Next Article in Journal
From Biosensors to Robotics: Pioneering Advances in Breast Cancer Management
Previous Article in Journal
Compact VHF/UHF Ultrawideband Discone Antenna with Consistent Pattern
Previous Article in Special Issue
Dual-Pulse Repeated Frequency Waveform Design of Time-Division Integrated Sensing and Communication Based on a 5G New Radio Communication System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Statistics of the Sum of Double Random Variables and Their Applications in Performance Analysis and Optimization of Simultaneously Transmitting and Reflecting Reconfigurable Intelligent Surface-Assisted Non-Orthogonal Multi-Access Systems

1
Faculty of Engineering and Technology, Nguyen Tat Thanh University, Ho Chi Minh City 754000, Vietnam
2
Wireless Communications Research Group, Faculty of Electrical & Electronics Engineering, Ton Duc Thang University, Ho Chi Minh City 70000, Vietnam
3
Faculty of Electrical & Electronics Engineering, Ton Duc Thang University, Ho Chi Minh City 70000, Vietnam
4
Faculty of Electrical Engineering and Computer Science, VSB-Technical University of Ostrava, 17. Listopadu 2172/15, 708 00 Ostrava, Czech Republic
*
Author to whom correspondence should be addressed.
Sensors 2024, 24(18), 6148; https://doi.org/10.3390/s24186148
Submission received: 6 August 2024 / Revised: 18 September 2024 / Accepted: 20 September 2024 / Published: 23 September 2024
(This article belongs to the Special Issue 5G/6G Networks for Wireless Communication and IoT)

Abstract

:
For the future of sixth-generation (6G) wireless communication, simultaneously transmitting and reflecting reconfigurable intelligent surface (STAR-RIS) technology is emerging as a promising solution to achieve lower power transmission and flawless coverage. To facilitate the performance analysis of RIS-assisted networks, the statistics of the sum of double random variables, i.e., the sum of the products of two random variables of the same distribution type, become vitally necessary. This paper applies the statistics of the sum of double random variables in the performance analysis of an integrated power beacon (PB) energy-harvesting (EH)-based NOMA-assisted STAR-RIS network to improve its outage probability (OP), ergodic rate, and average symbol error rate. Furthermore, the impact of imperfect successive interference cancellation (ipSIC) on system performance is also analyzed. The analysis provides the closed-form expressions of the OP and ergodic rate derived for both imperfect and perfect SIC (pSIC) cases. All analyses are supported by extensive simulation results, which help recommend optimized system parameters, including the time-switching factor, the number of reflecting elements, and the power allocation coefficients, to minimize the OP. Finally, the results demonstrate the superiority of the proposed framework compared to conventional NOMA and OMA systems.

1. Introduction

Fifth-generation and sixth-generation wireless networks must meet demanding specifications, including high spectral efficiency, huge connectivity, and low latency [1]. Non-orthogonal multiple access (NOMA) has emerged as a promising multiple-access candidate [2,3]. The ability of NOMA to support numerous users via a single resource block effectively increases spectral efficiency. This contrasts with traditional orthogonal multiple access (OMA), in which different users are allocated to different resource blocks (time, frequency, and code). In power-domain (PD) NOMA, users are given varying power levels, and the base station (BS) transmits a single message that combines the signals of all users. Successive interference cancellation (SIC) is used at the receiver side to recover the broadcast symbol and eliminate inter-user interference [4]. The authors of [5] provided a controllable analysis framework for assessing the security and dependability of underlay cognitive radio networks (CRs) that use non-orthogonal multiple access (NOMA), in which a secondary BS broadcasts private data to several uniformly distributed secondary users while an external eavesdropper is present nearby. In [6], the authors examined an Internet-of-Things (IoT) network with two-way relaying non-orthogonal multiple access enabled, where two NOMA users connect via an IoT access point relay using the decode-and-forward (DF) protocol. For an IoT network configuration operating in real time, a deep learning solution with low computational complexity and accurate outage probability (OP) prediction has been established. The performance of the energy-harvesting NOMA system’s uplink and downlink was reviewed by the authors of [7], where they also investigated and tested the closed-form expressions for the OP in a group of two users. The authors of [8] presented a multi-carrier-based method that improves reliability and sum-rate performance by combining transmit diversity with the NOMA protocol. In [9], the authors analyzed the outage performance and optimized simultaneous wireless information and power transfer (SWIPT) in an energy-harvesting wireless sensor network deploying NOMA. The authors of [10] investigated uplink and downlink transmissions in NOMA networks aided by unmanned aerial vehicles. Numerous other papers have also focused on analyzing the OP in cooperative relaying systems employing NOMA [11,12,13,14]. While NOMA has many benefits for communication systems, it cannot address the problems caused by the randomness of wireless channels, which severely limits the performance improvements that NOMA can achieve.
Additionally, the reconfigurable intelligent surface (RIS), a viable option for next-generation wireless communication networks, has also garnered attention for its ability to alter the wireless propagation environment desirably [15,16,17,18]. An RIS is made up of a large number of low-cost passive elements that, by varying their individual phase shifts and amplitudes, reconfigure the propagation of incident wireless signals. By simultaneously adjusting the RIS passive beamforming vector, the transfer time scheduling, and the power splitting ratio under the energy harvest-then-transmit policy protocol, the authors of [19] maximized the sum throughput for an RIS-assisted IoT network. Reconstructing the wireless environment using RIS technology is a unique approach to enhancing NOMA system performance, which is why it is highly advisable to apply RIS technology to the NOMA system [20]. In the context of an RIS-assisted NOMA network combined with the IoT, the authors of [21] explored physical layer security and addressed the issue of non-conclusive eavesdroppers. Two distinct phase shift designs have been investigated to improve the performance of RIS-assisted NOMA systems [22]. In addition, a unique method based on reflection amplitude and phase shift was proposed to find the maximum total rate of all users, including both ideal and non-ideal conditions [23]. By using RISs to produce coordinated multipoint broadcasts, [24] examined how successfully NOMA cellular networks exploit the spectrum. The energy efficiency, OP, and coverage probability of an RIS-assisted two-user NOMA network were examined in [25,26]. In [27], the authors proposed an RIS–NOMA architecture to maximize user service in each orthogonal spatial direction while accounting for hardware constraints. Furthermore, the effects of incorrect successive interference cancellation on the ergodic rate and OP were examined in [28]. The study suggests looking into an RIS-assisted cooperative NOMA network as a means of bridging the gap between RIS-assisted NOMA and user-relaying cooperation.
With the ongoing progress of the technological era, simultaneously transmitting and reflecting reconfigurable intelligent surface (STAR-RIS) technology has emerged as superior to conventional RIS due to its simultaneous transmission and reflection capabilities, enhanced flexibility, and improved energy efficiency [29]. Specifically, STAR-RIS represents a significant advancement over conventional RIS in terms of wireless signal manipulation capabilities. While traditional RIS can only reflect signals, STAR-RIS can both reflect and transmit signals simultaneously. This dual functionality enables STAR-RIS to achieve wider coverage, enhanced signal quality, and reduced interference compared to its conventional counterpart. The enhanced flexibility of STAR-RIS is evident in its ability to independently adjust reflection and transmission coefficients for each element, allowing for precise control over signal direction and intensity. This adaptability enables STAR-RIS to effectively optimize system performance in diverse environments. Moreover, STAR-RIS can leverage both reflected and transmitted signals to enhance signal strength, leading to improved energy efficiency. Numerous papers have demonstrated that STAR-RIS can outperform conventional RIS in various scenarios, including multi-user communication, physical layer security, and wireless power transfer [29,30,31,32,33].
A variety of works integrating STAR-RIS and NOMA in wireless networks have recently been described [31,33,34,35,36,37,38,39,40,41,42,43,44]. While NOMA is designed for massive user access, it experiences problems with cell-edge users due to additional intra-NOMA cluster interference and non-line-of-sight links from the source [45]. This problem can be solved by RIS, especially STAR-RIS. By adding STAR-RISs to NOMA networks, the authors of [31] showed how to greatly increase coverage. STAR-RIS technology was utilized by the authors of [34] to simultaneously improve the required signals and remove inter-cell interference in NOMA-enhanced coordinated multi-point transmission networks. For the STAR-RIS-assisted multiple-input multiple-output system, the weighted sum rate, the sum secrecy rate, and the attainable sum rate of the STAR-RIS–NOMA system were all optimized by the authors of [35,36,37]. To increase the coverage of heterogeneous networks and effectively prevent mutual interference between users, STAR-RIS was used in [33] to modify the decoding order of users. The authors of [38,39] presented several estimated mathematical channel models to examine the OP performance of STAR-RIS–NOMA and STAR-intelligent omni-surfaces-based NOMA multi-cell networks. A realistic transmission and reflection-coupled phase-shift model for STAR-RISs was considered by the authors of [40]. A STAR-RIS partitioning technique for a STAR-RIS–NOMA network was suggested by the authors of [41]. This technique aims to assign the appropriate amount of STAR-RIS components to each user in order to optimize the sum rate and ensure the quality-of-service criteria. For uplink STAR-RIS–NOMA networks, optimization challenges in minimizing power consumption and maximizing OP secrecy were examined in [42,43]. To optimize the sum rate, the authors of [44] examined the resource allocation mechanism in STAR-RIS-assisted multi-carrier OMA and NOMA networks. Table 1 shows the comparison of our work with related work.
In all the above-mentioned studies, the authors did not consider energy-harvesting technology from power beacons. Energy harvesting (EH) from power beacons (PBs) is an emerging technology that enables wireless devices to scavenge energy from ambient radio frequency (RF) signals [46,47,48,49]. This technology has the potential to prolong the lifetime of battery-powered devices and even enable battery-less operation in some cases. Additionally, EH from PBs is an environmentally friendly solution, as it reduces the need for battery disposal and promotes the use of renewable energy sources. Furthermore, power beacons (PBs) have been proposed as a means to provide wireless power transfer to energy-constrained devices in NOMA systems [6,50,51,52]. By integrating PBs with STAR-RIS, it is possible to achieve both wireless information and power transfer, enabling self-sustainable and energy-efficient NOMA networks. For this reason, in this paper, we propose a novel energy-harvesting (EH) framework that integrates PBs, STAR-RISs, and NOMA to achieve enhanced performance in terms of the OP and ergodic rate. We consider both ipSIC and perfect SIC (pSIC) scenarios and derive closed-form expressions for the OP and ergodic rate in both cases. Our analysis reveals the impact of ipSIC on system performance and provides insights into the design and optimization of EH–PB–STAR-RIS–NOMA systems.
One may be worried about the feasibility of this combination of multiple complicated techniques in a single system. But, in general, there are no significant problems in implementing this integrated system in practice. First, the combination of NOMA and STAR-RIS is natural due to their characteristics, and their integration has been confirmed to be feasible and more effective than the conventional NOMA system [45,53]. Second, each contributed technique is primarily processed on independent devices; for example, NOMA operates at the user level, the channel gains of STAR-RIS elements can be controlled at the STAR-RIS, and EH occurs at the source node.
The following is a summary of this manuscript’s main contributions in light of the previously indicated explanations:
  • We provide a comprehensive review of the statistics (including the probability density function (PDF) and/or the cumulative distribution function (CDF)) of the sum of double random variables, which plays an important role in the derivation of the performance factors of RIS-assisted networks.
  • We propose a novel STAR-RIS-assisted NOMA framework that integrates power beacons (PBs) for energy harvesting, enabling self-sustainable and energy-efficient communication. The probability density function (PDF) and cumulative distribution function (CDF) are considered to validate the accuracy of the approximation method used to model the channel characteristics.
  • We derive closed-form expressions for the OP, ergodic rate, and average symbol error rate of the proposed system, considering both perfect and imperfect SIC scenarios. In order to shed light on the design and optimization of EH–PB–STAR-RIS–NOMA systems, we examine the effects of SIC on system performance.
  • We provide a comprehensive performance analysis and optimization framework for STAR-RIS-assisted NOMA systems, paving the way for their practical implementation in future wireless networks. Furthermore, analysis and simulation results evaluate the impact of important parameters on performance. To confirm the correctness of our analytical model and to show the advantages of the suggested framework over traditional NOMA and OMA systems, we verify our analytical results through Monte Carlo simulations.
Table 1. Comparison of the proposed scheme with other works.
Table 1. Comparison of the proposed scheme with other works.
Our Scheme[54][55][56][57]
NOMAXXX
STAR-RISXXXX
Power BeaconXX
Energy Harvesting
Generalized ChannelsXXXX
Outage ProbabilityXX
Ergodic CapacityXXX
Average SERXXXX
ThroughputXX
OptimizationXXXX

