Active Simultaneously Transmitting and Reflecting Surface Assisted NOMA Networks
Active Simultaneously Transmitting and Reflecting Surface Assisted NOMA Networks
Abstract
The novel active simultaneously transmitting and reflecting surface (ASTARS) has recently received
a lot of attention due to its capability to conquer the multiplicative fading loss and achieve full-space
smart radio environments. This paper introduces the ASTARS to assist non-orthogonal multiple access
(NOMA) communications, where the stochastic geometry theory is used to model the spatial positions
of pairing users. We design the independent reflection/transmission phase-shift controllers of ASTARS
to align the phases of cascaded channels at pairing users. We derive new closed-form and asymptotic
expressions of the outage probability and ergodic data rate for ASTARS-NOMA networks in the presence
of perfect/imperfect successive interference cancellation (pSIC). The diversity orders and multiplexing
gains for ASTARS-NOMA are derived to provide more insights. Furthermore, the system throughputs
of ASTARS-NOMA are investigated in both delay-tolerant and delay-limited transmission modes. The
numerical results are presented and show that: 1) ASTARS-NOMA with pSIC outperforms ASTARS
assisted-orthogonal multiple access (ASTARS-OMA) in terms of outage probability and ergodic data
rate; 2) The outage probability of ASTARS-NOMA can be further reduced within a certain range by
X. Yue and J. Xie are with the Key Laboratory of Information and Communication Systems, Ministry of Information Industry
and also with the Key Laboratory of Modern Measurement & Control Technology, Ministry of Education, Beijing Information
Science and Technology University, Beijing 100101, China (email: {xinwei.yue and jin.xie}@bistu.edu.cn).
Chongjun Ouyang is with the School of Information and Communication Engineering, Beijing University of Posts and
Telecommunications, Beijing, 100876, China (e-mail: [email protected]).
Y. Liu is with the School of Electronic Engineering and Computer Science, Queen Mary University of London, London E1
4NS, U.K. (email: [email protected]).
X. Shen is with the China Academy of Information and Communications Technology (CAICT), Beijing 100191, China (email:
[email protected]).
Z. Ding is with the Department of Electrical Engineering And Computer Science, Khalifa University, Abu Dhabi, UAE, and
also with Department of Electrical Engineering, Princeton University, Princeton, USA (e-mail: [email protected]).
2
increasing the power amplification factors; 3) The system throughputs of ASTARS-NOMA are superior
to that of ASTARS-OMA in both delay-limited and delay-tolerant transmission modes.
Index terms— Active simultaneously transmitting and reflecting surface, non-orthogonal multiple
access, stochastic geometry.
I. I NTRODUCTION
As the number of device in wireless networks explosively grows, the sixth-generation (6G)
wireless networks are facing unprecedented challenges for providing high-speed, low-latency data
services for massive users [1, 2]. From the standpoint of expanding capacity, signal strength, and
coverage range, the reconfigurable intelligent surface (RIS) boasts remarkable capabilities. It
has been considered as one of promising technologies of 6G networks [3]. Essentially, passive
RIS (PRIS) is a planar surface comprising abundant inexpensive passive reflecting components,
which is able to modify the phase and amplitude of incident signals to achieve smart radio
environments [4]. In addition, PRIS has been shown to be able to improve the performance of
physical layer security [5], user localization [6] and unmanned aerial vehicle communications
[7].
Despite of the aforementioned advantages of PRIS, it only provides the half-space smart
transmissions [8, 9]. The innovative simultaneously transmitting and reflecting surfaces (STARS),
which can achieve full spatial coverage, was proposed to get around this restriction [10]. Specif-
ically, a passive STARS (PSTARS) integrates many passive simultaneously transmitting and
reflecting elements that can transmit and reflect the incident signals [11]. Based on hardware
architecture and physical principles, the authors in [12] further studied mode switching, energy
splitting and time switching protocols for PSTARS networks. For these three protocols, the
authors of [13] investigated the minimization of power consumption in PSTARS networks. The
impact of separate codebooks on PSTARS was examined in terms of detection capabilities and
system performance without complete channel state information (CSI) [14]. To enhance channel
conditions, the authors in [15] incorporated PSTARS into the over-the-air computation system,
enabling excellent learning accuracy and privacy perservation over large coverage areas. In [16],
the authors employed a diversity-preserving phase-shift strategy to attain complete diversity
order of PSTARS networks by taking into account coupled phase-shift models. As a further
development, the authors in [17] evaluated the secrecy capacity of PSTARS networks with
coupled phase-shift scheme.
3
Widespread interests have also been drawn to non-orthogonal multiple access (NOMA) which
is a potential multiple access method for the next generation of wireless communication networks
[18, 19]. NOMA has the ability to boost system throughput, capacity and energy efficiency
in comparison to orthogonal multiple access (OMA), delivering an improved communication
service for large numbers of users [20]. The notion of cooperative NOMA was introduced in
[21], where one cell-centred user is utilized as a relay to enhance the quality of service for an
edge user. Inspired by this work, the ergodic data rate and outage probability of full/half-duplex
cooperative NOMA networks were studied in [22]. The authors integrated PRIS into NOMA
networks [23], where the effect of stochastic discrete and coherent phase-shifting designs was
researched for PRIS-NOMA networks. With the focus on green communications, the authors of
[24] revealed the tradeoff between maximizing the sum rate and minimizing the power budget in
PRIS-assisted NOMA networks. Considering complexity expansion and error propagation issues,
the authors in [25] studied the ergodic data rate and outage performances of PRIS-NOMA with
perfect/imperfect successive interference cancellation (pSIC/ipSIC) schemes. Recently, a new
concept of near-field NOMA communication was introduced in [26], which benefits from the
beamforming characteristics of near-field to enable NOMA in both angular and distance domains.
Moreover, the authors of [27] utilized pre-configured spatial beams to serve both near-field and
far-field users, confirming that NOMA can effectively support the coexistence of near-field and
far-field communications.
As mentioned above, the integration of NOMA with other technologies is flourishing, and the
PSTARS-assisted NOMA (PSTARS-NOMA) networks naturally becomes a promising direction.
The superiority of NOMA related on the differentiated channel conditions among users [28],
and thus the establishment of channel condition differences was essential for NOMA. With
the help of PSTARS, the users can be deployed to different half-spaces with vastly disparate
channel conditions, thereby augmenting the performance of NOMA [29]. From the perspective of
performance analysis, the outage performance of PSTARS-NOMA was evaluated by utilizing the
central limit theorem and curve fitting model [30]. On the basis of these models, the authors of
[31] analyzed the ergodic data rate, outage behaviors and system throughput of PSTARS-NOMA
with pSIC/ipSIC schemes. The coverage characteristics of PSTARS networks were surveyed
in [32], where the coverage of PSTARS-NOMA can be significantly extended compared to
PSTARS-OMA. In the presence of Nakagami-m cascade channels, the secrecy outage probability
of PSTARS-NOMA was researched in [33] by considering the residual hardware impairments.
4
The authors of [34] researched a matching theory based channels allocation scheme to achieve
the maximum sum rate of PSATRS-NOMA systems. In [35], over-the-air federated learning and
PSTARS-NOMA were integrated into an unified framework, which achieves both high spectral
efficiency and learning performance.
While PRIS/PSTARS bring the enhanced performance of wireless networks, they also cause
multiplicative fading loss. Specifically, the small-scale fading of transmitter-PSTARS/PRIS link
and PSTARS/PRIS-receiver link were multiplied, which is usually worse than direct-link fading
[36]. To eliminate this effect, an active RIS (ARIS) with integrated reflection-type amplifiers has
been proposed [37], which magnifies the signals’ power, and then reflect to the desired users. The
simulation results demonstrated that the service area coverage and spectral efficiency of ARIS
were superior to those of PRIS [38]. Condition on the same power consumption, the authors of
[39] revealed that ARIS outperforms PRIS in terms of the achievable data rate if the number of
elements is small. In energy-constrained internet-of-things systems, ARIS-NOMA was proven to
achieve higher system throughput than ARIS-OMA [40]. A subarray-based ARIS structure was
designed to improve energy efficiency [41], where each subarray can be independently controlled.
