Sunteți pe pagina 1din 17

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/262677756

Power Scaling of Uplink Massive MIMO


Systems with Arbitrary-Rank Channel Means
Article in IEEE Journal of Selected Topics in Signal Processing May 2014
DOI: 10.1109/JSTSP.2014.2324534 Source: arXiv

CITATIONS

READS

47

98

5 authors, including:
Qi Zhang

Kai-Kit Wong

Nanjing University of Posts and Telecomm

University College London

4 PUBLICATIONS 52 CITATIONS

232 PUBLICATIONS 4,353 CITATIONS

SEE PROFILE

SEE PROFILE

Michail Matthaiou
Queen's University Belfast
125 PUBLICATIONS 1,087 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Michail Matthaiou


Retrieved on: 11 October 2016

966

IEEE JOURNAL OF SELECTED TOPICS IN SIGNAL PROCESSING, VOL. 8, NO. 5, OCTOBER 2014

Power Scaling of Uplink Massive MIMO Systems


With Arbitrary-Rank Channel Means
Qi Zhang, Shi Jin, Member, IEEE, Kai-Kit Wong, Senior Member, IEEE, Hongbo Zhu, and
Michail Matthaiou, Senior Member, IEEE

AbstractThis paper investigates the uplink achievable rates of


massive multiple-input multiple-output (MIMO) antenna systems
in Ricean fading channels, using maximal-ratio combining (MRC)
and zero-forcing (ZF) receivers, assuming perfect and imperfect
channel state information (CSI). In contrast to previous relevant
works, the fast fading MIMO channel matrix is assumed to have an
arbitrary-rank deterministic component as well as a Rayleigh-distributed random component. We derive tractable expressions for
the achievable uplink rate in the large-antenna limit, along with
approximating results that hold for any finite number of antennas.
Based on these analytical results, we obtain the scaling law that
the users transmit power should satisfy, while maintaining a desirable quality of service. In particular, it is found that regardless
of the Ricean -factor, in the case of perfect CSI, the approximations converge to the same constant value as the exact results, as
the number of base station antennas,
, grows large, while the
transmit power of each user can be scaled down proportionally to
. If CSI is estimated with uncertainty, the same result holds
true but only when the Ricean -factor is non-zero. Otherwise,
if the channel experiences Rayleigh fading, we can only cut the
transmit power of each user proportionally to
. In addition, we show that with an increasing Ricean -factor, the uplink
rates will converge to fixed values for both MRC and ZF receivers.
Index TermsMassive MIMO, Ricean fading channels, uplink
rates.

I. INTRODUCTION

ULTIPLE-INPUT multiple-output (MIMO) antenna


technology has emerged as an effective technique for
significantly improving the capacity of wireless communication systems [1], [2]. Recently, multiuser MIMO (MU-MIMO)
Manuscript received September 30, 2013; revised January 18, 2014; accepted April 30, 2014. Date of publication May 16, 2014; date of current
version September 11, 2014. This work was supported in part by the China 973
project under Grant 2013CB329005, the National Natural Science Foundation
of China under Grant 6127123, and the China 863 Program under Grant
2014AA01A705. The work of S. Jin was supported by the National Natural
Science Foundation of China under Grant 61222102 and the Natural Science
Foundation of Jiangsu Province under Grant BK2012021. This paper was
presented in part at the IEEE Globe Communication Conference, Atlanta,
GA, USA, December 2013. The guest editor coordinating the review of this
manuscript and approving it for publication was Prof. A. Lee Swindlehurst.
(Corresponding author: S. Jin.)
Q. Zhang and H. Zhu are with Jiangsu Key Laboratory of Wireless Communications, Nanjing University of Posts and Telecommunications, Nanjing 210003,
China (e-mail: zhangqiqi_1212@126.com; zhuhb@njupt.edu.cn).
S. Jin is with the National Mobile Communications Research Laboratory,
Southeast University, Nanjing 210096, China (e-mail: jinshi@seu.edu.cn).
K.-K. Wong is with the Department of Electronic and Electrical Engineering, University College London, London WC1E 7JE, U.K. (e-mail:
kai-kit.wong@ucl.ac.uk).
M. Matthaiou is with the School of Electronics, Electrical Engineering and
Computer Science, Queens University Belfast, Belfast, BT3 9DT, U.K., and
also with the Department of Signals and Systems, Chalmers University of Technology, 412 96 Gothenburg, Sweden (e-mail: m.matthaiou@qub.ac.uk).
Digital Object Identifier 10.1109/JSTSP.2014.2324534

systems, where a base station (BS) equipped with multiple


antennas serves a number of users in the same time-frequency
resource, have gained much attention because of their considerable spatial multiplexing gains even without multiple antennas
at the users [3][6]. To reap all the benefits of MIMO at a
greater scale, the paradigm of massive MIMO, which considers
the use of hundreds of antenna elements to serve tens of users
simultaneously, has recently come at the forefront of wireless
communications research [7].
Great efforts have been made to understand the spectral
and energy efficiency gains of massive MIMO systems, e.g.,
[7][14]. In particular, [8] indicates that the high number of
degrees-of-freedom can average out the effects of fast fading.
It has further been revealed in [10] that, when the number
of antennas increases without bound, uncorrelated noise, fast
fading and intracell interference vanish. The only impairment
left is pilot-contamination. Another merit of massive MIMO
is that the transmit power can be greatly reduced. In [12], the
power-scaling law was investigated and it was shown that, as
the number of BS antennas grows without limit, the uplink
rate can be maintained while the transmit power can be substantially cut down. For example, ideally, to maintain the same
quality-of-service as with a single-antenna BS, the transmit
power of a 100-antenna BS would be only almost 1% of the
power of the single-antenna one.
The increasing physical size of massive MIMO arrays is a
fundamental problem for practical deployment and utilization
and the millimeter-wave operating from 30 300 GHz finds a
way out, since the small wavelengths make possible for many
antenna elements to be packed with a finite volume. On top of
this, due to the highly directional and quasi-optical nature of
propagation at millimeter-waves, line-of-sight (LOS) propagation is dominating [15][17]. Therefore, massive MIMO systems operating in LOS conditions is expected to be a novel paradigm. Unfortunately, many of the existing pioneering works
simply assume Rayleigh fading conditions [8], [10], [12]. While
this assumption simplifies extensively all mathematical manipulations, it falls short of capturing the fading variations when
there is a specular or LOS component between the transmitter
and receiver. As such, more general fading models need to be
considered.
In this paper, we extend the results in [12] to Ricean fading
channels with arbitrary-rank mean matrices. In [12], the authors
studied the potential of power savings in massive MU-MIMO
systems, assuming that the fast fading channel matrix has
zero-mean unit-variance entries. In our analysis, the fast fading
channel matrix consists of an arbitrary-rank deterministic
component, and a Rayleigh-distributed random component accounting for the scattered signals [18]. We consider a single-cell
MU-MIMO system in the uplink, where both maximal-ratio
combining (MRC) and zero-forcing (ZF) receivers are assumed

1932-4553 2014 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.