2. Review of the Statistics of the Sum of Double Random Variables

In this section, we present a review of the statistics of the sum of double random variables (RVs). The exact or approximated forms of the PDF and/or CDF of the sum of double random variables are derived for different distribution types.
The problem of finding the PDF or CDF of double random variables has been a popular topic in wireless communications [58]. A double random variable has the form of Z = X Y , where X and Y follow the same distribution type but may have different distribution parameters. Recently, with the invention of RIS, the interest has extended to the sum of double random variables [59,60], i.e., the sum
U = i = 1 n X i Y i ,
where X i and Y i follow the same distribution but may have different parameters. Each term of this sum corresponds to the end-to-end channel gain from the source to the destination through a specific reflecting element on the RIS. As a result, U corresponds to the total received signal at the receiver.

2.1. Sum of Double Rayleigh Random Variables

First, let us denote two RVs, X and Y, and define Z = n = 1 N Z n = n = 1 N X n Y n . Next, we present the PDF and CDF of Z for various RVs. For the Rayleigh RV, the PDF and CDF of T = { X n , Y n } are, respectively, given as
f T ( x ) = 2 x e x 2 ,
F T ( x ) = 1 e x 2 ,
Furthermore, the k-th moment of T is obtained as
μ ¯ T ( k ) = E [ T k ] = Γ k 2 + 1 .
Next, the k-th moment of Z n is obtained as
μ ¯ Z n ( k ) = μ ¯ X n ( k ) μ ¯ Y n ( k ) .
Making use of a multinomial expansion [61], the k-th moment of n = 1 N Z n can be obtained as
μ ¯ Z ( k ) = k 1 = 0 k k 2 = 0 k 1 k Q 1 = 0 k Q 2 k k 1 k 1 k 2 k Q 2 Q 1 × μ ¯ Z n ( k k 1 ) μ ¯ Z n ( k 1 k 2 ) μ ¯ Z n ( k Q 1 ) ,
Next, we can fit Z as the Gamma distribution, with the shape parameter ξ and the inverse scale parameter ζ , which are given by
ξ = E Z 2 Var Z = μ ¯ Z ( 2 ) 2 μ ¯ Z ( 4 ) μ ¯ Z ( 2 ) 2 ,
ζ = E Z Var Z = μ ¯ Z ( 2 ) μ ¯ Z ( 4 ) μ ¯ Z ( 2 ) 2 .
Furthermore, the corresponding PDF and CDF of A i are given by
f Z ( x ) = x ξ 1 ζ i ξ e x ζ i Γ ( ξ ) ,
F Z ( x ) = γ ( ξ , ζ x ) Γ ( ξ ) = 1 Γ ( ξ , ζ x ) Γ ( ξ ) ,
where Γ ( . ) is the Gamma function, and γ ( . , . ) and Γ ( . , . ) are the lower and upper incomplete Gamma functions, respectively.

2.2. Sum of Double Nakagami-m Random Variables

In this subsection, the PDF and CDF of T can be represented as
f T ( x ) = 2 m T Γ ( m T ) x 2 m T 1 e m T x 2 ,
F T ( x ) = 1 Γ ( m T ) γ m T , m T x 2 ,
where m T is the shape parameter and Γ ( . ) is the Gamma distribution [62]. Furthermore, the k-th moment of T is obtained as   
μ ¯ T ( k ) = Γ ( m T + k / 2 ) Γ ( m T ) m T k .
Next, the k-th moment of Z n and Z in this case can be obtained similarly to those in (5) and (6).
Similar to the Rayleigh RV, the corresponding PDF and CDF of Z are, respectively, given by
f Z ( x ) = x ξ 1 ζ i ξ e x ζ i Γ ( ξ ) ,
F Z ( x ) = γ ( ξ , ζ x ) Γ ( ξ ) = 1 Γ ( ξ , ζ x ) Γ ( ξ ) .
where ξ and ζ defined given similarly to (7a) and (7b).

2.3. Sum of Double κ - μ Random Variables

We can express the PDF and CDF of T as [63]
f T x = μ 1 + κ μ + 1 2 κ μ 1 2 e μ κ x μ 1 2 e μ 1 + κ x I μ 1 2 μ κ 1 + κ x , x > 0 ,
and
F T ( z ) = 1 Q μ 2 κ μ , 2 1 + κ x ,
where Q c a , b = a 1 c b x c e x 2 + a 2 x 2 + a 2 2 2 I c 1 a x d x specifies the generalized Marcum Q-function and I v . is the modified Bessel function of the first kind and order v. Furthermore, the k-th moment of T is obtained as [63]
μ ¯ T ( k ) = Γ ( μ + k / 2 ) exp ( κ μ ) Γ ( μ ) [ ( 1 + κ ) μ ] k / 2 F 1 1 ( μ + k / 2 ; μ ; κ μ ) ,
Next, the k-th moment of Z n and Z in this case can be obtained similarly to those in (5) and (6). Similar to the Rayleigh RV, the corresponding PDF and CDF of Z are, respectively, given by
f Z ( x ) = x ξ 1 ζ i ξ e x ζ i Γ ( ξ ) ,
F Z ( x ) = γ ( ξ , ζ x ) Γ ( ξ ) = 1 Γ ( ξ , ζ x ) Γ ( ξ ) .
where ξ and ζ are defined similarly to (7a) and (7b).

