |g00 h0 |2 + |g00 |2 σ 2
00 0
(,k) =(0,1)
 
(1) 2(1−)
(M + 1) β00
SINR =     (10)
(1) (1) 1− (1)  K  (k) 1−  K (k)

(1) 1− (1)
M 2 + β00 1 + k=2 β 00 +
k=1 1 β 00 + 1

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
4596 IEEE TRANSACTIONS ON COMMUNICATIONS, VOL. 64, NO. 11, NOVEMBER 2016

where N is the number of terms used in the cal- Corollary 1: With  = 1 and σ 2 = 0, the approximate SIR
σ2
culation, η = N(N!)− N , Cσ 2 =
1
α(1−) , distribution can be computed as
 α   K Pt C 1− (λb π) 2
N   
C1 = M+1 K +1 2 ( 2 +1) M α (+1)  C5 K + α (1.5)
(α−2) + Cσ 2 , C2 = (M+1)(α−1) + N n+1 −T η
1
+ α−1
 α  2 P̄(SIR > T ) = (−1) e M+1
,
2 ( 2 +1) n
(α−2) + Cσ 2 ,
K n=1
M+1
(12)
 ∞ α
K −1 4 2α (1.5)+(α 2 −4) α (1.5)
−u−u − 2 (1−) T ηC 4 (t ) where C5 = .
C3 (t) = e du (α−2)2
0
  K −1 Based on Corollary 1, a linear scaling law between the
∞ e−u
≈ 1 − ηT C4 (t) du , number of users and antennas is observed as follows.
α
0 T ηC4 (t) + u − 2 (1−) Corollary 2: With  = 1 and σ 2 = 0, to maintain the
 α    uplink SIR distribution unchanged, the scaling law between
2 ( 2 +1) α
C4 (t) = M+1
1
α−2 + Cσ 2 t α(1−) + t 2 (1−) , and the number of base station antennas M and users per cell K
 ∞ −t α−1 is approximately
(α) = 0 e t dt is the gamma function. 
α (1.5)
Proof: See Appendix B. (M + 1) ∼ K + ≈ K. (13)
C5
Besides the exclusion ball approximation, the main approx-
Note that when  = 1, the linear scaling law matches prior
imation in Theorem 1 is to replace certain out-of-cell interfer-
results in [8, Sec. IV], where the path loss to all associated
ence terms by their means in (23) and (26). The approximation
users in the typical cell was assumed to be identical. The linear
results in a minor error in the SINR distribution, as (i) with
scaling law, however, does not apply to other cases with  < 1,
K users in a cell, the intra-cell interference dominates the
e.g. in the following case without power control.
out-of-cell interference with high probability; (ii) with large
Case 2 (No Power Control,  = 0): When assuming no
antenna arrays, the ratio of the signal power to certain out-of-
(k) power control and MRC receivers, the uplink SIR can be
cell interference power terms, e.g. the terms in 1 , decays evaluated as follows.
as M . In Section V, using N ≥ 5 terms, the distribution
1
Theorem 2: With  = 0 and σ 2 = 0, an approximate uplink
P̄(SINR > T ) from Theorem 1 is shown to be a good SIR distribution can be calculated as
match with the SINR distribution P(SINR > T ) from Monte N 
Carlo simulations. In addition, the error of the approximation  N
P̄(SIR > T ) = (−1)n+1
becomes more prominent with a smaller noise power, as all n
n=1
the approximations are made with respect to the interference  ∞ nηT α
e−(μ (1−2/α)(nηT ) +1)t − α−1 t dt, (14)
2/α
distribution. The expression is intended for the massive MIMO
regime when M 1, as the error of the approximations 0
1 where N is the number of terms used in the computation, and
decays with M . In simulations, we find that the results
generally applies to the multi-user MIMO networks with not- μ = (M+1)
K
2/α .
so-large M, e.g. the case of (M, K ) = (10, 2). Proof: The proof is similar to that in [1, Appendix A].
In Theorem 1, the noise power is taken account by the We will show in Section V that Theorem 2 provides a tight
σ2 approximation of the exact SIR distribution P(SIR > T ), when
parameter Cσ 2 = α(1−) , which shows that the
K Pt C 1− (λb π) 2
N ≥ 5 terms are used. Moreover, note that in (14), the number
impact of noise on the SINR is reduced with a larger number of antennas M and the number of scheduled users per cell K
of scheduled users per cell K , a higher base station density λb , only affect the value of μ. Therefore, by Theorem 2, in the
a smaller path loss α, and a larger power control parameter . no power control case, we observe the following scaling law
Besides, the impact of noise goes down with larger M, as in to maintain SIR.
the expressions for C1 and C2 , the noise parameter Cσ 2 is Corollary 1: Assuming no power control, the approximate
divided by (M + 1). scaling law to maintain the same uplink SIR distribution is
We will show in Section V that the impact of noise is negli- (M + 1) ∼ K α/2 , which is a superlinear polynomial scaling
gible in certain urban macro-cell cases with M = 64 antennas when α > 2.
at base stations. Therefore, we focus on the performance of In the case of no power control, the difference in the path
interference-limited networks, where the general expression in losses between the typical user and the intra-cell interferers
Theorem 1 can be further simplified in the following special affect the SIR distribution, and thus the scaling law to maintain
cases. the SIR becomes a function of the path loss exponent. The
Case 1 (Full Power Control,  = 1): In this case, super-linearity in the scaling law can be explained by the
the transmitting power at scheduled user Y(k) is adjusted near-far effect of the intra-cell interference from multiple
(k) (k)
to compensate for the full path loss, i.e., P = Pt β , users in a cell. With no power control, the cell edge users
such that a base station receives equal signal powers will receive weaker signals than the cell center user. With
from all of its associated users. When  = 1, the a uniform user distribution in a cell, the typical user will be
SIR distribution can be simplified as in the following more likely to be located at the cell edge. When increasing the
corollary. number of scheduled users K in a cell, the probability that the

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
BAI AND HEATH: ANALYZING UPLINK SINR AND RATE IN MASSIVE MIMO SYSTEMS 4597