ZHANG et al.: POWER SCALING OF UPLINK MASSIVE MIMO SYSTEMS WITH ARBITRARY-RANK CHANNEL MEANS

at the BS, with perfect and imperfect channel state information


(CSI). Some relevant works on Ricean fading in massive
MIMO systems are [19][22]. However, [19] considers the
single-user scenario and in [20][22], the numbers of antennas
at the transmitters and the receiver go to infinity with a constant
ratio. Our main contributions include new, tractable expressions
for the achievable uplink rate in the large-antenna limit, along
with approximating results that hold for any finite number
of antennas. We also elaborate on the power-scaling laws as
follows:
We reveal that under Ricean fading, with perfect CSI, if the
number of BS antennas, , grows asymptotically large,
we can cut down the transmit power of each user proporto maintain a desirable rate. In addition,
tionally to
, the sum rates of both MRC and ZF receivers
as
converge to the same constant value, indicating that in the
large-system limit, intracell interference disappears.
gets
If CSI is estimated with uncertainty, then when
asymptotically large, massive MIMO will still bring considerate power savings for each user. In particular, if the
Ricean -factor is non-zero, the transmit power for each
to obtain the same rate,
user can be scaled down by
while the uplink rates will tend to a fixed value as a function
of Ricean -factor. However, when the Ricean -factor
is zero, the transmit power can be scaled down only by
and the uplink rates will again approach to a fixed
.
value if
The remainder of the paper is organized as follows. Section II
describes the MU-MIMO system model in Ricean fading channels, and provides the definition of the uplink rate with perfect
and imperfect CSI. Section III derives closed-form approximations for the achievable uplink rates and also investigates the
power-scaling laws. In Section IV, we provide a set of numerical results, while Section V summarizes the main results of this
paper.
Notation
Throughout the paper, vectors are expressed in lowercase
boldface letters while matrices are denoted by uppercase boldand
to denote the conjuface letters. We use
gate-transpose, transpose, conjugate and inverse of , respectively. Moreover,
denotes an
identity matrix,
equals 1 when
and 0 otherwise, and
or
gives
is the expectation operator,
the ( )th entry of . Finally,
is the Euclidean norm and
denotes that
is a complex Gaussian matrix with mean matrix and covariance matrix .
II. SYSTEM MODEL
We consider a MU-MIMO system with
single-antenna
users and an -antenna BS, where users transmit their signals
to the BS in the same time-frequency channel. The system is
single-cell with no interference from neighboring cells. The reat the BS can be written as [11]
ceived vector
(1)
where
denotes the
MIMO channel matrix between
users,
denotes the
1 vector conthe BS and the
is the average
taining the transmitted signals from all users,
transmitted power of each user, and represents the vector of
zero-mean additive white Gaussian noise (AWGN). To facilitate

967

our analysis and without loss of generality, the noise variance is


assumed to be 1.
A. Channel Model
We denote the channel coefficient between the th user and
the th antenna of the BS as
, which embraces
independent fast fading, geometric attenuation and log-normal
shadow fading [10] and can be expressed as
(2)
where
is the fast fading element from the th user to
is the large-scale fading
the th antenna of the BS, while
coefficient to model both the geometric attenuation and shadow
fading, which is assumed to be constant across the antenna
array. Under this model, we can write
(3)
where
denotes the
channel matrix modeling fast
and
fading between the users and the BS, i.e.,
is the
diagonal matrix with
. The fast
fading matrix consists of two parts, namely a deterministic component corresponding to the LOS signal and a Rayleigh-distributed random component which accounts for the scattered
signals. Moreover, the Ricean factor, represents the ratio of the
power of the deterministic component to the power of the scattered components. Here, we assume that the Ricean -factor of
each user is different and the th users -factor is denoted by
. Then, the fast fading matrix can be written as [23]
(4)
where is a
diagonal matrix with
,
denotes the random component, the entries of which are independent and identically distributed (i.i.d.) Gaussian random
variables with zero-mean, independent real and imaginary parts,
, and denotes the deterministic comeach with variance
ponent, which was usually assumed to be rank-1 in previous
studies, e.g., [18], [19], [24], [25]. In this paper, due to the assumption of multiple geographically distributed users, this constraint is relaxed and we let have an arbitrary rank as [26]1
(5)
where is the antenna spacing, is the wavelength, and is
the arrival angle of the th user. For convenience, we will set
in the rest of this paper.2
B. Achievable Uplink Rate
1) Perfect CSI: We first consider the case that the BS has
linear receiver matrix which
perfect CSI. Let be the
depends on the channel matrix . The BS processes its received
signal vector by multiplying it with the conjugate-transpose of
the linear receiver as [10]
(6)
Then, substituting (1) into (6) gives
(7)
1Note

that with the change of

2Since

, (5) can become arbitrary-rank.

the physical size of the antenna array depends on the operating freat high frequencies (e.g., 60
quency, it can be very small even with large
GHz communications [27]).

968

The th element of

IEEE JOURNAL OF SELECTED TOPICS IN SIGNAL PROCESSING, VOL. 8, NO. 5, OCTOBER 2014

can be written as
(8)

where
is the th column of
plication, we further get

. By the law of matrix multi-

where represents the


AWGN matrix with i.i.d. zeromean and unit-variance elements. With (4), we can remove the
LOS component, which is assumed to be already known from
(13), and the remaining terms of the received matrix are
(14)

(9)
which can be further written as
where
denotes the th element of and
is the th column
of . Assuming an ergodic channel, the achievable uplink rate
of the th user is given by [12]3

(15)
The minimum mean-square-error (MMSE) estimate of
from
is [28]
(16)

(10)
where
get

. From (13), we can easily

Therefore, the uplink sum rate per cell, measured in bits/s/Hz,


can be defined as
(11)
2) Imperfect CSI: In real situations, the channel matrix is
estimated at the BS. For the considered Ricean fading channel
model, we assume that both the deterministic LOS component
and the Ricean -factor matrix are perfectly known at both
the transmitter and receiver,4 such that the estimate of can be
expressed as

(17)
where
. Noting that
, the entries of
are i.i.d. Gaussian random variables with zero-mean and unitdenote the channel estimation error
variance. Let
and be the linear receiver matrix which depends on . Then,
after linear reception, we have
(18)
As such, the estimated signal for the th user is given by

(12)
where

(19)

denotes the deterministic component of , i.e.,


and
represents the estimate of the random compo. In this paper, we consider that the channel
nent
is estimated using uplink pilots. Let an interval of length symbols be used for uplink training, where is smaller than the coherence time of the channel. In the training stage, all users simultaneously transmit orthogonal pilot sequences of symbols,
matrix
,
which can be stacked into a

where
is the
rewritten as

, where
which satisfies
and
is the transmit pilot power. As a result, the BS
noisy pilot matrix as
receives the

where
and are the th columns of and , respectively.
Note that the last three terms in (20) correspond to intracell interference, channel estimation error and noise, respectively. According to the classical assumption of worst-case uncorrelated
Gaussian noise [12], along with the fact that the variance of elements of estimation error vector is

(13)
3Hereafter, the subscript will correspond to the case of perfect CSI, while
the subscript to the imperfect CSI case.
4Since the LOS channels are hardly changing, the BS may estimate the LOS
components during the previous transmission from the users to the BS. While
factor in massive MIMO systems is an interthe estimation of the Ricean
esting topic for additional research, we do not pursue it herein due to space
constraints.

th column of

, and (19) can be easily

(20)

(21)
the achievable uplink rate of the th user can be given by (22)
(at the bottom of the page).