3. System Model and Channel Characteristics

3.1. System Model

We consider a wireless communication system, as depicted in Figure 1, where a signal is transmitted from the source (S) to the destination ( D i with i 1 , 2 ). The source’s transmitted signal is reflected by the STAR-RIS, which is made up of 2 Q elements after the power beacon (PB) sends wireless energy to S. Furthermore, we adopt the switching mode used in [64], i.e., 2 Q elements are divided into two equal parts to serve D 1 and D 2 . Furthermore, using the received RF signal from the PB node, we assume that source S harvests energy for α τ seconds, where τ is the frame period and 0 < α < 1 . The energy captured by S is utilized to send data to the destination for the remainder of the frame, i.e., 1 α τ seconds. We calculate the throughput for various IRS setups when used as a reflector. We also recommend optimizing the harvesting period to enhance throughput. A short harvesting period results in little gathered energy and poor throughput. If the harvesting length is sufficiently long, there may not be enough time for data transmission during the remainder of the frame, i.e., 1 α τ .
S gathers energy from the RF signal received from the PB. The amount of energy collected is equal to
E = η α τ P P B h ¯ 2 ,
where α is the fraction of the block time in which S harvests energy from the PB’s information signal, 0 < η < 1 denotes the energy conversion efficiency, P P B is the power of the PB node, h is the channel coefficient between the PB and S, and h ¯ = Δ h e j θ h denotes the channel from the PB to S, where θ h represents the channel phases that are uniformly distributed between 0 and 2 π . The average power of the channel coefficient h is E h 2 = λ h , where E X is the expectation operation of X.
As in several prior works, we assume that all the harvested energy is used during the information transmission phase. Hence, the transmit power of the source is given by
P S = E 1 α τ = η α P P B 1 α h ¯ 2 .
S uses the gathered energy to transmit the separate signal D i via the STAR-RIS paradigm during the wireless information transfer (WIT) phase. As a result, the signals received at D 1 and D 2 are represented as
y 1 = g 1 H Φ 1 g 0 a 1 P S x 1 + a 2 P S x 2 + n 1 ,
y 2 = g 2 H Φ 2 g 0 a 1 P S x 1 + a 2 P S x 2 + n 2 ,
where P S denotes the transmit power of S, and x i , i 1 , 2 is the normalized transmit signal with unit energy such that E x i 2 = 1 . We have a 1 and a 2 , which are the power allocation coefficients satisfying a 2 > a 1 and a 2 + a 1 = 1 . We define g 0 = [ g ¯ 0 , 1 , , g ¯ 0 , q , , g ¯ 0 , Q ] , g 1 = [ g ¯ 1 , 1 , , g ¯ 1 , n , , g ¯ 1 , Q ] H , and g 2 = [ g ¯ 2 , 1 , , g ¯ 2 , n , , g ¯ 2 , Q ] H , where g ¯ 0 , q = g 0 , q e j θ q g 0 , g ¯ 1 , q = g 1 , q e j θ q g 1 , and g ¯ 2 , q = g 2 , q e j θ q g 2 denote the channel coefficients from S to the STAR-RIS, from the STAR-RIS to D 1 and from the STAR-RIS to D 2 , respectively. Meanwhile, the phases of the channels, i.e., θ q g 0 , θ q g 1 , and θ q g 2 are uniformly distributed from 0 to 2 π .
Let Φ 1 = d i a g β 1 g 1 e j ϕ 1 g 1 , , β q g 1 e j ϕ q g 1 , , β Q g 1 e j ϕ Q g 1 and Φ 2 = d i a g β 1 g 2 e j ϕ 1 g 2 , , β q g 2 e j ϕ q g 2 , , β Q g 2 e j ϕ Q g 2 denote the reflection and transmission coefficient matrices of the STAR-RIS, where β q g 1 , β q g 2 0 , 1 denote the amplitude of the qth STAR-RIS element and ϕ q g 1 , ϕ q g 2 0 , 2 π denote the phase shift of the qth STAR-RIS element. The additive Gaussian noise is given by n i C N 0 , σ i 2 , which has a mean of 0 and variance of σ i 2 .
For the STAR-RIS–NOMA network, D 1 first decodes D 2 ’s signal and then decodes its own signal by eliminating D 2 ’s signal with SIC. Assuming perfect channel state information (CSI), the signal-plus-interference-to-noise ratio (SINR) of these two processes is expressed as
γ D 2 D 1 = a 2 P S g 1 H Φ 1 g 0 2 a 1 P S g 1 H Φ 1 g 0 2 + σ 1 2 ,
γ D 1 = a 1 P S g 1 H Φ 1 g 0 2 a 2 P S g 1 H Φ 1 g 0 2 + σ 1 2 , 0 < 1 a 1 P S g 1 H Φ 1 g 0 2 σ 1 2 , = 0
where , 0 1 , represents the efficiency of SIC for x 1 at D 1 . The cases = 0 and = 1 correspond to perfect SIC (pSIC) and imperfect SIC (ipSIC), respectively.
Submitting (21) into (23a) and (23b), γ D 2 D 1 and γ D 1 are given by
γ D 2 D 1 = a 2 δ P P B h ¯ g 1 H Φ 1 g 0 2 a 1 δ P P B h g 1 H Φ 1 g 0 2 + σ 1 2 ,
γ D 1 = a 1 δ P P B h ¯ g 1 H Φ 1 g 0 2 a 2 δ P P B h ¯ g 1 H Φ 1 g 0 2 + σ 1 2 , 0 < 1 a 1 δ P P B h ¯ g 1 H Φ 1 g 0 2 σ 1 2 , = 0
where δ = η α ρ η α ρ 1 α 1 α .
It should be noted that we can rewrite (24a) and (24b) as
γ D 2 D 1 = a 2 δ ρ h ¯ g 1 H Φ 1 g 0 2 a 1 δ ρ h ¯ g 1 H Φ 1 g 0 2 + 1 ,
γ D 1 = a 1 δ ρ h ¯ g 1 H Φ 1 g 0 2 a 2 δ ρ h ¯ g 1 H Φ 1 g 0 2 + σ 1 2 , 0 < 1 a 1 δ ρ h ¯ g 1 H Φ 1 g 0 2 , = 0
where ρ = P P B σ 1 2 = P P B σ 2 2 is the signal-to-noise ratio (SNR) at the PB.
Unlike D 1 , D 2 directly detects its own signal; thus, its SINR is given as
γ D 2 = a 2 δ ρ h ¯ g 2 H Φ 2 g 0 2 a 1 δ ρ h ¯ g 2 H Φ 2 g 0 2 + 1 .
We denote the phases of the cascaded channels as ϕ 1 = θ h + θ q g 0 + θ q g 1 and ϕ 2 = θ h + θ q g 0 + θ q g 2 . Furthermore, we assume that ϕ q g 1 = ϕ 1 and ϕ q g 2 = ϕ 2 , and then the effective cascade channels from S to users are expressed as h q = 1 Q g 0 , q g 1 , q 2 and h q = 1 Q g 0 , q g 2 , q 2 , respectively. In this case, Equations (25a), (25b) and (26) are rewritten as   
γ D 2 D 1 = a 2 δ ρ A 1 a 1 δ ρ A 1 + 1 ,
γ D 1 = a 1 δ ρ A 1 a 2 δ ρ A 1 + 1 , 0 < 1 a 1 δ ρ A 1 , = 0
γ D 2 = a 2 δ ρ A 2 a 1 δ ρ A 2 + 1 ,
where A 1 = h q = 1 Q g 0 , q g 1 , q 2 and A 2 = h q = 1 Q g 0 , q g 2 , q 2 .

3.2. Channel Characteristics

First, we define B i = q = 1 Q g 0 , q g i , q and present the distribution for the sum of double RVs in Section 2. Thus, the k-th moment of B i is given by
μ ¯ B i ( k ) = Γ ( ξ + k ) γ ( ξ ) ζ k ,
Based on the independence of B i and h, the k-th moment of A i can be obtained as
μ ¯ A i ( k ) = μ ¯ B i ( k ) μ ¯ h ( k ) ,
where μ ¯ h ( k ) is presented in Section 2. Next, we can fit A i as the Gamma distribution, with the shape parameter ξ i and the inverse scale parameter ζ i , which are given by
ξ i = E A i 2 Var A i = μ A i 2 2 μ A i 4 μ A i 2 2 ,
ζ i = E A i Var A i = μ A i 2 μ A i 4 μ A i 2 2 .
Furthermore, the corresponding PDF and CDF of A i are given by
f A i ( x ) = x ξ i 1 ζ i ξ i e x ζ i Γ ( ξ i ) ,
F A i ( x ) = γ ( ξ i , ζ i x ) Γ ( ξ i ) = 1 Γ ( ξ i , ζ i x ) Γ ( ξ i ) ,
where Γ ( . ) is the Gamma function, and γ ( . , . ) and Γ ( . , . ) are the lower and upper incomplete Gamma functions, respectively.

4. Outage Probability Analysis

The performance of the STAR-RIS-assisted NOMA system is characterized in terms of the OP. The next section discusses the OP caused by the channel reciprocity of D 2 and D 1 .

4.1. Outage Probability D 2

The following scenario depicts NOMA failure incidents at D 2 . D 2 can reliably locate x 2 . The OP of D 2 can therefore be expressed as
P D 2 = 1 Pr γ D 2 > γ t h 2 = 1 Pr A 2 > υ 2 = F A 2 υ 2 ,
where γ t h 2 = 2 R 2 1 α 1 , with R 2 being the target rate at D 2 to detect x 2 , and υ 2 = γ t h 2 δ ρ a 2 γ t h 2 a 1 .
Substituting (32) into (33), the OP at D 2 is represented as
P D 2 = 1 Γ ( ξ 2 , ζ 2 υ 2 ) Γ ( ξ 2 ) .

4.2. Outage Probability D 1

In a STAR-RIS-assisted NOMA system, D 1 outage events are defined as follows: (i) D 1 is unable to detect the information x 2 ; and (ii) D 1 is unable to detect x 2 but can successfully decode x 1 . To simplify the current analysis, the complementary events of x 1 reflect the OP. As a result, for 0 < 1 , the OP at D 1 with ipSIC for a STAR-RIS-supported NOMA system is given as
P D 1 i p S I C = 1 Pr γ D 2 D 1 > γ t h 2 , γ D 1 > γ t h 1 = 1 Pr A 1 > υ 2 , A 1 > υ 3 = 1 Pr A 1 > υ max = F A 1 υ max ,
where γ t h 1 = 2 R 1 1 α 1 , with R 1 being the target rate at D 1 to detect x 1 , υ 3 = γ t h 1 δ ρ a 1 γ t h 1 a 2 , and υ max = max υ 2 , υ 3 .
Substituting (32) into (35), the OP at D 1 with ipSIC is expressed as
P D 1 i p S I C = 1 Γ ( ξ 1 , ζ 1 υ max ) Γ ( ξ 1 ) .
Similar to (36), for the case = 0 , the OP at D 1 with pSIC is calculated as follows:
P D 1 p S I C = 1 Γ ( ξ 1 , ζ 1 υ ˜ max ) Γ ( ξ 1 ) ,
where υ ˜ max = max υ 2 , υ 1 and υ 3 = γ t h 1 δ ρ a 1 .

4.3. Diversity Analysis

To shed light on the performance of the proposed STAR-RIS-assisted NOMA system, the asymptotic OP in the high SNR regime ( ρ ) is shown. The diversity order is defined as follows:
d 1 = lim ρ log P D 1 , ρ log ρ , i p S I C , p S I C ,
d 2 = lim ρ log P D 2 ρ log ρ ,
where P D 1 , ρ and P D 2 ρ denote the asymptotic OPs of D 1 and D 2 , respectively.
When ρ , the asymptotic form of F A i ( z ) is as follows [62]:
F A i ( x ) 1 ξ i ! ζ i x ξ i .
Plugging (37), (36) and (34) into (39), the asymptotic OPs of P D 1 , i p S I C , P D 1 , p S I C , and P D 2 can be expressed as   
P D 1 , i p S I C = ζ 1 υ max ξ 1 ξ 1 ! ,
P D 1 , p S I C = ζ 1 υ ˜ max ξ 1 ξ 1 ! ,
P D 2 = ζ 2 υ 2 ξ 2 ξ 2 ! .

5. Evaluation of Ergodic Rate

Other capacity performance gaps arise because the outage performance of the destination users D 1 and D 2 is expected to exhibit a performance gap due to differing power allocation parameters in the NOMA system. As a result, this inspires us to conduct an additional evaluation of the suggested STAR-RIS-assisted NOMA system in terms of the ergodic rate.