Recently, a novel hardware model for active STARS (ASTARS) was proposed [42], which has
the ability to offset the multiplicative fading loss and achieve full-space coverage. In [43], the
authors confirmed that ASTRAS-aided communication systems outperform ARIS in terms of the
sum-rate improvement and power consumption reduction. Moreover, the maximum secrecy rate
of ASTARS assisted wireless networks was achieved by jointly optimizing the configuration of
elements and beamforming of access points [44].
As a new topic, only a few works have been researched for ASTARS networks, where the
hardware model design [42], sum rate maximisation [43] and system security [44] have been the
focus of the previous works. Since ASTARS is able to provide different channel differences and
amplify the desired signals for non-orthogonal users, the physical layer performance analysis
of ASTARS assisted NOMA networks is necessary to gain valuable insights. To the best of
our knowledge, the integration of ASTARS with NOMA networks have not been researched
yet, and the critical questions require further exploration. In particularly, considering the issues
of complexity scaling and error propagation, it is important to analyse the effect of ipSIC on
ASTARS-NOMA networks. The impact of the ASTARS elements’ configuration affects on the
5
can be further improved within a certain range by increasing the power amplification
factors.
The rest of this article is divided into the following sections. Section II presents the system
model of ASTARS-NOMA in terms of hardware architecture, network deployment, and channel
statistics. The outage probability expressions of ASTARS-NOMA are derived in Section III, in
which the diversity orders for Ur and Ut are provided. Section IV evaluates the ergodic data rate
of Ur and Ut . The simulation results and the corresponding analyses are presented in Section V.
Then the conclusions of this paper are given in Section VI, and Appendix contains a collection
of mathematical proofs.
The main symbols used in this article are as follows: The probability density function (PDF)
of a random variable X is denoted as fX (·), and its cumulative distribution function (CDF)
is denoted as FX (·). E{·} denotes the expectation and D{·} denotes variance operations; (·)H
stands for conjugate-transpose operation.
The complex channel coefficients from the BS to ASTARS, and then from ASTARS to Uφ are
denoted by hs ∈ CL×1 and hφ ∈ CL×1 with φ ∈ {r, t}, respectively. For practical considerations,
the ASTARS-NOMA networks’ wireless communication links undergo Rician fading. hH
r Θr hs
and hH
t Θt hs separately stand for the cascade complex channel coefficients from the BS to
√ r r r √
ASTARS, and then to Ur and Ut , where Θr = λβr diag ejθ1 , ..., ejθl , ..., ejθL = λβr Φr and
√ t t t √
Θt = λβt diag ejθ1 , ..., ejθl , ..., ejθL = λβt Φt denote the reflection and transmission phase-
shifting amplification matrixes of ASTARS, respectively. To facilitate analysis, assume that all
ASTARS elements have the same amplification factor λ and λ > 1. βr and βt are denoted by the
reflection and transmission amplitude coefficients, where βr + βt ≤ 1. θlr , θlt ∈ [0, 2π) represent
the transmission and reflection response’s phase-shift of the l-th element, respectively. Since the
transmission and reflection phase shifts are controlled by two different phase-shifters, θlr and
θlt can be tuned independently. The perfect CSI is required for the users to carry out coherent
demodulation.
ASTARS
Amplifier
Incident Signals +
hs -
ht
hr Amplified Signals
Ut
Phase Shifter Phase Shifter
BS
Reflected
Transmission Signals Transmitted
Ur D region Signals
Power Divider
Reflection
region
ASTARS An active element
(a) Network devices and users deployment. (b) Hardware structure of an ASTARS element.
thermal noise generated by active components. At this moment, the received signal expressions
at Ur and Ut can be separately written as
p
yr = hH
r Θr hs Psact XΣ + hH
r Θr ns + ñr , (1)
and
p
yt = hH
t Θt hs Psact XΣ + hH
t Θt ns + ñt , (2)
√ √
where XΣ = ar x r + at xt , Psact denote BS’s transmit power, xr and xt denote the signals
of Ur and Ut , respectively. ar and at stands for the power allocation factor of Ur and Ut ,
respectively. For the sake of fairness, ar and at satisfy the relation ar ≤ at and ar + at = 1. ns =
1 H
ns , · · · , nls , · · · , nLs is denoted by the thermal noise matrix generated by ASTARS elements
and nls ∼ CN (0, σs2 ). ñφ ∼ CN (0, σ02 ) stands for white Gaussian noise with average power
p 1 H 1 H
σ02 . Let hs = η0 d−α hs , · · · , hls , · · · , hLs , hφ = η0 d−α hφ , · · · , hlφ , · · · , hLφ
p
s φ denote the
channel coefficients from the BS to ASTARS, and then from ASTARS to Uφ , where hls =
p κ q q
1 l l l
p κ 1
κ+1
+ h̃
κ+1 s
, h̃φ ∼ CN (0, 1), hφ = κ+1
+ h̃l , h̃ls ∼ CN (0, 1), κ denotes the
κ+1 φ
Rician factor and α is path loss exponent, η0 expresses the path loss, ds stands for the distances
from BS to ASTARS, and dφ stand for the distance from ASTARS to Uφ . Ur has better channel
conditions and carried out the SIC to firstly detect the signal xt of Ut . Hence, the signal-plus-
interference-to-noise ratio (SINR) for Ur to decode xt can be expressed as
2
at λβr Psact hH
r Φr hs
γr→t = 2 2 2
. (3)
ar λβr Psact |hH H
r Φr hs | + λβr |hr Φr ns | + σ0
After decoding and deleting xt , the SNR for Ur to decode its own information can be expressed
as
2
ar λβr Psact hH
r Φr hs
γr = 2 2 act , (4)
λβr |hH
r Φr ns | + ε|hre | Ps + σ02
2
where hre ∼ CN (0, σre ) stands for the residual interference caused by ipSIC. In particular, ε= 0
stands for pSIC and ε= 1 denotes ipSIC, respectively.
Ut has weak channel conditions and thus regard xt and thermal noise as interference. The
SINR for detecting xt can be given by
2
at λβt Psact hH
t Φt hs
γt = 2 2 . (5)
ar λβt Psact |hH H 2
t Φt hs | + λβt |ht Φt ns | + σ0
9
given by [49]
∞ X
∞
X (κ + 1)u+v+2 xu+v+1
fXφl (x) = 4 Ku−v [2x (κ + 1)] , (6)
u=0 v=0
(u!)2 (v)2 e2κ κ−u−v
where Kx (·) indicates the modified Bessel function of the second kind with order x. We can
separately express the mean and variance of Xφl as
l
π h i2
E Xφ = L 1 (−κ) , (7)
4 (κ + 1) 2
and
l
π2 h i4
D Xφ = 1 − L (−κ) ,
1 (8)
16(κ + 1)2 2
x
where L 1 (x) = e 2 (1 − x) I0 − x2 − xI1 − x2 is the Laguerre polynomial.
2
By applying Laguerre polynomial series [50, Eq. (2.76)], the PDF and CDF for Xφ are
separately approximated as
pφ √
−1
x 2 −qx
fXφ (x) = p e φ , (9)
2qφφ Γ (pφ )
and
√
x
FXφ (x) = γ pφ , Γ(pφ )−1 , (10)
qφ
Rx
where γ (a, x) = 0
ta−1 e−t dt is the lower incomplete Gamma function [51, Eq. (8.350.1)] and
KE2 (X l ) D(X l )
Γ (·) is the gamma function [51, Eq. (8.310.1)], pφ = D X l φ , qφ = E Xφl .
( φ) ( φ)
φ 2
2) Thermal noise intensity: Since θl can only be phase-aligned with the hH φ Φφ hs , the phase
2 −α 2
PL l 2
in each term of hHφ Φ φ n s = η 0 d φ σ s l=1 hφ is considered to be randomly distributed. Let
PL l
H = l=1 hφ , since H is obtained by adding up L independent identically distributed hlφ , it can
p κ q
L
be calculated as H = L κ+1 + κ+1 H̃, where H̃ follows a complex Gaussian distribution and
H̃ ∼ CN (0, 1). We use the mean to characterize the channel power of H, which is expressed
as E |H|2 = L Lκ+1
κ+1
.
10
3) User’s location characteristics: For the path-loss experienced by the users, the PDFs of dφ
can be obtained by using the fact that the locations of paring users are stochastically distributed
within ASTARS’s serving area OD . In this case, the PDFs of dr and dt are written as [30]
Z xZ π
∂ 2r 2x
fdr (x) = 2
drdθ = 2 , (11)
∂x 0 0 πD D
and
Z x Z 2π
∂ 2r 2x
fdt (x) = 2
drdθ = 2 , (12)
∂x 0 π πD D
respectively.