(22)

ZHANG et al.: POWER SCALING OF UPLINK MASSIVE MIMO SYSTEMS WITH ARBITRARY-RANK CHANNEL MEANS

Similar to the perfect CSI case, the uplink sum rate per cell
can be defined as

969

Proof: See Appendix III.


1) MRC Receivers: For MRC, the linear receiver matrix is
given by
(30)

(23)
where represents the coherence time of the channel, in terms
symbols are
of the number of symbols, during which
used as pilots for channel estimation.

which yields
user is

. From (10), the achievable rate of the th

(31)

III. ANALYSIS OF ACHIEVABLE UPLINK RATE


In this section, we derive closed-form expressions for the
achievable uplink rates in the large-antenna limit along with
tractable approximations that hold for any finite number of antennas. Note that our results are tight and apply for systems with
arbitrary-rank Ricean fading channel mean matrices. We also
quantify the power-scaling laws in the cases of perfect and imperfect CSI. A key preliminary result is given first.
and
are both
Lemma 1: If
sums of nonnegative random variables
and , then we get
the following approximation

Next, we will investigate the power-scaling properties of the


uplink rate in (31); this is done by first presenting the exact rate
limit in the theorem below.
Theorem 1: Using MRC receivers with perfect CSI, if the
,
transmit power of each user is scaled down by a factor of
for
and a fixed
, we have
i.e.,
(32)

(24)
Proof: See Appendix I.
Note that the approximation in (24) does not require the
and
to be independent and becomes
random variables
more accurate as and increase. Thus, in massive MIMO
systems, due to the large number of BS antennas, this approximation will be particularly accurate.

where
Proof: Let
into (31), we obtain

.
, where

. Substituting it

A. Perfect CSI
We begin by considering the case with perfect CSI and establishing some key preliminary results which will be useful in
deriving our main results.
is asympLemma 2: By the law of large numbers, when
totically large, we have

(33)
We can also get

(25)
where -a.s. denotes almost sure convergence.
Proof: See Appendix II.
Lemma 3: The expectation for the inner product of two same
columns in can be found as

(34)
From Lemma 2, it is know that

(26)
and the expectation of the norm-square of the inner product of
any two columns in is given by

(35)
Then, (34) becomes5

(27)
(28)
where

(36)
Using

is defined as
(37)
(29)

5Note that the convergence of logarithms is sure, not almost sure as was
shown in [11].

970

IEEE JOURNAL OF SELECTED TOPICS IN SIGNAL PROCESSING, VOL. 8, NO. 5, OCTOBER 2014

and Lemma 2 again, it can be easily shown that6


(38)
Applying (38) in (36) yields the desired result.
Note that we find a deterministic equivalent for
in
varies depending on . If
Theorem 1 and the value of
,
will converge to zero, which means that the
transmit power of each user has been cut down too much. In
contrast, if
,
will grow without bound, which means
that the transmit power of each user can be reduced more to
maintain the same performance. These observations illustrate
can make
become a fixed non-zero
that only
value. Hence, by setting
in (32), we have the following
corollary:
Corollary 1: With no degradation in the th users rate, using
MRC receivers and perfect CSI, the transmit power of each user
. Then, the achievable uplink
can be cut down at most by
rate becomes
(39)
Corollary 1 shows that when
grows without bound, we
to
can scale down the transmit power proportionally to
maintain the same rate, which aligns with the conclusion in [12]
for the case of Rayleigh fading. It also indicates that the exact
limit of the uplink rate with MRC processing and perfect CSI
is irrelevant to the underlying fading model, since the Ricean
-factor does not appear in the limiting expression.
The following theorem presents a closed-form approximation
for the achievable uplink rate with MRC receivers and perfect
CSI. This approximation hold for any finite number of antennas;
it can also reveal more effectively the impact of the Ricean
-factor on the rate performance.
Theorem 2: Using MRC receivers with perfect CSI, the
achievable uplink rate of the th user can be approximated as

(40)
where
Proof: Applying Lemma 1 on (31) yields

The three expectations in (41) can be evaluated using the re, the substitution of
sults in Lemma 3. Now, with
(26)(28) into (41) leads to the final result.
When
, (40) reduces to the special case of
Rayleigh fading channels:

(42)

It is known that

(43)

and the right hand side of (43) is the uplink rate lower bound
with perfect CSI in Rayleigh fading channels given by [12,
Proposition 2].
We use the expression in Theorem 2 to analyze how the uplink
rate changes with the Ricean -factor. The following corollary
presents the uplink rate limit when the Ricean -factor grows
without bound. To the best of our knowledge, this result is also
new.
, the approxCorollary 2: If for any and ,
imation in (40) converges to

(44)

It is interesting to note from Corollary 2 that as


grows,
the uplink rate of MRC will also approach a constant value.
With the power-scaling law uncovered in Corollary 1, we now
investigate the impact of power reduction on the uplink rate
approximation.
, when
grows without
Corollary 3: With
bound while the Ricean -factor is fixed, the result in Theorem
2 tends to
(45)

Comparing Corollary 3 with Corollary 1, we find that when


is large, the uplink rate approximation becomes the same
, things that are
as the exact rate. This is because as
random become deterministic and (24) will hold with equality.
, the linear receiver
2) ZF Receivers: For ZF with
so that [29]
is given by
(46)
(41)

6Note that
istic equivalent of

is not a constant value and is actually the determin. Therefore, it is a notational abuse to write
.

Substituting (46) into (10) yields [25]


(47)

ZHANG et al.: POWER SCALING OF UPLINK MASSIVE MIMO SYSTEMS WITH ARBITRARY-RANK CHANNEL MEANS

Similar to the case with MRC receivers, we first quantify the


power-scaling law as follows.
Theorem 3: Using ZF receivers with perfect CSI, if the
transmit power of each user is scaled down by a factor of
, i.e.,
for
and a fixed
, we have
(48)
where
Proof: With

.
, we can write (47) as

971

variance matrix of the row vectors of


and is the mean matrix of , i.e.,

, i.e.,

Now, we can approximate


by a central Wishart distribution with covariance matrix as [31]7
(57)

(49)
which leads to
(50)
With the aid of Lemma 2, we can easily get the desired result.
It is interesting to note from Theorem 3 that with increasing
, the deterministic equivalent for the uplink rate of ZF
, is the same as that of MRC receivers,
.
receivers,
Hence, we can draw the conclusion that with no degradation
in the th users rate, in the case of ZF receivers and perfect
CSI, the transmit power of each user can be cut down at most
; then, the achievable uplink rate becomes
by
(51)
The following theorem gives a closed-form approximation
for the uplink rate using ZF receivers with perfect CSI. Note
that the steps to derive this expression are more complicated
than in the case with MRC receivers, since we need to deal with
complex non-central Wishart matrices.
Theorem 4: Using ZF receivers with perfect CSI, the achievable uplink rate of the th user can be approximated by
(52)
where
(53)

(58)
. Under the above assumpNow, let
tion, this is a Chi-squared random variable with distribution [34]
(59)
where
pressing the integral

is the gamma function. By ex, we have8


(60)

From (60) and (56), (52) can be easily obtained.