5.1. Ergodic Rate of D 2

In this section, we investigate the system’s ergodic rate. The ergodic rate of D 2 for the NOMA downlink is given by
C D 2 = E log 2 1 + γ D 2 .
By the definition of the expectation operator and after integration by parts, C D 2 can be expressed as
C D 2 = 1 ln 2 0 a 2 a 1 1 1 + x F ¯ A 2 x δ ρ a 2 x a 1 d x ,
where F ¯ A 2 x denotes the complementary CDF of A 2 , i.e., F ¯ A 2 x = 1 F A 2 x . By changing variables to t x δ ρ a 2 x a 1 and after a few steps, (42) can be further derived as
C D 2 = 1 ln 2 0 1 t + Θ 1 1 t + Θ 2 F ¯ A 2 t d t ,
where Θ 1 = δ ρ a 2 + a 1 1 and Θ 2 = δ ρ a 1 1 .
By substituting (34) into (43), C D 2 can be written as
C D 2 = 1 ln 2 Γ ( ξ 2 ) 0 1 t + Θ 1 1 t + Θ 2 Γ ( ξ 2 , ζ 2 t ) d t .
To solve the integrals in (44), we use the following transformations with the Meijer G-function [6]
Γ α , x = G 1 , 2 2 , 0 x 1 α , 0 .
By converting the required functions in (44) into Meijer G-functions using the aforementioned equalities in (45), we can solve the integral as follows:
C D 2 = 1 ln 2 Γ ( ξ 2 ) 0 1 t + Θ 1 1 t + Θ 2 G 1 , 2 2 , 0 ζ 2 t 1 ξ 2 , 0 d t = 1 ln 2 Γ ( ξ 2 ) 1 t + Θ 1 G 1 , 2 2 , 0 ζ 2 t 1 ξ 2 , 0 d t 1 t + Θ 2 G 1 , 2 2 , 0 ζ 2 t 1 ξ 2 , 0 d t .
Then, with the aid of [62], Equation (7.811.5), C D 2 is obtained as
C D 2 = 1 ln 2 Γ ( ξ 2 ) G 2 , 3 3 , 1 ζ 2 Θ 1 0 , 1 0 , ξ 2 , 0 G 2 , 3 3 , 1 ζ 2 Θ 2 0 , 1 0 , ξ 2 , 0 .

5.2. Ergodic Rate of D 1 with Imperfect SIC

The achievable rate of the considered system at D 1 with 0 < 1 is given as
C D 1 i p S I C = E log 2 1 + γ D 1 = E log 2 1 + a 1 δ ρ A 1 a 2 δ ρ A 1 + 1 = 1 ln 2 0 a 1 a 2 1 1 + x 1 F A 1 x δ ρ a 1 x a 2 d x .
By replacing the variable t = x δ ρ a 1 x a 2 in (48) and after a few steps, C D 1 i p S I C can be further derived as
C D 1 i p S I C = 1 ln 2 0 1 t + Θ 3 1 t + Θ 4 1 F A 1 t d t ,
where Θ 3 = δ ρ a 1 + a 2 1 and Θ 4 = δ ρ a 2 1 .
Based on [62], Equation (7.811.5), and applying some polynomial expansion manipulations, C D 1 i p S I C is given by
C D 1 i p S I C = 1 ln 2 Γ ( ξ 1 ) G 2 , 3 3 , 1 ζ 1 Θ 3 0 , 1 0 , ξ 1 , 0 G 2 , 3 3 , 1 ζ 1 Θ 4 0 , 1 0 , ξ 1 , 0 .

5.3. Ergodic Rate of D 1 with Perfect SIC

The ergodic rate of device D 1 with = 0 for the pSIC case is similar to that of the ipSIC case:
C D 1 p S I C = E log 2 1 + γ D 1 = E log 2 1 + a 1 δ ρ A 1 = 1 ln 2 0 1 1 + x 1 F A 1 x d x .
Plugging (27b) into (51), we have
C D 1 p S I C = 1 ln 2 Γ ( ξ 1 ) 0 1 1 + x G 1 , 2 2 , 0 ζ 1 x δ ρ a 1 1 ξ 1 , 0 d x .
With the help of [62], Equation (7.811.5), C D 1 p S I C can be further computed by
C D 1 p S I C = 1 ln 2 Γ ( ξ 1 ) G 2 , 3 3 , 1 ζ 1 δ ρ a 1 0 , 1 0 , ξ 1 , 0 .

5.4. Average Symbol Error Rate

Let us denote p and q as constants. In particular, the modulation types depend on the values of p and q. We treat the binary phase-shift keying (BPSK) modulation corresponding to p = 1 , q = 2 . If the values are p = 2 , q = 1 , they represent quadrature phase-shift keying (QPSK) and 4-quadrature amplitude modulation (4-QAM), as shown in [65], where Q . is the Gaussian error function. For the RIS-aided point-to-point system, the average symbol error rate (SER) needs to be computed as [65]
S j = p E Q q γ j = p 2 π 0 F γ j x 2 q e x 2 2 d x = y = Δ x 2 x 2 q q p q 2 2 π 0 e q 2 y y F γ j y d y , j D 1 , D 2

5.5. Average SER of D 2

From (32), the average SER of D 2 is calculated as
S D 2 = p q 2 2 π 0 e q 2 x x F γ D 2 x d x = p q 2 2 π 0 e q 2 x x d x 1 Γ ( ξ 2 ) 0 a 2 a 1 e q 2 x x Γ ξ 2 , ζ 2 x δ ρ a 2 x a 1 d x .
For the first integral in (55), using [62], Equation (3.361.2), we have
0 e q 2 x x d x = 2 π q .
For the second integral, let t = 2 a 1 x a 2 1 a 2 t + 1 2 a 1 = x a 2 2 a 1 d t = d x , and we have
1 Γ ( ξ 2 ) 0 a 2 a 1 e q 2 x x Γ ξ 2 , ζ 2 x δ ρ a 2 x a 1 d x = a 2 2 a 1 Γ ( ξ 2 ) 1 1 e q 2 Λ t Λ t Γ ξ 2 , ζ 2 Λ t δ ρ a 2 Λ t a 1 d t ,
where Λ t = a 2 t + 1 a 2 t + 1 2 a 1 2 a 1 . Unfortunately, finding a closed-form expression for the second integral is a challenging task, but an accurate approximation can be obtained for it. By using the Gaussian–Chebyshev quadrature ([66], Equation (25.4.38)), it can be achieved by
a 2 2 a 1 Γ ( ξ 2 ) 1 1 e q 2 Λ t Λ t Γ ξ 2 , ζ 2 Λ t δ ρ a 2 Λ t a 1 d t π a 2 2 W a 1 Γ ( ξ 2 ) w = 1 W 1 χ w 2 e q 2 Λ χ w Λ χ w Γ ξ 2 , ζ 2 Λ χ w δ ρ a 2 Λ χ w a 1 ,
where χ w = cos 2 w 1 2 W π and W is a complexity–accuracy tradeoff parameter.
Now, substituting (58) and (56) into (55), the average SER of D 2 is given by
S D 2 p 2 p q π a 2 2 2.5 W a 1 Γ ( ξ 2 ) w = 1 W 1 χ w 2 e q 2 Λ χ w Λ χ w Γ ξ 2 , ζ 2 Λ χ w δ ρ a 2 Λ χ w a 1 .

5.6. Average SER of D 1

The average SER in the ipSIC case for D 1 can be computed as follows:
S D 1 p S I C = p q 2 2 π 0 a 1 a 2 e q 2 x x F A 1 x δ ρ a 1 x a 2 d x = p q 2 2 π 0 e q 2 x x d x 1 Γ ( ξ 1 ) 0 a 2 a 1 e q 2 x x Γ ξ 1 , ζ 1 x δ ρ a 1 x a 2 d x .
Similarly, by solving S D 2 , S D 1 i p S I C can be obtained as
S D 1 p S I C p 2 p q π a 1 2 2.5 W a 2 Γ ( ξ 1 ) w = 1 W 1 χ w 2 e q 2 Ψ χ w Ψ χ w Γ ξ 2 , ζ 2 Ψ χ w δ ρ a 1 Ψ χ w a 2 .
where Ψ t = a 1 t + 1 a 1 t + 1 2 a 2 2 a 2 .
Next, the average SER in the pSIC case for D 1 is given by
S D 1 p S I C = p q 2 2 π 0 e q 2 x x F γ D 1 x d x = p q 2 2 π 0 e q 2 x x 1 1 Γ ξ 1 Γ ξ 1 , ζ 1 x δ ρ a 1 d x = p q 2 2 π 0 e q 2 x x d x 1 Γ ξ 1 0 e q 2 x x G 1 , 2 2 , 0 ζ 1 x δ ρ a 1 1 ξ 1 , 0 d x .
For the first integral, it is solved in formula as follows: (56) 0 e q 2 x x d x = 2 π q . For the second integral, using similar methods as in [62,67], (7.813.1), we have
1 Γ ξ 1 0 e q 2 x x G 1 , 2 2 , 0 ζ 1 x δ ρ a 1 1 ξ 1 , 0 d x = 1 Γ ξ 1 2 q 1 2 G 2 , 2 2 , 1 2 ζ 1 δ q ρ a 1 0.5 , 1 ξ 1 , 0 .
Substituting (63) and (56) into (62) and performing some mathematical manipulations, we obtain the ASER of D 1 with pSIC as
S D 1 p S I C = p 2 p 2 π Γ ξ 1 G 2 , 2 2 , 1 2 ζ 1 δ q ρ a 1 0.5 , 1 ξ 1 , 0 .