In this section, the outage behaviors of ASTARS-NOMA are evaluated by invoking the
stochastic geometry. To be more specific, the closed-form outage probability expressions of
Ur with pSIC/ipSIC and Ut are derived for ASTARS-NOMA. To acquire further insight, the
asymptotic expressions of outage probability and diversity orders for Ur and Ut are obtained.
With the help of ASTARS, Ur , i.e., the user with strong channel conditions, needs to decode the
information of Ut , and then decode its own signals. As a consequence, the Ur outage occurrences
may be described as follows: 1) An outage event occurs when the signal xt of Ut cannot be
successfully decoded by Ur ; and 2) The signal xt is successfully decoded, while the signal xr of
Ur fails to detect. Based on these explanations, the outage probability at Ur in ASTARS-NOMA
networks is shown as
Pout,r = Pr (γr→t > γ̂t , γr < γ̂r ) + Pr (γr→t < γ̂t ) , (13)
where γ̂r = 2R̂r − 1 and γ̂t = 2R̂t − 1 separately represent the target SNR for decoding xr
and xt . The corresponding target rates of Ur and Ut are defined as R̂r and R̂t , respectively.
The outage probability expression of Ur with ipSIC for ASTARS-NOMA is illustrated in the
following theorem.
Theorem 1. Condition on at > γ̂t ar , the closed-form expression of Ur ’s outage probability with
ipSIC for ASTARS-NOMA is written as
11
( s )
K X U
σ02
2
ipSIC
X πAk (xu +1) p 1 γ̂r dαs σs χαu εyk Psact
Pout,r = 1 − x2u γ pr , ζ + 2 −2
+ ,
k=1 u=1
2U Γ (p r ) qr ar Psact η0 η0 βr λσre βr λ
(14)
(xu +1)D
where ε = 1, ζ = L Lκ+1 2u−1
κ+1
, xu = cos 2U
π , χu = 2
, yk is the k-th zero point of
(K!)2 yk
Laguerre polynomial LK (yk ) and the k-th weight is expressed as Ak = [LK+1 (yk )]2
. In addition,
a trade-off between complexity and accuracy is also guaranteed by the parameters K and U .
Remark 1. If at < γ̂t ar , and by substituting (3) and (4) into (13), the expression of Ur ’s outage
probability with ipSIC can be written as
σ2
ipSIC 2 ∂ 2
Pout,r = Pr hH r Φr hs ≥ act hH
r Φr ns + 0 , (15)
Ps βr λ
γ̂t
where ∂ = (at −γ̂t ar )
. Due to condition at < γ̂t ar , the right side of the inequality (15) is less than
zero, making the inequality always hold. At this time, the outage probability of Ur with ipSIC
will always equal to one.
Corollary 1. For case ε = 0, the closed-form expression of Ur ’s outage probability with pSIC
for ASTARS-NOMA is written as
" s #
U α 2
pSIC
X 1 γ̂r dαs χu σ0 σs2 π (xu +1) p
Pout,r = γ pr , act 2
+ ζ 1 − x2u . (16)
u=1
q r a r P s η 0 β r λ η0 2U Γ (p r )
The following can be applied to indicate the outage occurrence at Ut : the SINR of decoded
signal xt is lower the target SINR. The corresponding outage probability is shown as
Theorem 2. Condition on at > γ̂t ar , the closed-form expression of Ut ’s outage probability for
ASTARS-NOMA is written as
s
U −2 2
X ∂dαs qt yuα σ0 σs2 π (xu +1) p
Pout,t = γ pt , act 2
+ ζ 1 − x2u , (18)
u=1
P s η0 β t λ η0 2U Γ (p t )
γ̂t
where ∂ = (at −γ̂t ar )
. Similar to Remark 1, if at < γ̂t ar , the outage probability of Ut will always
equal to one.
12
Proof. By substituting (5) into (17), the outage probability expression of Ut is further expressed
as
2
" !#
2 hH
t Φt ns β −1 σ 2
Pout,t = Pr hH
t Φt hs <∂ + t act 0 . (19)
Psact Ps λ
By configuring the reflection phase-shift to align the phases of cascaded channels, the above
expression can be rewritten as
2
L
∂dαs ζσs2 dαt σ02
X
Pout,t = Pr hls hlt ≤ + . (20)
l=1
Psact η0 η02 βt λ
Proposition 1. When an outage occurs for at least one user in the system, it is considered as
a system outage event. Hence, the ASTARS-NOMA’s system outage probability with pSIC/ipSIC
is written as
PNAST ARS τ
OM A,τ = 1 − 1 − Pout,r (1 − Pout,t ) , (21)
ipSIC pSIC
where τ ∈ {ipSIC, pSIC}. Pout,r , Pout,r and Pout,t is given by (14), (16) and (18), respectively.
C. Diversity Analysis
The diversity order is an essential performance metric in wireless networks, which determines
the robustness and fading resistance of networks. Specifically, a system with a larger diversity
order means that the outage probability decays faster and it is more robust to fading [52],
particularly at high SNR. The analysis of diversity order provides a basis for optimizing networks
performance and designing more efficient diversity mechanisms. The expression of diversity order
is shown as
∞
log (Pout (Psact ))
Dorder = − act
lim , (22)
Ps →∞ log Psact
∞
where Pout (Psact ) denotes the asymptotic expression of outage probability within high SNR
region (Psact → ∞).
The expression for the asymptotic outage probability of Ur with ipSIC can be directly derived
from (14), and it is provided by the following corollary.
13
s !
K X U
ipSIC,∞
X πAk χu yk γ̂r dαs χαu σre
2
Pout,r = γ pr , , (23)
k=1 u=1
2U Γ (pr ) η02 ar βr λqr2
p
where χu = 1 − x2u (xu +1).
Remark 2. As can be observed that the outage probability of Ur with ipSIC is almost constant
under the assumptions of Psact → ∞, i.e., there is an error floor for the outage probability
achieved for Ur with ipSIC. By substituting (23) into (22), the diversity order of Ur with ipSIC
can be calculated as zero. This is attributed to the effect of residual interference generated by
the ipSIC scheme on networks outage performance.
For Ur with pSIC and Ut , the precise outage probability diversity orders can be calculated by
employing the Laplace transform.
Remark 3. The Ur ’s diversity order with pSIC can be calculated as L by substituting (24) into
(22), which is proportional to the number of ASTARS elements L. The Ur ’s asymptotic outage
probability under the pSIC scheme is an oblique line rather than a constant. This indicates that
the reflection user Ur with pSIC for ASTARS-NOMA is able to achieve a diversity order of L,
which is the maximal diversity for the considered scenario.
14
(25)
Proof. The following procedures resemble those in Appendix B.
Remark 4. The Ut ’s diversity order is calculated as L by substituting (25) into (22), which is
proportional to the numbers of ASTARS elements. This indicates that Ut of ASTARS-NOMA is
also able to achieve the full diversity order.
D. Delay-limited Transmission
Rτlimited = 1 − Pout,r
τ
R̂r + (1 − Pout,t ) R̂t , (26)
ipSIC pSIC
where Pout,r , Pout,r and Pout,t are obtained from (14), (16) and (18), respectively.
This section analyzes the ergodic data rate of Ut and Ur with pSIC/ipSIC to reveal the data
transmission rate and capability of ASTARS-NOMA networks. The definition of ergodic data
rate is shown as
which indicates that the high ergodic data rate achieves high channel capacity. On this basis, we
further derive the asymptotic expressions of ergodic data rate for Ur with pSIC/ipSIC and Ut .
Furthermore, the multiplexing gains for Ur and Ut are discussed in detail.
15
Assuming that Ur can effectively detect information xt by using the ipSIC method, the
expression of Ur ’s ergodic data rate with ipSIC is then obtained as the following theorem.