interference from a cell-center interferer dominates the desired We will examine the average cell throughput as a function of
signal increases. Therefore, compared with the linear scaling M and K in Section V. Before that, we continue to present
law with full power control ( = 1) where such near-far effect the results for ZF receivers in the next section.
is mitigated, more antennas will be needed in the no power
control case to reduce the intra-cell interference, and preserve
IV. P ERFORMANCE A NALYSIS W ITH ZF R ECEIVERS
the SIR distribution, when increasing K .
Next, we focus on the scaling law in the general fractional In this section, we first derive the SINR distributions
power control case with  ∈ (0, 1). It is difficult to derive with ZF receivers. Then, we apply the analytical results to
the exact scaling law directly from the expression (11), due to compare the SIR performance of MRC and ZF receivers
the integral form. Since with the fractional power control, the in an interference-limited network. In particular, we aim to
equivalent path loss exponent in the signal link linearly scales answer the question: compared with MRC receivers, how
with , we propose the following conjectured scaling law by many antennas can be saved by applying ZF receivers, while
linearly fitting the exponent s of the scaling law (M +1) ∼ K s , keeping the same uplink SIR distribution.
based on two special cases of : by Corollary 2, when  = 1,
s = 1; and by Theorem 2, when  = 0, s = α2 . Therefore,
for general 0 <  < 1, the linearly fitted exponent s in the A. SINR Analysis of ZF Receivers
conjectured scaling law is given as follows. Now we begin to investigate the performance of
Scaling law 1: With fractional power control and MRC ZF receivers in an uplink massive MIMO network. For
receivers, a conjectured scaling law between M and K is ZF receivers, we still focus on the case of IID fading. To cancel
(M + 1) ∼ K s , where the exponent of the scaling law is the intra-cell interference, base station X  will apply the
(k)
s = α2 (1 − ) + . combining vector g for user Y(k) as
Scaling law 1 indicates that a (superlinear) polynomial
(k)
scaling law between K and M is required to maintain uplink g = H† [:, k], (18)
SIR distribution, for a general  < 1. The conjecture in Scaling

(1) (2) (K )
law 1 are verified by numerical simulations in Section V. where H = u , u , . . . , u ∈ C M×K is the matrix of
all estimated channels to the associated users in cell X  , and
B. Rate Analysis  (1) (1) 2
(k)  P β +σ (k)
u =    K h̄ is a scaled version of the channel
In this section, we apply the SINR results to compute (k) (k)
P β
the achievable rate. First, we define the average achievable
estimate. The scaling in the channel estimates will not change
spectrum efficiency at a typical user as
  the uplink SINR distribution, as it will only cause certain
τ0 = E log2 (1 + min{SINR, Tmax }) , (15) scaling in the corresponding combining vector. Similar to the
where Tmax is a SINR distortion threshold determined by case of MRC receivers, the exact uplink SINR distribution is
limiting factors like distortion in the radio frequency front-end. difficult to derive, as due to pilot contamination, the combining
By [30, Sec. III-C], given the SINR distribution P(SINR > T ), vector is correlated with certain interference channel vectors.
the average achievable spectrum efficiency can be computed Therefore, applying the same approximations in (23) and (26),
 Tmax P(SINR>x) we derive an approximate distribution for the uplink SINR
as τ0 = ln(2)1
0 1+x dx. To take account for the expression in (8) for the typical user X 0(1) in the following
overhead, let ψ be the fraction time for overhead. In this theorem.
paper, for simplicity, we only consider the overhead due to Theorem 3: With M K and ZF receivers, an approx-
uplink channel training, and compute the overhead fraction ψ imate uplink SINR distribution for the typical user can be
as ψ = TTct = TKc , where Tt and Tc are the length of channel calculated by
training period and coherent time, in terms of symbol time.
N 
The length of channel training is assumed to be equal to the N
number of scheduled users in a cell, as we assumed full reuse P̄(SINR > T ) = (−1)n+1
n
of orthogonal pilots throughout the network. Then the average n=1
 ∞  α 
achievable rate with the overhead penalty τ̄0 equals −nηT C 6 t 2 (1−) +C 7 t α(1−) −t
e dt, (19)
0
τ̄0 = (1 − ψ)τ0 , (16)
where the constant
Note that when ignoring thermal noise, the scaling law to 
maintain SIR distribution also maintains the average achiev- 1 1 M(K − 1)
C6 = C9 + +
able rate τ0 . When taking account for the training over- M −K +1 M + 1 (M − K + 1)2
head penalty ψ, however, the scaling law will not keep τ̄0 M(K − 1)C8
unchanged, as 1 − ψ linearly decreases with K , unless ψ is + ,
(M − K + 1)2
negligible, e.g. when the coherence time Tc K . Next, we 
M α ( + 1) 1 (K − 1)M
define the average cell throughput τcell , in terms of spectrum C7 = + + C92
M +1 α−1 M + 1 (M − K + 1)2
efficiency, as
(K − 1)M
τcell = K (1 − ψ)τ0 . + C8 C9 ,
(17) (M − K + 1)2

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
4598 IEEE TRANSACTIONS ON COMMUNICATIONS, VOL. 64, NO. 11, NOVEMBER 2016

2 α ( 2 +1)+(α−2)C σ 2 2 α ( 2 +1)
C8 = (α−2)(1+C σ 2 )+2 α ( 2 +1)
, C9 = α−2 + Cσ 2 , same uplink SIR distribution as MRC receivers with MMRC
σ2 antennas?
Cσ 2 = α(1−) , N is the number of terms used
K Pt C 1− (λb π) 2 Based on Scaling law 1 and Scaling law 2, we have the
1
in the computation, and η = N(N!)− N . following proposition to determine MZF to match the SIR
Proof: See Appendix C. coverage with MRC receivers.
Note that when K = 1, the SINR distribution in (19) for Proposition 1: Assuming MZF K , ZF receivers with
ZF receivers is the same as that for MRC receivers in (12). (MZF +1) = ξ(MMRC +1) antennas approximately provide the
We will verify the tightness of the approximation P(SINR > same uplink SIR distribution as MRC receivers with MMRC
T ) ≈ P̄(SINR > T ) by numerical simulation in Section V. antennas in a massive MIMO networks, where the scaling
α
We have the following remark on the applicable regime for factor ξ = K −( 2 −1)(1−), and K is the number of scheduled
Theorem 3. users in a cell.
Remark 1: We need the condition M K in the proof, Proof: For the ease of notation, let ZF(M, K ) and
as the error in the approximation in (28) decays as M−K 1 MRC(M, K ) denote the uplink SIR distributions with ZF and
+1 .
In numerical simulations, we find that the approximate SINR MRC receivers of M antennas, when serving K users in a
distribution in Theorem 3 shows a good match with the cell. By Scaling law 2, when MZF K , ZF(MZF , K ) ≈
ZF( MZFK+1 , 1). Next, note that when K = 1, i.e., with a single
simulations when M K ≥ 3 and M ≥ 10. The same comment
applies to Scaling law 2 as below. scheduled user in a cell, MRC and ZF receivers provide the
Next, we focus on the interference-limited case. Based on same SIR coverage. Thus, it follows that ZF(MZF , K ) =
Theorem 3, we can derive an approximate scaling law between ZF( MZFK+1 , 1) = MRC( MZFK+1 , 1). Last, by Scaling law 1,
M and K to maintain the SIR distribution in the region of ZF(MZF , K ) ≈ MRC( MZFK+1 , 1) ≈ MRC((MZF +
α
M K as follows. 1)K ( 2 −1)(1−) − 1, K ).
Scaling law 2: With ZF receivers and σ 2 = 0, the uplink The condition MZF K in the proposition is required
SIR distribution of the typical user remains approximately to ensure the applicability of Scaling law 2. In numerical
unchanged when the number of antennas M linearly scales simulations, the result is found to be a good approximation
with the number of users per cell K as (M + 1) ∼ K . with MKZF > 3. Note that the exponent of the scaling factor
Proof: Note that when σ 2 = 0, Cσ = 0. The dependence −( α2 − 1)(1 − ) is non-positive, which indicates we need
on M and K in (19) only occurs in the constants C6 and C7 . MMRC ≥ MZF to provide the same SIR coverage. Further, the
Therefore, it is sufficient to show that a linear scaling between scaling factor ξ increases with the number of the scheduled
M and K (approximately) maintains the values of C6 and C7 . user K , which reveals that the performance gap between
Note that when M → ∞, and M K , the following limits MRC and ZF receivers grows with K . When K increases, the
hold: M+11
→ 0, M−K 1
+1 → 0, and M+1 → 1. Therefore, it
M
mitigation of the intra-cell interference from (K − 1) users
follows that when keeping M+1 K
=t, lim M→∞ C6 = (C81−t +C 9 )t
,
by ZF receivers becomes more prominent to improve SIR
α (+1) 4 2α ( 2 +1) 2C 8 α ( 2 +1) coverage. In addition, Proposition 1 also shows that in terms of
and lim M→∞ C7 = α−1 + 1−t
t
(α−2)2
+ α−2 , the SIR distribution, the performance gap reduces with larger
which are invariant when (M + 1) linearly scales with K .  in the power control scheme, as the scaling factor ξ is a
Compared with MRC receivers, the near-far effect for users decreasing function of . Simulations show that with  = 1,
in a cell becomes minor with ZF receivers, as the intra-cell only a minor gap exists between the SIR coverage curves for
interference is largely suppressed. Therefore, a linear scaling ZF and MRC receivers.
law applies for ZF receivers even without power control. Based Last, we note that Proposition 1, which is drawn based on
on the SINR coverage results, the achievable rate per user and the SIR distribution, need not extend to a general SINR dis-
sum throughput can be computed following the same line as tribution that is not dominated by interference; prior work [6]
in Section III-B. In the next section, we will use the derived showed that when the noise is not negligible, MRC receivers
results to compare the SIR coverage performance between would provide a comparable or even better SINR coverage,
MRC and ZF receivers in an interference-limited network. compared with the ZF receivers. In the following section, we
will present numerical results to validate our analytical results.
B. Comparison of SIR Coverage Performance
V. N UMERICAL R ESULTS
Now assuming the network is interference-limited, we
compare the SIR coverage between ZF and MRC receivers. In this section, we verify our analytical results with numeri-
Prior work [8] showed that ZF and MRC receivers have the cal simulations, which follow the procedure as: (1) generating
same asymptotic performance, both which are limited by the the base station process as a PPP of density λb ; (2) generating
pilot contamination. The analysis in [24] showed that by the overall user process as a PPP of density λu,o , where we
suppressing intra-cell interference, which turns to be more use λu,o = 60λb , unless otherwise specified; (3) associating
dominant than the out-of-cell interference, the ZF receivers the points in the overall user process to base stations, based
suffers from less interference than MRC receivers. In this on the minimal path loss rule, and then randomly scheduling
section, we make a quantitative comparison by answering the K out of the associated users in each cell as their scheduled
following question: in IID fading channels, how many base users; (4) picking the base station closest to the origin as the
(1)
station antennas MZF is needed for ZF receivers to provide the tagged base station X 0 , and its first scheduled user Y0 as the

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
BAI AND HEATH: ANALYZING UPLINK SINR AND RATE IN MASSIVE MIMO SYSTEMS 4599

Fig. 1. Comparison of SINR and SIR distributions. In the figures, we use


markers to represent SINR curves, solid lines for SIR. We assume K = 10 Fig. 2. SINR coverage for MRC receivers. In the simulations, we assume
users per cell,  = 0, and α = 4 in all cases. The gap between the SIR α = 4. In (a), the analytical curves are drawn based on Theorem 1, which
and SINR distributions becomes minor when ISD=500 meters, which is the are shown a good fit with simulation. The difference in the curves for
typical size for the urban macro cells [31]. (M, K , ) = (64, 10, 0) and (M, K , ) = (128, 20, 0) indicates that linear
scaling between M and K does not generally preserve SIR for MRC receivers.
In (b), we compare the SINR distributions with and without the maximum
typical user; (5) generating channel vectors as IID Gaussian transmit power constraint. We assume (M, K ) = (128, 10), the average
vectors, and computing the SINR for the iteration; (6) repeat- ISD= 500 m, the maximum uplink transmit power is 23 dBm, the system
ing the step (1)-(5) for 10,000 iterations, and computing the bandwidth is 20 MHz. The open loop transmit power Pt is chosen, such that
(1) the user located at the edge of an average-size cell has 0 dB uplink SNR.
empirical distribution of the SINR at Y0 . For the simulations
using hexagonal grids, we follow the same procedure except network with ISD=500 meters is shown to be interference-
that the base station process is generated as a 19-cell hexagonal limited with M = 64 antennas for both MRC and ZF receivers,
grid, and the tagged base station is the center cell. In addition, as the SIR curves almost coincide with the SINR curves,
we will use N = 5 terms when evaluating the analytical which justifies the interference-limited assumption in urban
expressions. marco cells. In the sparse network with ISD=1000 meters,
however, simulations show that even with M=64 antennas,
A. Impact of the Noise
notable gaps exist between the SINR and SIR distributions,
To begin with, we examine the impact of noise by com- especially for ZF receivers. In addition, the results in Fig. 1(b)
paring the SINR and SIR distributions for both MRC and ZF shows that when the noise power is high, even with no
receivers in different scenarios in Fig. 1. In the simulations, power control, ZF and MRC receivers have a similar SINR
we assume Pt =23 dBm, and the bandwidth is 20 MHz as coverage performance, which indicates that the SIR com-
in the current LTE standards [31]. We examine the case of parison results in Proposition 1 need not extends to general
 = 0, which maximizes the impact of the noise. We sim- SINR comparisons.
ulate with two inter-site distances (ISDs): an average ISD
of 500 meters in Fig. 1(a), and 1000 meters in Fig. 1(b). B. SINR Coverage for MRC Receivers
Note that a typical ISD of 500 meters is assumed for urban In Fig. 2(a), we verify the analytical results for the SIR
macro-cells in the 3GPP standards [31]. In Fig. 1(a), the distribution with MRC receivers. Numerical results show that

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
4600 IEEE TRANSACTIONS ON COMMUNICATIONS, VOL. 64, NO. 11, NOVEMBER 2016

Fig. 4. Comparison of SIR coverage with MRC and ZF receivers. We assume


α = 4. As the double arrays display, when fixing  = 0, the performance gap
in SIR coverage is shown to increase with K ; when fixing K = 20, the gap
Fig. 3. SIR distributions with ZF receivers. We assume α = 4, and IID fading diminishes when  → 1.
channel. The analytical curves are plotted based on Theorem 3. Simulations
verify the analytical results, and show that when both M and K double, the
SIR curves remain almost unchanged.

the SIR coverage is sensitive to the compensation fraction 


in the fractional power control: a large compensation fraction
 improves the SIR coverage in the low SIR regime at the
expense of sacrificing the coverage in the high SIR regime.
Besides, a comparison of the curves for (M, K ) = (64, 10)
and (M, K ) = (128, 20) shows that the linear scaling
law does not maintain the SIR distribution when  = 0.
In Fig. 2(b), we consider the case when the uplink transmit
power after the fractional power control is clipped by a
maximum power constraint. In the simulations, the maximum
power constraint is assumed to be 23 dBm [31]; the open
loop transmit power Pt is determined, such that the uplink
SNR at a cell-edge user is 0 dB in a cell of the average Fig. 5. Verification of Proposition 1. In the simulation, α = 4. In the
size. The results indicate that ignoring the maximum transmit simulation, we use the SIR curve of MRC(64, 5) as a baseline for comparison.
power constraint has a minor impact on the SINR distribution We use Proposition 1 to compute the required number of antennas for ZF
receivers that provides the same SIR distribution as the baseline curve.
in massive MIMO networks; as with large antenna arrays,
mobile stations can apply a lower transmit power than the the performance gap decreases with ; when  = 1, the
single-antenna case, to meet the target SNR. SIR coverage gap becomes minimal between MRC and ZF
receivers. With full compensation of path loss in power
C. SIR Coverage for ZF Receivers control, linear scaling laws between M and K apply to both
We verify the analysis for ZF receivers in Fig. 3. The analyt- MRC and ZF receivers, as the near-far effect for users in
ical curves generally match well with numerical simulations. a cell is mitigated. When M K , the difference in the
A comparison of the curves for (M, K , ) = (64, 10, 0) and average (residue) intra-cell interference between MRC and ZF
1
(M, K , ) = (128, 20, 0) shows that unlike the case of MRC receivers becomes minor, as it decays with M .
receivers, a linear scaling law between M and K maintains In Fig. 5, we verify our theoretical results in Proposition 1.
the SIR distribution, even when there is no fractional power In the simulation, we fix the number of antennas for the MRC
control implemented. receivers to be MMRC = 64, and use Proposition 1 to calculate
the required MZF , to maintain the same SIR distribution.
Numerical results show a good match with our analysis; the
D. SIR Comparison Between MRC and ZF Receivers
minor mismatch in the case  = 0 is because Proposition 1
We compare the uplink SIR distributions for MRC and theoretically requires MKZF 1, while we use MKZF = 13 5 in
ZF receivers in Fig. 4. Simulations show that ZF receivers the simulation.
provide better SIR coverage, due to the suppression of intra-
cell interference. Moreover, for the same , the performance
gap between MRC and ZF receivers increases with the number E. Verification With Hexagonal Grid Model
of scheduled users K , as the strength of total intra-cell We verify the scaling laws derived from stochastic geometry
interference also increases with K . When fixing M and K , with the hexagonal grid model in Fig. 6. In the simulations,

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
BAI AND HEATH: ANALYZING UPLINK SINR AND RATE IN MASSIVE MIMO SYSTEMS 4601

Fig. 7. Comparison of different scaling laws. We plot the required number


of antennas to provide the same SIR as that of the case (M, K ) = (16, 5) as
a function of K with different system parameters.

Fig. 6. Verification of the scaling laws in the hexagonal model. We use Fig. 8. Average spectrum efficiency per user in an interference-limited
(M, K ) = (32, 5) as the baseline curves. When increasing the number of network. In the simulation, we assume Tmax = 21 dB, which sets the
users to K =10, we compute the required M to preserve the SIR distribution maximum spectrum efficiency per data stream as 7 bps/Hz. Training overhead
as baseline curves, according to Scaling law 1 and Scaling law 2. Simulations is not taken account in this figure.
indicates that the scaling law results apply to the hexagonal model.

that in the case of (M, K ) = (16, 5), as a function of K .


we use a layout of 19 hexagonal cells with inter-site distance As shown in the plot, for MRC receivers, given the path loss
of 300 meters; only the scheduled users in the central cell are exponent α, the slope of the scaling law is determined by
counted for the SIR statistics, to avoid edge effect. In Fig. 6(a), the fraction of path loss compensation : the linear scaling
for MRC receivers, we use a (M, K ) = (32, 5) as the baseline law proposed in prior work [8], [24] is only achieved when
curve for comparison. When doubling the number of scheduled  = 1. Although the choice of  = 1 makes the system with
users to K = 10, we use Scaling law 1 to compute the required MRC receivers linearly scalable, it need not maximize the per-
M to maintain the same SIR distribution. Scaling law 1 is user rate, as will be shown in Fig. 8. On the contrary, for ZF
shown to be almost accurate with extensive combinations of receivers, the linear scaling applies for all  ∈ [0, 1].
the system parameters in the hexagonal grid model. Similarly,
results in Fig. 6(b) verifies the linear scaling law for ZF G. Rate Performance
receivers in Scaling law 2. This indicates that the stochastic
geometry model provides reasonable predictions even for the We illustrate the results on the average spectrum efficiency
hexagonal model. per user in Fig. 8. In the simulation, the average ISD is 500
meters, and K = 10, which is shown to be interference-
limited in Fig. 1. Consistent with the SIR results, in a
F. Comparison of Scaling Laws interference-limited network, ZF receivers provide a higher
We compare scaling laws to maintain the uplink SIR dis- spectrum efficiency per user. Numerical results also show that
tribution in different scenarios in Fig. 7. We plot the required the average spectrum efficiency is sensitive to the fraction of
number of antennas to maintain the same SIR distribution as the path loss compensation ; the optimum  for per user

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
4602 IEEE TRANSACTIONS ON COMMUNICATIONS, VOL. 64, NO. 11, NOVEMBER 2016

station density, to simulate the case with large K . In Fig. 9(a),


in the high mobility case, when Tc < M, the optimal K ∗
for cell throughput is limited by the duration of Tc , and the
optimal value generally is K ∗ ≤ T2c . In the example of low
mobility case, when Tc > M, the results in Fig. 9(b) show
that the optimal K ∗ depends much on M: for ZF receivers,
K
the cell throughput drops fast when M approaches to 1, and the
∗ M
optimal K is around 2 for maximum throughput; for MRC
receivers, the cell throughput becomes saturated approximately
when K > M3 . In addition, ZF receivers generally achieve
better cell throughput than MRC receivers; the only exception
is the case of M K ≈ 1, where the cell throughput of ZF
receivers drops below that of MRC. In addition, compared
with the single user per cell case (K = 1), the results confirm
that massive MIMO improves the cell throughput by serving
multiple users simultaneously.
VI. C ONCLUSIONS
In this paper, we proposed a stochastic geometry framework
to analyze the spatial average SINR coverage and rate in mas-
sive MIMO networks. We applied the analysis and numerical
results to draw the following important system design insights.
• The uplink massive MIMO networks can be interference-
limited in urban marco cells (ISD=500 meters) with
M = 64 antennas at base stations.
• With MRC receivers, the number of antennas M should
scale super-linearly with the number of scheduled users
α
per cell K as (M + 1) ∼ K 2 (1−)+ , to maintain the
uplink SIR distribution; a linear scaling law only applies
to the case of full path loss compensation in the power
control, i.e., when  = 1.
• With ZF receivers, a linear scaling between the number
of antennas M and users per cell K maintains the uplink
SIR distribution in massive MIMO.
Fig. 9. Uplink cell throughput as a function of K . The overhead due • When noise is negligible, ZF receivers provide better SIR
to channel training is taken account in the simulations. We simulate an
interference-limited network with ISD=500 meters. We use  = 0.5 for MRC coverage rate than MRC receivers. The performance gap
receivers, and  = 0.2 for ZF receivers, which are shown to optimize the per increases with K , and decreases with path loss compen-
user rate. sation faction . The gap becomes minor when  = 1.
• The SIR coverage and rate are sensitive to the fraction 
TABLE I of path loss compensation in power control. Larger 
C OHERENCE T IME Tc IN THE E XAMPLES improves coverage in the low SIR regime while reducing
coverage probability at high SIR. Numerical results show
that the optimal  for rate is around 0.5 for MRC,
and 0.2 for ZF receivers in certain cases.
A PPENDIX
rate is generally around 0.5 for MRC receiver, and 0.2 for A. Derivation of (10):
ZF receivers. In addition, we also observe that there is a With the combining vector in (9), the SINR expression
minor performance gap in rate between ZF and MRC receivers equals (20), as shown at the top of the next page
under full channel compensation power control, as predicted In the numerator, the signal power can be computed as
by Proposition 1.  
(1) (1) 2
Last, we examine the cell throughput in a system operated P β
(1) (1)∗ (1) (a) 0 00 (1)
P0 |u00 h̄00 |2 =  |u |4
at 2 GHz in Fig. 9. As an example, we consider an OFDM  (1) (1) σ 2 2 00
system, where the symbol time is 66.7 μs. We consider  P β0 + K
two cases with different mobilities as listed in Table I; the  
(1) (1) 2
coherence time Tc is computed as Tc = 4 1fD [32], where (b) P0 β00
(1) 4
≈  E|u00 |
f D is the maximum doppler frequency. In this simulation, we  (1) (1) σ 2 2
assume the density of overall users to be 100 times the base P
  β 0 + K

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
BAI AND HEATH: ANALYZING UPLINK SINR AND RATE IN MASSIVE MIMO SYSTEMS 4603

 
 (1)∗ (1) 2
P0(1) u00 h̄00 
SINR =     2 . (20)
(1)  (1)∗ (1) 2  (k)  ∗ (k)  (1) (1)∗ (k)  (1) 2 2
P0 E u00 ĥ00  + (,k)=(0,1) P E nt h0 +  ≥0 P h0 h0  + |u00 | σ

 
(c) (1) 2
= P0(1) β00 (M 2 + M) Therefore, we can express (22) as
   σ2  
(k) − (k)
(1) 2(1−) M Pt β β0
= Pt2 β00 (M 2 + M),
K
(,k) =(0,1)
where (a) follows from the MMSE estimator in (5), 


(1) M→∞ (1) (k) (1) (1)∗ (k)
(b) follows from the fact that |u00 |4 → E|u00 |4 , and the + P P E |h0 h0 |2

approximation error decays as M12 [33], and (c) follows from
 2 ⎛  >0 ⎞
(1) 4 (1) (1) σ2   −
the fact that E|u00 | = (M 2 +M)  P β0 + K . Next, = M Pt2 ⎝ (k)
β (k) ⎠
β0
we compute the first term in the denominator of (20) as (k,) =(1,0)
⎛ ⎞
P0(1) E|u00
(1)∗ (1) 2
ĥ00 | =
1 (1) (1) 2
P0 E|u00 (1) 2
| E|ĥ00 | σ2   (1) − (1)
M ×⎝ + β    β0 ⎠
  K Pt
(a) 1 (1)  (1) (1) σ 2 (1)
 ≥0
= P0 M P β0 + Mβ00   (1) −2  (1) 2
M

K + M 2 Pt2 β β0 .
  >0
P0(1) β00
(1)
× 1 −  (1) (1) Last, the thermal noise term in the denominator can be
σ2
 P β0 + K simplified as
   
    (1)− (1)
(1) 1− σ2   (1)− (1) σ2
= M Pt β00
2
β β0 + , (1) 2 2
|u00 | σ = Pt Mσ 2
β β0 + .
Pt K Pt K
>0 
where (a) follows from the fact that the channel Then the expression in (10) is obtained through algebraic
(1)
estimation
  error ĥ00 follows
the distribution manipulation in the denominator.
(1) (1)
(1) P0 β00
CN 0, β00 1 −  (1) (1) σ 2
IM .
 P β0 + K B. Proof of Theorem 1:
Next, we simplify the second term in the denominator. Note To allow for tractable computation and decouple the cor-
that unless (,  , k) = (n, n  , m), h(1)∗ (k) (1)∗ (m)
 0 h0 and hn  0 hn0 are related terms in the denominator of (10), we propose the
uncorrelated zero-mean random variables. Therefore, we can following approximations on the out-of-cell interference terms:
simplify the second term in the denominator of (20) as for k ∈ [1, K ],
   (1) (1)∗ (k)  
∗ (k) (k) (a)
  (k) − (k) σ2
E|nt h0 + P h0 h0 |2 (21) 1 ≈ E β β0 +
(,k) =(0,1)  ≥0 K Pt
     >0 
σ2 (k) − (k)   (k) − (k) σ2
= M Pt β β0 = E E β β0 +
K K Pt
(,k) =(0,1) >0
  

 (k)
+
(k) (1)
P P E
(1)∗ (k)
|h0 h0 |2 . (22)
(b) − − α α  σ2
≈ C (λb π) 2 ( + 1)E β0 +
 >0 2 K Pt
>0
Note that for k = 1, =  = 0, the expression is simplified (c) 2C
1− (λ π) α(1−)
2   σ2
α
b
as = +1 + , (23)


α−2 2 K Pt
(k) (1) (1)∗ (k) (1)2 (1) (k)
P P E |h0 h0 |2 = P E |h0 |4 where in (a), we approximate 1 by its mean; step (b)
  follows from
(1)2 (1) 2   
  
= (M 2 + M)P β0 (k) − (k) α α
    E β = C − E R = (λb π)− 2 α +1 ;
(1) −2 (1) 2 2
= Pt2 (M 2 + M) β β0 ; (24)
for k > 1 or k = 1,  =  > 0, it follows that and step (c) follows from the exclusion ball model in

Approximation 1 and the Campbell’s theorem [29] as
P(k) P(1)
 E |h(1)∗ (k) 2
h
0 0 | = M P(1) (k) (1) (k)
 P β0 β0
   ∞
 (k) 2C(λb π)α/2
  E β0 = 2πλb C x −α xdx = . (25)
(k) − (1) (k)
= M Pt2 β(1)
  β β0 β0 . Re α−2
>0

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
4604 IEEE TRANSACTIONS ON COMMUNICATIONS, VOL. 64, NO. 11, NOVEMBER 2016

Similarly, we can approximate (1)


2 by its mean as: where in (a) C1 , C2 and C4 (x) are defined in Theorem 1;
(k) C 2(1−) (λ π)α(1−) in (b) we use a “dummy” gamma variable g with unit
2 ≈ α−1
b
α ( + 1), (26) mean and shape parameter N to approximate the con-
  

(k) −2 (k)2 stant number one, and the approximation follows from the
where follows from E β = C −2 E R =
fact that g converges to one when N goes to infinity,
(λb π)−α α ( + 1). n n−1 −nx
i.e., lim n→∞ n x (n)e = δ(x − 1) [34], where δ(x) is
Next, applying the approximation in (23) and (26) and
(1) the Dirac delta function; in (c), the approximation follows
conditioning on R00 = x, we simplify the approximate
conditional uplink SINR as from Alzer’s inequality [30], [35, Appendix A], where
 (k)
η = N(N!)− N ; (d) follows from the fact that R00 is
1
α
SINR ≈ T C1 (λb π x 2 )α(1−) + C2 (λb π x 2 ) 2 (1−) assumed to be IID Rayleigh random variable; in (e) we change
−1
  α
(k)2 − 2 (1−)
variable as u = λb πs 2 ; in (f), we apply the approximation
+ C4 (λb π x )
2
λb π R00 , (27) exp(−x) ≈ 1+x 1
inside the integral, to allow for faster numer-
k>0 ical evaluations. Last, we obtain the uplink SIR distribution
(1)
where in (a) C1 , C2 and C4 (x) are defined in Theorem 1. Next, by de-conditioning on R00 = x, which√is assumed to be a
(1)
conditioning on R00 = x, the approximate SINR distribution Rayleigh random variable with mean 0.5 1/λb , and changing
can be computed as the variable as t = πλb x 2 .
 
(1)
P SIR > T |R00 =x

(a) T  α
C. Proof of Theorem 3
≈ P 1> C1 (λb π x 2 )α(1−) + C2 (λb π x 2 ) 2 (1−) Before proving the theorem, we present a useful lemma on
M +1  (1)
  α
(k)2 − 2 (1−)
the distribution of the combining vector g00 as follows.
+ C4 (λb π x )
2
λb π R00 Lemma 1 (From [36]): The square norm of the
(1) 2
 k>0 combining vector |g00 | follows the distribution of
(b) T  α   −1
≈ P g> C1 (λb π x 2 )α(1−) + C2 (λb π x 2 ) 2 (1−) χ2(K −M+1) (1) (1) σ2
 P β0 + K , where χ2(K
2 2
M +1  −M+1)
  α
(k)2 − 2 (1−) represents a Chi-square random variables with 2(K − M + 1)
+ C4 (λb π x )
2
λb π R00
degrees of freedom.
 k>0 χ2
(c) ηT  Note that when (M K + 1) → ∞, M−K 2(K −M+1)
+1 → 1. Therefore,
≈ 1 − E 1 − exp − C1 (λb π x 2 )α(1−) by Lemma 1, when M K , the following approximation
M +1
α
+ C2 (λb π x 2 ) 2 (1−) holds as
 N ⎤
  α
(k)2 − 2 (1−)
(1)
|g00 |2 ≈
1
, (28)
+ C4 (λb π x )
2
λb π R00 ⎦ (M − K + 1)S (1)
k>0 
N   
α(1−)  α (1−)  where for ease of notation, we define S (k) =  P(1) (1)
 β +
N −nT η C 1 λb π x 2 +C 2 λb π x 2 2  
= (−1)n+1 e σ2 (k) 1− (k)
n K = β00 + 1 , and the approximation error decays
n=1
  K −1 (1) (1) 2
+1 . Noting that |g00 u00 | = 1, the nominator of the
  α 1
 (k)2 − 2 (1−) as M−K
−nT ηC 4 λb π x 2 λb π R00
×E e SINR expression in (8) can be computed as

N   
α(1−)  α (1−) 
(1) (1)∗ (1) P0(1)2 β00
(1)2
(1) (1)
(d)
=
N
(−1)n+1 e
−nT η C 1 λb π x 2 +C 2 λb π x 2 2 P0 |g00 h̄00 |2 =  2
|g00 u00 |2
S (1)
n
n=1
 ∞ K −1    2
(1) 2(1−)

−nT ηC 4 λb π x 2 s −α(1−) −λb πs 2 = β00 / S (1) . (29)
× e λb πsds
0
N    
2 α(1−) +C λ π x 2
α (1−)  Applying the results in (28), the first term in the denomi-
(e) N n+1 −nT η C 1 λb π x 2
= (−1) e 2 b nator of (8) is computed as
n
P0(1) E|g00
(1)∗ (1) 2
n=1
 ∞  α
K −1 ĥ00 |
−nT ηC 4 λb π x 2 u − 2 (1−) −u  
e du × P β
(1) (1)
1
0 ≈ P0(1) β00
(1)
1 − 0 (1)00
N 
  
α(1−)  α (1−)  S (M − K + 1)S (1)
(f) N n+1 −nT η C 1 λb π x +C 2 λb π x 2
2 2
≈ (−1) e  1−
n (1) (1)

n=1 1 β00
  =  2
. (30)
× 1 − nT ηC4 λb π x 2 (M − K + 1) S (1)

  K −1 Now we simplify the second sum in the denominator of (8):


∞ e−u du
×  , for k = 1, and  > 0, as (31), shown at the top of the next
α
0 1 + nT ηC4 λb π x 2 u − 2 (1−) page.

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
BAI AND HEATH: ANALYZING UPLINK SINR AND RATE IN MASSIVE MIMO SYSTEMS 4605

(1)∗ (1) 2 (1) −4 (1)∗ (1) (1)∗ (1) 2 (1) −4 (1)∗ (1) 2
E|g00 h0 | = E|u00 | |g00 u00 u00 h0 | = E|u00 | |u00 h0 |
(1)2 (1)2  (1) (1) (1) (1)
(M + M)P β0 + M  = P P β0 β0 +
2 Mσ 2
K P(1) β0
(1)
=  2
(31)
(M 2 + M) S (1)

For k > 1, it follows that [6] H. Yang and T. L. Marzetta, “Performance of conjugate and zero-
  forcing beamforming in large-scale antenna systems,” IEEE J. Sel. Areas
(1) 2 (k) 2 (k) −2 (k)∗ (k) 2
E|g00 | E |h0 | − |u00 | |u00 h0 | Commun., vol. 31, no. 2, pp. 172–179, Feb. 2013.
(1)∗ (k) 2
E|g00 h0 | = [7] N. Krishnan, R. D. Yates, and N. B. Mandayam, “Uplink linear receivers
 M −K +1 
for multi-cell multiuser MIMO with pilot contamination: Large system
(k) (k) (k)∗ (k) analysis,” IEEE Trans. Wireless Commun., vol. 13, no. 8, pp. 4360–4373,
E |h0 |2 − |u00 |−2 |u00 h0 |2 Aug. 2014.
= [8] J. Hoydis, S. ten Brink, and M. Debbah, “Massive MIMO in the UL/DL
(M − K + 1)2 S (1)
 of cellular networks: How many antennas do we need?” IEEE J. Sel.
(k) (k) P(k) β0
(k) Areas Commun., vol. 31, no. 2, pp. 160–171, Feb. 2013.
M P β0 1 − S (k) [9] Y. Li, Y.-H. Nam, B. L. Ng, and J. Zhang, “A non-asymptotic throughput
≈ . for massive MIMO cellular uplink with pilot reuse,” in Proc. IEEE
(M − K + 1)2 S (1) Globecom Conf., Dec. 2012, pp. 4500–4504.
[10] H. Q. Ngo, M. Matthaiou, T. Q. Duong, and E. G. Larsson, “Uplink per-
Then, in the case of k > 1, for  = 0, we approximate the formance analysis of multicell MU-SIMO systems with ZF receivers,”
residue intra-cell interference of ZF receivers as IEEE Trans. Veh. Technol., vol. 62, no. 9, pp. 4471–4483, Nov. 2013.
[11] E. Björnson, E. G. Larsson, and M. Debbah. (2014). “Massive MIMO
α(1−) (k)
(1)∗ (k) MC 1− (λb π) 2 1 for maximal spectral efficiency: How many users and pilots should be
E|g00 h00 |2 ≈  α(1−)
 , allocated?” [Online]. Available: http://arxiv.org/abs/1412.7102
(k)
(M − K +1)2 C 1− (λb π) 2 + 1 S (1) [12] I. Atzeni, J. Arnau, and M. Debbah, “Fractional pilot reuse in mas-
sive MIMO systems,” in Proc. IEEE ICC Workshops, Jun. 2015,
(32) pp. 1030–1035.
[13] P. Madhusudhanan, X. Li, Y. Liu, and T. X. Brown, “Stochastic geo-

 α(1−)
(k)2 − 2 metric modeling and interference analysis for massive MIMO systems,”
which follows from P0(k) β00
(k)
≈ E R00 = in Proc. Modeling Optim. Mobile, Ad Hoc Wireless Netw. (WiOpt),
α(1−) May 2013, pp. 15–22.
(λb π) ; for  > 0, the out-of-cell interference is upper
2
[14] T. Bai and R. W. Heath, “Asymptotic coverage and rate in mas-
bounded (and approximated) as sive MIMO networks,” in Proc. IEEE Global Conf. Signal Inf.
Process. (GlobalSIP), Atlanta, GA, USA, 2014, pp. 602–606.
(k)
(1)∗ (k) M P1 β0 [15] J. G. Andrews, F. Baccelli, and R. K. Ganti, “A tractable approach to
E|g00 h0 |2  . (33) coverage and rate in cellular networks,” IEEE Trans. Commun., vol. 59,
(M − K + 1)2 S (1) no. 11, pp. 3122–3134, Nov. 2011.
[16] H. S. Dhillon, M. Kountouris, and J. G. Andrews, “Downlink MIMO
The noise term in the denominator is HetNets: Modeling, ordering results and performance analysis,” IEEE
(1) σ2 Trans. Wireless Commun., vol. 12, no. 10, pp. 5208–5222, Oct. 2013.
|g00 |2 σ 2 = . (34) [17] V. Chandrasekhar, M. Kountouris, and J. G. Andrews, “Coverage
(M − K + 1)S (1) in multi-antenna two-tier networks,” IEEE Trans. Wireless Commun.,
vol. 8, no. 10, pp. 5314–5327, Oct. 2009.
Next, substituting (23), (26) and (30)-(34) for (8), conditioning [18] Y. Wu, Y. Cui, and B. Clerckx, “Analysis and optimization of
(1)
on R00 = x, and after some algebraic manipulation, the SINR inter-tier interference coordination in downlink multi-antenna HetNets
expression is simplified as with offloading,” IEEE Trans. Wireless Commun., vol. 14, no. 12,
pp. 6550–6564, Dec. 2015.
 α
−1 [19] C. Li, J. Zhang, and K. B. Letaief, “Performance analysis of SDMA
SINR = C6 (λb π x 2 ) 2 (1−) + C7 (λb π x 2 )α(1−) . in multicell wireless networks,” in Proc. Globecom, Dec. 2013,
pp. 3867–3872.
The rest of the proof follows the same line as [20] M. Kountouris and J. G. Andrews, “Downlink SDMA with limited feed-
back in interference-limited wireless networks,” IEEE Trans. Wireless
in Appendix B. Commun., vol. 11, no. 8, pp. 2730–2741, Aug. 2012.
[21] T. D. Novlan, H. S. Dhillon, and J. G. Andrews, “Analytical modeling
of uplink cellular networks,” IEEE Trans. Wireless Commun., vol. 12,
R EFERENCES no. 6, pp. 2669–2679, Jun. 2013.
[1] T. Bai and R. W. Heath, Jr., “Uplink massive MIMO SIR analysis: [22] H. ElSawy and E. Hossain, “On stochastic geometry modeling of
How do antennas scale with users?” in Proc. IEEE Globecom Conf., cellular uplink transmission with truncated channel inversion power
Dec. 2015, pp. 1–6. control,” IEEE Trans. Wireless Commun., vol. 13, no. 8, pp. 4454–4469,
[2] T. L. Marzetta, “Noncooperative cellular wireless with unlimited num- Aug. 2014.
bers of base station antennas,” IEEE Trans. Wireless Commun., vol. 9, [23] S. Singh, X. Zhang, and J. G. Andrews, “Joint rate and SINR
no. 11, pp. 3590–3600, Nov. 2010. coverage analysis for decoupled uplink-downlink biased cell associa-
[3] E. G. Larsson, O. Edfors, F. Tufvesson, and T. L. Marzetta, “Massive tions in HetNets,” IEEE Trans. Wireless Commun., vol. 14, no. 10,
MIMO for next generation wireless systems,” IEEE Commun. Mag., pp. 5360–5373, Oct. 2015.
vol. 52, no. 2, pp. 186–195, Feb. 2014. [24] N. Liang, W. Zhang, and C. Shen, “An uplink interference analysis for
[4] L. Lu, G. Y. Li, A. L. Swindlehurst, A. Ashikhmin, and R. Zhang, massive MIMO systems with MRC and ZF receivers,” in Proc. WCNC,
“An overview of massive MIMO: Benefits and challenges,” IEEE J. Sel. 2015, pp. 310–315.
Topics Signal Process., vol. 8, no. 5, pp. 742–758, Oct. 2014. [25] W. Xiao, R. Ratasuk, A. Ghosh, R. Love, Y. Sun, and R. Nory, “Uplink
[5] F. Boccardi, R. W. Heath, A. Lozano, T. L. Marzetta, and P. Popovski, power control, interference coordination and resource allocation for
“Five disruptive technology directions for 5G,” IEEE Commun. Mag., 3GPP E-UTRA,” in Proc. IEEE 64th Veh. Technol. Conf., Sep. 2006,
vol. 52, no. 2, pp. 74–80, Feb. 2014. pp. 1–5.

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.
4606 IEEE TRANSACTIONS ON COMMUNICATIONS, VOL. 64, NO. 11, NOVEMBER 2016

[26] T. Bai and R. W. Heath, “Asymptotic SINR for millimeter wave massive Robert W. Heath, Jr. (S’96–M’01–SM’06–F’11)
MIMO cellular networks,” in Proc. IEEE 16th Int. Workshop Signal received the B.S. and M.S. degrees from the Univer-
Process. Adv. Wireless Commun. (SPAWC), Stockholm, Sweden, 2015, sity of Virginia, Charlottesville, VA, USA, in 1996
pp. 620–624. and 1997, respectively, and the Ph.D. degree from
[27] A. Goldsmith, Wireless Communications. Cambridge, U.K.: Cambridge Stanford University, Stanford, CA, USA, in 2002,
Univ. Press, 2005. all in electrical engineering. From 1998 to 2001,
[28] E. Björnson, L. Sanguinetti, and M. Kountouris. (2015). “Deploying he was a Senior Member with the Technical Staff
dense networks for maximal energy efficiency: Small cells meet massive and a Senior Consultant with Iospan Wireless Inc.,
MIMO.” [Online]. Available: http://arxiv.org/abs/1505.01181 San Jose, CA, USA, where he worked on the design
[29] F. Baccelli and B. Błaszczyszyn, Stochastic Geometry and Wireless and implementation of the physical and link layers of
Networks: Volume I Theory. Boston, MA, USA: NOW Publishers, 2009. the first commercial MIMO-OFDM communication
[30] T. Bai and R. W. Heath, Jr., “Coverage and rate analysis for millimeter- system. Since 2002, he has been with the Department of Electrical and
wave cellular networks,” IEEE Trans. Wireless Commun., vol. 14, no. 2, Computer Engineering, The University of Texas at Austin, where he is
pp. 1100–1114, Feb. 2015. currently a Cullen Trust for Higher Education Endowed Professor and a
[31] 3GPP, “Further advancements for E-UTRA physical layer aspects member of the Wireless Networking and Communications Group. He is also
(release 9),” 3rd Generat. Partnership Project, Cedex, France, the President and CEO of MIMO Wireless Inc., and the Chief Innovation
Tech. Rep. TR 36.814, Mar. 2010. Officer of Kuma Signals LLC. His research interests include several aspects of
[32] D. Tse and P. Viswanath, Fundamentals of Wireless Communication. wireless communication and signal processing: limited feedback techniques,
Cambridge, U.K.: Cambridge Univ. Press, 2005. multihop networking, multiuser and multicell MIMO, interference alignment,
[33] H. Q. Ngo, E. G. Larsson, and T. L. Marzetta, “Aspects of favorable adaptive video transmission, manifold signal processing, and millimeter wave
propagation in massive MIMO,” in Proc. 22nd Eur. Signal Process. communication techniques. He has co-authored the book entitled Millimeter
Conf. (EUSIPCO), Sep. 2014, pp. 76–80. Wave Wireless Communications (Prentice Hall, 2014).
[34] R. Aris, Mathematical Modeling: A Chemical Engineer’s Perspective. He has been an Editor for the IEEE T RANSACTIONS ON C OMMUNICATION,
San Diego, CA, USA: Academic, 1999. an Associate Editor for the IEEE T RANSACTIONS ON V EHICULAR T ECH -
[35] H. Alzer, “On some inequalities for the incomplete gamma function,” NOLOGY , and a Lead Guest Editor for the IEEE J OURNAL ON S ELECTED
Math. Comput., vol. 66, pp. 771–778, Apr. 1997. A REAS IN C OMMUNICATIONS of the Special Issue on Limited Feedback
[36] M. Kiessling and J. Speidel, “Analytical performance of MIMO zero- Communication, and for the IEEE J OURNAL ON S ELECTED T OPICS IN
forcing receivers in correlated Rayleigh fading environments,” in Proc. S IGNAL P ROCESSING of the Special Issue on Heterogenous Networks. He was
SPAWC, Jun. 2003, pp. 383–387. on the Steering Committee for the IEEE T RANSACTIONS ON W IRELESS
C OMMUNICATIONS from 2011 to 2014. He was a member of the Signal
Processing for Communications Technical Committee in the IEEE Signal
Processing Society and was the Chair of the IEEE COMSOC Communications
Technical Theory Committee. He was the Technical Co-Chair for the 2007
Fall Vehicular Technology Conference, the General Chair of the 2008 Com-
munication Theory Workshop, the General Co-Chair, Technical Co-Chair, and
Co-Organizer of the 2009 IEEE Signal Processing for Wireless Communica-
tions Workshop, the Local Co-Organizer for the 2009 IEEE CAMSAP Con-
ference, the Technical Co-Chair for the 2010 IEEE International Symposium
on Information Theory, the Technical Chair for the 2011 Asilomar Conference
on Signals, Systems, and Computers, the General Chair for the 2013 Asilomar
Conference on Signals, Systems, and Computers, the Founding General
Co-Chair for the 2013 IEEE GlobalSIP conference, and the Technical
Co-Chair for the 2014 IEEE GLOBECOM conference.
Dr. Heath was a co-author of best paper awards at the IEEE VTC 2006
Spring, WPMC 2006, the IEEE GLOBECOM 2006, the IEEE VTC 2007
Spring, the IEEE RWS 2009, and the IEEE GLOBECOM 2015, and a
co-recipient of the Grand Prize in the 2008 WinTech WinCool Demo
Tianyang Bai (S’11) received the B.Eng. degree
from the Harbin Institute of Technology, Harbin, Contest. He was a co-recipient of the 2010 and 2013 EURASIP Journal
China, in 2011, and the M.S.E. and Ph.D. degrees on Wireless Communications and Networking Best Paper Awards, the 2012
Signal Processing Magazine Best Paper Award, a 2013 Signal Processing
from The University of Texas at Austin, Austin, TX,
USA, in 2013 and 2016, respectively, all in electrical Society Best Paper Award, the 2014 EURASIP Journal on Advances in
engineering. He is currently a Senior System Engi- Signal Processing Best Paper Award, the 2014 Journal of Communications
neer with Qualcomm Flarion Technologies, Inc., and Networks Best Paper Award, the 2016 IEEE Communications Society
Fred W. Ellersick Prize, and the 2016 IEEE Communications Society and
Bridgewater, NJ, USA. His research interests include
applications of stochastic geometry, mmWave com- Information Theory Society Joint Paper Award. He was a Frontiers in
munications, and massive MIMO. Education New Faculty Fellow in 2003. He is also a Licensed Amateur Radio
Operator and a Registered Professional Engineer in Texas.

Authorized licensed use limited to: Nanjing University. Downloaded on January 12,2021 at 00:26:06 UTC from IEEE Xplore. Restrictions apply.

You might also like