6. Optimal Performance Analysis

We can see from (37), (36) and (34) that the OP is related to the time-switching (TS) factor α i . As a result, in order to minimize the likelihood of an outage, we must discover the best solution of α i , which can be expressed as
α i * = arg min α i P D i s . t . 0 < α i < 1
where i 1 , 2 and α i * denotes the optimal solution of α i for minimizing the system’s OP.
We know from (35) and (33) that R i is a fixed number and that there is a positive correlation between the OP and γ t h i . This suggests that the solution to (65) is unique and can be solved using a linear search strategy, such as the golden section method. So, we use the golden section method to solve the optimization problem, and the specific process is presented in Algorithm 1.
Algorithm 1: Optimization algorithm to find α i * based on the golden section method.
  Input: Initialize ψ min i = 0 , ψ max i = 1 , the golden section search ϖ i = 5 1 2 and a
     stopping threshold Δ i = 10 3
  Output: The optimal of α i * that minimum the OP P D i α i *
  Sensors 24 06148 i001

7. Numerical Results

In this section, we quantitatively assess our theoretical conclusions regarding the OP and ergodic rate performance. We define the fading parameters as m = m h = m g 0 = m g 1 = m g 2 . Monte Carlo simulation [68,69,70,71,72] results are averaged across 10 6 independent trials. The target rate is represented in BPCU (bits per channel use). In the figures below, we represent “Ana.”, “Sim.”, and “Asymp.” as analytical analysis, simulation, and asymptotic computation, respectively. Table 2 summarizes the other major parameters. The parameters of the κ μ random variables are presumed to be k = k h = k g 0 = k g 1 = k g 2 and μ = μ h = μ g 0 = μ g 1 = μ g 2 (see Table 3 for examples). In addition, the Gauss–Chebyshev parameter is chosen as W = 100 to obtain a near approximation.
Figure 2 illustrates the probability density function (PDF) and cumulative distribution function (CDF) for different numbers of STAR-RIS elements (Q = 8, 12, 16). The top graph shows the CDF, where the curves shift to the right as the number of elements increases, indicating a higher probability of achieving larger values of the random variable x. The bottom graph shows the PDF, where the curves become narrower and taller as the number of elements increases, suggesting a more concentrated distribution around the mean value of x. The simulation results, represented by circles, closely match the approximation results, represented by lines, validating the approximation method’s accuracy.
Figure 3 demonstrates the OPs of users D 1 (in subfigure a) and D 2 (in subfigure b) in a STAR-RIS-assisted NOMA system under various classical channel models (Rayleigh, Nakagami-m, and Rician) and transmit power levels ρ . The OP is plotted on a logarithmic scale against the transmit power in dB. The results show that the OP decreases as the transmit power increases for all channel models. Additionally, the Rician channel model consistently exhibits the lowest OP, followed by Nakagami-m and Rayleigh. The analytical results closely match the asymptotic results, validating the accuracy of the analysis.
Figure 4 provides a comparative analysis of the OPs of two users (User 1 and User 2) in both STAR-RIS-assisted NOMA and conventional OMA systems. The OP is depicted as a function of the transmit power ( ρ ) in dB, with varying numbers of STAR-RIS elements (Q) ranging between 4 and 16. The results reveal several key insights. First, the OP for both NOMA and OMA users decreases as the transmit power increases, highlighting the positive impact of higher power levels on signal quality and reception. Second, the incorporation of STAR-RIS elements significantly enhances performance for both NOMA and OMA users. This improvement can be attributed to the superior signal focusing and interference mitigation capabilities of STAR-RIS, which become more pronounced with an increasing number of elements. Furthermore, the STAR-RIS-assisted NOMA system consistently outperforms the conventional OMA system in terms of the OP, particularly for User 2. This advantage stems from NOMA’s ability to serve multiple users simultaneously on the same time-frequency resource, leading to improved spectral efficiency. However, within the NOMA scenario, User 1 experiences a higher OP compared to User 2 due to the successive interference cancellation (SIC) process. In this process, User 1 decodes its own signal while treating User 2’s signal as interference, leading to residual interference and a higher OP. In addition, User 1 experiences different outage probabilities under imperfect successive interference cancellation (ipSIC) and perfect successive interference cancellation (pSIC). With ipSIC, residual interference from User 2’s signal degrades User 1’s signal quality, leading to a higher OP. Conversely, with pSIC, User 1 perfectly cancels this interference, resulting in a lower OP. This difference highlights the importance of effective SIC techniques in NOMA systems. The close alignment between the simulation and analytical results validates the accuracy of the analytical model employed to derive the OP expressions.
Figure 5 illustrates the OPs of users D 1 (top subfigure) and D 2 (bottom subfigure) in a STAR-RIS-assisted NOMA system versus the target rates R 1 and R 2 . The OP is plotted on a logarithmic scale. The results show that the OP increases with the target rates for both users. Additionally, the OP of user D 1 is significantly higher than that of user D 2 , which is due to the SIC process in NOMA. The simulation results closely match the analytical results, thereby validating the accuracy of the analysis.
Observing Figure 6, it is evident that the simulation and analytical results closely align, thus validating the accuracy of the analysis in illustrating the OP of users within a STAR-RIS-assisted NOMA system. The simulation specifically focuses on the impact of the power allocation factor ( a 2 ) on the OP. It reveals that increasing a 2 initially decreases the OP for both users, but beyond a certain point, the OP of User 1 begins to rise again, while that of User 2 continues to decline. This can be attributed to the increased power allocation to User 2, which improves its performance but negatively affects User 1 due to heightened interference. The optimal value of a 2 , which minimizes the OP for both users, is contingent upon specific system parameters and channel conditions. Furthermore, an increase in the number of STAR-RIS elements significantly enhances performance for both users. Notably, the OP for User 1 under ipSIC conditions is higher than that under pSIC due to residual interference, as elucidated in Figure 4.
Figure 7 clearly shows the OPs of users in a NOMA system as a function of the fraction of block time ( α ), considering both perfect SIC (pSIC) and imperfect SIC (ipSIC) scenarios, with varying numbers of reflecting elements (8 or 16). The simulation reveals that increasing the number of STAR-RIS elements significantly enhances performance for both users. Notably, the OP for User 1 under ipSIC conditions is higher than the OP under pSIC due to residual interference, as illustrated in Figure 4. In addition, as α increases, the OP for both users initially decreases, reaching an optimal point, and then starts to increase. Specifically, when α = 0.4, User 1 in the ipSIC scenario achieves the best performance. Similarly, for User 1 in the pSIC scenario and User 2, the optimal α values are α = 0.5 and α = 0.6, respectively. This facilitates the determination of the optimal α value to minimize the OP for a given system configuration.
In Figure 8, the STAR-RIS network shows significantly better performance in terms of the outage probability compared to the relay network. The STAR-RIS network curves drop much faster, reaching lower outage probabilities as ρ increases. Both networks show that User 1 experiences a higher outage probability than User 2 for the same value of ρ . However, the performance gap between User 1 and User 2 is more pronounced in the relay network. At higher values of ρ (15–20 dB), the STAR-RIS network for User 2 approaches extremely low outage probabilities (below 10 4 ), while the relay network does not reach such low probabilities even at the highest ρ . The plot indicates that the STAR-RIS network provides a substantial improvement in reducing the outage probability compared to the relay network, particularly at higher SNRs. This is consistent for both User 1 and User 2, with User 2 benefiting more from both networks across the range of ρ . This figure demonstrates the clear advantage of the STAR-RIS network in improving reliability, especially in high-SNR conditions.
Figure 9 illustrates the ergodic rate (BPCU) of users in a STAR-RIS-assisted NOMA system and a conventional OMA system versus ρ under varying conditions of pSIC and ipSIC, as well as different numbers of reflecting elements (Q = 16 or 32). The analysis reveals several key findings. First is NOMA’s superiority. Across all scenarios, the NOMA system consistently outperforms the OMA system in terms of the ergodic rate for both users. This highlights NOMA’s inherent advantage in spectral efficiency, as it allows multiple users to share the same time-frequency resources. Second is the impact of SIC. The ergodic rate of User 1 under pSIC is significantly higher than that under ipSIC. This underscores the critical role of effective interference cancellation in NOMA. In pSIC, User 1 can perfectly decode and remove User 2’s signal before decoding its own, while ipSIC leaves residual interference, degrading User 1’s performance. This disparity widens as the transmit ρ increases, indicating that ipSIC becomes more detrimental at higher ρ levels. In addition, increasing the number of reflecting elements in the STAR-RIS from 8 to 16 substantially boosts the ergodic rate for both users in the NOMA system. This demonstrates the significant potential of the STAR-RIS to enhance the spectral efficiency of NOMA by intelligently manipulating the wireless environment.
Figure 10 shows the ergodic rate, a measure of the average achievable data rate, for a two-user NOMA system as a function of the number of reflecting elements (Q). The number of reflecting elements is equal for both users, and the ergodic rate is plotted for both users in the NOMA system. A key observation is that the ergodic rate for both users increases as the number of reflecting elements increases. This is due to the increased signal strength and diversity gain provided by the additional signal propagation paths created by the reflecting elements. User 1 consistently achieves a higher ergodic rate than User 2 due to the power allocation strategy in NOMA, which allocates more power to the stronger user (User 1) to ensure reliable communication. The analytical results, obtained through mathematical modeling, closely match the simulation results, validating the accuracy of the analytical model in predicting the ergodic rate performance of the NOMA system. Additionally, for User 1, the ergodic rate with ipSIC is slightly lower than that with pSIC, as expected. This is because ipSIC introduces some residual interference, which degrades the performance of User 1, highlighting the importance of accurate interference cancellation in NOMA systems.
Figure 11 plots the average symbol error rate (SER) performance of a STAR-RIS-assisted NOMA system with two users, employing both BPSK and 4-QAM modulation schemes. The SER, a measure of the rate at which symbols are incorrectly detected, is plotted against the transmit power ( ρ ) in dB. The results are presented for both analytical and simulation scenarios, with and without perfect successive interference cancellation (pSIC and ipSIC, respectively). As expected, the SER for both users decreases as the transmit power increases due to the stronger received signal resulting from higher power. The 4-QAM modulation scheme, which carries more bits per symbol than BPSK, exhibits a higher SER for the same transmit power, indicating its greater susceptibility to errors in the presence of noise and interference. A significant observation is the impact of perfect SIC on User 1’s SER. The SER is notably lower when perfect SIC is employed compared to the case with imperfect SIC, highlighting the importance of accurate interference cancellation in NOMA systems for the effective removal of interference from User 2’s signal. The close alignment between the analytical and simulation results validates the accuracy of the analytical model in predicting the SER performance of the STAR-RIS-assisted NOMA system. Overall, the figure provides valuable insights into the influence of transmit power, modulation scheme, and SIC on the average SER, emphasizing the importance of optimizing these parameters to achieve the desired SER performance in practical NOMA deployments.