Theorem 3. Conditioned on the stochastic geometry model and cascade Rician fading channels,
a closed-form expression of Ur ’s ergodic data rate with ipSIC for ASTARS-NOMA is written as
h i
Q −1 2
XXX K U π ln 1 + (x q ϑ ) Ak Aq (xu +1)
erg
Rr,ipSIC = 1−pr , (28)
q=1 k=1 u=1
2U ln 2Γ (p r ) x q
r h i
1 dα σs2 χα εyk Psact σ02
where ϑ = qr ar P act ζ η0 + η2 β λσ−2 + βr λ
s u
and Q is the parameter that guarantee a
s 0 r re
Corollary 5. For case ε = 0, a closed-form expression of Ur ’s ergodic data rate with pSIC for
ASTARS-NOMA is written as
Q U
!
ar βr λ(xq qr η0 )2 Psact
p
erg
X X πAq (xu +1) 1 − x2u
Rr,pSIC = ln 1 + α α 2 . (29)
q=1 u=1
ds (χu σ0 + ζη0 βr λσs2 ) 2U ln 2Γ (pr ) xq1−pr
Theorem 4. Conditioned on the stochastic geometry model and cascade Rician fading channels,
a closed-form expression of Ut ’s ergodic data rate for ASTARS-NOMA is written as
( s
N p
2 U −2 2
1 − xn α
α 2
πat X X πχu y n ds qt yu σ0 σs
Rterg = 1− γ pt , act (a − y a ) 2
+ ζ ,
2N ar ln 2 n=1
1 + yn u=1
2U Γ (p t ) P s t n r η0 βt λ η0
(30)
(xn +1)at
and xn = cos 2u−1
where yn = 2ar 2U
π .
We now analyze the multiplexing gains of ASTARS-NOMA networks at high SNRs (Psact →
∞) to reveal the variation of the ergodic data rate with transmit power [54], which is defined as
erg
log (R∞ (Psact ))
S= lim , (31)
act
Ps →∞ log Psact
erg
where R∞ (Psact ) is the asymptotic expression of the ergodic data rate within high SNR areas.
16
1) The Ur ’s multiplexing gain with ipSIC: Based on (28), when Psact → ∞ the asymptotic
expression of Ur ’s ergodic data rate with ipSIC for ASTARS-NOMA is written as
Q K X U
erg,∞
X X 2 λβr ar πAk Aq (xu +1) pr −1
Rr,ipSIC = ln 1 + (η0 xq qr ) α α 2
xq . (32)
q=1 k=1 u=1
d s χu y k σre 2U ln 2Γ (p r )
Remark 5. By substituting (32) into (31), a multiplexing gain of zero for Ur with ipSIC is
obtained. It implies that even when the Psact is sufficiently high, the ergodic data rate will not
increase.
2) The Ur ’s multiplexing gain with pSIC: We obtain an upper bound on the ergodic data rate
by invoking Jensen’s inequality, which is written as
erg
Rr,pSIC = E [log2 (1 + γr )] ≤ log2 [1 + E (γr )] . (33)
Based on the above inequality, when Psact → ∞ an upper bound expression of Ur ’s ergodic
data rate with pSIC for ASTARS-NOMA networks is written as
( )
act l 2 l
bound ΞLλP s D X r + LE X r
Rr,erg = log 1 + α , (34)
ds [λβr η0 σs2 ζ (2 + α) + 2D2+α σ02 ]
where Ξ = η02 ar βr (2 + α).
Remark 6. The multiplexing gain for Ur with pSIC is equal to one by inserting (34) into (31),
which is due to the limitation of power allocation factors ratio at high SNRs.
3) The Ut ’s multiplexing gain: When Psact → ∞ the asymptotic ergodic data rate expression
of Ut for ASTARS-NOMA networks can be obtained directly from (30) as
N p
erg,∞ πat X 1 − x2n
Rt = . (35)
2N ar ln 2 n=1 1 + yn
Remark 7. The multiplexing gain for Ut is equal to zero, by inserting (35) into (31), which has
the same conclusion as in Remark 5.
C. Delay-tolerant Transmission
The data transmission is capped at the ergodic data rate in delay-tolerant transmission situations
because the codeword can experience all channel realizations [53]. In this scenario, the system
throughput of ASTARS-NOMA with pSIC/ipSIC schemes in delay-tolerated transmission mode
are defined as
Rϑtolerant = Rr,τ
erg
+ Rterg , (36)
17
erg erg
where Rr,ipSIC , Rr,pSIC and Rterg can be obtained from (28), (29) and (30).
V. S IMULATION R ESULTS
This section provide the computer simulation results to verify the correctness of the theoretical
formulas in Sections III and IV. The simulation settings used, unless otherwise specified, are
displayed in Table I. The complexity-accuracy trade-off parameters K, Q and U are set to 103 . To
highlight the performance of ASTARS-NOMA networks, the ARIS-NOMA, ASTARS-OMA and
PSTARS-NOMA networks are selected as benchmarks. In particular, the total power budgets of
ASTARS and PSTARS-aided networks are respectively given by Qact act act
tot = Ps +Pr +L (Pc + Pd )
A. Outage Probability
In Fig. 2(a), we plot the outage probability of Ur and Ut versus the system power budget.
The outage probability curves of Ut and Ur with pSIC/ipSIC curves are plotted by (14), (16)
18
100 10 0
U r -ipSIC with
2
re
= -90, -85, -80 (dBm)
10-1
Outage Probability
10 -5
Outage Probability
10-2
L=3
L=6
10-3
Error floor
Simulation
Asymptotic 10 -10 Error floor
10-4 ASTARS-OMA-Un Simulation
ASTARS-OMA-Um Asymptotic
L=9
Ut ASTARS-NOMA-U t
ASTARS-NOMA-Um
U r -pSIC ASTARS-NOMA-U r -pSIC
10-5 ASTARS-NOMA-Un-pSIC
ASTARS-NOMA-Un-ipSIC ASTARS-NOMA-U r -ipSIC
-15
0 5 10 15 20 25 30 35 40 45 50
10
-10 0 10 20 30 40
The system power budget Q tot (dBm) P sact (dBm)
(a) Outage probability versus system power budget Qtot . (b) Outage probability versus BS’s transmit power Psact .
and (18), respectively. This figure demonstrates that the obtained analytical expressions match
the simulation results exactly, which validates the accuracy of the analytical methods applied.
The blue dotted lines for asymptotic outage probability are ploted based on (23), (24) and (25),
respectively. They perfectly match the outage probability curves of ASTARS-NOMA within high
SNR region, proving that our asymptotic approach is accurate. One phenomenon is that the outage
performance of Ur with pSIC and Ut for ASTARS-NOMA outperforms that of ASTARS-OMA,
which is due to the following two reasons. 1) NOMA is able to achieve better fairness in outage
performance between paring users; and 2) The performance of ASTAR-NOMA can be further
enhanced by the better compatibility between ASTARS and NOMA. Another phenomenon is
that, after Qtot exceeds 30 dBm, the outage performance of Ur with ipSIC for ASTARS-NOMA
is worse than that of ASTARS-OMA. In high SNR area, it converges to an error floor. This
is attributed that the Ur in ASTARS-NOMA networks suffers from residual interference caused
by ipSIC, which confirms the conclusions made in Remark 2. Moreover, Fig. 2(b) displays the
outage probability of Ur and Ut for ASTARS-NOMA networks versus transmit power of BS
under L = {3, 6, 9}. The figure shows that as the number of ASTARS elements rises, the outage
probability falls and its slope rises. This phenomenon is caused by the fact that the diversity
orders of Ut and Ur with pSIC are proportional to the number of ASTARS elements, which
confirms the conclusions made in Remark 3 and Remark 4.
In Fig. 3(a), we plot the outage probability of ASTARS-NOMA versus number of ASTARS
19
100 100
ipSIC scheme
10-2 10-1
10-4
10-2
Outage Probability
Outage Probability
10-6
10-3 pSIC scheme
Error floor
-8
10 Simulation
ASTARS-NOMA-Ut
10-4 ASTARS-OMA
ASTARS-NOMA-Ur-pSIC ARIS-NOMA-pSIC
-10 ASTARS-NOMA-pSIC
10 ASTARS-NOMA-Ur-ipSIC
PSTARS-NOMA-Ut 10-5 PSTARS-NOMA-pSIC
ARIS-NOMA-ipSIC
-12 PSTARS-NOMA-Ur-pSIC
10 ASTARS-NOMA-ipSIC
PSTARS-NOMA-Ur-ipSIC PSTARS-NOMA-ipSIC
10-6
10 20 30 40 50 60 70 80 0 5 10 15 20 25 30 35 40 45 50
Number of elements L The system power budget Q tot (dBm)
(a) Outage probability versus number of ASTARS/ARIS elements(b) System outage probability versus system power budget Qtot .
pas
L, with Qact 2
tot = Qtot = 20 dBm, λ = 10 and σs = −30 dBm.
pas
elements L, with Qact 2
tot = Qtot = 20 dBm, λ = 10 and σs = −30 dBm. As can be observed
that with ASTARS elements increase, the outage probability first decreases and then gradually
increases. This trend can be attributed to the complex interactions between various factors. On
the one hand, the introduction of more ASTARS elements can enhance the spatial degrees of
freedom and thus reduce the outage probability. This is because that the increased degrees of
freedom enable a more efficient use of spatial domain, resulting in better signal quality and
stronger channel gains. On the other hand, using too many ASTARS elements can lead to a
large amount of thermal noise, which severely hinders the user’s ability to decode the signal.