The following corollary presents the exact limit of the approximation in (52) for ZF receivers with perfect CSI as the
Ricean -factor becomes large.
, the approxCorollary 4: If for any and ,
imation in (52) converges to9
(61)
By the discussion after Theorem 3, we know that the transmit
power of each user with ZF receivers and perfect CSI can be
. Applying this result into the apat most scaled down by
proximation (52) gives the following corollary.
, when
grows without
Corollary 5: With
bound while is fixed, the result in Theorem 4 tends to
(62)

(54)

into (52), we obtain

Proof: Substituting

Proof: By the convexity of


and Jensens inequality, we obtain the following lower bound on the achievable
uplink rate in (47)
(55)
With

(63)

The entries of

if
if

, we can rewrite (55) as


(56)
follows a non-central Wishart distribution,
[30], where is the co-

.
.
(64)

7Note that this approximation has been often used in the context of MIMO
systems, e.g., [32], [33].
8The

Note that
denoted by

are known as

first negative moment of

9Note

exists only if

[25].

that Corollary 4 requires


to be full-rank and this can be guaranteed with great probability in practice, since randomly distributed users are very
likely to have different angles of arrival.

972

IEEE JOURNAL OF SELECTED TOPICS IN SIGNAL PROCESSING, VOL. 8, NO. 5, OCTOBER 2014

When

grows large, we asymptotically have

Lemma 5: The expectation for the inner product of two same


columns in can be found as

(65)
becomes asymptotically an identity
Therefore,
matrix. Applying this result into (58) and then combining it
with (63), we obtain the desired result.
Corollary 5 shows that when
is large, similar to MRC, the
approximation for ZF also converges to the exact limit.
B. Imperfect CSI
In real scenarios, CSI is estimated at the receiver. A typical
channel estimation process using pilots is assumed in this paper
and has been described in Section II. Following a similar approach as for the case with perfect CSI, we first give some key
preliminary results.
Lemma 4: By the law of large numbers, when is asymptotically large, the inner product of any two columns in the channel
matrix
can be found as
(66)
(67)
where

denotes the th column of

, i.e.,
(68)

(69)
and the expectation for the norm-square of the inner product of
any two columns in is expressed as (70) (at the bottom of the
page).
Proof: See Appendix V.
1) MRC Receivers: For MRC, we have
. As a result,
from (22), the uplink rate of the th user with MRC receivers
can be written as (71) (at the bottom of the page).
Same as before, we will investigate the power-scaling law
but this time with imperfect CSI. We will first derive the exact
uplink rate and then an approximation for further analysis.
Theorem 5: Using MRC receivers with imperfect CSI from
MMSE estimation, if the transmit power of each user is scaled
, i.e.,
for
and a
down by a factor of
, we have
fixed
(72)
where

Proof: Let
, where
. Substituting it
into (71) yields (73) (at the bottom of the page). Let the numer, we obtain
ator and denominator of (73) be multiplied by
, using (67) in
(74) (at the bottom of the page). When
Lemma 4, it can be easily obtained that
(75)

Proof: See Appendix IV.

if
if

(70)

(71)

(73)

(74)

ZHANG et al.: POWER SCALING OF UPLINK MASSIVE MIMO SYSTEMS WITH ARBITRARY-RANK CHANNEL MEANS

and with
(76)
we can further simplify (74) as

The remaining task is to evaluate the limit of


Using (66) in Lemma 4, along with the fact that

(77)
.
(78)

and

, we get

973

be cut down by
. This is reasonable since LOS propagation reduces fading fluctuations, thereby increasing the received
signal-to-interference-plus-noise ratio. As a consequence, with
no reduction in the rate performance, the transmit power for
non-zero Ricean -factor can be cut down more aggressively.
The following theorem presents a closed-form approximation
for the uplink rate using MRC with imperfect CSI.
Theorem 6: Using MRC receivers with imperfect CSI from
MMSE estimation, the achievable uplink rate of the th user
can be approximated by (82) (at the bottom of the page). where
.
Proof: Follow the same procedure as Theorem 2. After utilizing Lemma 1 and (69)(70), we can obtain the desired result
after some simplifications.
Note that when
,
reduces to the special
case of Rayleigh fading channel. After performing some simplifications, we have

(79)
Since
,
and we can further simplify (79) to
complete the proof.
It is important to note from Theorem 5 that the value of the deis dependent on both the Ricean
terministic equivalent for
-factor and the scaling parameter , which can be more pre, with increasing ,
cisely described as follows. When
only
can make
approach a non-zero constant
value. Otherwise, it will tend to zero with
and grow
, which will change the rate perwithout bound with
, we find that only
can
formance. For the case
lead
to a fixed value, and when
and
, it will
approximately become zero and , respectively. These observations can be summarized as the power-scaling law in the following corollary.
Corollary 6: With no degradation in the th users rate, using
MRC receivers with imperfect CSI and for a fixed , we obtain
the following power-scaling law. When the th users Ricean
-factor is zero, we can at most scale down the transmit power
and the achievable uplink rates becomes
to
(80)
On the other hand, with a non-zero Ricean
power can be at most scaled down to

-factor, the transmit


and we have

(83)
It is known that

(84)
and the right hand side of (84) is the uplink rate lower bound
with imperfect CSI in Rayleigh fading channels given by [12,
Proposition 6].
Now, we consider the particular case when the Ricean
-factor grows infinite for the approximation in Theorem 6.
, the approxCorollary 7: If for any and ,
imation in (82) converges to

(81)
Corollary 6 reveals that with imperfect CSI, the amount of
.
power reduction achievable on the th user depends on
, the transmit power
For Rayleigh fading channels, i.e.,
, which agrees
can be at most cut down by a factor of
with the result in [12, Proposition 5]; however, for channels
, the transmit power can
with LOS components, i.e.,

(85)

This conclusion indicates that when the Ricean -factor


grows large, the approximation of the uplink rate using MRC
with imperfect CSI will approach a fixed value, which is the

(82)

974

IEEE JOURNAL OF SELECTED TOPICS IN SIGNAL PROCESSING, VOL. 8, NO. 5, OCTOBER 2014

same as the constant value in the case of perfect CSI given by


(44). That is, in a purely deterministic channel, the approximation for MRC receivers will tend to the same constant value
regardless of the CSI quality at the BS.
With the power-scaling law already derived in Corollary 6,
we now perform the same analysis but on the uplink rate approximation in Theorem 6.
for a fixed , when grows
Corollary 8: If
without bound, the limit of (82) will exist only when
and (82) becomes

Proof: Let
, where
. Then, (89) becomes (91) (at the bottom of the page). By Lemma 4, we know
that
(92)
where

is a diagonal matrix containing

(86)
for a fixed
, when
grows without bound,
If
(82) will converge to a non-zero constant value only when
and (82) becomes

along its main diagonal. The inverse matrix of can be easily


obtained by calculating the reciprocals of its main diagonal. As
, we have
a consequence, using the fact

(87)
Proof: With the fact that
(88)
the result then follows by performing some basic algebraic manipulations.
Again, it is noted that when
, the approximation
of the uplink rate becomes identical to the exact limit obtained
from Corollary 6.
.
2) ZF Receivers: For ZF receivers, we have
According to (22), the achievable uplink rate of ZF receivers
with imperfect CSI is given by

(93)
Due to

, as
into (93) yields

, the substitution of

(94)
and we can also easily get
(95)

(89)
Similar to the case with MRC receivers, we next investigate
the power-scaling law for ZF receivers.
Theorem 7: Using ZF receivers with imperfect CSI from
MMSE estimation, if the transmit power of each user is scaled
, i.e.,
for
and a
down by a factor of
, we have
fixed
(90)
where

The desired result is obtained by substituting (94) and (95) into


(91).
From Theorem 7, it is worth pointing out that with imperfect
CSI and increasing , the deterministic equivalent for the uplink rate of ZF receivers becomes the same as that of MRC receivers in (72), which agrees with the conclusion in the case of
perfect CSI. Therefore, we can easily obtain the conclusion that
with no degradation in the th users rate performance, using
, when
,
ZF receivers with imperfect CSI for a fixed
we can at most scale down the transmit power of each user to
and the achievable uplink rates becomes
(96)
On the other hand, with a non-zero
, the transmit power of
and we
each user can be at most scaled down to
have
(97)

(91)

ZHANG et al.: POWER SCALING OF UPLINK MASSIVE MIMO SYSTEMS WITH ARBITRARY-RANK CHANNEL MEANS

975

It is also important to point out that for both perfect and imperfect CSI, the exact limit of the uplink rate of ZF receivers
equals that of MRC receivers, which is consistent with the conclusion given in [12]. The following theorem presents a closedform approximation for the achievable uplink rate using ZF receivers with imperfect CSI.
Theorem 8: Using ZF receivers with imperfect CSI from
MMSE estimation, the achievable uplink rate of the th user
can be approximated by

where

(98)

(99)
is
a
diagonal
matrix
consisting
of
and
.
Proof: Applying Jensens inequality in (89), we get

(100)
The proof then follows the same procedure as in Theorem 4.
Corollary 9: If for any and ,
, the approximation in (98) converges to
(101)
Note that the fixed value in this case is the same as the limit
in Corollary 4 given by (61), which shows that for both MRC
and ZF receivers, as the Ricean -factor grows without bound,
the approximations of the uplink rate will tend to the same fixed
value regardless of the CSI quality.
for a fixed
, when
Corollary 10: If
grows without bound, the limit of (98) will exist only when
and (98) becomes
(102)
for a fixed
, when
grows without bound,
If
(98) will converge to a non-zero constant value only when
and (98) becomes
(103)
Proof: From the proof of Corollary 5, we know that when
is large,
becomes asymptotically an identity
matrix. With the fact that
(104)
the result then follows by some basic algebraic manipulations.

Fig. 1. Uplink sum rate per cell versus the number of BS antennas, with
users and transmit power
.

It is interesting to note from Corollary 10 that for both perfect


grows large, the exact limit equals the
and imperfect CSI, as
approximating expression regardless of the type of the receiver.
IV. NUMERICAL RESULTS
For our simulations, we consider a cell with a radius of
users distributed randomly and uniformly
1000 m and
over the cell, with the exclusion of a central disk of radius
. For convenience, we assume that every user
has the same Ricean -factor, denoted by . The large-scale
, where
channel fading is modelled using
is a log-normal random variable with standard deviation
, is the path loss exponent, and
is the distance between
the th user and the BS. We have assumed that
and
. Also, the transmitted data are modulated using
orthogonal frequency-division multiplexing (OFDM). The
parameters were chosen according to the LTE standard: an
, a subOFDM symbol interval of
, a useful symbol duration
carrier spacing of
. We choose the channel coherence time
. As a result, the coherence time of the channel
to be
.
becomes
In Fig. 1, the simulated uplink sum rates per cell of MRC and
ZF receivers in (31), (47), (71) and (89) are compared with their
corresponding analytical approximations in (40), (52), (82) and
(98), respectively. Results are presented for two different scenarios perfect CSI and imperfect CSI. In this example, the
number of users is assumed to be 10 with the transmit power
. For each receiver, results are shown for different
, 3 dB, 6
values of the Ricean -factor, which are
dB, 10 dB, respectively. Clearly, in all cases, we see a precise
agreement between the simulation results and our analytical reincreases, the sum rate grows
sults. For MRC receivers, as
obviously for both perfect and imperfect CSI. For ZF receivers,
the sum rate has a noticeable growth only with imperfect CSI.
We also find that both rates grow without bound, when no power
normalization is being performed.
Fig. 2 investigates the power-scaling law when the transmit
. In the simulations, we
power of each user is scaled down by
and has two different values,
choose
and 6 dB. With
and when
, results show
that the analytic approximations are the same as the exact results.
In the case of perfect CSI, the sum rates of MRC and ZF receivers
approach the same value, independent of the Ricean -factor,

976

IEEE JOURNAL OF SELECTED TOPICS IN SIGNAL PROCESSING, VOL. 8, NO. 5, OCTOBER 2014

Fig. 2. Uplink sum rate per cell versus the number of BS antennas, with
users and scaled-down power
, where
.

as predicted in Section III-B. For imperfect CSI, the sum rates


of MRC and ZF receivers always tend to the same value as
becomes large, which depends on . Note that when
,
the sum rates do not have a monotonic behavior. This is because
when is small, has not been cut down so much. Therefore,
is
the improvement of the sum rate caused by the increased
greater than the loss of the sum rate caused by the reduced .
gets larger,
is cut down more aggressively.
However, as
with imperfect CSI, the decay of the sum
For the case
rate brought by the reduced
exceeds the growth of the sum
rate brought by the increased . Hence, the sum-rate curve will
first rise and then drop as increasing . In all other cases, with
large , there is a balance between the increase and decrease of
and scaling-down ,
the sum rate brought by the increased
respectively. Therefore, these curves eventually saturate with an
increased . Note that all these observations agree with (81) and
are consistent with the
(97) and the curves for the case
simulation results given by [12, Fig. 2].
In Fig. 3, we consider the same setting as in Fig. 2, but the
transmit power for each user is scaled down by a factor of
. For convenience, we only show the results in the case
with imperfect CSI. As can be seen, the sum rates of both MRC
and ZF receivers converge to deterministic constants when
but increase with
when
, which implies that
can be cut down more, just as predicted
the transmit power
in Section III-C. Compared with Fig. 2, we see that ZF performs
much better than MRC. This is because ZF performs inherently
better than MRC at high signal to noise ratio (SNR) and the
, which
transmit power here is only scaled down as
makes the SNR relatively high. Again, all these observations
are consistent with the conclusions in Section III-C. Interestingly, the convergence speed of the curves corresponding to the
is rather slow (approximately
times slower
case
than that in the case considered in Fig. 2).
To conclude this section, we study how the uplink sum rate
varies with the Ricean -factor in Fig. 4. Due to the tightness
between the simulated values and the approximations, here we
only use the approximations in (40), (52), (82) and (98) for anal,
,
ysis. The parameters were assumed to be
. In all cases, as expected, the sum rates approxiand
. The two black dots represent
mate the fixed value as
the values obtained by (44) and (61), respectively. The two black
stars correspond to the values of the black dots multiplied by a
for the calculation of the sum rate with
coefficient

Fig. 3. Uplink sum rate per cell versus the number of BS antennas with
users and scaled-down power
, where
.

Fig. 4. Uplink sum rate per cell versus the Ricean -factor with
users, number of BS antennas
and transmit power

imperfect CSI, as defined in (23). It is observed that every dot is


indistinguishable from the curve limit, which not only means that
our analytical results are accurate, but also that for both MRC and
ZF receivers, the sum rates will tend to the same value as increases for both CSI cases. We also find that the sum rates rise
as increases with , except for the case of ZF with perfect CSI.
This is because the performance of ZF receivers can be severely
limited if the channel matrix is ill-conditioned [35], [36]. For the
case of ZF receivers with perfect CSI, we experience no intergrows, the
user interference and no estimation-error. When
channel matrix becomes identical to , whose singular values
have a large spread. Then, the condition number10 of attains
very high values, which implies that is ill-conditioned. However,
for the case of ZF with imperfect CSI, the main limiting factor
is the estimation error. When grows large, channel estimation
becomes far more robust, since quantities that were random before become deterministic. The same holds true for both cases of
MRC receivers, where a higher Ricean -factor reduces the effects of inter-user interference.
V. CONCLUSION
This paper worked out the achievable uplink rate of massive
MIMO systems over Ricean fading channels with arbitrary-rank
10Note that we are working with the 2-norm condition number, which is the
ratio of the smallest to the largest singular value.

ZHANG et al.: POWER SCALING OF UPLINK MASSIVE MIMO SYSTEMS WITH ARBITRARY-RANK CHANNEL MEANS

mean. We deduced new, tractable expressions for the uplink rate


in the large-antenna limit along with exact approximations. Our
analysis incorporated both ZF and MRC receivers and the cases
of both perfect and imperfect CSI. These results were used to
pursue a detailed analysis of the power-scaling law and of how
the uplink rates change with the Ricean -factor. We observed
that with no degradation in the rate performance, the transmit
,
power of each user can be at most cut down by a factor of
except for the case of Rayleigh fading channels with imperfect
CSI where the transmit power can only be scaled down up to a
. With scaling power and the same CSI quality,
factor of
it was also shown that the uplink rates of MRC and ZF receivers
will tend to the same constant value. Finally, with increasing
Ricean -factor and the same receiver, the uplink rates for perfect and imperfect CSI will also converge to the same fixed
value.

977

The above equation can be further written as

(110)
Comparing

(108)

and
(110),
we
find
that
is a quantity that lies between the upper and lower bound of
.
Therefore, it can be used as an approximation.
Moreover, we can quantify the offset between these two
bounds as

APPENDIX I
PROOF OF LEMMA 1
We know that

By applying a Taylor series expansion of


we get that

around

(111)
,

(105)
With the Jensens inequality, we can get the following bounds:

(112)
(106)
and

The substitution of (112) into (111) yields

(107)
Combing (106) and (107), a lower and upper bound on (105)
can be obtained as follows

(108)
Utilizing

and
and combining them with the left and right hand side
of (108), respectively, we get that

(113)
Since both and are nonnegative, as and grow large,
and
will increase, and according to the law
and will approach their means, that
of large numbers,
is, their variances will become small. Hence, the offset decreases with the increase of and , that is, this approximation will be more and more accurate as the number of random
variables of and increase.
APPENDIX II
PROOF OF LEMMA 2
According to the law of large numbers, we know that
(114)
The entries of can be obtained from (4) and (5) described in
Section II-A. Hence,

(109)

(115)

978

IEEE JOURNAL OF SELECTED TOPICS IN SIGNAL PROCESSING, VOL. 8, NO. 5, OCTOBER 2014

with
and
representing the independent real and imag, respectively, each with zero-mean and
inary parts of
variance of 1/2. After substituting (115) into (114) and with the
definition

parts. Hence, its norm-square can be obtained by squaring (120)


directly as (122) (at the bottom of the page).
Applying
in (122) and
removing all the terms with zero mean, we can get

(116)
we have, as

(123)
(117)

If

, from (117) we know that

, it can be easily obtained that


. If
, the last three terms of (117) all become zero and the only
remaining term is
If

(124)
To obtain the norm-square of (124), both the real and imaginary
parts should be extracted. Recalling (118), we can rewrite (124)
as
(118)
where
by using

is defined in (29) and


[26, Eq. (14)]. Due to

the

is obtained
fact that

is bounded, we get as
(125)
(119)
Hence, we arrive at the desired result in (25).
APPENDIX III
PROOF OF LEMMA 3
Using (117) in Appendix II, we first consider the case
and we have
(126)
Then, with
(120)

(127)
substituting (125) and (126) into (127) and removing the terms
with zero expectation, we get the final result as

Then, it is easily obtained that


(121)
To evaluate the norm-square of a vector, we first should exhas only real
tract its real and imaginary parts. Note that

(128)
As we have obtained all expectations used in Lemma 3, we now
conclude the proof.

(122)

ZHANG et al.: POWER SCALING OF UPLINK MASSIVE MIMO SYSTEMS WITH ARBITRARY-RANK CHANNEL MEANS

APPENDIX IV
PROOF OF LEMMA 4

979

which yields that

Following the similar procedure as in the proof of Lemma 2,


we start by giving the entries of the channel matrix. The imperfect CSI model has been described in (12) of Section II-B2,
which yields

(135)
Now consider
(133), with

. Following the same operations as in

(129)
The first term which does not need to be estimated is the same as
, we use the MMSE estimain the case with perfect CSI. For
tion to get it and the details of the method have been introduced
in Section II-B2. Therefore, we have

(136)
we find (137) (at the bottom of the page). Since the entries of
and
are all i.i.d. zero-mean with unit-variance, many terms
in (137) have zero expectation. The only remaining term is the
element of the channel mean matrix. Hence, (137) can be written
as

(130)
where
,
and
have all been defined in (116) of
and
represent the independent real
Appendix III,
with zero-mean and variance 1/2,
and imaginary part of
is defined in (68). By the law of large numbers, it can be
and
got that

(138)
which can be simplified as

(131)
With (131) and the definition
(132)
we can obtain (133) (at the bottom of the page), as
.
Evaluating the expectation of all terms in (133) and removing
the terms with zero expectation, (133) can be simplified as

(139)
According to (119), we know that, as
(140)
As a consequence, (139) can be further simplified as

(134)

(141)

(133)

(137)

980

IEEE JOURNAL OF SELECTED TOPICS IN SIGNAL PROCESSING, VOL. 8, NO. 5, OCTOBER 2014

(143)

(144)

APPENDIX V
PROOF OF LEMMA 5
The expectation of
can be easily obtained by (134) in
Appendix IV. Hence, we have
(142)
Following the same procedure as in the proof of Lemma 3, we
can obtain the remaining three norm-square expectations. First
consider the case
. According to (133), we have (143)
(at the top of the page). After expanding the above equation
and removing all the terms with zero expectation, the remaining
terms are written as (144) (at the top of the page).
, we use the same method as in the proof
For the case
of Lemma 3. With (137), it is noted that the inner product
has non-zero real and imaginary parts. The result then follows
trivially by some algebraic manipulations.
REFERENCES
[1] A. Paulraj, R. Nabar, and D. Gore, Introduction to Space-time Wireless
Communications. Cambridge, U.K.: Cambridge Univ. Press, 2003.
[2] D. N. C. Tse and P. Viswanath, Fundamentals of Wireless Communication. Cambridge, U.K.: Cambridge Univ. Press, 2005.
[3] D. Gesbert, M. Kountouris, R. W. Heath Jr., C.-B. Chae, and T. Salzer,
Shifting the MIMO paradigm, IEEE Signal Process. Mag., vol. 24,
no. 5, pp. 3646, Sep. 2007.
[4] P. Viswanath and D. N. C. Tse, Sum capacity of the vector Gaussian
broadcast channel and uplink-downlink duality, IEEE Trans. Inf.
Theory, vol. 49, no. 8, pp. 19121921, Aug. 2003.
[5] G. Caire, N. Jindal, M. Kobayashi, and N. Ravindran, Multiuser
MIMO achievable rates with downlink training and channel state
feedback, IEEE Trans. Inf. Theory, vol. 56, no. 6, pp. 28452866,
Jun. 2010.
[6] J. Jose, A. Ashikhmin, T. L. Marzetta, and S. Vishwanath, Pilot contamination and precoding in multi-cell TDD systems, IEEE Trans.
Wireless Commun., vol. 10, no. 8, pp. 26402651, Aug. 2011.
[7] F. Rusek, D. Persson, B. K. Lau, E. G. Larsson, T. L. Marzetta, O.
Edfors, and F. Tufvesson, Scaling up MIMO: Opportunities and challenges with very large arrays, IEEE Signal Process. Mag., vol. 30, no.
1, pp. 4060, Jan. 2013.
[8] I. E. Telatar, Capacity of multi-antenna Gaussian channels, Europ.
Trans. Commun., pp. 585595, Nov.-Dec. 1999.
[9] T. L. Marzetta, How much training is required for multiuser
MIMO?, in Proc. Asilomar Conf. Signals, Syst., Comput., Oct. 2006,
pp. 359363.
[10] T. L. Marzetta, Noncooperative cellular wireless with unlimited numbers of base station antennas, IEEE Trans. Wireless Commun., vol. 9,
no. 11, pp. 35903600, Nov. 2010.

[11] 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. 160171, Feb. 2013.
[12] H. Q. Ngo, E. G. Larsson, and T. L. Marzetta, Energy and spectral efficiency of very large multiuser MIMO systems, IEEE Trans. Commun.,
vol. 61, no. 4, pp. 14361449, Apr. 2013.
[13] A. Pitarokoilis, S. K. Mohammed, and E. G. Larsson, On the optimality of single-carrier transmission in large-scale antenna systems,
IEEE Wireless Commun. Lett., vol. 1, no. 4, pp. 276279, Aug. 2012.
[14] S. Wagner, R. Couillet, D. T. M. Slock, and M. Debbah, Large system
analysis of zero-forcing precoding in MISO broadcast channels with
limited feedback, in Proc. IEEE Int. Work. Signal Process. Adv. Wireless Commun. (SPAWC), Jun. 2010.
[15] A. M. Sayeed and N. Behdad, Continuous aperture phased MIMO: A
new architecture for optimum line-of-sight links, in Proc. IEEE Int.
Symp. Ant. Propag. (APS), Jul. 2011, pp. 293296.
[16] J. Brady, N. Behdad, and A. M. Sayeed, Beamspace MIMO for millimeter-wave communications: System architecture, modeling, analysis, and measurements, IEEE Trans. Antennas Propag., vol. 61, no.
7, pp. 38143827, Jul. 2013.
[17] A. M. Sayeed and J. Brady, Beamspace MIMO for high-dimensional
multiuser communication at millimeter-wave frequencies, in Proc.
IEEE Global Telecommun. Conf. (GLOBECOM), Dec. 2013.
[18] A. Lozano, A. M. Tulino, and S. Verd, Multiple-antenna capacity in
the low-power regime, IEEE Trans. Inf. Theory, vol. 49, no. 10, pp.
25272544, Oct. 2003.
[19] S. Jin, X. Q. Gao, and X. H. You, On the ergodic capacity of rank-1
Ricean fading MIMO channels, IEEE Trans. Inf. Theory, vol. 53, no.
2, pp. 502517, Feb. 2007.
[20] J. Zhang, C. K. Wen, S. Jin, X. Gao, and K.-K. Wong, On capacity
of large-scale MIMO multiple access channels with distributed sets of
correlated antennas, IEEE J. Sel. Areas Commun., vol. 31, no. 2, pp.
133148, Feb. 2013.
[21] C. K. Wen, S. Jin, and K.-K. Wong, On the sum-rate of multiuser
MIMO uplink channels with jointly-correlated Rician fading, IEEE
Trans. Commun., vol. 59, no. 10, pp. 28832895, Oct. 2011.
[22] C. K. Wen, K.-K. Wong, and J. C. Chen, Asymptotic mutual information for Rician MIMO-MA channels with arbitrary inputs: A replica
analysis, IEEE Trans. Commun., vol. 58, no. 10, pp. 27822788, Oct.
2010.
[23] J. G. Proakis, Digital Communications, 4th ed. New York, NY, USA:
McGraw-Hill, 2001.
[24] G. Alfano, A. Lozano, A. M. Tulino, and S. Verd, Mutual information
and eigenvalue distribution of MIMO Ricean channels, in Proc. IEEE
Int. Symp. Inf. Theory Appl. (ISITA), Oct. 2004, pp. 1013.
[25] M. Matthaiou, C. Zhong, and T. Ratnarajah, Novel generic bounds on
the sum rate of MIMO ZF receivers, IEEE Trans. Signal Process., vol.
59, no. 9, pp. 43414353, Sep. 2011.
[26] N. Ravindran, N. Jindal, and H. C. Huang, Beamforming with finite rate feedback for LOS MIMO downlink channels, in Proc. IEEE
Global Commun. Conf. (GLOBECOM), Nov. 2007, pp. 42004204.
[27] E. Torkildson, U. Madhow, and M. Rodwell, Indoor millimeter wave
MIMO: Feasibility and performance, IEEE Trans. Wireless Commun.,
vol. 10, no. 12, pp. 41504160, Dec. 2011.

ZHANG et al.: POWER SCALING OF UPLINK MASSIVE MIMO SYSTEMS WITH ARBITRARY-RANK CHANNEL MEANS

[28] S. M. Kay, Fundamentals of Statistical Signal Processing: Estimation


Theory. Englewood Cliffs, NJ, USA: Prentice-Hall, 1993, vol. 1.
[29] C. Wang, E. Au, R. Murch, W. Mow, R. Cheng, and V. Lau, On the
performance of the MIMO zero-forcing receiver in the presence of
channel estimation error, IEEE Trans. Wireless Commun., vol. 6, no.
3, pp. 805810, Mar. 2007.
[30] A. M. Tulino and S. Verd, Random matrix theory and wireless communications, Foundat. Trends Commun.. Inf. Theory, vol. 1, no. 1, pp.
1182, Jun. 2004.
[31] H. Steyn and J. Roux, Approximations for the non-central Wishart
distribution, South African Statist J., vol. 6, pp. 165173, 1972.
[32] M. Matthaiou, M. R. McKay, P. J. Smith, and J. A. Nossek, On the
condition number distribution of complex Wishart matrices, IEEE
Trans. Commun., vol. 58, no. 6, pp. 17051717, Jun. 2010.
[33] C. Siriteanu, Y. Miyanaga, S. D. Blostein, S. Kuriki, and X. Shi,
MIMO zero-forcing detection analysis for correlated and estimated Rician fading, IEEE Trans. Veh. Technol., vol. 61, no. 7, pp.
30873099, Sep. 2012.
[34] D. Gore, R. W. Heath Jr., and A. Paulraj, Transmit selection in spatial multiplexing systems, IEEE Commun. Lett., vol. 6, no. 11, pp.
491493, Nov. 2002.
[35] E. G. Larsson, MIMO detection methods: How they work, IEEE
Signal Process. Mag., vol. 26, no. 3, pp. 9195, May 2009.
[36] H. Artes, D. Seethaler, and F. Hlawatsch, Efficient detection algorithms for MIMO channels: A geometrical approach to approximate
ML detection, IEEE Trans. Signal Process., vol. 51, no. 11, pp.
28082820, Nov. 2003.

Qi Zhang received the B.S. degree in Electrical and


Information Engineering from Nanjing University
of Posts and Telecommunications, Nanjing, China,
in 2010. She is currently working toward the Ph.D.
degree in Communication and Information System
at the Nanjing University of Posts and Telecommunications, China. Her research interests include
massive MIMO systems and space-time wireless
communications.

Shi Jin (S06M07) received the B.S. degree in


communications engineering from Guilin University
of Electronic Technology, Guilin, China, in 1996, the
M.S. degree from Nanjing University of Posts and
Telecommunications, Nanjing, China, in 2003, and
the Ph.D. degree in communications and information
systems from the Southeast University, Nanjing, in
2007. From June 2007 to October 2009, he was a
Research Fellow with the Adastral Park Research
Campus, University College London, London, U.K.
He is currently with the faculty of the National
Mobile Communications Research Laboratory, Southeast University. His
research interests include space time wireless communications, random matrix
theory, and information theory. He serves as an Associate Editor for the IEEE
TRANSACTIONS ON WIRELESS COMMUNICATIONS, and IEEE COMMUNICATIONS
LETTERS, and IET Communications. Dr. Jin and his co-authors have been
awarded the 2011 IEEE Communications Society Stephen O. Rice Prize Paper
Award in the field of communication theory and a 2010 Young Author Best
Paper Award by the IEEE Signal Processing Society.

981

Kai-Kit Wong (S99M01SM08) received the


B.Eng., the M.Phil., and the Ph.D. degrees, all in
Electrical and Electronic Engineering, from the
Hong Kong University of Science and Technology,
Hong Kong, in 1996, 1998, and 2001, respectively.
Since August 2006, he has been with University
College London, first at Adastral Park Campus
and at present the Department of Electronic and
Electrical Engineering, where he is a Reader in
Wireless Communications.
Dr. Wong is a Senior Member of IEEE and Fellow
of the IET. He is on the editorial board of IEEE Wireless Communications
Letters, IEEE Communications Letters, IEEE ComSoc/KICS Journal of
Communications and Networks, and IET Communications. He also previously
served as Editor for IEEE TRANSACTIONS ON WIRELESS COMMUNICATIONS
from 20052011 and IEEE SIGNAL PROCESSING LETTERS from 20092012.

Hongbo Zhu received the bachelor degree in


Telecommunications Engineering from the Nanjing
University of Posts and Telecommunications, Nanjing, China, and the Ph.D. degree in Information
and Communications Engineering from Beijing
University of Posts and Telecommunications, Beijing, China, in 1982 and 1996, respectively. He is
presently working as a Professor and Vice-president
in Nanjing University of Posts and Telecommunications, Nanjing, China. He is also the head of the
Coordination Innovative Center of IoT Technology
and Application (Jiangsu), which is the first governmental authorized Coordination Innovative Center of IoT in China. He also serves as referee or expert in
multiple national organizations and committees.
He has published more than 200 papers on information and communication
area, such as IEEE Transactions Presently, he is leading a big group and multiple
funds on IoT and wireless communications with current focus on architecture
and enabling technologies for Internet of Things.

Michail Matthaiou (S05M08SM13) was born


in Thessaloniki, Greece, in 1981. He obtained the
Diploma degree (5 years) in Electrical and Computer
Engineering from the Aristotle University of Thessaloniki, Greece in 2004. He then received the M.Sc.
(with distinction) in Communication Systems and
Signal Processing from the University of Bristol,
U.K. and Ph.D. degrees from the University of
Edinburgh, U.K. in 2005 and 2008, respectively.
From September 2008 through May 2010, he was
with the Institute for Circuit Theory and Signal
Processing, Munich University of Technology (TUM), Munich, Germany,
working as a Postdoctoral Research Associate. He is currently a Senior Lecturer at Queens University Belfast, U.K., and also holds an adjunct Assistant
Professor position at Chalmers University of Technology, Sweden. His research
interests span signal processing for wireless communications, massive MIMO,
hardware-constrained communications, and performance analysis of fading
channels.
Dr. Matthaiou is the recipient of the 2011 IEEE ComSoc Young Researcher
Award for the Europe, Middle East and Africa Region and a co-recipient
of the 2006 IEEE Communications Chapter Project Prize for the best M.Sc.
dissertation in the area of communications. He was an Exemplary Reviewer
for IEEE COMMUNICATIONS LETTERS for 2010. He has been a member of
Technical Program Committees for several IEEE conferences such as ICC,
GLOBECOM, VTC, etc. He currently serves as an Associate Editor for
the IEEE TRANSACTIONS ON COMMUNICATIONS, IEEE COMMUNICATIONS
LETTERS, and was the Lead Guest Editor of the special issue on Large-scale
multiple antenna wireless systems of the IEEE JOURNAL ON SELECTED
AREAS IN COMMUNICATIONS. He is an associate member of the IEEE Signal
Processing Society SPCOM and SAM technical committees.

S-ar putea să vă placă și