8. Conclusions

In this paper, we have investigated the performance of a STAR-RIS-assisted NOMA system with PB energy harvesting. We derived closed-form expressions for the OP and ergodic rate of the system, considering both perfect and imperfect SIC scenarios. Our analysis revealed the impact of imperfect SIC on system performance and provided insights into the design and optimization of EH–PB–STAR-RIS–NOMA systems. We also presented extensive numerical results to validate the accuracy of our analytical model and demonstrate the superiority of the proposed framework compared to conventional NOMA and OMA systems. Our results show that the proposed framework can significantly improve the OP and ergodic rate of NOMA systems, especially in the presence of blockages. To achieve the best performance for the STAR-RIS-assisted NOMA system, it is crucial to determine the optimal values for various system parameters. These parameters include the power allocation coefficients ( a 1 and a 2 ), the time-switching factor ( α ), and the number of reflecting elements (Q). The optimal values can be determined by analyzing the derived closed-form expressions for the OP and ergodic rate and by conducting simulations under different scenarios. For instance, Figure 6 illustrates the impact of the power allocation factor ( a 2 ) on the OPs of both users. By finding the value of a 2 that minimizes the OP for both users, we can determine the optimal power allocation strategy. Similarly, Figure 7 shows the effect of the time-switching factor ( α ) on the OP. The optimal α value can be determined by identifying the point where the OP is minimized for both users. By carefully optimizing these parameters, we can maximize the performance of the STAR-RIS-assisted NOMA system in terms of the OP and ergodic rate while also ensuring its practicality and cost-effectiveness. In conclusion, the proposed EH–PB–STAR-RIS–NOMA framework offers a promising solution to overcome the challenges of blockages and limited energy in wireless communication systems. By integrating STAR-RIS with NOMA and PB energy harvesting, we can achieve self-sustainable, energy-efficient, and high-performance wireless networks. The analytical and simulation results presented in this paper provide valuable insights for the design and optimization of such networks.

Author Contributions

P.T.T. proposed the idea and system model. B.V.M. implemented the methodology and software. A.-T.L. and T.-H.T.P. investigated the data curation. S.-P.L. wrote the original draft preparation and P.P. supervised the entire work. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Union within the REFRESH project—Research Excellence For Region Sustainability and High-Tech Industries—ID No. CZ.10.03.01/00/22_003/0000048 of the European Just Transition Fund, and by the Ministry of Education, Youth, and Sports of the Czech Republic (MEYS CZ) through the project SGS, ID No. SP 061/2024, conducted by VSB, Technical University of Ostrava.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare they have no conflicts of interest related to the publication of this paper.