This counteracts the channel gains generated by spatial degrees of freedom and leads to a surge in
outage probability. Hence, the optimization of ASTARS-NOMA networks is a balance between
the numbers of active component and the spatial degrees of freedom, with the ultimate goal of
minimizing outage probability.
Fig. 3(b) displays the system outage probability of ASTARS-NOMA and different benchmarks
versus the system power budget. As can be shown that the outage performance of ASTARS-
NOMA perform better than the PSTARS-NOMA. This can be interpreted that the ASTARS
elements allocate a portion of power budget to amplify the input radio signals, which enhances
the received SNR at paring users. This also confirms that ASTARS is an effective technology for
combating multiplicative fading loss. One occurrence is that ASTARS-NOMA achieves better
20
outage performance than ARIS-NOMA. In addition, the outage curves of ASTARS-NOMA are
steeper than the OMA ones. This can be explained by the fact that ASTARS is able to achieve
a better diversity order than ARIS.
100
Simulation
ASTARS-NOMA-pSIC
10-1 ASTARS-NOMA-ipSIC
100
ARIS-NOMA-pSIC
10-2
Outage Probability
Outage Probability
10-2 D = 45 (m)
10-3
10-4 D = 35 (m)
10-4
D = 25 (m)
10-5
0.8
0.8
0.6 ASTARS-NOMA-pSIC
0.6
0.4 10-6
ar 0.4
0.2 r
0.2
10 20 30 40 50 60 70 80 90 100
Amplification factor
(a) System outage probability versus reflection amplitude coeffi-(b) Outage probability versus amplification factor λ, with Psact =
cient and power allocation factor, with Psact = 20 dBm. 25 dBm and σs2 = −50 dBm.
In Fig. 4(a), we show the system outage probability of ASTARS-NOMA versus reflection
amplitude coefficient and power allocation factor under Qact = 20 dBm. The curved surface for
outage probability of ASTARS-NOMA is plotted according to (21). Due to the effects of both
amplitude coefficients and power allocation factors, the curved surface for outage probability
of ASTARS-NOMA takes on a valley-like shape. This suggests the existence of a solution that
satisfies the minimum system outage probability for the given parameter. Additionally, it can be
shown that by adjusting ar , the ARIS-NOMA’s outage probability can be reduced. However, in
most cases of parameter setting, the outage performance of ARIS-NOMA is inferior to that of
ASTARS-NOMA. This is because that ASTARS can introduce more spatial degrees of freedom
for NOMA networks, which enhances the outage performance. It is worth noting that when with
some extreme parameter choices, such as βr = 0.1 and ar = 0.1, the performance of ASTARS-
NOMA networks may be even worse than that of ARIS-NOMA networks. This indicates that
optimizing the amplitude coefficients and power allocation factors is essential to reducing the
outage probability of ASTARS-NOMA.
In Fig. 4(b), we plot the system outage probability of ASTARS-NOMA versus amplification
21
factor λ, with Psact = 25 dBm and σs2 = −50 dBm. One phenomenon is that the outage probability
of ASTARS-NOMA first reduces dramatically as the amplification factor steadily rises and then
tends to stabilize. This is due to the fact that the larger amplification factors helps to improve
the users’ received SNR, thus enhancing the outage performance. However, while increasing the
received signal strength at the users, also introduces a large amount of thermal noise, which
interferes with the decoding of user signals. As the amplification factor increases, a balance is
achieved between the gain of the enhanced signal and the loss of the enhanced noise such that
the outage probability remains constant. Another observation is that reducing the deployment
range of users improves outage performance. This is because that a smaller deployment range
diminishes the effect of path loss on the ASTARS-NOMA’s outage probability.
In Fig. 5(a), we present ergodic data rate of ASTARS-NOMA and PSTARS-NOMA versus
system power budget with ar = 0.2, at = 0.8. The ergodic data rate curves for ASTARS-NOMA
networks are drawn from (28), (29) and (30), respectively. According to (32), (34) and (35),
the asymptotic ergodic data rates are illustrated. This figure indicates that the Ut ’s ergodic data
rate converges towards the upper limit of the throughput, resulting in zero high SNR slope.
The ergodic data rate of ipSIC stops increasing with a rise in transmit power at high SNRs due
to the effects of residual interference, which in accordance with the discussion in Remark 5.
One phenomenon is that the Ur ’s ergodic data rates with pSIC/ipSIC of ASTARS-NOMA are
higher than those of PSTARS-NOMA. This is due to the fact that ASTARS is able to increase
the strength of users’ received signals, which further increase the average data transmission rate
of networks over an extended period. Another phenomenon is that the ergodic data rates of
ASTARS-NOMA with a larger power amplification factor is more efficient. This suggests that
increasing the power amplification factor can enhance the receiving SNR and thus improve the
ergodic data rate of ASTARS-NOMA.
Fig. 5(b) compares the ergodic data rates of ASTARS-NOMA with ARIS-NOMA and ASTARS-
OMA. This figure indicates that the Ur ’s ergodic data rate with pSIC/ipSIC of ASTARS-NOMA
outperform that of ARIS-NOMA and ASTARS-OMA. For the OMA transmission, it takes twice
as long to serve two users as NOMA transmission. As a result, the slope of OMA transmission is
only half of that of NOMA transmission, which is the reason for its lower ergodic data rate. The
reason why ARIS-NOMA networks have lower ergodic data rate compared to ASTARS-NOMA
22
networks is that they cannot provide the same spatial degrees of freedom as ASTARS networks
do. Another phenomenon is that the ergodic data rate of Ut for ASTARS-NOMA outperforms
ARIS-NOMA at low SNRs, while they reach the same upper limit of rate within high SNR
region. This can be explained by using the conclusion of Remark 7 that the upper limit of Ut ’s
ergodic data rate in the NOMA network is related to the power allocation factors.
Additionally, Fig. 6 plots the ergodic data rate of ASTARS-NOMA networks versus system
power budget with different path loss exponents, i.e., α. In ASTARS-NOMA networks, α is a
parameter to describe the signal power attenuation as it propagates through the wireless channels.
A few real-world channel models are to be adopted, depending on the choice of α. For example,
α = 2 denotes the free space propagation case, α = 2.5 denotes the scenario with obstacles and
α = 3 denotes the urban cellular networks. One can observe that as α increases, the channel
conditions of ASTARS-NOMA networks become progressively worse, resulting in a deterioration
of outage performance. This indicates that a reasonable α should be selected when studying the
ASTARS-NOMA networks in different practical scenarios.
6 6
Simulation Simulation
Throughput ceiling Throughput ceiling
ASTARS-NOMA-Ut ARIS-NOMA-Ut
5 PSTARS-NOMA-Ut 5 ASTARS-OMA-Ut
ASTARS gain
ASTARS-NOMA-Ur -pSIC ASTARS-NOMA-Ut
Ergodic Data Rate (BPCU)
2 2
=5
1 1
0 0
-5 0 5 10 15 20 25 30 35 40 45 -5 0 5 10 15 20 25 30 35 40 45
The system power budget Q tot (dBm) The system power budget Q tot (dBm)
(a) Ergodic data rate versus system power budget Qtot with ar =(b) Ergodic data rate versus system power budget Qtot with ar =
0.2, at = 0.8. 0.2, at = 0.8.
C. System Throughput
In Fig. 7(a), we present the system throughput of ASTARS-NOMA with pSIC/ipSIC versus
system power budget in the delay-limited transmission mode. According to (26), the system
23
6
Simulation
Throughput ceiling
ASTARS-NOMA-Ut
5
ASTARS-NOMA-Ur-pSIC
ASTARS-NOMA-Ur-ipSIC
3 =2
2
= 2.5
1
=3
0
-5 0 5 10 15 20 25 30 35 40 45
The system power budget Q tot (dBm)
Fig. 6: Ergodic data rate versus system power budget Qtot with ar = 0.2, at = 0.8.
2 8
Simulation
Throughput ceiling
1.8
7 ASTARS-NOMA-pSIC
ipSIC scheme ASTARS-NOMA-ipSIC
1.6 pSIC scheme
System Throughput (BPCU)
5
1.2
1 4
0.8
3
ASTARS-OMA
0.6
ARIS-NOMA-pSIC
2
ASTARS-NOMA-pSIC 2
0.4 = -90, -80, -70 (dBm)
PSTARS-NOMA-pSIC re
ARIS-NOMA-ipSIC 1
0.2
ASTARS-NOMA-ipSIC
PSTARS-NOMA-ipSIC
0 0
0 5 10 15 20 25 30 35 40 45 50 -5 0 5 10 15 20 25 30 35 40 45
The system power budget Q tot (dBm) The system power budget Q tot (dBm)
throughput curves of ASTARS-NOMA with pSIC/ipSIC are drawn. One phenomenon is that
the system throughput of ASTARS-NOMA outperform other comparison baselines under pSIC
scheme. This is attributed to the fact that outage probability under delay-limited transmission
model determines the throughput of ASTARS-NOMA networks. It can also be seen that the
NOMA networks with ipSIC fail to reach the target rate even when the transmit power is large.
This is due to the fact that residual interference limits the performance gains from increasing
transmit power at high SNRs.
Fig. 7(b) shows the system throughput of ASTARS-NOMA versus system power budget in
24
the delay-tolerant transmission model, with ar = 0.2, at = 0.8 and α = 2.3. According to
(36), the system throughput curves of ASTARS-NOMA with pSIC/ipSIC schemes are shown.
One phenomenon is that the system throughput ceiling of ASTARS-NOMA under ipSIC scheme
increases as residual interference strength diminishes. Another phenomenon is that reducing the
noise intensity generated by active devices can improve the system throughput of ASTARS-
NOMA with pSIC. This indicates that the design of low-power and low-interference hardware
architecture is essential to improve the performance of ASTARS-NOMA.
VI. C ONCLUSION
In this article, we have studied the novel ASTARS-NOMA networks with randomly deployed
users, which can mitigate the multiplicative fading loss and achieve full-space smart radio
environments. Specifically, we have obtained analytical expressions of outage probability and
ergodic data rate for ASTARS-NOMA networks with pSIC/ipSIC scheme. To gain further
insights, the asymptotic expressions of outage probability and ergodic data rate were also obtained
within high SNR region. On this basis, we have analyzed the diversity orders and multiplexing
gains for paring users. Simulation results demonstrated that the performance of ASTARS-NOMA
outperforms the PSTARS-NOMA, ARIS-NOMA and ASTARS-OMA for the same power con-
sumption. For hardware configuration, the ASTARS-NOMA networks require an appropriate
power amplification factors and number of ASTARS elements to ensure that the thermal noise
interference is within a reasonable range.
The top of the next page displays the expression of Ur ’s outage probability with ipSIC by
substituting (3) and (4) into (13). After combining the probability events, (A.1) can be simplified
to
!#
ipSIC
h
2 γ̂r 2 ε|hre |2 act σ2
Pout,r = Pr hH
r Φr hs ≤ hH
r Φr ns + Ps + 0 . (A.2)
ar Psact βr λ βr λ
With help of the topics discussed in Section II-B2, Ur ’s outage probability with ipSIC can be
further calculated as
α 2 2
2
ipSIC
XL γ̂r ds η0 σs α ε 2 σ0
Pout,r = Pr hls hlr ≤ ζ + d r |hre | + , (A.3)
| l=1
{z } ar η02 Psact |{z} βr λ | {z } βr λPsact
Z Y
Xr
25
2
" #
ipSIC at λβr Psact hH
r Φr hs
Pout,r = Pr 2 2 2
≤ γ̂t
ar λβr Psact |hH H
r Φr hs | + λβr |hr Φr ns | + σ0
2 2
" #
at λβr Psact hH
r Φr hs ar λβr Psact hH
r Φr hs
+ 2 2 2
> γ̂t , 2 2 act ≤ γ̂r .
ar λβr Psact |hH H
r Φr hs | + λβr |hr Φr ns | + σ0 λβr |hH
r Φr ns | + ε|hre | Ps + σ02
(A.1)
Lκ+1
where ζ = L κ+1
.
y
1 − σre
2
The PDF of Y can be express as fY (y) = 2
σre
e , and Xr ’s CDF and Z’s PDF denoted by
(10) and (11), respectively. Combining (10), (11) and PDF of Y , (A.3) can be converted into
integral form as
( s )
∞ D
γ̂r dα σ02
2
z α εyPsact
Z Z
ipSIC 1 − σy2 2z 1 s σs
Pout,r = 2
e re 2 γ pr , ζ + 2 + dzdy, (A.4)
0 0 σre D Γ (pr ) qr ar Psact η0 η0 βr λ βr λ
By applying Gauss-Chebyshev quadrature [56, Eq. (8.8.4)], the definite integral of above
expression can be calculated as
( s )
U Z
π X ∞ − σy2 (xu +1) p γ̂r dα σ02
2
χα εyPsact
ipSIC 1 s σs u
Pout,r = e re 2 1 − x2u γ pr , ζ + 2 + dy. (A.5)
2U u=1 0 σre Γ (pr ) qr ar Psact η0 η0 βr λ βr λ
It can be seen that the above equation contains the term e−at and the limits of integration
are 0 to infinity. This type of integral equation can be calculated by applying Gauss-Laguerre
R∞ K
quadrature formula [56, Eq. (8.6.5)], i.e., 0 e−at f (t) dt = a1 Ak f xak . After some algebraic
P
k=1
manipulations, we can obtain (14). The proof is complete.
To obtain the accurate asymptotic outage probability, the Laplace transform is applied in the
following proof process. The Laplace transform formula of the PDF for Xφl is calculated by [51,
Eq. (6.621.3)] as
∞ X ∞
h i √ X (1 + κ)2(u+1) 4u−v+1 Γ (2 + 2u) Γ (2 + 2v)
L fXφl (x) (s) = π
(u!) (v!) e [s + 2 (κ + 1)]2u+2 Γ u + v + 25
2 2 2κ
κ −u−v
u=0 v=0
1 5 −2 (κ + 1) + s
×2 F1 2 + 2u, + u − v; + u + v; . (B.1)
2 2 2 (κ + 1) + s
26
When Psact → ∞, s in the above equation goes to infinity. At the same time, the first term
(i = 0, j = 0) of above series dominates the whole expression, thus the Laplace transform is
eventually simplified as
16(1 + κ)2
h i 1 5
L fXφl (x) (s) =2 F1 2, ; ; 1 . (B.2)
2 2 3e2κ s2
Xφ = Ll=1 Xφl and by applying the convolution theorem, the Laplace transform for
p P
As
p
the PDF of Xφ can be given by
" #L
16(1 + κ)2 −2
0+ 1 5
L f√ (x) (s) = 2 F1 2, ; ; 1 s . (B.3)
Xφ 2 2 3e2κ
After using the inverse Laplace transform, the above expression can be derived as
2 L
" #
2L−1
0+ x 1 5 16(1 + κ)
f√ (x) = 2 F1 2, ; ; 1 . (B.4)
Xφ (2L − 1)! 2 2 3e2κ
R √x
Furthermore, by applying equation Fχ (x) = 0 fχ (x) dx, the approximated CDF of Xφ at
high SNRs can be finally expressed as
L
Λ L xL
1 5
FX0+φ (x) = 2 F1 2, ; ; 1 . (B.5)
(2L)! 2 2
Combining (B.5) and (A.3), (24) can be obtained after using Gauss-Laguerre quadrature and
Gauss-Chebyshev quadrature. The proof is complete.
By taking the derivative of above equation, we can obtain the PDF expression for γr as
U X K p
X ϑAk (xu +1) 1 − x2u √ pr −1
fγr (x) = π √ √ ϑ x , (C.2)
u=1 k=1
2U Γ (pr ) 2 xeϑ x
r h 2 i
1 dα σs χα εyk Psact σ02
where ϑ = qr ar P act ζ η0 + η2 β λσ−2 + βr λ .
s u
s 0 r re
By substituting (4) into (27), the ergodic data rate expression of Ur with ipSIC scheme is
calculated as
2
" !#
∞
ar λβr Psact hH
s Φr hr
Z
erg 1
Rr,ipSIC = log2 1+ 2 2 act = ln (1 + x) fγr (x)dx.
λβr |nH
s Φr hr | + ε|hre | Ps + σ02 ln 2 0
(C.3)
27
By substituting (C.2) into (C.3), the expression of Ur ’s ergodic data rate with ipSIC for
ASTARS-NOMA networks can be given by
Z ∞X K X U p
erg πAk (xu +1) ln (1 + x2 ) 1 − x2u
Rr,ipSIC = dx. (C.4)
0 k=1 u=1
2U ln 2Γ (pr ) eϑx ϑ−pr x1−pr
Also by using Gauss-Laguerre quadrature, (28) can be obtained. The proof is complete.
at
Different from (C.1), the x in (D.1) needs to satisfy the inequality x < ar
. Thus the ergodic
data rate expression of Ut is calculated as
Z at Z at
erg 1 ar 1 ar 1
Rt = fγt (x) ln (1 + x)dx = [1 − Fγt (x)]dx. (D.2)
ln 2 0 ln 2 0 1 + x
By substituting (D.1) into (D.2), the expression of Ut ’s ergodic data rate for ASTARS-NOMA
networks is given by
at ( " s #)
Z U α
α 2 2
1 ar 1 X π (xu +1) p 1 xd y σ σ
Rterg = 1− 1 − x2u γ pt , r
act
u 0
2β λ + ζ η
s
dx.
ln 2 0 1+x u=1
2U Γ (pt ) qt (at − xa r ) P s η 0 t 0
(D.3)
R EFERENCES
[1] Z. Zhang, Y. Xiao, Z. Ma, M. Xiao, Z. Ding, X. Lei, G. K. Karagiannidis, and P. Fan, “6G wireless networks: Vision,
requirements, architecture, and key technologies,” IEEE Veh. Technol. Mag., vol. 14, no. 3, pp. 28–41, Sept. 2019.
[2] M. Giordani, M. Polese, M. Mezzavilla, S. Rangan, and M. Zorzi, “Toward 6G networks: Use cases and technologies,”
IEEE Commun. Mag., vol. 58, no. 3, pp. 55–61, Mar. 2020.
[3] E. Basar, M. Di Renzo, J. De Rosny, M. Debbah, M.-S. Alouini, and R. Zhang, “Wireless communications through
reconfigurable intelligent surfaces,” IEEE Access, vol. 7, pp. 116 753–116 773, Aug. 2019.
[4] Q. Wu and R. Zhang, “Towards smart and reconfigurable environment: Intelligent reflecting surface aided wireless network,”
IEEE Commun. Mag., vol. 58, no. 1, pp. 106–112, Jan. 2020.
[5] L. Yang, J. Yang, W. Xie, M. O. Hasna, T. Tsiftsis, and M. D. Renzo, “Secrecy performance analysis of RIS-aided wireless
communication systems,” IEEE Trans. Veh. Technol., vol. 69, no. 10, pp. 12 296–12 300, Oct. 2020.
[6] K. Keykhosravi, M. F. Keskin, S. Dwivedi, G. Seco-Granados, and H. Wymeersch, “Semi-passive 3D positioning of
multiple RIS-enabled users,” IEEE Trans. Veh. Technol., vol. 70, no. 10, pp. 11 073–11 077, Oct. 2021.
28
[7] L. Yang, F. Meng, J. Zhang, M. O. Hasna, and M. D. Renzo, “On the performance of RIS-assisted dual-hop UAV
communication systems,” IEEE Trans. Veh. Technol., vol. 69, no. 9, pp. 10 385–10 390, Sept. 2020.
[8] Y. Liu, X. Liu, X. Mu, T. Hou, J. Xu, M. D. Renzo, and N. Al-Dhahir, “Reconfigurable intelligent surfaces: Principles
and opportunities,” IEEE Commun. Surveys Tutorials, vol. 23, no. 3, pp. 1546–1577, Thirdquarter 2021.
[9] M. Zeng, X. Li, G. Li, W. Hao, and O. A. Dobre, “Sum rate maximization for IRS-assisted uplink NOMA,” IEEE Commun.
Lett., vol. 25, no. 1, pp. 234–238, Jan. 2021.
[10] H. Zhang, S. Zeng, B. Di, Y. Tan, M. D. Renzo, M. Debbah, Z. Han, H. V. Poor, and L. Song, “Intelligent omni-surfaces
for full-dimensional wireless communications: Principles, technology, and implementation,” IEEE Commun. Mag., vol. 60,
no. 2, pp. 39–45, Feb. 2022.
[11] H. Zhang and B. Di, “Intelligent omni-surfaces: Simultaneous refraction and reflection for full-dimensional wireless
communications,” IEEE Commun. Surveys Tutorials, vol. 24, no. 4, pp. 1997–2028, Fourthquarter. 2022.
[12] Y. Liu, X. Mu, J. Xu, R. Schober, Y. Hao, H. V. Poor, and L. Hanzo, “STAR: Simultaneous transmission and reflection
for 360o coverage by intelligent surfaces,” IEEE Wireless Commun., vol. 28, no. 6, pp. 102–109, Dec. 2021.
[13] X. Mu, Y. Liu, L. Guo, J. Lin, and R. Schober, “Simultaneously transmitting and reflecting (STAR) RIS aided wireless
communications,” IEEE Trans. Wireless Commun., vol. 21, no. 5, pp. 3083–3098, May. 2022.
[14] Y. Zhang, B. Di, H. Zhang, M. Dong, L. Yang, and L. Song, “Dual codebook design for intelligent omni-surface aided
communications,” IEEE Trans. Wireless Commun., vol. 21, no. 11, pp. 9232–9245, Fourthquarter. 2022.
[15] X. Zhai, G. Han, Y. Cai, Y. Liu, and L. Hanzo, “Simultaneously transmitting and reflecting (STAR) RIS assisted over-the-air
computation systems,” IEEE Trans. Veh. Technol., vol. 71, no. 3, pp. 1309–1322, Mar. 2023.
[16] J. Xu, Y. Liu, X. Mu, R. Schober, and H. V. Poor, “STAR-RISs: A correlated T&R phase-shift model and practical
phase-shift configuration strategies,” IEEE Trans. Wireless Commun., vol. 16, no. 5, pp. 1097–1111, Aug. 2022.
[17] Z. Zhang, Z. Wang, Y. Liu, B. He, L. Lv, and J. Chen, “Security enhancement for coupled phase-shift STAR-RIS networks,”
IEEE Trans. Veh. Technol., Early Access, 2023.
[18] W. U. Khan, J. Liu, F. Jameel, V. Sharma, R. Jantti, and Z. Han, “Spectral efficiency optimization for next generation
NOMA-enabled IoT networks,” IEEE Trans. Veh. Technol., vol. 69, no. 12, pp. 15 284–15 297, Dec. 2020.
[19] Y. Liu, S. Zhang, X. Mu, Z. Ding, R. Schober, N. Al-Dhahir, E. Hossain, and X. Shen, “Evolution of NOMA toward next
generation multiple access (NGMA) for 6G,” IEEE J. Sel. Areas Commun., vol. 40, no. 4, pp. 1037–1071, Apr. 2022.
[20] Y. Yuan, Y. Wu, Z. Ding, X. You, H. V. Poor, and L. Hanzo, “NOMA for next-generation massive IoT: Performance
potential and technology directions,” IEEE Commun. Mag., vol. 59, no. 7, pp. 115–121, Jul. 2021.
[21] Z. Ding, M. Peng, and H. V. Poor, “Cooperative non-orthogonal multiple access in 5G systems,” vol. 19, no. 8, pp.
1462–1465, Aug. 2015.
[22] X. Yue, Y. Liu, S. Kang, A. Nallanathan, and Z. Ding, “Exploiting full/half-duplex user relaying in NOMA systems,”
vol. 66, no. 2, pp. 560–575, Feb. 2018.
[23] Z. Ding, R. Schober, and H. V. Poor, “On the impact of phase shifting designs on IRS-NOMA,” vol. 9, no. 10, pp.
1596–1600, Oct. 2020.
[24] F. Fang, Y. Xu, Q.-V. Pham, and Z. Ding, “Energy-efficient design of IRS-NOMA networks,” IEEE Trans. Veh. Technol.,
vol. 69, no. 11, pp. 14 088–14 092, Nov. 2020.
[25] X. Yue and Y. Liu, “Performance analysis of intelligent reflecting surface assisted NOMA networks,” vol. 21, no. 4, pp.
2623–2636, Apr. 2022.
[26] J. Zuo, X. Mu, and Y. Liu, “Non-orthogonal multiple access for near-field communications,” May. 2023. [Online].
Available: https://arxiv.org/abs/2304.13185v2
29
[27] Z. Ding, R. Schober, and H. V. Poor, “NOMA-based coexistence of near-field and far-field massive MIMO communications,”
IEEE Wireless Commun. Lett., Early Access, 2023.
[28] S. M. R. Islam, N. Avazov, O. A. Dobre, and K. sup Kwak, “Power-domain non-orthogonal multiple access (NOMA) in
5G systems: Potentials and challenges,” IEEE Commun. Surveys Tutorials, vol. 19, no. 2, pp. 721–742, Secondquarter.
2017.
[29] M. Aldababsa, A. Khaleel, and E. Basar, “STAR-RIS-NOMA networks: An error performance perspective,” IEEE Commun.
Lett., vol. 26, no. 8, pp. 1784–1788, Aug. 2022.
[30] C. Zhang, W. Yi, Y. Liu, Z. Ding, and L. Song, “STAR-IOS aided NOMA networks: Channel model approximation and
performance,” IEEE Trans. Wireless Commun., vol. 21, no. 9, pp. 6861–6876, Sept. 2022.
[31] X. Yue, J. Xie, Y. Liu, R. Liu, Z. Han, and Z. Ding, “Simultaneously transmitting and reflecting reconfigurable intelligent
surface assisted NOMA networks,” IEEE Trans. Wireless Commun., vol. 22, no. 1, pp. 189–204, Jan. 2023.
[32] C. Wu, X. Mu, X. Gu, and O. A. Dobre, “Coverage characterization of STAR-RIS networks: NOMA and OMA,” IEEE
Commun. Lett., vol. 25, no. 9, pp. 3036–3040, Sept. 2021.
[33] X. Li, Y. Zheng, M. Zeng, Y. Liu, and O. A. Dobre, “Enhancing secrecy performance for STAR-RIS NOMA networks,”
IEEE Trans. Veh. Technol., vol. 72, no. 2, pp. 2684–2688, Feb. 2023.
[34] C. Wu, X. Mu, Y. Liu, X. Gu, and X. Wang, “Resource allocation in STAR-RIS-aided networks: OMA and NOMA,” IEEE
Trans. Wireless Commun., vol. 21, no. 9, pp. 7653–7667, Sept. 2022.
[35] W. Ni, Y. Liu, Y. C. Eldar, Z. Yang, and H. Tian, “STAR-RIS integrated no-northogonal multiple access and over-the-air
federated learning: Framework, analysis, and optimization,” IEEE Internet Things J., vol. 9, no. 18, pp. 17 136–17 156,
Sept. 2022.
[36] Z. Zhang, L. Dai, X. Chen, C. Liu, F. Yang, R. Schober, and H. V. Poor, “Active RIS vs. passive RIS: Which will prevail
in 6G?” IEEE Trans. Commun., vol. 71, no. 3, pp. 1707–1725, Mar. 2023.
[37] Y. P. R. Long, Y.-C. Liang and E. G. Larsson, “Active reconfigurable intelligent surface aided wireless communications,”
IEEE Trans. Wireless Commun., vol. 20, no. 8, pp. 4962–4975, Aug. 2021.
[38] E. Basar and H. V. Poor, “Present and future of reconfigurable intelligent surface-empowered communications,” IEEE
Signal Process. Mag., vol. 38, no. 6, pp. 146–152, Nov. 2021.
[39] K. Zhi, C. Pan, H. Ren, K. Chai, and M. Elkashlan, “Active RIS versus passive RIS: Which is superior with the same
power budget?” IEEE Commun. Lett., vol. 26, no. 5, pp. 1150–1154, May. 2022.
[40] G. Chen, Q. Wu, C. He, W. Chen, J. Tang, and S. Jin, “Active IRS aided multiple access for energy-constrained IoT
systems,” IEEE Trans. Wireless Commun., vol. 22, no. 3, pp. 1677–1694, Mar. 2023.
[41] Y. Zhu, Y. Liu, M. Li, Q. Wu, and Q. Shi, “A flexible design for active reconfigurable intelligent surfaceła sub-array
architecture,” IEEE Trans. Veh. Technol., Early Access, 2023.
[42] J. Xu, J. Zuo, J. T. Zhou, and Y. Liu, “Active simultaneously transmitting and reflecting (STAR)-RISs: Modelling and
analysis,” Feb. 2023. [Online]. Available: https://arxiv.org/abs/2302.04432
[43] Y. Ma, M. Li, Y. Liu, Q. Wu, and Q. Liu, “Optimization for reflection and transmission dual-functional active RIS-assisted
systems,” Early Access, 2023.
[44] Y. Guo, Y. Liu, Q. Wu, Q. Shi, and Y. Zhao, “Enhanced secure communication via novel double-faced active RIS,” IEEE
Trans. Wireless Commun., vol. 71, no. 6, pp. 3497– 3512, Jun. 2023.
[45] K. K. Kishor and S. V. Hum, “An amplifying reconfigurable reflectarray antenna,” IEEE Trans. Antennas Propag., vol. 60,
no. 1, pp. 197–205, Jun. 2012.
[46] J. Bousquet, S. Magierowski, and G. G. Messier, “A 4-GHz active scatterer in 130-nm CMOS for phase sweep amplify-
and-forward,” IEEE Trans. Circuits Syst. I, Reg. Papers, vol. 59, no. 3, pp. 529–540, Mar. 2012.
30
[47] J. Loncar, Z. Sipus, and S. Hrabar, “Ultrathin active polarizationselective metasurface at X-band frequencies,” Physical
Review B, vol. 100, no. 7, p. 075131, Oct. 2019.
[48] M. Jereminov, A. Pandey, D. M. Bromberg, X. Li, G. Hug, and L. Pileggi, “Steady-state analysis of power system
harmonics using equivalent split-circuit models,” IEEE PES Innovative Smart Grid Tech. Conf. Europe (ISGT-Europe), pp.
1–6, Ljubljana, Slovenia, Otc. 2016.
[49] M. K. Simon, Probability Distributions Involving Gaussian Random Variables. Springer US, 2006.
[50] S. Primak, V. Kontorovich, and V. Lyandres, Stochastic Methods and their Applications to Communications: Stochastic
Differential Equations Approach, West Sussex, U.K.: Wiley, 2004.
[51] I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series and Products, 6th ed. New York, NY, USA: Academic
Press, 2000.
[52] J. Laneman, D. Tse, and G. Wornell, “Cooperative diversity in wireless networks: Efficient protocols and outage behavior,”
vol. 50, no. 12, pp. 3062–3080, Dec. 2004.
[53] C. Zhong, H. A. Suraweera, G. Zheng, I. Krikidis, and Z. Zhang, “Wireless information and power transfer with full
duplex relaying,” IEEE Trans. Commun., vol. 62, no. 10, pp. 3447–3461, Oct. 2014.
[54] L. Zheng and D. N. C. Tse, “Diversity and multiplexing: A fundamental tradeoff in multiple-antenna channels,” IEEE
Trans. Inf. Theory, vol. 45, no. 5, pp. 1073– 1096, May. 2003.
[55] J. Xu, Y. Liu, X. Mu, and O. A. Dobre, “STAR-RISs: Simultaneous transmitting and reflecting reconfigurable intelligent
surfaces,” IEEE Commun. Lett., vol. 25, no. 9, pp. 3134–3138, Sept. 2022.
[56] E. Hildebrand, Introduction to numerical analysis, New York, NY, USA: Dover, 1987.