References

  1. Lin, Z.; Lin, M.; de Cola, T.; Wang, J.B.; Zhu, W.P.; Cheng, J. Supporting IoT With Rate-Splitting Multiple Access in Satellite and Aerial-Integrated Networks. IEEE Internet Things J. 2021, 8, 11123–11134. [Google Scholar] [CrossRef]
  2. Khan, W.U.; Liu, J.; Jameel, F.; Sharma, V.; Jäntti, R.; Han, Z. Spectral Efficiency Optimization for Next Generation NOMA-Enabled IoT Networks. IEEE Trans. Veh. Technol. 2020, 69, 15284–15297. [Google Scholar] [CrossRef]
  3. Nguyen, T.T.T.; Do, D.T. Exploiting Full-duplex and Fixed Power Allocation Approaches for Dual-hop Transmission in Downlink NOMA. Adv. Electr. Electron. Eng. 2021, 19, 212–221. [Google Scholar] [CrossRef]
  4. Islam, S.M.R.; Avazov, N.; Dobre, O.A.; Kwak, K.s. Power-Domain Non-Orthogonal Multiple Access (NOMA) in 5G Systems: Potentials and Challenges. IEEE Commun. Surv. Tutorials 2017, 19, 721–742. [Google Scholar] [CrossRef]
  5. Nguyen, H.; Nguyen, T.N.; Minh, B.V.; Pham, T.H.T.; Le, A.T.; Voznak, M. Security-Reliability Analysis in CR-NOMA IoT Network Under I/Q Imbalance. IEEE Access 2023, 11, 119045–119056. [Google Scholar] [CrossRef]
  6. Le, A.T.; Tran, D.H.; Le, C.B.; Tin, P.T.; Nguyen, T.N.; Ding, Z.; Poor, H.V.; Voznak, M. Power Beacon and NOMA-Assisted Cooperative IoT Networks With Co-Channel Interference: Performance Analysis and Deep Learning Evaluation. IEEE Trans. Mob. Comput. 2024, 23, 7270–7283. [Google Scholar] [CrossRef]
  7. Thuan, D.H.; Nguyen, N.T.; Nguyen, X.T.; Hau, N.T.; Minh, B.V.; Nguyen, T.N. Uplink and downlink of energy harvesting NOMA system: Performance analysis. J. Inf. Telecommun. 2024, 8, 92–107. [Google Scholar] [CrossRef]
  8. Wan, D.; Huang, R.; Wen, M.; Chen, G.; Ji, F.; Li, J. A Simple Multicarrier Transmission Technique Combining Transmit Diversity and Data Multiplexing for Non-Orthogonal Multiple Access. IEEE Trans. Veh. Technol. 2021, 70, 7216–7220. [Google Scholar] [CrossRef]
  9. Nguyen, H.S.; Ly, T.T.H.; Nguyen, T.S.; Huynh, V.V.; Nguyen, T.L.; Voznak, M. Outage Performance Analysis and SWIPT Optimization in Energy-Harvesting Wireless Sensor Network Deploying NOMA. Sensors 2019, 19, 613. [Google Scholar] [CrossRef]
  10. Do, D.T.; Nguyen, T.T.T.; Nguyen, T.N.; Li, X.; Voznak, M. Uplink and downlink NOMA transmission using full-duplex UAV. IEEE Access 2020, 8, 164347–164364. [Google Scholar] [CrossRef]
  11. Ly, T.T.H.; Nguyen, H.S.; Nguyen, T.S.; Huynh, V.V.; Nguyen, T.L.; Voznak, M. Outage Probability Analysis in Relaying Cooperative Systems with NOMA Considering Power Splitting. Symmetry 2019, 11, 72. [Google Scholar] [CrossRef]
  12. Nguyen, V.V.; Tran, T.T.; Tri, V.V.; Huynh, V.V.; Nguyen, H.S.; Voznak, M. Power-splitting protocol non-orthogonal multiple access (NOMA) in 5G systems: Outage performance. In Proceedings of the 10th International Symposium on Information and Communication Technology, Hanoi Ha Long Bay, Vietnam, 4–6 December 2019; pp. 224–228. [Google Scholar] [CrossRef]
  13. Hoc, L.T.T.; Nguyen, H.S.; Ma, Q.P.; Huynh, V.V.; Nguyen, T.L.; Phuoc, H.T.; Voznak, M. Outage and Bit Error Probability Analysis in Energy Harvesting Wireless Cooperative Networks. Elektron. Ir Elektrotechnika 2019, 25, 69–74. [Google Scholar] [CrossRef]
  14. Van, H.T.; Nguyen, H.S.; Nguyen, T.S.; Huynh, V.V.; Nguyen, T.L.; Sevcik, L.; Voznak, M. Outage Performance Analysis of Non-Orthogonal Multiple Access with Time-Switching Energy Harvesting. Elektron. Ir Elektrotechnika 2019, 25, 85–91. [Google Scholar] [CrossRef]
  15. Phan, V.D.; Nguyen, B.C.; Hoang, T.M.; Nguyen, T.N.; Tran, P.T.; Minh, B.V.; Voznak, M. Performance of Cooperative Communication System With Multiple Reconfigurable Intelligent Surfaces Over Nakagami-m Fading Channels. IEEE Access 2022, 10, 9806–9816. [Google Scholar] [CrossRef]
  16. Quang, P.; Kien, N.; Duy, T.; An, N.; Tung, N.; Le, A.V. Performance evaluation of reconfigurable intelligent surface aided multi-hop relaying schemes with short packet communication. Adv. Electr. Electron. Eng. 2024, 22, 97–106. [Google Scholar] [CrossRef]
  17. Lin, Z.; Niu, H.; An, K.; Wang, Y.; Zheng, G.; Chatzinotas, S.; Hu, Y. Refracting RIS-Aided Hybrid Satellite-Terrestrial Relay Networks: Joint Beamforming Design and Optimization. IEEE Trans. Aerosp. Electron. Syst. 2022, 58, 3717–3724. [Google Scholar] [CrossRef]
  18. Niu, H.; Chu, Z.; Zhou, F.; Zhu, Z.; Zhen, L.; Wong, K.K. Robust Design for Intelligent Reflecting Surface-Assisted Secrecy SWIPT Network. IEEE Trans. Wirel. Commun. 2022, 21, 4133–4149. [Google Scholar] [CrossRef]
  19. Zhu, Z.; Li, Z.; Chu, Z.; Sun, G.; Hao, W.; Liu, P.; Lee, I. Resource Allocation for Intelligent Reflecting Surface Assisted Wireless Powered IoT Systems With Power Splitting. IEEE Trans. Wirel. Commun. 2022, 21, 2987–2998. [Google Scholar] [CrossRef]
  20. Yang, G.; Xu, X.; Liang, Y.C. Intelligent Reflecting Surface Assisted Non-Orthogonal Multiple Access. In Proceedings of the 2020 IEEE Wireless Communications and Networking Conference (WCNC), Seoul, Republic of Korea, 25–28 May 2020; pp. 1–6. [Google Scholar] [CrossRef]
  21. Le, A.T.; Hieu, T.D.; Nguyen, T.N.; Le, T.L.; Nguyen, S.Q.; Voznak, M. Physical layer security analysis for RIS-aided NOMA systems with non-colluding eavesdroppers. Comput. Commun. 2024, 219, 194–203. [Google Scholar] [CrossRef]
  22. Ding, Z.; Schober, R.; Poor, H.V. On the Impact of Phase Shifting Designs on IRS-NOMA. IEEE Wirel. Commun. Lett. 2020, 9, 1596–1600. [Google Scholar] [CrossRef]
  23. Mu, X.; Liu, Y.; Guo, L.; Lin, J.; Al-Dhahir, N. Exploiting Intelligent Reflecting Surfaces in NOMA Networks: Joint Beamforming Optimization. IEEE Trans. Wirel. Commun. 2020, 19, 6884–6898. [Google Scholar] [CrossRef]
  24. Elhattab, M.; Arfaoui, M.A.; Assi, C.; Ghrayeb, A. Reconfigurable Intelligent Surface Assisted Coordinated Multipoint in Downlink NOMA Networks. IEEE Commun. Lett. 2021, 25, 632–636. [Google Scholar] [CrossRef]
  25. Hou, T.; Liu, Y.; Song, Z.; Sun, X.; Chen, Y.; Hanzo, L. Reconfigurable Intelligent Surface Aided NOMA Networks. IEEE J. Sel. Areas Commun. 2020, 38, 2575–2588. [Google Scholar] [CrossRef]
  26. Zhang, C.; Yi, W.; Liu, Y.; Wang, Q. Multi-cell NOMA: Coherent Reconfigurable Intelligent Surfaces Model With Stochastic Geometry. In Proceedings of the ICC 2021—IEEE International Conference on Communications, Montreal, QC, Canada, 14–23 June 2021; pp. 1–6. [Google Scholar] [CrossRef]
  27. Ding, Z.; Vincent Poor, H. A Simple Design of IRS-NOMA Transmission. IEEE Commun. Lett. 2020, 24, 1119–1123. [Google Scholar] [CrossRef]
  28. Yue, X.; Liu, Y. Performance Analysis of Intelligent Reflecting Surface Assisted NOMA Networks. IEEE Trans. Wirel. Commun. 2022, 21, 2623–2636. [Google Scholar] [CrossRef]
  29. Xu, J.; Liu, Y.; Mu, X.; Dobre, O.A. STAR-RISs: Simultaneous Transmitting and Reflecting Reconfigurable Intelligent Surfaces. IEEE Commun. Lett. 2021, 25, 3134–3138. [Google Scholar] [CrossRef]
  30. Liu, Y.; Mu, X.; Xu, J.; Schober, R.; Hao, Y.; Poor, H.V.; Hanzo, L. STAR: Simultaneous Transmission and Reflection for 360° Coverage by Intelligent Surfaces. IEEE Wirel. Commun. 2021, 28, 102–109. [Google Scholar] [CrossRef]
  31. Wu, C.; Liu, Y.; Mu, X.; Gu, X.; Dobre, O.A. Coverage Characterization of STAR-RIS Networks: NOMA and OMA. IEEE Commun. Lett. 2021, 25, 3036–3040. [Google Scholar] [CrossRef]
  32. Mu, X.; Liu, Y.; Guo, L.; Lin, J.; Schober, R. Simultaneously Transmitting and Reflecting (STAR) RIS Aided Wireless Communications. IEEE Trans. Wirel. Commun. 2022, 21, 3083–3098. [Google Scholar] [CrossRef]
  33. Ni, W.; Liu, Y.; Eldar, Y.C.; Yang, Z.; Tian, H. STAR-RIS Integrated Nonorthogonal Multiple Access and Over-the-Air Federated Learning: Framework, Analysis, and Optimization. IEEE Internet Things J. 2022, 9, 17136–17156. [Google Scholar] [CrossRef]
  34. Hou, T.; Wang, J.; Liu, Y.; Sun, X.; Li, A.; Ai, B. A Joint Design for STAR-RIS Enhanced NOMA-CoMP Networks: A Simultaneous-Signal-Enhancement-and-Cancellation-Based (SSECB) Design. IEEE Trans. Veh. Technol. 2022, 71, 1043–1048. [Google Scholar] [CrossRef]
  35. Niu, H.; Chu, Z.; Zhou, F.; Xiao, P.; Al-Dhahir, N. Weighted Sum Rate Optimization for STAR-RIS-Assisted MIMO System. IEEE Trans. Veh. Technol. 2022, 71, 2122–2127. [Google Scholar] [CrossRef]
  36. Niu, H.; Chu, Z.; Zhou, F.; Zhu, Z. Simultaneous Transmission and Reflection Reconfigurable Intelligent Surface Assisted Secrecy MISO Networks. IEEE Commun. Lett. 2021, 25, 3498–3502. [Google Scholar] [CrossRef]
  37. Zuo, J.; Liu, Y.; Ding, Z.; Song, L.; Poor, H.V. Joint Design for Simultaneously Transmitting and Reflecting (STAR) RIS Assisted NOMA Systems. IEEE Trans. Wirel. Commun. 2023, 22, 611–626. [Google Scholar] [CrossRef]
  38. Zhang, C.; Yi, W.; Liu, Y.; Ding, Z.; Song, L. STAR-IOS Aided NOMA Networks: Channel Model Approximation and Performance Analysis. IEEE Trans. Wirel. Commun. 2022, 21, 6861–6876. [Google Scholar] [CrossRef]
  39. Xie, Z.; Yi, W.; Wu, X.; Liu, Y.; Nallanathan, A. STAR-RIS Aided NOMA in Multicell Networks: A General Analytical Framework With Gamma Distributed Channel Modeling. IEEE Trans. Commun. 2022, 70, 5629–5644. [Google Scholar] [CrossRef]
  40. Liu, Y.; Mu, X.; Schober, R.; Poor, H.V. Simultaneously Transmitting and Reflecting (STAR)-RISs: A Coupled Phase-Shift Model. In Proceedings of the ICC 2022—IEEE International Conference on Communications, Seoul, Republic of Korea, 16–20 May 2022; pp. 2840–2845. [Google Scholar] [CrossRef]
  41. Aldababsa, M.; Khaleel, A.; Basar, E. Simultaneous transmitting and ReflectingIntelligent surfaces-empowered NOMA networks. arXiv 2021, arXiv:2110.05311. [Google Scholar] [CrossRef]
  42. Zuo, J.; Liu, Y.; Ding, Z.; Wang, X. Uplink noma for star-ris networks. arXiv 2021, arXiv:2110.05686. [Google Scholar] [CrossRef]
  43. Zhang, Z.; Chen, J.; Liu, Y.; Wu, Q.; He, B.; Yang, L. On the Secrecy Design of STAR-RIS Assisted Uplink NOMA Networks. IEEE Trans. Wirel. Commun. 2022, 21, 11207–11221. [Google Scholar] [CrossRef]
  44. Wu, C.; Mu, X.; Liu, Y.; Gu, X.; Wang, X. Resource Allocation in STAR-RIS-Aided Networks: OMA and NOMA. IEEE Trans. Wirel. Commun. 2022, 21, 7653–7667. [Google Scholar] [CrossRef]
  45. Zhao, B.; Zhang, C.; Yi, W.; Liu, Y. Ergodic Rate Analysis of STAR-RIS Aided NOMA Systems. IEEE Commun. Lett. 2022, 26, 2297–2301. [Google Scholar] [CrossRef]
  46. Do, D.T. Optimal Energy Har vesting Scheme for Power Beacon-Assisted Wireless-Powered Networks. Indones. J. Electr. Eng. Comput. Sci. 2017, 7, 802–808. [Google Scholar]
  47. Tin, P.T.; Dinh, B.H.; Nguyen, T.N.; Ha, D.H.; Trang, T.T. Power Beacon-Assisted Energy Harvesting Wireless Physical Layer Cooperative Relaying Networks: Performance Analysis. Symmetry 2020, 12, 106. [Google Scholar] [CrossRef]
  48. Nguyen, T.N.; Tu, L.T.; Fazio, P.; Van Chien, T.; Le, C.V.; Binh, H.T.T.; Voznak, M. On the Dilemma of Reliability or Security in Unmanned Aerial Vehicle Communications Assisted by Energy Harvesting Relaying. IEEE J. Sel. Areas Commun. 2024, 42, 52–67. [Google Scholar] [CrossRef]
  49. Minh, B.; Tran, M.; Phan, V.; Hieu, N. D2D Communication Network with the Assistance of Power Beacon under the Impact of Co-channel Interferences and Eavesdropper: Performance Analysis. Adv. Electr. Electron. Eng. 2024, 21, 351–359. [Google Scholar] [CrossRef]
  50. Vu, T.H.; Kim, S. Performance Evaluation of Power-Beacon-Assisted Wireless-Powered NOMA IoT-Based Systems. IEEE Internet Things J. 2021, 8, 11655–11665. [Google Scholar] [CrossRef]
  51. Do, D.T.; Nguyen, M.S.V.; Nguyen, T.N.; Li, X.; Choi, K. Enabling Multiple Power Beacons for Uplink of NOMA-Enabled Mobile Edge Computing in Wirelessly Powered IoT. IEEE Access 2020, 8, 148892–148905. [Google Scholar] [CrossRef]
  52. Zhang, Y.; Feng, S.; Tang, W. Performance Analysis and Optimization for Power Beacon-Assisted Wireless Powered Cooperative NOMA Systems. IEEE Access 2020, 8, 198436–198450. [Google Scholar] [CrossRef]
  53. Lei, J.; Zhang, T.; Mu, X.; Liu, Y. NOMA for STAR-RIS Assisted UAV Networks. IEEE Trans. Commun. 2024, 72, 1732–1745. [Google Scholar] [CrossRef]
  54. Alanazi, F. Intelligent reflecting surfaces with energy harvesting for Nakagami fading channels. Telecommun. Syst. 2021, 78, 351–361. [Google Scholar] [CrossRef]
  55. Zhang, B.; Yang, K.; Wang, K.; Zhang, G. Performance Analysis of RIS-Assisted Wireless Communications With Energy Harvesting. IEEE Trans. Veh. Technol. 2023, 72, 1325–1330. [Google Scholar] [CrossRef]
  56. Alanazi, F. Non orthogonal multiple access with energy harvesting using reconfigurable intelligent surfaces for Rayleigh channels. Wirel. Pers. Commun. 2022, 12, 2161–2181. [Google Scholar] [CrossRef]
  57. Alakoca, H.; Babaei, M.; Durak-Ata, L.; Basar, E. RIS-Empowered Non-Linear Energy Harvesting Communications Over Nakagami-m Channels. IEEE Commun. Lett. 2022, 26, 2215–2219. [Google Scholar] [CrossRef]
  58. Badarneh, O.S. Statistics of the Product of Two α-F Variates With Applications. IEEE Commun. Lett. 2021, 25, 1761–1765. [Google Scholar] [CrossRef]
  59. Tegos, S.A.; Tyrovolas, D.; Diamantoulakis, P.D.; Liaskos, C.K.; Karagiannidis, G.K. On the Distribution of the Sum of Double-Nakagami-m Random Vectors and Application in Randomly Reconfigurable Surfaces. IEEE Trans. Veh. Technol. 2022, 71, 7297–7307. [Google Scholar] [CrossRef]
  60. Silva, H.S.; Lourenço, G.B.L.C.; Queiroz, W.J.L.; Oliveira, A.S.R.; Madeiro, F.; Badarneh, O.S.; Dias, U.S.; de Souza, R.A.A. On the Sum of α-F and Double α-F VariatesWith Application to MRC and RIS. IEEE Commun. Lett. 2022, 26, 2894–2898. [Google Scholar] [CrossRef]
  61. Do, T.N.; Kaddoum, G.; Nguyen, T.L.; da Costa, D.B.; Haas, Z.J. Multi-RIS-Aided Wireless Systems: Statistical Characterization and Performance Analysis. IEEE Trans. Commun. 2021, 69, 8641–8658. [Google Scholar] [CrossRef]
  62. Gradshteyn, I.S.; Ryzhik, I.M. Table of Integrals, Series, and Products; Academic Press: Cambridge, MA, USA, 2014. [Google Scholar]
  63. Yacoub, M.D. The κ-μ distribution and the η-μ distribution. IEEE Antennas Propag. Mag. 2007, 49, 68–81. [Google Scholar] [CrossRef]
  64. Li, X.; Zheng, Y.; Zeng, M.; Liu, Y.; Dobre, O.A. Enhancing Secrecy Performance for STAR-RIS NOMA Networks. IEEE Trans. Veh. Technol. 2023, 72, 2684–2688. [Google Scholar] [CrossRef]
  65. Goldsmith, A. Wireless Communications; Cambridge University Press: Cambridge, UK, 2005. [Google Scholar]
  66. Abramowitz, M.; Stegun, I.A. Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables; US Government Printing Office: Washington, DC, USA, 1948; Volume 55.
  67. An, K.; Lin, M.; Ouyang, J.; Zhu, W.P. Secure Transmission in Cognitive Satellite Terrestrial Networks. IEEE J. Sel. Areas Commun. 2016, 34, 3025–3037. [Google Scholar] [CrossRef]
  68. Nguyen, T.N.; Tran, D.H.; Chien, T.V.; Phan, V.D.; Voznak, M.; Tin, P.T.; Chatzinotas, S.; Ng, D.W.K.; Poor, H.V. Security–Reliability Tradeoff Analysis for SWIPT- and AF-Based IoT Networks With Friendly Jammers. IEEE Internet Things J. 2022, 9, 21662–21675. [Google Scholar] [CrossRef]
  69. Nguyen, T.N.; Tu, L.T.; Tran, D.H.; Phan, V.D.; Voznak, M.; Chatzinotas, S.; Ding, Z. Outage Performance of Satellite Terrestrial Full-Duplex Relaying Networks With co-Channel Interference. IEEE Wirel. Commun. Lett. 2022, 11, 1478–1482. [Google Scholar] [CrossRef]
  70. Nguyen, T.N.; Duy, T.T.; Tran, P.T.; Voznak, M.; Li, X.; Poor, H.V. Partial and Full Relay Selection Algorithms for AF Multi-Relay Full-Duplex Networks With Self-Energy Recycling in Non-Identically Distributed Fading Channels. IEEE Trans. Veh. Technol. 2022, 71, 6173–6188. [Google Scholar] [CrossRef]
  71. Huynh, V.V.; Tan-Loc, N.; Quoc-Phu, M.; Sevcik, L.; Nguyen, H.S.; Voznak, M. Energy Efficiency Maximization of Two-Time-Slot and Three-Time-Slot Two-Way Relay-Assisted Device-to-Device Underlaying Cellular Networks. Energies 2020, 13, 3422. [Google Scholar] [CrossRef]
  72. Al-qudah, Z.; Darabkh, K. A simple Encoding Scheme to Achieve the Capacity of Half-Duplex Relay Channel. Adv. Electr. Electron. Eng. 2022, 20, 33–42. [Google Scholar] [CrossRef]
  73. Paris, J.F. Statistical Characterization of κμ Shadowed Fading. IEEE Trans. Veh. Technol. 2014, 63, 518–526. [Google Scholar] [CrossRef]
Figure 1. STAR-RIS-assisted NOMA system.
Figure 1. STAR-RIS-assisted NOMA system.
Sensors 24 06148 g001
Figure 2. PDF and CDF for different numbers of STAR-RIS elements ( Q = 8 , 12 , 16 ), with κ = 1 and μ = 4 .
Figure 2. PDF and CDF for different numbers of STAR-RIS elements ( Q = 8 , 12 , 16 ), with κ = 1 and μ = 4 .
Sensors 24 06148 g002
Figure 3. Comparison of OPs in a classic channel with varying numbers of STAR-RIS elements Q = 16 .
Figure 3. Comparison of OPs in a classic channel with varying numbers of STAR-RIS elements Q = 16 .
Sensors 24 06148 g003
Figure 4. OPs of two users versus ρ with varying numbers of STAR-RIS elements Q = 8 , 32 , κ = 0 , and μ = 2 .
Figure 4. OPs of two users versus ρ with varying numbers of STAR-RIS elements Q = 8 , 32 , κ = 0 , and μ = 2 .
Sensors 24 06148 g004
Figure 5. OP versus the target rates with Q = 16 , a 2 = 0.9 , a 1 = 0.1 , κ = 2 , μ = 0 , = 0.01 , and ρ = 11 dB.
Figure 5. OP versus the target rates with Q = 16 , a 2 = 0.9 , a 1 = 0.1 , κ = 2 , μ = 0 , = 0.01 , and ρ = 11 dB.
Sensors 24 06148 g005
Figure 6. OP versus a 2 with = 0.05 , κ = 2 , μ = 1 , R 1 = 1 , R 2 = 0.5 , and ρ = 5 dB.
Figure 6. OP versus a 2 with = 0.05 , κ = 2 , μ = 1 , R 1 = 1 , R 2 = 0.5 , and ρ = 5 dB.
Sensors 24 06148 g006
Figure 7. OP of the destination versus α with = 0.01 , κ = 2 , μ = 1 , a 2 = 0.9 , a 1 = 0.1 , R 1 = 1 , R 2 = 0.5 , and ρ = 5 dB.
Figure 7. OP of the destination versus α with = 0.01 , κ = 2 , μ = 1 , a 2 = 0.9 , a 1 = 0.1 , R 1 = 1 , R 2 = 0.5 , and ρ = 5 dB.
Sensors 24 06148 g007
Figure 8. OP of the STAR-RIS-aided NOMA: Relay networks versus ρ with = 0.01 , κ = 2 , μ = 0 , a 2 = 0.8 , a 1 = 0.2 , R 1 = 2 , R 2 = 1 , and Q = 8 .
Figure 8. OP of the STAR-RIS-aided NOMA: Relay networks versus ρ with = 0.01 , κ = 2 , μ = 0 , a 2 = 0.8 , a 1 = 0.2 , R 1 = 2 , R 2 = 1 , and Q = 8 .
Sensors 24 06148 g008
Figure 9. Ergodic rate versus ρ with = 0.01 , κ = 0 , μ = 4 , a 2 = 0.8 , and a 1 = 0.2 .
Figure 9. Ergodic rate versus ρ with = 0.01 , κ = 0 , μ = 4 , a 2 = 0.8 , and a 1 = 0.2 .
Sensors 24 06148 g009
Figure 10. Ergodic rate versus Q with = 0.01 , κ = 1 , μ = 3 , a 2 = 0.8 , a 1 = 0.2 , and ρ = 3 dB.
Figure 10. Ergodic rate versus Q with = 0.01 , κ = 1 , μ = 3 , a 2 = 0.8 , a 1 = 0.2 , and ρ = 3 dB.
Sensors 24 06148 g010
Figure 11. The average SERs of the considered STAR-RIS assisted NOMA system versus ρ for BPSK and 4QAM modulations with = 0.05 , κ = 1 , μ = 2 , a 2 = 0.8 , a 1 = 0.2 , and Q = 8 .
Figure 11. The average SERs of the considered STAR-RIS assisted NOMA system versus ρ for BPSK and 4QAM modulations with = 0.05 , κ = 1 , μ = 2 , a 2 = 0.8 , a 1 = 0.2 , and Q = 8 .
Sensors 24 06148 g011
Table 2. Main parameters for our simulations.
Table 2. Main parameters for our simulations.
ParameterNotationValues
Total reflecting elementsQ32
Power allocation factors a 1 , a 2 0.2 , 0.8
Target rates used to decode x 1 and x 2 R 1 , R 2 2 , 1 BPCU
The fading parameterm2
The energy conversion efficiency η 1
The fraction of the block time α 0.5
The efficiency of SIC for x 2 0.01
Table 3. Common fading distributions derived from the κ μ distribution [73].
Table 3. Common fading distributions derived from the κ μ distribution [73].
ChannelParameters of the κ μ Distribution
One-sided Gaussian μ ̲ = 0.5 , κ ̲ 0
Rayleigh μ ̲ = 1 , κ ̲ 0
Nakagami-m μ ̲ = m , κ ̲ 0
Rician with parameter K μ ̲ = 1 , κ ̲ = K
κ μ μ ̲ = μ , κ ̲ = κ
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Minh, B.V.; Tran, P.T.; Pham, T.-H.T.; Le, A.-T.; Le, S.-P.; Partila, P. Statistics of the Sum of Double Random Variables and Their Applications in Performance Analysis and Optimization of Simultaneously Transmitting and Reflecting Reconfigurable Intelligent Surface-Assisted Non-Orthogonal Multi-Access Systems. Sensors 2024, 24, 6148. https://doi.org/10.3390/s24186148

AMA Style

Minh BV, Tran PT, Pham T-HT, Le A-T, Le S-P, Partila P. Statistics of the Sum of Double Random Variables and Their Applications in Performance Analysis and Optimization of Simultaneously Transmitting and Reflecting Reconfigurable Intelligent Surface-Assisted Non-Orthogonal Multi-Access Systems. Sensors. 2024; 24(18):6148. https://doi.org/10.3390/s24186148

Chicago/Turabian Style

Minh, Bui Vu, Phuong T. Tran, Thu-Ha Thi Pham, Anh-Tu Le, Si-Phu Le, and Pavol Partila. 2024. "Statistics of the Sum of Double Random Variables and Their Applications in Performance Analysis and Optimization of Simultaneously Transmitting and Reflecting Reconfigurable Intelligent Surface-Assisted Non-Orthogonal Multi-Access Systems" Sensors 24, no. 18: 6148. https://doi.org/10.3390/s24186148

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop