This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Novel Outage-Aware NOMA Protocol for Secrecy Fairness Maximization Among Untrusted Users

Sapna Thapar,  Deepak Mishra, 
and Ravikant Saini
Copyright (c) 2015 IEEE. Personal use of this material is permitted. However, permission to use this material for any other purposes must be obtained from the IEEE by sending a request to pubs-permissions@ieee.org.S. Thapar and R. Saini are with the Department of Electrical Engineering, Indian Institute of Technology Jammu, Jammu, Jammu &\& Kashmir 181 221, India (e-mail: 2018ree0019@iitjammu.ac.in; ravikant.saini@iitjammu.ac.in). D. Mishra is with the School of Electrical Engineering and Telecommunications, University of New South Wales Sydney, NSW 2052, Australia (e-mail: d.mishra@unsw.edu.au).This research work was supported by the Tata Consultancy Services Research Scholar Program (TCS RSP) Fellowship.A preliminary version [1] of this work has been published at IEEE GLOBECOM, Hawaii, USA, Dec. 2019.
Abstract

Observing the significance of spectrally-efficient secure non-orthogonal multiple access (NOMA), this paper proposes a novel quality of service (QoS) aware secure NOMA protocol that maximizes secrecy fairness among untrusted users. Considering a base station (BS) and two users, a novel decoding order is designed that provides security to both users. With the objective of ensuring secrecy fairness between users, while satisfying their QoS requirements under BS transmit power budget constraint, we explore the problem of minimizing the maximum secrecy outage probability (SOP). Closed-form expression of pair outage probability (POP) and optimal power allocation (PA) minimizing POP are obtained. To analyze secrecy performance, analytical expressions of SOP for both users are derived, and individual SOP minimization problems are solved using concept of generalized-convexity. High signal-to-noise ratio approximation of SOP and asymptotically optimized solution minimizing this approximation is also found. Furthermore, global-optimal solution from secrecy fairness standpoint is obtained at low computational complexity, and tight approximation is derived to get analytical insights. Numerical results present useful insights on globally optimized PA which ensure secrecy fairness and provide performance gain of about 55.12%55.12\%, 69.30%69.30\%, and 19.11%19.11\%, respectively, compared to fixed PA and individual users’ optimal PAs. Finally, a tradeoff between secrecy fairness performance and QoS demands is presented.

Index Terms- 5G communications, non-orthogonal multiple access, physical layer security, power allocation, optimization.

I Introduction

Exponentially increasing network traffic, imposed by internet-enabled applications, poses severe challenges of supporting massive connections loaded with high data rate demands for upcoming fifth generation (5G) communication networks. Non-orthogonal multiple access (NOMA) has been recognized as a potential breakthrough because of the possibility of accommodating a number of users within the same subcarrier allocation [2]. This exhibits a significant spectral efficiency enhancement compared to traditional orthogonal multiple access techniques. But, the broadcast nature of communication at the transmitter makes NOMA vulnerable to the eavesdropping. Therefore, research on security issues in NOMA networks is of great significance. The amalgamation of NOMA and physical layer security (PLS) has been observed as a promising solution to provide spectrally-efficient secure wireless communication [3]. Despite merits, the design process has secrecy challenge of wiretapping in the presence of untrusted NOMA users.

I-A Related Art

To address the demand for massive connections, modern wireless networks strive for better spectral efficiency. Motivated by the potential of NOMA in providing spectrally-efficient networks, survey papers [2], [4] have summarized recent research contributions in power-domain NOMA. Besides, the broadcast nature of wireless transmission leads to security issues on the information-carrying signal. Triggered by these security concerns, researchers have considered the PLS technique which was first recommended by Wyner [5] as a complement to cryptographic approaches. To highlight the capabilities of PLS, [6] has provided the latest survey of PLS research on diverse 5G technologies, such as multiple input multiple output, full-duplex, and millimeter wave.

Even though a great amount of research has been done in NOMA and PLS independently, researchers have recently focused attention on using PLS as a security measure in NOMA. An investigation of PLS in large-scale networks by invoking stochastic geometry has been done in [3], where a base station (BS) communicates with randomly distributed NOMA users and eavesdroppers. The authors have proposed a protected zone around BS to maintain an eavesdropper free zone. Secure NOMA transmission for multiple users against eavesdropper in single-input single-output network has been studied in [7], where the authors have derived a closed-form expression of optimal power allocation (PA) that maximizes sum secure rate of NOMA users. In [8], the authors have investigated the optimal decoding order, individual transmission rates, and PA to each user to design secure NOMA system against external eavesdropper under secrecy outage constraint. The authors have proved that optimal decoding order for secure NOMA scheme against external eavesdropper is same as that for conventional NOMA. Secrecy of a cooperative NOMA system with a decode-and-forward and an amplify-and-forward relay against external eavesdropper has been analyzed in [9]. The authors in [10] have proposed a NOMA-inspired jamming and relaying scheme for enhancing the PLS of untrusted relay networks. In [11], the authors considered a multiple-input single-output NOMA cognitive relay radio network in the presence of multiple eavesdroppers and proposed an artificial noise (AN) aided cooperative jamming scheme for secure transmission in considered network.

It is noteworthy that the aforementioned works have focused on securing NOMA against external eavesdropper only. Besides external eavesdroppers, secure data transmission among NOMA users is much more challenging due to the underlying SIC based decoding at receivers. Actually, due to the sharing of same resource blocks among users in NOMA systems, users may not trust each other in real scenario, and hence at least some level of secrecy must be provided against internal eavesdropping. Based on the idea, assuming far users as untrusted, sum secrecy rate of trusted near users is investigated in [12] for a multiple-input single-output NOMA system. In [13] also, a NOMA system is considered where near user is assumed to be trusted, whereas far user as untrusted, and SOP of trusted user against untrusted user has been investigated. Note that in [12] and [13], it is assumed that the untrusted far user first decodes its own data and then can access the message of the trusted near user with the aid of the SIC technique.

I-B Research Gap and Motivation

Note that existing works [3], [7]-[11] have considered potential PLS techniques such as AN aided strategy [3], [11], optimal PA [7], [8], and cooperative relaying [9], [10] to enhance secrecy of NOMA transmission against external eavesdroppers assuming all NOMA users as trusted. Since NOMA users share same resource block and have access to decode the message of other paired user to perform SIC, there exists an inherent security issue in NOMA in the case of untrusted users. Therefore, at least some level of secrecy must be provided against internal eavesdropping. Note that the untrusted scenario is a hostile situation, where users have no mutual trust, and each user focuses on achieving secure communication from BS in the presence of other untrusted users, which leads to more complex and constrained resource allocations [14], [15]. It is noteworthy that the existing works [12] and [13] only considered the security problem of trusted near user in a NOMA system against untrusted far untrusted user. However, near user also decodes the data of far user, which causes a serious security risk for far user in the case of untrusted near user. Therefore, from security point of view, it is more desirable to design a secure NOMA system for worst-case scenario, where both near and far users are assumed to be untrusted. Thus, with the objective of designing a smart secure communication network, our work focuses on investigating whether secure communication is possible in NOMA in the presence of untrusted users.

In this direction, [16] considered symbol-level SIC (SLSIC) receiver where other user’s signal is demodulated but not decoded to perform SIC. In contrast, codeword-level SIC (CLSIC) receiver exists in literature [17], where other user’s signal is demodulated and decoded. Since channel decoding is conducted during signal detection in CLSIC receiver, the probability of successful recovery of other user’s signal increases in comparison to SLSIC, and hence, the impact of error propagation can be reduced drastically [17]. Nevertheless, securing all users’ data in an untrusted scenario from a decoding perspective, which is a better system design aspect, has not been considered yet in literature.

It also needs to noted that when both users are assumed to be untrusted in NOMA, one might think that if any of these users is capable of decoding other user’s signal then security and privacy of decoded users are compromised. But, decoding of users’ signals can even happen in orthogonal multiple access because of the broadcast nature of wireless channels [18]. Actually, we focus on achieving secure communication by utilizing the concept of PLS. According to the definition of PLS, information leakage of the legitimate user may occur only when data rate over the desired link is lesser than that of the potential eavesdropper’s link because the secrecy rate for a legitimate user is defined as the difference of the rates when a legitimate user decodes itself, and the rate that another user achieves while decoding data of legitimate user [5], [6]. Therefore, considering the definition of PLS into account for an untrusted NOMA system, our work is focused on designing such a secure NOMA transmission protocol that can be feasible in providing positive secrecy rate for all NOMA users.

Taking this challenge into account, a decoding order has been proposed for a two-user untrusted NOMA scenario in the preliminary version [1] of this work that is feasible in providing positive secrecy rate for both near and far NOMA users. As an extension of [1], in this work, we study all possible decoding orders for a two user NOMA system including conventional one and investigate the best decoding order under secrecy constraint for both users. Besides, in [1], we have considered ideal SIC receivers in which interference from the decoded user is completely removed while decoding the later user. Though better spectral efficiency can be achieved through perfect SIC, but it is not realistic due to various implementation issues such as decoding error and complexity scaling [4], [19]. Therefore, a practical scenario would be, when residual interference from the inaccurately decoded user is considered while decoding later user. To this end, we study secure NOMA to provide positive secrecy rates at all untrusted users with consideration of imperfect SIC based decoding which, to the best of our knowledge, is an open problem.

It should be emphasized that, with the vision of ensuring secure communication to all users, secrecy performance of the far user against the untrusted near user, and rather secrecy fairness between these users cannot be ignored. Fulfilling users’ quality of service (QoS) requirement is also an essential parameter for spectrally-efficient communication. Also, PA to users may play deciding role in obtaining optimal pair outage and secrecy outage performance of a system. Inspired by these solid observations, this paper focuses on investigating numerical and analytical global-optimal PA solutions for a secure NOMA system with untrusted users from a secrecy fairness maximization point of view while satisfying users’ QoS requirements.

I-C Key Contributions

The contributions of this work are summarized as follows:

  • Considering a NOMA system with one BS, two untrusted users and imperfect SIC model, a novel optimal decoding order is proposed that provides security to both users.

  • Pair outage probability (POP), as a QoS measure, has been derived analytically. Optimal PA minimizing POP has been obtained using generalized-convexity of POP.

  • Analytical expressions of SOPs have been derived for both users. SOP minimization problems have been solved using pseudoconvexity of SOP at both users and optimal PAs are obtained numerically. Asymptotic approximations of SOP, and optimal PA for both users are also derived to get analytical insights.

  • To maximize the secrecy fairness between users while satisfying users’ QoS demands, we formulate the problem of minimizing the maximum SOP between users under the POP and PA constraints. Global-optimal PA solution is obtained using a low complexity algorithm. Tight analytical approximation for optimal PA is also derived.

  • Extensive simulations are conducted that validate the accuracy of analysis, present insights on optimal performance, and evaluate performance gains by the proposed solution. Tradeoff between secrecy fairness performance of system and users’ QoS demands is also investigated.

I-D Organization

The system model, proposed NOMA protocol and investigation on optimal decoding order have been presented in Section II. In Section III, we provide analysis of pair outage performance of the proposed system and derive an optimal PA policy to minimize POP. In section IV, we present SOP analysis for both users, along with their closed-form asymptotic expressions, and also investigate their optimality. Analysis of optimal PA for secrecy fairness maximization between untrusted users under POP constraint is described in Section V. Performance evaluation of proposed analytical model via simulation results is discussed in Section VI, and Section VII presents concluding remarks and future directions.

II Secure NOMA Protocol for Untrusted Users

In this section, we first present the system model, which is followed by the NOMA transmission protocol. Then, an optimal decoding order to secure untrusted NOMA is investigated.

Refer to caption
Figure 1: Illustration of downlink secure NOMA protocol with proposed optimal decoding order for two untrusted users where decoding order is changed for the far user in comparison to the conventional approach.

II-A System Model

We consider the downlink NOMA system with one BS and two untrusted users (Fig. 1). The near and far users are denoted as U11 and U22, respectively. Our consideration for two-user model is motivated by two reasons: firstly, NOMA system is an interference limited system because co-channel interference is strong in NOMA due to sharing of same resource by multiple users [20], [21]; and secondly, the implementation complexity at transmitter and receiver sides increases with an increase in number of users because the users perform SIC to remove messages of other users before decoding their own messages [22]. Note that due to these reasons, asking all the users in the system to perform NOMA jointly is not preferable in practice [21] and therefore, the users can be divided into multiple groups, where NOMA is implemented within each group. The users are ordered according to their distance to BS. As a result, U11 and U22 are considered as strong and weak users, respectively [23], [24]. The Rayleigh fading channel gain coefficients from BS to Uii is denoted by hih_{i} for i{1,2}i\in\{1,2\}. Channels between BS and users are assumed to be independent and suffer both small scale fading and path loss effects, such that channel power gains |hi|2|h_{i}|^{2} follows exponential distribution with mean parameter λi=Lpdin\lambda_{i}=L_{p}d_{i}^{-n}. Here LpL_{p} is path loss constant, nn is path loss exponent, and did_{i} is the distance from BS to Uii. The channel power gains have been sorted such that |h1|2>|h2|2|h_{1}|^{2}>|h_{2}|^{2} [23], [24]. All three nodes are assumed to be equipped with a single antenna [13]. PtP_{t} denotes total power transmitted by BS to users, and α\alpha represents the PA coefficient denoting the fraction of PtP_{t} allocated to U11. Remaining (1α)Pt(1-\alpha)P_{t} is allocated to U22.

II-B Proposed Secure NOMA Protocol

In NOMA, the message signals dedicated to the users are superimposed at the BS and then transmitted to the users. The signal transmitted by BS is αPtx1+(1α)Ptx2\sqrt{\alpha P_{t}}x_{1}+\sqrt{(1-\alpha)P_{t}}x_{2} [13] where x1x_{1} and x2x_{2} are unit power signals which contains the message required by U11 and U22, respectively. The signals received by U11 and U22, respectively, from BS are given as

y1=h1(αPtx1+(1α)Ptx2)+n1,y_{1}=h_{1}(\sqrt{\alpha P_{t}}x_{1}+\sqrt{(1-\alpha)P_{t}}x_{2})+n_{1}, (1)
y2=h2(αPtx1+(1α)Ptx2)+n2,y_{2}=h_{2}(\sqrt{\alpha P_{t}}x_{1}+\sqrt{(1-\alpha)P_{t}}x_{2})+n_{2}, (2)

where n1n_{1} and n2n_{2} denote zero mean additive white Gaussian noise with variance σ2\sigma^{2} at U11 and U22, respectively. ρt=ΔPt\rho_{t}\stackrel{{\scriptstyle\Delta}}{{=}}P_{t}/σ2\sigma^{2} denotes BS transmit signal-to-noise ratio (SNR).

After obtaining received signals, two steps are followed at receivers in conventional decoding order: (1) far user decodes its own information signal first and after obtaining its own signal it may decode the information signal of the near user using SIC in case of the far user is an untrusted user [12], [13]; (2) near user first decodes the signal intended for far user to perform SIC, and then decodes its own message. As a result, near and far users both have access to the signals of far and near user, respectively, which is a critical issue in secure communication with untrusted users. Note that the SIC can be performed by all users in a system because SIC on the receivers’ side is actually a physical layer capability enabling the receiver to decode packets that arrive collectively. Thus, far user may also decode the information signal of near user using SIC which has been extensively studied in literature also [12], [13], [25], [26]. In this work, we consider imperfect SIC model, where decoding error may occur at both users, resulting in residual interference from the incompletely decoded user after SIC.

Since all users focus to safe their own data from others in untrusted scenario, next we investigate a decoding order that can provide security at both users. Even the conventional decoding order has also not yet been investigated in the literature to ensure positive secrecy for both users in an untrusted NOMA system. Hence, we discuss all possible decoding orders including conventional one to investigate the best decoding order under secrecy constraint.

II-C Optimal Decoding Order from Secrecy Perspective

In secure NOMA, U11 needs to be protected from U22 and, vice-versa. Decoding order identifies whether any user decodes its own data first or other user’s data. Note that both U11 and U22 are the multiplexed NOMA users which themselves are the part of the system. Now since BS decides the decoding order for both the users of the system, decoding order can also be changed [25]. Thus, for two users’ case, four decoding orders are possible. Let us denote the decoding order as (i,j)(i,j), where ii and jj, respectively, denotes users, i.e., Uii and Ujj (for i{1,2},j{1,2}i\in\{1,2\},j\in\{1,2\}) whose data will be decoded first by U11 and U22. For example, (2,1)(2,1) means U11 and U22 will decode other user’s data first, and then decode its own data. Specifically, the four possible cases are: (a) (1,1)(1,1), (b) (1,2)(1,2), (c) (2,1)(2,1) and (d) (2,2)(2,2). Under secrecy considerations, the idea is to study whether it is possible to achieve positive secrecy rate at both users. In this regard, investigation of feasible decoding orders out of total possible decoding orders is presented by Theorem 11.

Theorem 1

Positive secrecy rate for both near and far users in untrusted NOMA can be obtained in decoding orders (2,1)(2,1), (1,2)(1,2), and (1,1)(1,1).

Proof:

We consider each case of decoding orders one by one and investigate the feasibility under positive secrecy constraint. We first consider the conventional decoding order (2,2)(2,2), where each user first decodes U22’s data. Then, the proof continues with a consideration of remaining cases, i.e., (2,1)(2,1), (1,2)(1,2) and (1,1)(1,1). Adopting the imperfect SIC model, we use βij\beta_{ij} to denote the residual interference from the imperfectly decoded Uii when data of Uii is decoded by Ujj.

Case 1: Considering conventional decoding order (2,2)(2,2) [13], U22 first decodes its own signal by considering other signal as noise. After decoding its own signal, it may decode the signal of near user using SIC. Then, U11 decodes signal associated to U22, applies SIC, and then decode its own signal from the remaining signal. Let Γij\Gamma_{ij} denote the received signal-to-interference-plus-noise-ratio at Uii when data of Uii is decoded by Ujj. Various Γij\Gamma_{ij} can be given as [13]

Γ21=(1α)|h1|2α|h1|2+1ρt,Γ11=α|h1|2β21+1ρt,\Gamma_{21}=\frac{(1-\alpha)|h_{1}|^{2}}{\alpha|h_{1}|^{2}+\frac{1}{\rho_{t}}},\quad\Gamma_{11}=\frac{\alpha|h_{1}|^{2}}{\beta_{21}+\frac{1}{\rho_{t}}}, (3)
Γ22=(1α)|h2|2α|h2|2+1ρt,Γ12=α|h2|2β22+1ρt.\Gamma_{22}=\frac{(1-\alpha)|h_{2}|^{2}}{\alpha|h_{2}|^{2}+\frac{1}{\rho_{t}}},\quad\Gamma_{12}=\frac{\alpha|h_{2}|^{2}}{\beta_{22}+\frac{1}{\rho_{t}}}. (4)

The achievable secrecy rate Rs1R_{s1} of U11 can be given as

Rs1=R11R12,R_{s1}=R_{11}-R_{12}, (5)

where R11R_{11} and R12R_{12} are given by Shannon’s Theorem [27] as

R11=log2(1+Γ11),R12=log2(1+Γ12).R_{11}=\log_{2}(1+\Gamma_{11}),\quad R_{12}=\log_{2}(1+\Gamma_{12}). (6)

Positive secrecy rate condition R11>R12R_{11}>R_{12} (5), simplified as Γ11>Γ12\Gamma_{11}>\Gamma_{12} (6) for U11 gives a condition (ρtβ22+1)|h1|2>(ρtβ21+1)|h2|2(\rho_{t}\beta_{22}+1)|h_{1}|^{2}>(\rho_{t}\beta_{21}+1)|h_{2}|^{2}, which shows that based on the values of ρt\rho_{t}, β22\beta_{22}, β21\beta_{21}, |h1|2|h_{1}|^{2} and |h2|2|h_{2}|^{2}, a feasible condition can be obtained. This proves that positive secrecy rate can be achievable at U11. Similarly, the achievable secrecy rate Rs2R_{s2} of U22 is

Rs2=R22R21,R_{s2}=R_{22}-R_{21}, (7)

where

R22=log2(1+Γ22),R21=log2(1+Γ21).R_{22}=\log_{2}(1+\Gamma_{22}),\quad R_{21}=\log_{2}(1+\Gamma_{21}). (8)

Here the condition R22>R21R_{22}>R_{21} (7), simplified as Γ22>Γ21\Gamma_{22}>\Gamma_{21} (8) gives |h2|2>|h1|2|h_{2}|^{2}>|h_{1}|^{2} which is an infeasible condition because channel power gains are sorted as |h1|2>|h2|2|h_{1}|^{2}>|h_{2}|^{2}. Thus, positive secrecy rate cannot be achieved at U22.

Since our goal is to provide positive secrecy rate to both users, the conventional decoding order cannot be considered for untrusted NOMA because it cannot provide positive secrecy rate for the far user against the untrusted near user. Therefore, next we study other possible decoding orders one by one and investigate the optimal decoding order.

Case 2: Next, we consider decoding order (2,1)(2,1) where both users first decode signals associated to other user, and then decode their own signal after performing SIC. In decoding order (2,1), the order of decoding is different only for the far user in comparison to conventional decoding order. Note that the far user can even first decode the signal of near user, and then decodes its own signal after applying SIC [25], [26]. As a result, we obtain

Γ21=(1α)|h1|2α|h1|2+1ρt,Γ11=α|h1|2β21+1ρt,\Gamma_{21}=\frac{(1-\alpha)|h_{1}|^{2}}{\alpha|h_{1}|^{2}+\frac{1}{\rho_{t}}},\quad\Gamma_{11}=\frac{\alpha|h_{1}|^{2}}{\beta_{21}+\frac{1}{\rho_{t}}}, (9)
Γ12=α|h2|2(1α)|h2|2+1ρt,Γ22=(1α)|h2|2β12+1ρt.\Gamma_{12}=\frac{\alpha|h_{2}|^{2}}{(1-\alpha)|h_{2}|^{2}+\frac{1}{\rho_{t}}},\quad\Gamma_{22}=\frac{(1-\alpha)|h_{2}|^{2}}{\beta_{12}+\frac{1}{\rho_{t}}}. (10)

Positive secrecy rate condition Γ11>Γ12\Gamma_{11}>\Gamma_{12}, at U11 gives

α<1+|h1|2|h2|2|h2|2ρtβ21|h1|2|h2|2ρt.\alpha<1+\frac{|h_{1}|^{2}-|h_{2}|^{2}-|h_{2}|^{2}\rho_{t}\beta_{21}}{|h_{1}|^{2}|h_{2}|^{2}\rho_{t}}. (11)

Similarly, the feasibility condition Γ22>Γ21\Gamma_{22}>\Gamma_{21} for positive secrecy rate at U22, gives

α>|h1|2|h2|2+|h1|2ρtβ12|h1|2|h2|2ρt.\alpha>\frac{|h_{1}|^{2}-|h_{2}|^{2}+|h_{1}|^{2}\rho_{t}\beta_{12}}{|h_{1}|^{2}|h_{2}|^{2}\rho_{t}}.\\ (12)

Observing (11) and (12), we note that decoding order (2,1)(2,1) can provide positive secrecy rate to both the users in untrusted NOMA, provided α\alpha is constrained as (|h1|2|h2|2+|h1|2ρtβ12|h1|2|h2|2ρt<α<1+|h1|2|h2|2|h2|2ρtβ21|h1|2|h2|2ρt)\left(\frac{|h_{1}|^{2}-|h_{2}|^{2}+|h_{1}|^{2}\rho_{t}\beta_{12}}{|h_{1}|^{2}|h_{2}|^{2}\rho_{t}}<\alpha<1+\frac{|h_{1}|^{2}-|h_{2}|^{2}-|h_{2}|^{2}\rho_{t}\beta_{21}}{|h_{1}|^{2}|h_{2}|^{2}\rho_{t}}\right).

Case 3: In decoding order (1,2)(1,2), U11 and U22 first detect their own signals, apply SIC, and then decode signal of other user. Similar to the aforementioned analysis, we obtain Γ11=α|h1|2(1α)|h1|2+1ρt,Γ21=(1α)|h1|2β11+1ρt,Γ22=(1α)|h2|2α|h2|2+1ρt,Γ12=α|h2|2β22+1ρt\Gamma_{11}=\frac{\alpha|h_{1}|^{2}}{(1-\alpha)|h_{1}|^{2}+\frac{1}{\rho_{t}}},\Gamma_{21}=\frac{(1-\alpha)|h_{1}|^{2}}{\beta_{11}+\frac{1}{\rho_{t}}},\Gamma_{22}=\frac{(1-\alpha)|h_{2}|^{2}}{\alpha|h_{2}|^{2}+\frac{1}{\rho_{t}}},\Gamma_{12}=\frac{\alpha|h_{2}|^{2}}{\beta_{22}+\frac{1}{\rho_{t}}}. In this case, Γ11>Γ12\Gamma_{11}>\Gamma_{12} leads to α>1|h1|2|h2|2+|h1|2ρtβ22|h1|2|h2|2ρt\alpha>1-\frac{|h_{1}|^{2}-|h_{2}|^{2}+|h_{1}|^{2}\rho_{t}\beta_{22}}{|h_{1}|^{2}|h_{2}|^{2}\rho_{t}}, and Γ22>Γ21\Gamma_{22}>\Gamma_{21} provides α<|h2|2|h1|2+|h2|2ρtβ11|h1|2|h2|2ρt\alpha<\frac{|h_{2}|^{2}-|h_{1}|^{2}+|h_{2}|^{2}\rho_{t}\beta_{11}}{|h_{1}|^{2}|h_{2}|^{2}\rho_{t}} which are feasible. Thus, (1,2)(1,2) is concluded as feasible decoding order in providing positive secrecy rate to both users under constraint (1|h1|2|h2|2+|h1|2ρtβ22|h1|2|h2|2ρt<α<|h2|2|h1|2+|h2|2ρtβ11|h1|2|h2|2ρt)\left(1-\frac{|h_{1}|^{2}-|h_{2}|^{2}+|h_{1}|^{2}\rho_{t}\beta_{22}}{|h_{1}|^{2}|h_{2}|^{2}\rho_{t}}<\alpha<\frac{|h_{2}|^{2}-|h_{1}|^{2}+|h_{2}|^{2}\rho_{t}\beta_{11}}{|h_{1}|^{2}|h_{2}|^{2}\rho_{t}}\right).

Case 4: In (1,1)(1,1) decoding order, U11 first decodes its own signal, applies SIC and then decodes signal of other user, whereas U22 first decodes signal of other user, subtracts it via SIC and then decodes its own signal. As a result, Γ11=α|h1|2(1α)|h1|2+1ρt,Γ21=(1α)|h1|2β11+1ρt,Γ12=α|h2|2(1α)|h2|2+1ρt,Γ22=(1α)|h2|2β12+1ρt\Gamma_{11}=\frac{\alpha|h_{1}|^{2}}{(1-\alpha)|h_{1}|^{2}+\frac{1}{\rho_{t}}},\Gamma_{21}=\frac{(1-\alpha)|h_{1}|^{2}}{\beta_{11}+\frac{1}{\rho_{t}}},\Gamma_{12}=\frac{\alpha|h_{2}|^{2}}{(1-\alpha)|h_{2}|^{2}+\frac{1}{\rho_{t}}},\Gamma_{22}=\frac{(1-\alpha)|h_{2}|^{2}}{\beta_{12}+\frac{1}{\rho_{t}}}. Here Γ11>Γ12\Gamma_{11}>\Gamma_{12} gives |h1|2>|h2|2|h_{1}|^{2}>|h_{2}|^{2}, and Γ22>Γ21\Gamma_{22}>\Gamma_{21} required for positive secrecy rate at U22 gives (ρtβ11+1)|h2|2>(ρtβ12+1)|h1|2(\rho_{t}\beta_{11}+1)|h_{2}|^{2}>(\rho_{t}\beta_{12}+1)|h_{1}|^{2} which is also feasible. Thus, (1,1)(1,1) also can also provide security to both the users, when obtained feasibility conditions are satisfied. ∎

Thus, it can be concluded from the aforementioned analysis that utilizing the concept of PLS, three decoding orders, i.e., (2,1)(2,1), (1,2)(1,2), (1,1)(1,1), are feasible in providing positive secrecy rate for both near and far users, but with a suitable constraint on PA. Next, the investigation of optimal decoding order among these three feasible decoding orders is presented through Theorem 22.

Theorem 2

For untrusted NOMA scenario, the optimal decoding order among the feasible decoding orders is (2,1)(2,1) that gives the best secrecy rate in the system.

Proof:

See Appendix A. ∎

Remark 1

Note that the number of total possible decoding orders is enormous with more users due to which excessive computational complexity occurs in finding an optimal secure decoding order. Therefore, we have studied the system with two users, however, the investigation can be extended for more users easily.

III Pair Outage Performance of untrusted NOMA

In order to ensure users’ QoS demands for reliable communication over all the links, we first derive expression of POP for the optimal decoding order (2,1)(2,1). Next, we obtain optimal PA to minimize POP by utilizing the concept of generalized-convexity.

III-A Pair Outage Probability Analysis

Each user in a network has a predefined QoS demand, i.e., each user desires the transmitter to send data with a minimum information rate guarantees. POP, denoted as pop_{o}, ensures minimum rate guarantee to each user. In other words, pop_{o} is defined as the probability that a pair outage happens, i.e., the achievable rate at each user falls below a pre-determined threshold rate. Assuming RithR_{i}^{th} as the threshold rate for Uii and defining πi=Δ2Rith1\pi_{i}\stackrel{{\scriptstyle\Delta}}{{=}}2^{R_{i}^{th}}-1, pop_{o} can be given as

po\displaystyle p_{o} =1Pr{Γ11>π1,Γ21>π2,Γ12>π1,Γ22>π2},\displaystyle=1-\text{Pr}\{\Gamma_{11}>\pi_{1},\Gamma_{21}>\pi_{2},\Gamma_{12}>\pi_{1},\Gamma_{22}>\pi_{2}\},
=(g)1Pr{Γ11>π1,Γ21>π2}Pr{Γ12>π1,Γ22>π2},\displaystyle\stackrel{{\scriptstyle(\mathrm{g})}}{{=}}1-\text{Pr}\{\Gamma_{11}>\pi_{1},\Gamma_{21}>\pi_{2}\}\text{Pr}\{\Gamma_{12}>\pi_{1},\Gamma_{22}>\pi_{2}\},
=1Pr{|h1|2>max(ζ1,ζ2)}Pr{|h2|2>max(ζ3,ζ4)},\displaystyle=1-\text{Pr}\{|h_{1}|^{2}>\max(\zeta_{1},\zeta_{2})\}\text{Pr}\{|h_{2}|^{2}>\max(\zeta_{3},\zeta_{4})\},
=1(1F|h1|2(max(ζ1,ζ2)))(1F|h2|2(max(ζ3,ζ4))),\displaystyle=1-(1-F_{|h_{1}|^{2}}(\max(\zeta_{1},\zeta_{2})))(1-F_{|h_{2}|^{2}}(\max(\zeta_{3},\zeta_{4}))),
=1F¯|h1|2(max(ζ1,ζ2))F¯|h2|2(max(ζ3,ζ4)),\displaystyle=1-\bar{F}_{|h_{1}|^{2}}(\max(\zeta_{1},\zeta_{2}))\bar{F}_{|h_{2}|^{2}}(\max(\zeta_{3},\zeta_{4})), (13)

where Pr{.}\text{Pr}\{.\} denotes the probability measure. (g)(\mathrm{g}) follows from the property of independent events [28]. Here ζ1=Δπ1γ21ρtα,ζ2=Δπ2ρt(1ααπ2),ζ3=Δπ1ρt(α(1α)π1)\zeta_{1}\stackrel{{\scriptstyle\Delta}}{{=}}\frac{\pi_{1}\gamma_{21}}{\rho_{t}\alpha},\zeta_{2}\stackrel{{\scriptstyle\Delta}}{{=}}\frac{\pi_{2}}{\rho_{t}(1-\alpha-\alpha\pi_{2})},\zeta_{3}\stackrel{{\scriptstyle\Delta}}{{=}}\frac{\pi_{1}}{\rho_{t}(\alpha-(1-\alpha)\pi_{1})}, and ζ4=Δπ2γ12ρt(1α)\zeta_{4}\stackrel{{\scriptstyle\Delta}}{{=}}\frac{\pi_{2}\gamma_{12}}{\rho_{t}(1-\alpha)}. Note that γ21=Δρtβ21+1\gamma_{21}\stackrel{{\scriptstyle\Delta}}{{=}}\rho_{t}\beta_{21}+1 and γ12=Δρtβ12+1\gamma_{12}\stackrel{{\scriptstyle\Delta}}{{=}}\rho_{t}\beta_{12}+1. F|h1|2(x)F_{|h_{1}|^{2}}(x) and F¯|h1|2(x)\bar{F}_{|h_{1}|^{2}}(x) are cumulative distribution function (CDF) and complementary cumulative distribution function (CCDF), respectively, of channel power gain |h1|2|h_{1}|^{2}. Similarly, F|h2|2(x)F_{|h_{2}|^{2}}(x) and F¯|h2|2(x)\bar{F}_{|h_{2}|^{2}}(x) are CDF and CCDF, respectively, of channel power gain |h2|2|h_{2}|^{2}.

Note that F¯|h1|2(max(ζ1,ζ2))\bar{F}_{|h_{1}|^{2}}(\max(\zeta_{1},\zeta_{2})) in (III-A) can be rewritten for two cases ζ1>ζ2\zeta_{1}>\zeta_{2} and ζ1<ζ2\zeta_{1}<\zeta_{2}. The first case ζ1>ζ2\zeta_{1}>\zeta_{2}, results in a constraint on α\alpha as α<π1γ21π1γ21+π2+π1π2γ21\alpha<\frac{\pi_{1}\gamma_{21}}{\pi_{1}\gamma_{21}+\pi_{2}+\pi_{1}\pi_{2}\gamma_{21}}. Similarly, the constraint α>π1γ21π1γ21+π2+π1π2γ21\alpha>\frac{\pi_{1}\gamma_{21}}{\pi_{1}\gamma_{21}+\pi_{2}+\pi_{1}\pi_{2}\gamma_{21}} results in the second case ζ1<ζ2\zeta_{1}<\zeta_{2}. In addition, from the definition of CDF of exponential distribution, we observe that ζ1>0\zeta_{1}>0 and ζ2>0\zeta_{2}>0 which leads to α>0\alpha>0 and α<11+π2\alpha<\frac{1}{1+\pi_{2}}, respectively. Considering α1=Δπ1γ21π1γ21+π2+π1π2γ21\alpha_{1}\!\stackrel{{\scriptstyle\Delta}}{{=}}\!\frac{\pi_{1}\gamma_{21}}{\pi_{1}\gamma_{21}+\pi_{2}+\pi_{1}\pi_{2}\gamma_{21}} and α2=Δ11+π2\alpha_{2}\!\stackrel{{\scriptstyle\Delta}}{{=}}\!\frac{1}{1+\pi_{2}}, F¯|h1|2(max(ζ1,ζ2))\bar{F}_{|h_{1}|^{2}}(\max(\zeta_{1},\zeta_{2})) is given as

F¯|h1|2(max(ζ1,ζ2))={exp{ζ1λ1},0<α<α1exp{ζ2λ1},α1<α<α20,otherwise.\displaystyle\textstyle\bar{F}_{|h_{1}|^{2}}(\max(\zeta_{1},\zeta_{2}))\!=\!\begin{cases}\exp\{-\frac{\zeta_{1}}{\lambda_{1}}\},\!&0\!<\!\alpha\!<\!\alpha_{1}\\ \exp\{-\frac{\zeta_{2}}{\lambda_{1}}\},\!&\alpha_{1}\!<\!\alpha\!<\!\alpha_{2}\\ 0,&\text{otherwise}.\end{cases} (14)

Similarly, F¯|h2|2(max(ζ3,ζ4))\bar{F}_{|h_{2}|^{2}}(\max(\zeta_{3},\zeta_{4})) can be obtained as follows

F¯|h2|2(max(ζ3,ζ4))\displaystyle\textstyle\bar{F}_{|h_{2}|^{2}}(\max(\zeta_{3},\zeta_{4}))\! ={exp{ζ3λ2},α3<α<α4exp{ζ4λ2},α4<α<10,otherwise\displaystyle=\!\begin{cases}\exp\{-\frac{\zeta_{3}}{\lambda_{2}}\},&\alpha_{3}\!<\!\alpha\!<\!\alpha_{4}\\ \exp\{-\frac{\zeta_{4}}{\lambda_{2}}\},&\alpha_{4}\!<\!\alpha\!<\!1\\ 0,&\text{otherwise}\end{cases} (15)

where α3=π11+π1\alpha_{3}=\frac{\pi_{1}}{1+\pi_{1}}, and α4=π1+π1π2γ12π1+π2γ12+π1π2γ12\alpha_{4}=\frac{\pi_{1}+\pi_{1}\pi_{2}\gamma_{12}}{\pi_{1}+\pi_{2}\gamma_{12}+\pi_{1}\pi_{2}\gamma_{12}}. Using (III-A), (14), (15), the piecewise definition of pop_{o} as a function of α\alpha is given at the top of next page in (16)

po={1F¯|h1|2(ζ1)F¯|h2|2(ζ3),α3<α<α11F¯|h1|2(ζ2)F¯|h2|2(ζ3),{[α1<α<α2][α2<α4][α3<α1][α1<α<α4][α2>α4][α3<α1][α3<α<α2][α2<α4][α3>α1][α3<α<α4][α2>α4][α3>α1]1F¯|h1|2(ζ2)F¯|h2|2(ζ4),α4<α<α21,otherwise.p_{o}=\begin{cases}1-\bar{F}_{|h_{1}|^{2}}(\zeta_{1})\bar{F}_{|h_{2}|^{2}}(\zeta_{3}),\quad\quad\alpha_{3}<\alpha<\alpha_{1}\\ 1-\bar{F}_{|h_{1}|^{2}}(\zeta_{2})\bar{F}_{|h_{2}|^{2}}(\zeta_{3}),\quad\quad\begin{cases}[\alpha_{1}<\alpha<\alpha_{2}]\wedge[\alpha_{2}<\alpha_{4}]\wedge[\alpha_{3}<\alpha_{1}]\\ [\alpha_{1}<\alpha<\alpha_{4}]\wedge[\alpha_{2}>\alpha_{4}]\wedge[\alpha_{3}<\alpha_{1}]\\ [\alpha_{3}<\alpha<\alpha_{2}]\wedge[\alpha_{2}<\alpha_{4}]\wedge[\alpha_{3}>\alpha_{1}]\\ [\alpha_{3}<\alpha<\alpha_{4}]\wedge[\alpha_{2}>\alpha_{4}]\wedge[\alpha_{3}>\alpha_{1}]\\ \end{cases}\\ 1-\bar{F}_{|h_{1}|^{2}}(\zeta_{2})\bar{F}_{|h_{2}|^{2}}(\zeta_{4}),\quad\quad\alpha_{4}<\alpha<\alpha_{2}\\ 1,\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\text{otherwise.}\end{cases} (16)

 

where

F¯|h1|2(ζ1)=exp{π1γ21ρtαλ1},\bar{F}_{|h_{1}|^{2}}(\zeta_{1})=\exp\bigg{\{}-\frac{\pi_{1}\gamma_{21}}{\rho_{t}\alpha\lambda_{1}}\bigg{\}}, (17)
F¯|h1|2(ζ2)=exp{π2ρt(1ααπ2)λ1},\bar{F}_{|h_{1}|^{2}}(\zeta_{2})=\exp\bigg{\{}-\frac{\pi_{2}}{\rho_{t}(1-\alpha-\alpha\pi_{2})\lambda_{1}}\bigg{\}}, (18)
F¯|h2|2(ζ3)=exp{π1ρt(α(1α)π1)λ2},\bar{F}_{|h_{2}|^{2}}(\zeta_{3})=\exp\bigg{\{}-\frac{\pi_{1}}{\rho_{t}(\alpha-(1-\alpha)\pi_{1})\lambda_{2}}\bigg{\}}, (19)
F¯|h2|2(ζ4)=exp{π2γ12ρt(1α)λ2}.\bar{F}_{|h_{2}|^{2}}(\zeta_{4})=\exp\bigg{\{}-\frac{\pi_{2}\gamma_{12}}{\rho_{t}(1-\alpha)\lambda_{2}}\bigg{\}}. (20)
Remark 2

From the closed-form expression of pop_{o} in (16), it can be observed that pop_{o} depends on key system parameters such as threshold rates of users and SNR. It is noted that POP increases with an increase in threshold rates of users and decreases with an increase in the SNR.

III-B Pair Outage Probability Minimization

Observing pop_{o} as a function of α\alpha, the pair outage optimization problem can be formulated as

(J1):minimize𝛼\displaystyle(J1):\underset{\alpha}{\text{minimize}} po,\displaystyle p_{o}, s.t. (C1):0<α<1,\displaystyle(C1):0<\alpha<1, (21)

where (C1)(C1) indicates PA coefficient bounds.

Note that the closed-form piecewise expression of pop_{o} in (16) is a function of α\alpha. Therefore, in order to obtain optimal PA solution, next we need to study each case of piecewise expression one by one and find candidate optimal points to obtain the global-optimal solution. The global-optimal solution is given by Lemma 11.

Lemma 1

The unique global-optimal PA solution αpo\alpha_{p_{o}}^{*} is the feasible optimal point from the set of obtained optimal points that minimizes pop_{o}, which is given as

αpo=Δargminα{αc1,αr1,αr2,αc2}po,\displaystyle\!\alpha_{p_{o}}^{*}\!\stackrel{{\scriptstyle\Delta}}{{=}}\!\underset{\alpha\in\{\alpha_{c1},\alpha_{r1},\alpha_{r2},\alpha_{c2}\}}{\mathrm{argmin}}\!p_{o}, (22)

where αc1\alpha_{c1}, αr1\alpha_{r1}, αr2\alpha_{r2}, and αc2\alpha_{c2} are the candidate optimal points.

Proof:

In order to solve the optimization problem, closed-form piecewise expression of pop_{o} in (16) is considered. Each case is taken one by one to find the global-optimal solution.

Case 1 : In the first case, where po=1F¯|h1|2(ζ1)F¯|h2|2(ζ3)p_{o}=1-\bar{F}_{|h_{1}|^{2}}(\zeta_{1})\bar{F}_{|h_{2}|^{2}}(\zeta_{3}) and α3<α<α1\alpha_{3}<\alpha<\alpha_{1}, first-order derivative obtained by differentiating pop_{o} with respect to α\alpha is given as

dpodα=(π1γ21ρtα2λ1+π1s1ρt(αs1π1)2λ2)\displaystyle\frac{\mathrm{d}p_{o}}{\mathrm{d}\alpha}\!=\!-\bigg{(}\frac{\pi_{1}\gamma_{21}}{\rho_{t}\alpha^{2}\lambda_{1}}+\frac{\pi_{1}s_{1}}{\rho_{t}(\alpha s_{1}-\pi_{1})^{2}\lambda_{2}}\bigg{)}
exp{π1γ21ρtαλ1+π1ρt(αs1π1)λ2},\displaystyle\exp{\Big{\{}\frac{-\pi_{1}\gamma_{21}}{\rho_{t}\alpha\lambda_{1}}+\frac{-\pi_{1}}{\rho_{t}(\alpha s_{1}-\pi_{1})\lambda_{2}}\Big{\}}}, (23)

where s1=Δ(1+π1)s_{1}\stackrel{{\scriptstyle\Delta}}{{=}}(1+\pi_{1}). Since dpodα<0\frac{\mathrm{d}p_{o}}{\mathrm{d}\alpha}<0, pop_{o} is a monotonically decreasing function of α\alpha in this range. The optimal point is taken as the corner, given as αc1=α1\alpha_{c1}=\alpha_{1}.

Case 2 : For second case, where po=1F¯|h1|2(ζ2)F¯|h2|2(ζ3)p_{o}=1-\bar{F}_{|h_{1}|^{2}}(\zeta_{2})\bar{F}_{|h_{2}|^{2}}(\zeta_{3}), the derivative is given as

dpodα=(π2s2ρt(1αs2)2λ1π1s1ρt(αs1π1)2λ2)\displaystyle\frac{\mathrm{d}p_{o}}{\mathrm{d}\alpha}=\bigg{(}\frac{\pi_{2}s_{2}}{\rho_{t}(1-\alpha s_{2})^{2}\lambda_{1}}-\frac{\pi_{1}s_{1}}{\rho_{t}(\alpha s_{1}-\pi_{1})^{2}\lambda_{2}}\bigg{)}
exp{π2ρt(1αs2)λ1+π1ρt(αs1π1)λ2},\displaystyle\exp{\Big{\{}\frac{-\pi_{2}}{\rho_{t}(1-\alpha s_{2})\lambda_{1}}+\frac{-\pi_{1}}{\rho_{t}(\alpha s_{1}-\pi_{1})\lambda_{2}}\Big{\}}}, (24)

where s2=Δ(1+π2)s_{2}\stackrel{{\scriptstyle\Delta}}{{=}}(1+\pi_{2}). Note that (III-B) does not indicate any monotonic behavior. However, we observe that pop_{o} is monotonically increasing function of α\alpha if π2s2ρt(1αs2)2λ1>π1s1ρt(αs1π1)2λ2\frac{\pi_{2}s_{2}}{\rho_{t}(1-\alpha s_{2})^{2}\lambda_{1}}>\frac{\pi_{1}s_{1}}{\rho_{t}(\alpha s_{1}-\pi_{1})^{2}\lambda_{2}}, and monotonically decreasing function otherwise. The point of inflection can be obtained by solving π2s2ρt(1αs2)2λ1=π1s1ρt(αs1π1)2λ2\frac{\pi_{2}s_{2}}{\rho_{t}(1-\alpha s_{2})^{2}\lambda_{1}}=\frac{\pi_{1}s_{1}}{\rho_{t}(\alpha s_{1}-\pi_{1})^{2}\lambda_{2}} which can be simplified as a quadratic equation t1α2+t2α+t3=0t_{1}\alpha^{2}+t_{2}\alpha+t_{3}=0 where t1=Δr2s1s2r1s2,t2=Δ2r12r2π1,t3=Δπ12π2s2λ2π1s1λ1,r1=Δπ1λ1s1s2,t_{1}\stackrel{{\scriptstyle\Delta}}{{=}}r_{2}s_{1}s_{2}-r_{1}s_{2},t_{2}\stackrel{{\scriptstyle\Delta}}{{=}}2r_{1}-2r_{2}\pi_{1},t_{3}\stackrel{{\scriptstyle\Delta}}{{=}}\pi_{1}^{2}\pi_{2}s_{2}\lambda_{2}-\pi_{1}s_{1}\lambda_{1},r_{1}\stackrel{{\scriptstyle\Delta}}{{=}}\pi_{1}\lambda_{1}s_{1}s_{2}, and r2=Δπ2λ2s1s2r_{2}\stackrel{{\scriptstyle\Delta}}{{=}}\pi_{2}\lambda_{2}s_{1}s_{2}. Optimal solutions are the roots of the quadratic equation, given as

αr1,αr2=(r2π1r1)±(s2π1s1)2λ1λ2s1s2π1π2λ2s12s2π2λ1s1s22π1.\textstyle\alpha_{r1},\alpha_{r2}\!=\!\frac{(r_{2}\pi_{1}-r_{1})\pm\sqrt{(s_{2}\pi_{1}-s_{1})^{2}\lambda_{1}\lambda_{2}s_{1}s_{2}\pi_{1}\pi_{2}}}{\lambda_{2}s_{1}^{2}s_{2}\pi_{2}-\lambda_{1}s_{1}s_{2}^{2}\pi_{1}}. (25)

Case 3 : In the third case, where po=1F¯|h1|2(ζ2)F¯|h2|2(ζ4)p_{o}=1-\bar{F}_{|h_{1}|^{2}}(\zeta_{2})\bar{F}_{|h_{2}|^{2}}(\zeta_{4}), the derivative is given as

dpodα=(π2s2ρt(1αs2)2λ1+π2γ12ρt(1α)2λ2)\displaystyle\frac{\mathrm{d}p_{o}}{\mathrm{d}\alpha}=\bigg{(}\frac{\pi_{2}s_{2}}{\rho_{t}(1-\alpha s_{2})^{2}\lambda_{1}}+\frac{\pi_{2}\gamma_{12}}{\rho_{t}(1-\alpha)^{2}\lambda_{2}}\bigg{)}
exp{π2ρt(1αs2)λ1+π2γ12ρt(1α)λ2},\displaystyle\exp{\Big{\{}\frac{-\pi_{2}}{\rho_{t}(1-\alpha s_{2})\lambda_{1}}+\frac{-\pi_{2}\gamma_{12}}{\rho_{t}(1-\alpha)\lambda_{2}}\Big{\}}}, (26)

which is always greater than zero. Thus, pop_{o} is a monotonically increasing function of α\alpha. The optimal point, in this case, is given by the corner point of considered range as αc2=α4\alpha_{c2}=\alpha_{4}.

Case 4 : In fourth case, pop_{o} is a constant, i.e., po=1p_{o}=1, and it is the maximum feasible value.

From above analysis, we observe that αc1\alpha_{c1} and αc2\alpha_{c2} are two corner points due to monotonically decreasing and increasing property, respectively. αr1\alpha_{r1} and αr2\alpha_{r2} are roots as explained in Case 22. As a result, POP minimization problem has global-optimal solution αpo\alpha_{p_{o}}^{*}, which is feasible optimal point from set {αc1\alpha_{c1}, αr1\alpha_{r1}, αr2\alpha_{r2}, αc2\alpha_{c2}} at which pop_{o} is minimum. ∎

IV Secrecy Outage Performance

To analyze secrecy performance, we derive expressions of SOP for both users and investigate optimal PAs for optimizing SOPs. Next, we provide closed-form asymptotic expressions of SOPs, and optimal PAs to gain analytical insights.

IV-A Secrecy Outage Probability Analysis

For each user, an outage event happens when the received secrecy rate is below a pre-determined threshold. Particularly, SOP is defined as the probability that the maximum achievable secrecy rate is less than a target secrecy rate. Next, we derive analytical expressions of SOPs for both near and far users which have been denoted by sois_{oi} for Uii.

IV-A1 Near user

With Rs1R_{s1} and Rs1thR_{s1}^{th} as achievable and target secrecy rate for U11, so1s_{o1} is stated as

so1\displaystyle s_{o1} =Pr{Rs1<Rs1th}=Pr{1+Γ111+Γ12<Π1},\displaystyle=\text{Pr}\{R_{s1}<R_{s1}^{th}\}=\text{Pr}\Big{\{}\frac{1+\Gamma_{11}}{1+\Gamma_{12}}<\Pi_{1}\Big{\}},
=Pr{|h1|2<γ21Π1|h2|2ρt(1α)|h2|2+1+A1},\displaystyle=\text{Pr}\Big{\{}|h_{1}|^{2}<\frac{\gamma_{21}\Pi_{1}|h_{2}|^{2}}{\rho_{t}(1-\alpha)|h_{2}|^{2}+1}+A_{1}\Big{\}},
=0Fh12(γ21Π1|h2|2ρt(1α)|h2|2+1+A1)f|h2|2(y1)𝑑y1,\displaystyle=\int_{0}^{\infty}F_{\mid h_{1}\mid^{2}}\bigg{(}\frac{\gamma_{21}\Pi_{1}|h_{2}|^{2}}{\rho_{t}(1-\alpha)|h_{2}|^{2}+1}+A_{1}\bigg{)}f_{|h_{2}|^{2}}(y_{1})dy_{1},
=11λ20exp{γ21Π1y1(ρt(1α)y1+1)λ1y1λ2A1λ1}𝑑y1,\displaystyle=\!1-\!\frac{1}{\lambda_{2}}\int_{0}^{\infty}\!\exp\bigg{\{}\!\frac{-\gamma_{21}\Pi_{1}y_{1}}{(\rho_{t}(1-\alpha)y_{1}+1)\lambda_{1}}\!-\!\frac{y_{1}}{\lambda_{2}}-\frac{A_{1}}{\lambda_{1}}\bigg{\}}dy_{1}, (27)

where Π1=Δ2Rs1th\Pi_{1}\stackrel{{\scriptstyle\Delta}}{{=}}2^{R_{s1}^{th}}, A1=Δγ21(Π11)ρtαA_{1}\stackrel{{\scriptstyle\Delta}}{{=}}\frac{\gamma_{21}(\Pi_{1}-1)}{\rho_{t}\alpha}, and fh22(x)f_{\mid h_{2}\mid^{2}}(x) is probability density function (PDF) of channel power gain |h2|2|h_{2}|^{2}.

IV-A2 Far user

With Rs2R_{s2} and Rs2thR_{s2}^{th} as achievable and target secrecy rate for U22, so2s_{o2} is given as

so2\displaystyle s_{o2} =Pr{Rs2<Rs2th}=Pr{1+Γ221+Γ21<Π2},\displaystyle=\text{Pr}\{R_{s2}<R_{s2}^{th}\}=\text{Pr}\Big{\{}\frac{1+\Gamma_{22}}{1+\Gamma_{21}}<\Pi_{2}\Big{\}},
=Pr{|h2|2<γ12Π2|h1|2ρtα|h1|2+1+A2},\displaystyle=\text{Pr}\Big{\{}|h_{2}|^{2}<\frac{\gamma_{12}\Pi_{2}|h_{1}|^{2}}{\rho_{t}\alpha|h_{1}|^{2}+1}+A_{2}\Big{\}},
=0Fh22(γ12Π2|h1|2ρtα|h1|2+1+A2)f|h1|2(y2)𝑑y2,\displaystyle=\int_{0}^{\infty}F_{\mid h_{2}\mid^{2}}\bigg{(}\frac{\gamma_{12}\Pi_{2}|h_{1}|^{2}}{\rho_{t}\alpha|h_{1}|^{2}+1}+A_{2}\bigg{)}f_{|h_{1}|^{2}}(y_{2})dy_{2},
=11λ10exp{γ12Π2y2(ρtαy2+1)λ2y2λ1A2λ2}dy2,\displaystyle=\!1\!-\!\frac{1}{\lambda_{1}}\int_{0}^{\infty}\exp\bigg{\{}\frac{-\gamma_{12}\Pi_{2}y_{2}}{(\rho_{t}\alpha y_{2}+1)\lambda_{2}}-\frac{y_{2}}{\lambda_{1}}-\frac{A_{2}}{\lambda_{2}}\bigg{\}}dy_{2}, (28)

where Π2=Δ2Rs2th\Pi_{2}\stackrel{{\scriptstyle\Delta}}{{=}}2^{R_{s2}^{th}}, A2=Δγ12(Π21)ρt(1α)A_{2}\stackrel{{\scriptstyle\Delta}}{{=}}\frac{\gamma_{12}(\Pi_{2}-1)}{\rho_{t}(1-\alpha)}, and f|h1|2(x)f_{|h_{1}|^{2}}(x) denotes PDF of channel power gain |h1|2|h_{1}|^{2}.

IV-B Individual Secrecy Outage Probability Minimization

Next, we investigate optimality of SOPs of both the users.

IV-B1 Near User

Considering so1s_{o1} derived in (IV-A1) as a function of α\alpha, SOP minimization problem for U11 can be formulated as

(J2):minimize𝛼\displaystyle(J2):\underset{\alpha}{\text{minimize}} so1,\displaystyle s_{o1}, s.t. (C1).\displaystyle(C1). (29)

Noting the complexity of derived expression of so1s_{o1}, next we solve SOP minimization problems using pseudoconvexity of SOP. The feasibility of unique solution is asserted by Lemma 22.

Lemma 2

so1s_{o1} is a pseudoconvex function of α\alpha.

Proof:

so1s_{o1} in (IV-A1) can be rewritten as

so1=11λ20exp{γ21Π1y1(ρt(1α)y1+1)λ1y1λ2γ21(Π11)ρtαλ1}dy1.\displaystyle\textstyle s_{o1}\!=\!1\!-\!\frac{1}{\lambda_{2}}\int\limits_{0}^{\infty}\exp\bigg{\{}\!-\!\frac{\gamma_{21}\Pi_{1}y_{1}}{(\rho_{t}(1-\alpha)y_{1}+1)\lambda_{1}}\!-\!\frac{y_{1}}{\lambda_{2}}\!-\!\frac{\gamma_{21}(\Pi_{1}-1)}{\rho_{t}\alpha\lambda_{1}}\bigg{\}}dy_{1}. (30)

Denoting the integrand of so1s_{o1} as I1I_{1}, it can be defined as

I1=exp{γ21Π1y1(ρt(1α)y1+1)λ1y1λ2γ21(Π11)ρtαλ1}.\textstyle I_{1}=\exp\bigg{\{}-\frac{\gamma_{21}\Pi_{1}y_{1}}{(\rho_{t}(1-\alpha)y_{1}+1)\lambda_{1}}-\frac{y_{1}}{\lambda_{2}}-\frac{\gamma_{21}(\Pi_{1}-1)}{\rho_{t}\alpha\lambda_{1}}\bigg{\}}. (31)

Second-order derivative of the logarithm of the integrand function I1I_{1}, with respect to α\alpha is

d2log(I1)dα2=(2γ21(Π11)ρtλ1α3+2γ21Π1ρt2y13λ1((1α)ρty1+1)3),\begin{split}\textstyle\frac{\mathrm{d}^{2}\log(I_{1})}{\mathrm{d}\alpha^{2}}=-\Bigg{(}\frac{2\gamma_{21}(\Pi_{1}-1)}{\rho_{t}\lambda_{1}\alpha^{3}}+\frac{2\gamma_{21}\Pi_{1}\rho_{t}^{2}y_{1}^{3}}{\lambda_{1}((1-\alpha)\rho_{t}y_{1}+1)^{3}}\Bigg{)},\end{split} (32)

which is non-increasing. This indicates that the integrand I1I_{1} of the objective function so1s_{o1} is a logarithmically concave (log-concave) function. Since log-concavity is preserved under integration [29], we note that integral function in (30) is also log-concave function. Observing pseudoconcave property [30, Lemma 5] of log-concave function, we can state that integral function of (30) is pseudoconcave, and negative of a pseudoconcave function is a pseudoconvex function [31]. Hence, so1s_{o1} is a pseudoconvex function of α\alpha. ∎

Due to pseudoconvexity of so1s_{o1} of U11, there exists unique PA that minimizes so1s_{o1} [31, Chap. 3.5.9]. We apply a computationally efficient golden section search algorithm [32] to find α1\alpha_{1}^{*} (optimal value of α\alpha). The algorithm considers pseudoconvex function so1s_{o1}, αlb\alpha_{lb} and αub\alpha_{ub} as input, where αlb\alpha_{lb} and αub\alpha_{ub}, respectively, are the lower and upper bounds of α\alpha at initial stage. It gives optimal solution α1\alpha_{1}^{*} and corresponding minimum SOP as the output. Since (0<α<1)(0<\alpha<1), we first consider αlb=0\alpha_{lb}=0 and αub=1\alpha_{ub}=1 and then the algorithm searches along α\alpha with acceptable tolerance level of ϵ<<1\epsilon<<1. The algorithm works by reducing the search space with a fixed ratio of 0.6180.618 at the end of every iteration. Finally, the algorithm terminates after NN iterations if search length is less than the given tolerance [32].

IV-B2 Far User

SOP minimization problem for U22, considering expression of so2s_{o2} (IV-A2) is stated as

(J3):minimize𝛼\displaystyle(J3):\underset{\alpha}{\text{minimize}} so2,\displaystyle s_{o2}, s.t. (C1).\displaystyle(C1). (33)

In Lemma 33, we prove the feasibility of unique solution.

Lemma 3

so2s_{o2} is a pseudoconvex function of α\alpha.

Proof:

Similar to the proof of so1s_{o1} minimization problem, the pseudoconvexity of SOP for U22 can be proved. ∎

Here also, optimal PA solution of so2s_{o2} minimization problem can be evaluated numerically using golden section search algorithm. Now, the algorithm takes pseudoconvex function so2s_{o2}, αlb=0\alpha_{lb}=0, αub=1\alpha_{ub}=1 as input. It provides α2\alpha_{2}^{*}, and corresponding minimum SOP as optimal output.

IV-C Asymptotic SOP: Analysis and Optimization

Observing the complexity of derived expressions, and involved computations in optimizing SOPs, optimal PAs have been obtained numerically. In order to gain analytical insights, next we derive tight asymptotic expressions of SOPs and optimal PAs under high SNR scenario.

IV-C1 Near User

Asymptotic approximation of SOP for U11, s^o1\hat{s}_{o1}, can be obtained by setting ρt1\rho_{t}\gg 1 in (30), which leads to (ρt(1α)y1+1)(\rho_{t}(1-\alpha)y_{1}+1) \approx ρt(1α)y1\rho_{t}(1-\alpha)y_{1}. Thus, s^o1\hat{s}_{o1} is given as

s^o1\displaystyle\textstyle\hat{s}_{o1}\! =1exp{γ21Π1ρt(1α)λ1γ21(Π11)ρtαλ1}0exp{y1λ2}λ2dy1,\displaystyle=\!1\!-\!\exp\bigg{\{}\!\frac{-\gamma_{21}\Pi_{1}}{\rho_{t}(1\!-\!\alpha)\lambda_{1}}\!-\!\frac{\gamma_{21}(\Pi_{1}\!-\!1)}{\rho_{t}\alpha\lambda_{1}}\!\bigg{\}}\!\int_{0}^{\infty}\!\frac{\exp\{\frac{-y_{1}}{\lambda_{2}}\}}{\lambda_{2}}dy_{1},
=1exp{γ21(Π1+α1)ρtα(α1)λ1}.\displaystyle=\!1\!-\!\exp\bigg{\{}\frac{\gamma_{21}(\Pi_{1}+\alpha-1)}{\rho_{t}\alpha(\alpha-1)\lambda_{1}}\bigg{\}}. (34)

Considering (IV-C1), s^o1\hat{s}_{o1} minimization problem is given as

(J4):minimize𝛼\displaystyle(J4):\underset{\alpha}{\text{minimize}} s^o1,\displaystyle\hat{s}_{o1}, s.t. (C1).\displaystyle(C1). (35)

Asymptotic optimal PA for (J4)(J4) is given by Lemma 44.

Lemma 4

The optimal PA α^1\hat{\alpha}_{1} minimizing s^o1\hat{s}_{o1}, is given as

α^1=(Π11)+(Π1(Π11).\hat{\alpha}_{1}=-(\Pi_{1}-1)+\sqrt{(\Pi_{1}(\Pi_{1}-1)}. (36)
Proof:

See Appendix B. ∎

IV-C2 Far User

Similarly, the asymptotic expression s^o2\hat{s}_{o2} for U22, using (ρtαy2+1)(\rho_{t}\alpha y_{2}+1) \approx ρtαy2\rho_{t}\alpha y_{2} in (IV-A2), can be given as

s^o2\displaystyle\textstyle\hat{s}_{o2}\! =1exp{γ12Π2ρtαλ2γ12(Π21)ρt(1α)λ2}0exp{y2λ1}λ1dy2,\displaystyle=\!1\!-\!\exp\bigg{\{}\frac{-\gamma_{12}\Pi_{2}}{\rho_{t}\alpha\lambda_{2}}-\frac{\gamma_{12}(\Pi_{2}-1)}{\rho_{t}(1-\alpha)\lambda_{2}}\bigg{\}}\int_{0}^{\infty}\frac{\exp\{\frac{-y_{2}}{\lambda_{1}}\}}{\lambda_{1}}dy_{2},
=1exp{γ12(Π2α)ρtα(α1)λ2}.\displaystyle=\!1\!-\!\exp\bigg{\{}\frac{\gamma_{12}(\Pi_{2}-\alpha)}{\rho_{t}\alpha(\alpha-1)\lambda_{2}}\bigg{\}}. (37)

The s^o2\hat{s}_{o2} (IV-C2) minimization problem for U22 can be stated as

(J5):minimize𝛼\displaystyle(J5):\underset{\alpha}{\text{minimize}} s^o2,\displaystyle\hat{s}_{o2}, s.t. (C1).\displaystyle(C1). (38)
Lemma 5

The unique PA α^2\hat{\alpha}_{2} minimizing s^o2\hat{s}_{o2} is given as

α^2=Π2(Π2(Π21).\hat{\alpha}_{2}=\Pi_{2}-\sqrt{(\Pi_{2}(\Pi_{2}-1)}. (39)
Proof:

See Appendix C. ∎

Remark 3

It can be observed from the asymptotic approximations of SOPs, i.e., s^o1\hat{s}_{o1} and s^o2\hat{s}_{o2}, respectively, for U11 and U22, that the SOPs for near user and far user depend on target secrecy rates and SNR. Note that SOP increases with an increase in target secrecy rate of the user and decreases with an increase in the SNR.

V Secrecy Fairness Maximization

In order to ensure secrecy rate guarantee to both the users while satisfying users’ QoS requirements, next we formulate secrecy fairness optimization problem and investigate optimal PA. Also, we present asymptotic approximation of optimal PA to get deeper insights.

V-A Problem Formulation

QoS demands of users, in the form of maximum allowable POP forces an upper bound ξ\xi on pop_{o}, i.e., poξp_{o}\leq\xi. Hence, QoS constrained secrecy fairness maximization problem using (16), (IV-A1) and (IV-A2) that minimizes the maximum SOP between users under power budget and POP constraints can be formulated as

(J6):minimize 𝛼max[so1,so2],\displaystyle(J6):\underset{\alpha}{\text{minimize }}\max[s_{o1},s_{o2}],
s.t.(C1),(C2):poξ.\displaystyle\text{s.t.}\quad(C1),(C2):p_{o}\leq\xi. (40)

Using xc=Δmax[so1,so2]x_{c}\stackrel{{\scriptstyle\Delta}}{{=}}\max[s_{o1},s_{o2}], an equivalent formulation of (J6)(J6) can be obtained as follows

(J7):minimizeα,xcxc,s.t.(C1),(C2),\displaystyle(J7):\underset{\alpha,x_{c}}{\text{minimize}}\quad x_{c},\quad\text{s.t.}\quad(C1),(C2),
(C3):so1xc,(C4):so2xc,\displaystyle(C3):s_{o1}\leq x_{c},(C4):s_{o2}\leq x_{c}, (41)

where (C3)(C3) and (C4)(C4) appears from the definition of max[][\cdot].

Due to the presence of nonconvex constraints (C3)(C3) and (C4)(C4), (J7)(J7) is a nonconvex problem. We solve it by analyzing candidate optimal points that are characterized by Karush-Kuhn-Tucker (KKT) conditions [33].

V-B Power Control for optimizing Min-Max Secrecy Outage

First, we solve min-max SOP optimization problem without consideration of users’ rate demands. Next, QoS demand is taken into account and optimal NOMA protocol is analyzed.

V-B1 Without Pair Outage Constraint

The secrecy fairness maximization problem without (C2)(C2) can be formulated as

(J8):minimizeα,xc\displaystyle(J8):\underset{\alpha,x_{c}}{\text{minimize }} xc,\displaystyle x_{c}, s.t. (C1),(C3),(C4).\displaystyle(C1),(C3),(C4). (42)

The global-optimal solution is given by Lemma 66.

Lemma 6

The global-optimal solution αsop~\widetilde{\alpha_{sop}} of the nonconvex problem (J8)(J8), at which maximum SOP between users is minimized, is given as

αsop~=Δargminα{α1,α2,α3}max[so1,so2],\!\widetilde{\alpha_{sop}}\!\stackrel{{\scriptstyle\Delta}}{{=}}\!\underset{\alpha\in\{\alpha_{1}^{*},\alpha_{2}^{*},\alpha_{3}^{*}\}}{\mathrm{argmin}}\!\max[s_{o1},s_{o2}], (43)

where α1,α2,α3\alpha_{1}^{*},\alpha_{2}^{*},\alpha_{3}^{*} are, respectively, obtained by golden section search algorithm by minimizing so1,so2s_{o1},s_{o2} (cf. Section IV-B), and solving so1=so2s_{o1}=s_{o2}.

Proof:

Keeping the boundary constraint (C1)(C1) implicit, and associating Lagrange multipliers η1\eta_{1} with (C3)(C3) and η2\eta_{2} with (C4)(C4), the Lagrangian function \mathcal{L} can be stated as

=Δxc+η1[so1xc]+η2[so2xc].\mathcal{L}\stackrel{{\scriptstyle\Delta}}{{=}}x_{c}+\eta_{1}[s_{o1}-x_{c}]+\eta_{2}[s_{o2}-x_{c}]. (44)

The corresponding primal feasibility KKT conditions are given by the constraints (C3)(C3) and (C4)(C4). Following (42) and (44), the dual feasibility conditions are given as η10\eta_{1}\geq 0 and η20\eta_{2}\geq 0. The subgradient conditions are obtained as

ddxc=1η1η2=0,ddα=η1dso1dα+η2dso2dα=0.\frac{\mathrm{d}\mathcal{L}}{\mathrm{d}x_{c}}=1-\eta_{1}-\eta_{2}=0,\quad\frac{\mathrm{d}\mathcal{L}}{\mathrm{d}\alpha}=\eta_{1}\frac{\mathrm{d}s_{o1}}{\mathrm{d}\alpha}+\eta_{2}\frac{\mathrm{d}s_{o2}}{\mathrm{d}\alpha}=0. (45)

The two complementary slackness conditions are given as

η1[so1xc]=0,η2[so2xc]=0.\eta_{1}[s_{o1}-x_{c}]=0,\quad\eta_{2}[s_{o2}-x_{c}]=0. (46)

Note that there exists three cases. Case 1: η1>0\eta_{1}>0 and η2=0\eta_{2}=0, implies dso1dα=0\frac{\mathrm{d}s_{o1}}{\mathrm{d}\alpha}=0 which results in the same solution as SOP minimization problem (29) of U11, i.e., α=α1\alpha=\alpha_{1}^{*}. Case 2: η2>0\eta_{2}>0 and η1=0\eta_{1}=0, implies dso2dα=0\frac{\mathrm{d}s_{o2}}{\mathrm{d}\alpha}=0, which results in the solution of SOP minimization (33) of U22, i.e., α=α2\alpha=\alpha_{2}^{*}. Case 3: η1>0\eta_{1}>0 and η2>0\eta_{2}>0, implies so1=so2s_{o1}=s_{o2} using (46), which indicates equal SOP for both the users, and this case gives α=α3\alpha=\alpha_{3}^{*}. Hence, (J8)(J8) has three optimal points, i.e., α1\alpha_{1}^{*} and α2\alpha_{2}^{*} for minimizing so1s_{o1} and so2s_{o2}, respectively, and α3\alpha_{3}^{*} is obtained from the condition of so1=so2s_{o1}=s_{o2}. Finally, the global-optimal solution αsop~\widetilde{\alpha_{sop}} to this problem is obtained at the optimal point for which maximum SOP between users is minimum. ∎

Thus, we conclude that since the secrecy fairness maximization problem is nonconvex due to the presence of nonconvex constraints, we have successfully obtained all possible optimal points of this problem by KKT conditions which are the candidates for the global-optimal solution [33]. Lastly, the optimal point for which maximum SOP between users is minimum is considered as the global-optimal solution.

V-B2 With Pair Outage Constraint

Now we solve (J7)(J7) problem by taking QoS requirements into account. In this case, the global-optimal solution from secrecy perspective must satisfy the constraint (C2)(C2). Theoretically α[0,1]\alpha\in[0,1], but due to (C2)(C2), acceptable range of α\alpha gets limited, as obtained in Lemma 77.

Lemma 7

The lower and upper bound on α\alpha, required to achieve maximum allowable POP, are given as

αlb=max[αl1,αl2,αl3],αub=min[αu1,αu2,αu3],\alpha_{lb}^{*}=\max[\alpha_{l1}^{*},\alpha_{l2}^{*},\alpha_{l3}^{*}],\quad\alpha_{ub}^{*}=\min[\alpha_{u1}^{*},\alpha_{u2}^{*},\alpha_{u3}^{*}], (47)

respectively, where αl1,αl2,αl3,αu1,αu2,αu3\alpha_{l1}^{*},\alpha_{l2}^{*},\alpha_{l3}^{*},\alpha_{u1}^{*},\alpha_{u2}^{*},\alpha_{u3}^{*} are the roots of quadratic equations obtained by making (C2)(C2) an active constraint, i.e., po=ξp_{o}=\xi.

Refer to caption
Figure 2: Proposed solution methodology for pair outage constrained secrecy fairness maximization.
Proof:

Due to POP constraint, tight analytical bounds on α\alpha are obtained by using the expression of pop_{o} given in (16). Since pop_{o} is a piecewise function of α\alpha, we first consider each case of (16) one by one, substitute po=ξp_{o}=\xi, and find bounds in accordance with the respective α\alpha range. In first case of pop_{o}, where po=1F¯|h1|2(ζ1)×F¯|h2|2(ζ3)p_{o}=1-\bar{F}_{|h_{1}|^{2}}(\zeta_{1})\times\bar{F}_{|h_{2}|^{2}}(\zeta_{3}) and α3<α<α1\alpha_{3}<\alpha<\alpha_{1}, the lower and upper bound can be obtained by solving po=ξp_{o}=\xi which results in a quadratic equation given as log(1ξ)s1α2(π1log(1ξ)z1s1z2)απ1z1=0\log(1-\xi)s_{1}\alpha^{2}-(\pi_{1}\log(1-\xi)-z_{1}s_{1}-z_{2})\alpha-\pi_{1}z_{1}=0, where z1=Δγ21π1ρtλ1z_{1}\stackrel{{\scriptstyle\Delta}}{{=}}\frac{\gamma_{21}\pi_{1}}{\rho_{t}\lambda_{1}} and z2=Δπ1ρtλ2z_{2}\stackrel{{\scriptstyle\Delta}}{{=}}\frac{\pi_{1}}{\rho_{t}\lambda_{2}}. The roots of the quadratic equation, denoted as αl1\alpha_{l1} and αu1\alpha_{u1}, are given in (48) at next page.

αu1,αl1=(π1log(1ξ)z1s1z2)±(π1log(1ξ)z1s1z2)2+4log(1ξ)s1π1z12log(1ξ)s1.\alpha_{u1},\alpha_{l1}\!=\!\frac{(\pi_{1}\log(1-\xi)-z_{1}s_{1}-z_{2})\!\pm\!\sqrt{(\pi_{1}\log(1-\xi)-z_{1}s_{1}-z_{2})^{2}\!+\!4\log(1-\xi)s_{1}\pi_{1}z_{1}}}{2\log(1-\xi)s_{1}}.\\ (48)

 

The lower bound of PA in this case can be obtained as αl1=αl1\alpha_{l1}^{*}=\alpha_{l1} iff αl1\alpha_{l1} exists as a feasible point in the range else αl1=0\alpha_{l1}^{*}=0. Similarly, the upper bound of PA in this case can be given as αu1=αu1\alpha_{u1}^{*}=\alpha_{u1} iff α3<αu1<α1\alpha_{3}<\alpha_{u1}<\alpha_{1} otherwise αu1=1\alpha_{u1}^{*}=1. Similar to the above analysis, lower and upper bounds can be obtained for other cases also. Let us denote lower and upper bounds, respectively, for the second case as αl2\alpha_{l2}^{*} and αu2\alpha_{u2}^{*}, and for third case as αl3\alpha_{l3}^{*} and αu3\alpha_{u3}^{*}. After identifying lower and upper bounds of PA in each case, the αlb\alpha_{lb}^{*} in the whole range can be obtained as αlb=max[αl1,αl2,αl3]\alpha_{lb}^{*}=\max[\alpha_{l1}^{*},\alpha_{l2}^{*},\alpha_{l3}^{*}] and similarly, the upper bound αub\alpha_{ub}^{*} can be obtained as αub=min[αu1,αu2,αu3]\alpha_{ub}^{*}=\min[\alpha_{u1}^{*},\alpha_{u2}^{*},\alpha_{u3}^{*}]. ∎

Global-optimal solution of (J7)(J7) is given by Lemma 88.

Lemma 8

The global-optimal power control solution αsop\alpha_{sop}^{*} at which maximum SOP between users is minimized, is given as

αsop=Δargminα{α1,α2,α3}max[so1,so2],\!\alpha_{sop}^{*}\!\stackrel{{\scriptstyle\Delta}}{{=}}\!\underset{\alpha\in\{\alpha_{1}^{*},\alpha_{2}^{*},\alpha_{3}^{*}\}}{\mathrm{argmin}}\!\max[s_{o1},s_{o2}], (49)

subject to the feasibility of candidate optimal points as α1,α2,α3[αlb,αub]\alpha_{1}^{*},\alpha_{2}^{*},\alpha_{3}^{*}\in[\alpha_{lb}^{*},\alpha_{ub}^{*}].

Proof:

Since (J7)(J7) is a nonconvex optimization problem due to the presence of nonconvex constraints (C3)(C3) and (C4)(C4), the candidate optimal points (α1,α2,α3)(\alpha_{1}^{*},\alpha_{2}^{*},\alpha_{3}^{*}) are obtained similar to the proof of Lemma 66. In order to ensure users’ rates necessity, we check if the obtained candidate optimal points satisfy POP constraint (C2)(C2). Knowing αlb\alpha_{lb}^{*} and αub\alpha_{ub}^{*} as obtained in the proof of Lemma 77, feasibility of optimal points can be ensured. After investigating feasible optimal points, the global-optimal solution αsop\alpha_{sop}^{*} is obtained at that feasible optimal point for which maximum SOP between users is minimum. ∎

Fig. 22 presents the methodology to find optimal PA for pair outage constrained secrecy fairness maximization problem.

V-C Proposed Algorithm and Complexity Analysis

Algorithm 1 Algorithm to find global-optimal PA αsop\alpha_{sop}^{*}.
1:nn, LpL_{p}, d1d_{1}, d2d_{2}, ξ\xi, R1thR_{1}^{th}, R2thR_{2}^{th}, Rs1thR_{s1}^{th}, Rs2thR_{s2}^{th} and ρt\rho_{t}
2:αsop\alpha_{sop}^{*}
(A) Finding potential candidates
3:Obtain α1\alpha_{1}^{*}, α2\alpha_{2}^{*}, α3\alpha_{3}^{*} for minimizing so1s_{o1}, so2s_{o2}, so1=so2s_{o1}=s_{o2}, respectively, using golden section search algorithm
(B) Check Feasibility
4:Obtain αlb\alpha_{lb}^{*} and αub\alpha_{ub}^{*} by solving the equation po=ξp_{o}=\xi where pop_{o} is given in (16)
5:if αlb<α1<αub\alpha_{lb}^{*}<\alpha_{1}^{*}<\alpha_{ub}^{*} then
6:    α1\alpha_{1}^{*} is feasible optimal point
7:else
8:    update α1=αlb\alpha_{1}^{*}=\alpha_{lb}^{*}
9:end if
10:if αlb<α2<αub\alpha_{lb}^{*}<\alpha_{2}^{*}<\alpha_{ub}^{*} then
11:    α2\alpha_{2}^{*} is feasible optimal point
12:else
13:    update α2=αub\alpha_{2}^{*}=\alpha_{ub}^{*}
14:end if
15:if αlb<α3<αub\alpha_{lb}^{*}<\alpha_{3}^{*}<\alpha_{ub}^{*} then
16:    α3\alpha_{3}^{*} is feasible optimal point
17:else
18:    update α3\alpha_{3}^{*} = infeasible point
19:end if
(C) Global-optimal Solution
20:Calculate αsop=Δargminα{α1,α2,α3}max[so1,so2]\alpha_{sop}^{*}\stackrel{{\scriptstyle\Delta}}{{=}}\underset{\alpha\in\{\alpha_{1}^{*},\alpha_{2}^{*},\alpha_{3}^{*}\}}{\mathrm{argmin}}\max[s_{o1},s_{o2}]

Next, we present an algorithm that finds global-optimal PA which maximizes secrecy fairness between users while fulfilling their rate demands. For a given set of simulation parameters, αlb\alpha_{lb}^{*} and αub\alpha_{ub}^{*} need to be calculated just once. Also, golden section search is applied to find three candidate optimal points of nonconvex optimization problem (J7)(J7). Once they are calculated, αsop\alpha_{sop}^{*} is obtained by following the steps detailed in Algorithm 1. In order to investigate the complexity of proposed algorithm solving the optimization problem, we calculate total number of computations. The proposed algorithm is based on golden section search which terminates after NN iterations if (αubαlb)×0.618Nϵ(\alpha_{ub}^{*}-\alpha_{lb}^{*})\times 0.618^{N}\leq\epsilon. Hence, the total number of iterations in the proposed algorithm are 3N3N, where N=2ln(αubαlbϵ)N=2\ln\Big{(}\frac{\alpha_{ub}^{*}-\alpha_{lb}^{*}}{\epsilon}\Big{)}.

V-D Asymptotic Power Control for min-max Secrecy Outage

In above analysis, we solved the min-max SOP optimization problem numerically. In order to gain analytical insights, now we derive the asymptotic closed-form expression of global-optimal PA under high SNR region. Considering the users’ rate requirements, problem can be stated as

(J9):minimize𝛼\displaystyle(J9):\underset{\alpha}{\text{minimize }} max[s^o1,s^o2],\displaystyle\max[\hat{s}_{o1},\hat{s}_{o2}], s.t. (C1),(C2).\displaystyle(C1),(C2). (50)

Following x^c=Δmax[s^o1,s^o2]\hat{x}_{c}\stackrel{{\scriptstyle\Delta}}{{=}}\max[\hat{s}_{o1},\hat{s}_{o2}], an equivalent formulation of (J9)(J9) can be given as

(J10):minimizeα,x^cx^c,s.t.(C1),(C2),\displaystyle(J10):\underset{\alpha,\hat{x}_{c}}{\text{minimize}}\quad\hat{x}_{c},\quad\text{s.t.}\quad(C1),(C2),
(C5):s^o1x^c,(C6):s^o2x^c,\displaystyle(C5):\hat{s}_{o1}\leq\hat{x}_{c},(C6):\hat{s}_{o2}\leq\hat{x}_{c}, (51)

where (C5)(C5) and (C6)(C6) appears from the definition of max[][\cdot].

We first solve optimization problem without considering QoS constraint, which can be given as

(J11):minimizeα,x^c\displaystyle(J11):\underset{\alpha,\hat{x}_{c}}{\text{minimize }} x^c,\displaystyle\hat{x}_{c}, s.t. (C1),(C5),(C6).\displaystyle(C1),(C5),(C6). (52)

Global-optimal solution of (J11)(J11) is provided by Lemma 99.

Lemma 9

The global-optimal solution α^\hat{\alpha} of the optimization problem (J11)(J11) for maximizing the secrecy fairness is given by

α˙sop=Δargminα{α^1,α^2,α^3}max[s^o1,s^o2],\dot{\alpha}_{sop}\stackrel{{\scriptstyle\Delta}}{{=}}\underset{\alpha\in\{\hat{\alpha}_{1},\hat{\alpha}_{2},\hat{\alpha}_{3}\}}{\mathrm{argmin}}\max[\hat{s}_{o1},\hat{s}_{o2}], (53)

where α^1,α^2\hat{\alpha}_{1},\hat{\alpha}_{2}, respectively, have been obtained by minimizing s^o1,s^o2\hat{s}_{o1},\hat{s}_{o2} (cf. IV-C), and α^3\hat{\alpha}_{3} is obtained by solving s^o1=s^o2\hat{s}_{o1}=\hat{s}_{o2}.

Proof:

We associate Lagrange multipliers μ1\mu_{1} with (C5)(C5), μ2\mu_{2} with (C6)(C6), and keep the boundary constraint (C1)(C1) implicit. Then, Lagrangian function ^\mathcal{\hat{L}} can be written as

^=Δx^c+μ1[s^o1x^c]+μ2[s^o2x^c].\mathcal{\hat{L}}\stackrel{{\scriptstyle\Delta}}{{=}}\hat{x}_{c}+\mu_{1}[\hat{s}_{o1}-\hat{x}_{c}]+\mu_{2}[\hat{s}_{o2}-\hat{x}_{c}]. (54)

The corresponding KKT conditions are given by constraints (C5)(C5) and (C6)(C6). Dual feasibility conditions are obtained as μ10\mu_{1}\geq 0 and μ20\mu_{2}\geq 0 using (52) and (54). The subgradient conditions are given as d^dx^c=1μ1μ2=0,\frac{\mathrm{d}\mathcal{\hat{L}}}{\mathrm{d}\hat{x}_{c}}=1-\mu_{1}-\mu_{2}=0, and d^dα=μ1ds^o1dα+μ2ds^o2dα=0\frac{\mathrm{d}\mathcal{\hat{L}}}{\mathrm{d}\alpha}=\mu_{1}\frac{\mathrm{d}\hat{s}_{o1}}{\mathrm{d}\alpha}+\mu_{2}\frac{\mathrm{d}\hat{s}_{o2}}{\mathrm{d}\alpha}=0. The complementary slackness conditions are obtained as μ1[s^o1x^c]=0\mu_{1}[\hat{s}_{o1}-\hat{x}_{c}]=0 and μ2[s^o2x^c]=0\mu_{2}[\hat{s}_{o2}-\hat{x}_{c}]=0. Here also, three cases exist. Case 1: μ1>0\mu_{1}>0 and μ2=0\mu_{2}=0, implies ds^o1dα=0\frac{\mathrm{d}\hat{s}_{o1}}{\mathrm{d}\alpha}=0 and results α=α^1\alpha=\hat{\alpha}_{1} from s^o1\hat{s}_{o1} minimization for U11 (36). Case 2: μ2>0\mu_{2}>0 and μ1=0\mu_{1}=0, implies ds^o2dα=0\frac{\mathrm{d}\hat{s}_{o2}}{\mathrm{d}\alpha}=0 and provides the same solution as the optimal solution of s^o2\hat{s}_{o2} minimization, i.e., α=α^2\alpha=\hat{\alpha}_{2} from (39). Case 3: μ1>0\mu_{1}>0 and μ2>0\mu_{2}>0, implies s^o1=s^o2\hat{s}_{o1}=\hat{s}_{o2} and it gives α^3\hat{\alpha}_{3} which is given as

α^3=γ12Π2λ1+(ρtβ21γ21Π1)λ2γ12λ1+γ21λ2.\hat{\alpha}_{3}=\frac{\gamma_{12}\Pi_{2}\lambda_{1}+(\rho_{t}\beta_{21}-\gamma_{21}\Pi_{1})\lambda_{2}}{\gamma_{12}\lambda_{1}+\gamma_{21}\lambda_{2}}. (55)

Observing that the minimization problem has three optimal points, i.e., α^1\hat{\alpha}_{1}, α^2\hat{\alpha}_{2}, and α^3\hat{\alpha}_{3}, the global optimal solution α˙sop\dot{\alpha}_{sop} is obtained as the point where maximum SOP is minimum. ∎

Next, while taking the QoS constraint into account, the global-optimal solution of (J10)(J10) is given by Lemma 1010.

Lemma 10

Global-optimal solution α^sop\hat{\alpha}_{sop} of optimization problem (J10)(J10) for minimizing the maximum SOP is given by

α^sop=Δargminα{α^1,α^2,α^3}max[s^o1,s^o2],\hat{\alpha}_{sop}\stackrel{{\scriptstyle\Delta}}{{=}}\underset{\alpha\in\{\hat{\alpha}_{1},\hat{\alpha}_{2},\hat{\alpha}_{3}\}}{\mathrm{argmin}}\max[\hat{s}_{o1},\hat{s}_{o2}], (56)

subject to the feasibility of candidate optimal points as α^1,α^2,α^3[αlb,αub]\hat{\alpha}_{1},\hat{\alpha}_{2},\hat{\alpha}_{3}\in[\alpha_{lb}^{*},\alpha_{ub}^{*}].

Proof:

Similar to proof of Lemma 99, we first find the candidate optimal points α^1,α^2,\hat{\alpha}_{1},\hat{\alpha}_{2}, and α^3\hat{\alpha}_{3}. Next, the feasibility of these points is investigated between αlb\alpha_{lb}^{*} and αub\alpha_{ub}^{*}. Finally, the global optimal point is obtained as the feasible optimal point where maximum SOP is minimized. ∎

VI Numerical Investigations

Refer to caption
Figure 3: Validation of pair outage probability pop_{o} with users’ threshold rates, (a) R2th=R1thR_{2}^{th}=R_{1}^{th}, and (b) ρr=20\rho_{r}=20 dB.
Refer to caption
Figure 4: Validation of secrecy outage probability, (a) so1s_{o1} for U11, and (b) so2s_{o2} for U22.
Refer to caption
Figure 5: Variation of pair outage probability pop_{o} versus U22’s distance d2d_{2} with R2th=R1th=0.1R_{2}^{th}=R_{1}^{th}=0.1.
Refer to caption
Figure 6: Variation of secrecy outage probability versus U22’s distance d2d_{2}, (a) so1s_{o1} for U11, and (b) so2s_{o2} for U22.

Next, numerical results are presented to assess the performance of proposed protocol. We present simulation setup, and discuss the system performance under various system settings. Specifically, we validate accuracy of analytical results, provide insights on effect of users’ distance and optimal PAs, and identify the performance gains achieved by global-optimal PA. Finally, tradeoff between QoS parameter POP and secrecy performance SOP is presented.

VI-A Simulation Setup

Downlink of a NOMA system is considered, where BS communicates with two untrusted users. For the sake of simulation study, near user distance d1d_{1} and far user distance d2d_{2} from BS are taken as 5050 meter and 100100 meter, respectively. LpL_{p} and nn are taken as 11 and 2.52.5, respectively. Noise signal at both users follow Gaussian distribution with a noise power of 60-60 dBm. Small scale fading affecting both links is assumed to have exponential distribution with mean value 11 [13]. Residual interference terms β21\beta_{21} and β12\beta_{12} are assumed to be equal and 30-30 dBm [19]. Simulation results are averaged over 10610^{6} randomly generated channel realizations for both U11 and U22. The correctness of results is observed using the fact that root mean square error (RMSE) in approximating the derived results should be less than 0.00030.0003 [34]. For golden section search, ϵ=0.01\epsilon=0.01. ρr\rho_{r} is considered as the received SNR. POP and SOP are considered as performance metrics to evaluate the performance of system.

VI-B Validation of Analysis

We first validate analytical expressions of pop_{o} in Fig. 4, and so1s_{o1} and so2s_{o2} in Fig. 4. For these results, α\alpha is considered as 0.50.5. A close match between the simulation and analytical results validates the analysis of pop_{o}, so1s_{o1}, and so2s_{o2} with a RMSE of the order of 10410^{-4}. Accuracy of asymptotic results, marked as ‘Asy’, is also verified with simulation results in Fig. 4(b). We observe that pop_{o} increases with increase in the threshold rates R1thR_{1}^{th} or R2thR_{2}^{th}. Similarly, increasing target secrecy rates Rs1thR_{s1}^{th} and Rs2thR_{s2}^{th} increase so1s_{o1} and so2s_{o2}, respectively. Following the fact that the outage occurs when the users’ maximum achievable rate is below a threshold rate, it is natural that increasing threshold rates and target secrecy rates at users, respectively, increases POP and SOP.

From the results it is also observed that threshold rate pair (R1th,R2thR_{1}^{th},R_{2}^{th}) with lesser value of R2thR_{2}^{th} compared to R1thR_{1}^{th} results a performance improvement in POP. The reason is that the achievable information rate at U22 is always less in comparison to U11 due to the poorer channel condition of U22. On the other side, outage happens when the rate falls below a threshold rate. Therefore, in such a case, if more value of R2thR_{2}^{th} is considered, outage probability at U22 will be increased, which will further increase POP. Therefore, it can be concluded that the threshold rate pair (R1th,R2thR_{1}^{th},R_{2}^{th}) with R2th<R1thR_{2}^{th}<R_{1}^{th} has more practical significance to obtain better pair outage performance.

Further, we observe that increasing ρr\rho_{r} decreases pop_{o}, so1s_{o1} and so2s_{o2}. With increasing SNR, achievable data rates and secrecy rates at users increase, hence, for a given threshold rate and target secrecy rate, POP and SOP decreases, respectively. Noting the above observations, we can conclude that the performance of a system is highly dependent on system parameters.

Refer to caption
Figure 7: Pseudo convexity of POP through pop_{o} versus α\alpha analysis, (a) R2th=R1th=0.1R_{2}^{th}=R_{1}^{th}=0.1, and (b) ρr=20\rho_{r}=20 dB.
Refer to caption
Figure 8: Pseudo convexity of SOP through variation of SOP with α\alpha, (a) so1s_{o1} at Rs1th=1R_{s1}^{th}=1, and (b) so2s_{o2} at ρr=20\rho_{r}=20 dB.
Refer to caption
Figure 9: Optimal SOP max[so1,so2]\max[s_{o1},s_{o2}] and PA coefficient α\alpha analysis with Rs2th=Rs1th=1R_{s2}^{th}=R_{s1}^{th}=1, R2th=R1th=0.1R_{2}^{th}=R_{1}^{th}=0.1, and POP constraint ξ=0.5\xi=0.5.
Refer to caption
Figure 10: (a) Optimal PA with target secrecy rate Rs1thR_{s1}^{th}, and (b) secrecy fairness minmax[so1,so2]\min\max[s_{o1},s_{o2}] with Rs1thR_{s1}^{th}, ρr=30\rho_{r}=30 dB, R2th=R1th=0.1R_{2}^{th}=R_{1}^{th}=0.1, ξ=0.5\xi=0.5.

VI-C Impact of Relative Distance between Users

Fixing d1d_{1} as 5050 meter, now we observe the effect of variation of d2d_{2} from BS on achievable POP and SOP. In Fig. 6, increasing d2d_{2} implies decrease in achievable data rate and hence, increasing the outage probability of U22 which results in an increase in POP. On the contrary, the results presented in Fig. 6(a), show that so1s_{o1} monotonically decreases with the increase in d2d_{2}. The reason is, with increasing distance d2d_{2}, a decrease in achievable information rate at U22 improves secrecy rate at U11, and hence, SOP at U11 decreases. Also, decrease in information rate at U22 implies decrease in secrecy rate at U22 which increases SOP for U22 as shown in Fig. 6(b). Note that increasing the distance between BS to U22 has an opposite effect on pop_{o} and so1s_{o1}. Similarly, Fig. 6(a) and Fig. 6(b) also show a contradicting effect of d2d_{2} on so1s_{o1} and so2s_{o2}. Hence, it can be summarized that since POP and SOPs for both the users are important system parameters for reliable and secure NOMA communication and fulfilling them together is highly dependent on the relative distance between BS and users.

Refer to caption
Figure 11: Performance comparison of global-optimal PA, αsop\alpha_{sop}^{*}, with fixed PA α=0.33\alpha=0.33, and individual optimal PA α1\alpha_{1}^{*} and α2\alpha_{2}^{*}, R2th=R1th=0.1R_{2}^{th}=R_{1}^{th}=0.1, ξ=0.5\xi=0.5.
Refer to caption
Figure 12: Tradeoff between optimal SOP max[so1,so2]\max[s_{o1},s_{o2}] and POP threshold ξ\xi at SNR ρr=10\rho_{r}=10 dB for different secrecy rate targets.

VI-D Insights on optimality

Now we investigate optimal pair outage and secrecy outage performance by presenting the numerical proof of the generalized-convexity of POP and SOP with respect to α\alpha. Fig. 8(a) and Fig. 8(b) depict POP performance with α\alpha for different values of SNRs and threshold rates, respectively. The pseudoconvex nature of POP can be easily observed. It confirms the unique solution of PA that provides minimum POP. Analytical results have also been plotted as a validation to the analysis. Similarly, Fig. 8(a) and Fig. 8(b), respectively, validate pseudoconvex nature of so1s_{o1} and so2s_{o2} with α\alpha for different SNRs and threshold rates. The numerical optimal solution obtained using golden section algorithm has also been indicated. We observe that numerical results match with asymptotic results, which confirms the validity of asymptotic analysis. Further, to demonstrate secrecy fairness maximization under QoS constraint, Fig. 10 presents max[so1\max[s_{o1},so2]s_{o2}] with α\alpha. αlb\alpha_{lb}^{*} and αub\alpha_{ub}^{*} are shown as lower and upper bounds on α\alpha, respectively, for an example case of ξ=0.5\xi=0.5. A close match between numerical and asymptotically optimal results can be observed at ρr20\rho_{r}\geq 20 dB, which validates high SNR analysis. From the observations, it is worth noting that PA α\alpha decides allocated powers to users, which effects POP and SOPs. Hence, an appropriate PA for given system parameters can help to obtain an optimal secure and reliable communication system.

Next, Fig. 10(a) shows global-optimal αsop\alpha_{sop}^{*} that minimizes the maximum SOP between users as a function of target secrecy rates Rs1thR_{s1}^{th} for various values of Rs2thR_{s2}^{th}. Results show that for each target secrecy rate pair (Rs1thR_{s1}^{th}, Rs2thR_{s2}^{th}), there is one and only one α\alpha such that αPt\alpha P_{t} and (1α)Pt(1-\alpha)P_{t} are the optimal powers for U11 and U22, respectively. On increasing Rs1thR_{s1}^{th}, the optimal value of α\alpha decreases, whereas the optimal SOP obtained from min-max optimization problem, presented in Fig. 10(b) increases. Further, lower value of Rs2thR_{s2}^{th} in comparison to Rs1thR_{s1}^{th} provides improvement in SOP which results in a higher value of α\alpha. This is because the achievable data rate at U22 is less compared to the achievable data rate at U11 because of the weaker channel condition of U22, due to which the achievable secrecy rate at U22 decreases. Note that according to the definition of secrecy outage, outage happens when the rate falls below a threshold rate. Therefore, it is natural if more value of Rs2thR_{s2}^{th} is considered, SOP for U22 will be increased. Therefore, we can conclude that the threshold secrecy rate pair (Rs1th,Rs2thR_{s1}^{th},R_{s2}^{th}) with Rs2th<Rs1thR_{s2}^{th}<R_{s1}^{th} is a better choice to improve system performance.

VI-E Performance Comparison

To demonstrate the performance gain achieved by the proposed algorithm for secrecy fairness maximization, Fig. 12 presents the performance comparison of global-optimal PA αsop\alpha_{sop}^{*} with fixed PA α=0.33\alpha=0.33, and individual optimal PAs α1\alpha_{1}^{*} and α2\alpha_{2}^{*} obtained by minimizing so1s_{o1} and so2s_{o2}, respectively. Percentage gain depicts that αsop\alpha_{sop}^{*} achieves best SOP performance, because it ensures secrecy fairness between U11 and U22. The average percentage improvement by αsop\alpha_{sop}^{*} over fixed PA, individual optimal PAs α1\alpha_{1}^{*} and α2\alpha_{2}^{*} are respectively around 55.12%55.12\%, 69.30%69.30\% and 19.11%19.11\%. Thus, it can be concluded that the global optimal PA allocated to users plays an important role in providing the best system performance.

VI-F Tradeoff

Secrecy performance and users’ data rates demands both are important system performance parameters for secure and reliable communication. However, fulfilling them together is a challenge because when QoS is considered, system sacrifices secrecy rate to improve QoS demands. This interesting tradeoff between minimized SOP and POP, as shown in Fig. 12, emphasizes the effect of QoS demands on optimal SOP performance. Decrease in SOP with an increase in POP highlights that higher rate demands lead to poor secrecy performance.

VII Concluding Remarks

This paper has proposed a novel decoding order that is capable of providing positive secrecy rate for both near and far users in a two-user untrusted NOMA system. To analyze secrecy performance, SOP for both users are derived, and closed-form expressions are given for the high SNR regime. Individual PA optimization to minimize SOPs and asymptotic optimal solutions have also been presented. In order to ensure users’ QoS demands for reliable communication over all the links, POP has been derived as a QoS measure, and optimal PA minimizing POP has also been obtained. Further, with an intention to provide secrecy fairness between users while satisfying their predefined QoS demands, optimization of PA to minimize the maximum SOP between users is also investigated. Numerical results are provided to verify the correctness of analytical expressions, provide insights on generalized-convexity of POP and SOP, highlight significant performance gains achieved by global-optimal PA, and describe the tradeoff between different system performance requirements. While this work has considered a two user NOMA system, the work can be extended to the study of secrecy in a multi-user scenario and considering other variants like code-domain NOMA. Furthermore, for an untrusted users’ scenario, it would be interesting to compare the performance of a NOMA enabled system with jammer assisted orthogonal multiple access.

Appendix A Proof of Theorem 2

Having obtained feasible decoding orders for NOMA system with the positive secrecy rate consideration at all users, we compare these orders in terms of users’ secrecy rates to investigate the optimal decoding order. The key idea here is, to find out that one decoding order which gives best secrecy rates at users. In this context, we have observed that, for each decoding order, secrecy rates Rs1R_{s1} and Rs2R_{s2} at U11 and U22, respectively, are in the form of log21+(A/B)1+(C/D)\log_{2}\frac{1+(A/B)}{1+(C/D)}. Therefore, on comparing two decoding orders, we can easily show that a given secrecy rate is higher than its counterpart if either of AA or DD is higher, or either of BB or CC is lower, with other respective parameters being same. Thus, we can find out the optimal decoding order having maximum secrecy rate.

Considering Theorem 11, there exists three feasible decoding orders: (2,1)(2,1), (1,2)(1,2), and (1,1)(1,1). We first compare (2,1)(2,1) and (1,2)(1,2). In (2,1)(2,1), B=β21+1ρtB=\beta_{21}+\frac{1}{\rho_{t}} and D=(1α)|h2|2+1ρtD=(1-\alpha)|h_{2}|^{2}+\frac{1}{\rho_{t}} for Rs1R_{s1}, and, B=β12+1ρtB=\beta_{12}+\frac{1}{\rho_{t}} and D=α|h1|2+1ρtD=\alpha|h_{1}|^{2}+\frac{1}{\rho_{t}} for Rs2R_{s2}. Whereas, in (1,2)(1,2), B=(1α)|h1|2+1ρtB=(1-\alpha)|h_{1}|^{2}+\frac{1}{\rho_{t}} and D=β22+1ρtD=\beta_{22}+\frac{1}{\rho_{t}} for Rs1R_{s1}, and, B=α|h2|2+1ρtB=\alpha|h_{2}|^{2}+\frac{1}{\rho_{t}} and D=β11+1ρtD=\beta_{11}+\frac{1}{\rho_{t}} for Rs2R_{s2}. The remaining parameters, i.e., AA and CC are the same for the two considered orders for Rs1R_{s1}. The same holds true for Rs2R_{s2}. Observing Rs1R_{s1} in (2,1)(2,1) and (1,2)(1,2), we note that BB in (2,1)(2,1) is lower in comparison to BB in (1,2)(1,2), i.e., β21<(1α)|h1|2\beta_{21}<(1-\alpha)|h_{1}|^{2}. Also, DD in (2,1)(2,1) is higher compared to DD in (1,2)(1,2), i.e., (1α)|h2|2>β22(1-\alpha)|h_{2}|^{2}>\beta_{22}. This is because βij\beta_{ij} is the residual interefence from imperfectly decoded Uii by Ujj. Similarly, for Rs2R_{s2} also, BB is lower and DD is higher for (2,1)(2,1) in comparison to BB and DD, respectively, of (1,2)(1,2). Therefore, it can be concluded that secrecy rate in (2,1)(2,1) is higher than (1,2)(1,2) for both U11 and U22. In the similar manner, we compare (2,1)(2,1) and (1,1)(1,1). In (1,1)(1,1), B=(1α)|h1|2+1ρtB=(1-\alpha)|h_{1}|^{2}+\frac{1}{\rho_{t}} for Rs1R_{s1} and D=β11+1ρtD=\beta_{11}+\frac{1}{\rho_{t}} for Rs2R_{s2}. Here also, for Rs1R_{s1} in (2,1)(2,1) and (1,1)(1,1), BB is lower in (2,1)(2,1) compared to the BB in (1,1)(1,1), and, for Rs2R_{s2}, DD is higher in (2,1)(2,1) in comparison to the DD in (1,1)(1,1). Thus, (2,1)(2,1) will provide better secrecy rate than (1,1)(1,1) also, at both users.

Thus, we conclude that out of the three feasible decoding orders that can provide positive secrecy rate at both users, (2,1)(2,1) is a decoding order that ensures maximum secrecy rate at both the users as compared to other two decoding orders. Therefore, (2,1)(2,1) is considered as an optimal decoding order regardless of any PA, which is followed in analysis further.

Appendix B Proof of Lemma 4

Note that α^1\hat{\alpha}_{1} is obtained by minimizing s^o1\hat{s}_{o1} (IV-C1). The second-order derivative of s^o1\hat{s}_{o1} with respect to α\alpha does not imply monotonicity. By setting ds^o1dα=0\frac{\mathrm{d}\hat{s}_{o1}}{\mathrm{d}\alpha}=0, we obtain α^1=(Π11)±(Π1(Π11)\hat{\alpha}_{1}=-(\Pi_{1}-1)\pm\sqrt{(\Pi_{1}(\Pi_{1}-1)}. We observe that α^1=(Π11)(Π1(Π11)\hat{\alpha}_{1}=-(\Pi_{1}-1)-\sqrt{(\Pi_{1}(\Pi_{1}-1)} is negative, and hence, infeasible. Thus, asymptotic optimal PA minimizing s^o1\hat{s}_{o1} is given as (36).

Appendix C Proof of Lemma 5

Here also by setting, ds^o2dα=0\frac{\mathrm{d}\hat{s}_{o2}}{\mathrm{d}\alpha}=0 we obtain α^2=Π2±(Π2(Π21)\hat{\alpha}_{2}=\Pi_{2}\pm\sqrt{(\Pi_{2}(\Pi_{2}-1)}. We note α^2=Π2+(Π2(Π21)\hat{\alpha}_{2}=\Pi_{2}+\sqrt{(\Pi_{2}(\Pi_{2}-1)} is infeasible, because it forces Rs2th<0R_{s2}^{th}<0 which is not possible. Hence, asymptotic optimal PA for U22 is given as (39).

References

  • [1] S. Thapar, D. Mishra, and R. Saini, “Secrecy fairness aware NOMA for untrusted users,” in Proc. IEEE GLOBECOM, Hawaii, USA, Dec. 2019, pp. 1–6.
  • [2] Z. Ding, X. Lei, G. K. Karagiannidis, R. Schober, J. Yuan, and V. K. Bhargava, “A survey on non-orthogonal multiple access for 5G networks: Research challenges and future trends,” IEEE J. Sel. Areas Commun., vol. 35, no. 10, pp. 2181–2195, Oct. 2017.
  • [3] Y. Liu, Z. Qin, M. Elkashlan, Y. Gao, and L. Hanzo, “Enhancing the physical layer security of non-orthogonal multiple access in large-scale networks,” IEEE Trans. Wireless Commun., vol. 16, no. 3, pp. 1656–1672, Mar. 2017.
  • [4] Y. Liu, Z. Qin, M. Elkashlan, Z. Ding, A. Nallanathan, and L. Hanzo, “Nonorthogonal multiple access for 5G and beyond,” Proceedings of the IEEE, vol. 105, no. 12, pp. 2347–2381, Dec. 2017.
  • [5] A. D. Wyner, “The wire-tap channel,” Bell system technical journal, vol. 54, no. 8, pp. 1355–1387, 1975.
  • [6] Y. Wu, A. Khisti, C. Xiao, G. Caire, K. Wong, and X. Gao, “A survey of physical layer security techniques for 5G wireless networks and challenges ahead,” IEEE J. Sel. Areas Commun., vol. 36, no. 4, pp. 679–695, Apr. 2018.
  • [7] Y. Zhang, H.-M. Wang, Q. Yang, and Z. Ding, “Secrecy sum rate maximization in non-orthogonal multiple access,” IEEE Commun. Lett., vol. 20, no. 5, pp. 930–933, May 2016.
  • [8] B. He, A. Liu, N. Yang, and V. K. Lau, “On the design of secure non-orthogonal multiple access systems,” IEEE J. Sel. Areas Commun., vol. 35, no. 10, pp. 2196–2206, Oct. 2017.
  • [9] J. Chen, L. Yang, and M. Alouini, “Physical layer security for cooperative NOMA systems,” IEEE Trans. Veh. Technol., vol. 67, no. 5, pp. 4645–4649, May 2018.
  • [10] L. Lv, F. Zhou, J. Chen, and N. Al-Dhahir, “Secure cooperative communications with an untrusted relay: A NOMA-inspired jamming and relaying approach,” IEEE Trans. Inf. Forensics Security, vol. 14, no. 12, pp. 3191–3205, 2019.
  • [11] F. Zhou, Z. Chu, H. Sun, R. Q. Hu, and L. Hanzo, “Artificial noise aided secure cognitive beamforming for cooperative MISO-NOMA using SWIPT,” IEEE J. Sel. Areas Commun., vol. 36, no. 4, pp. 918–931, 2018.
  • [12] Y. Li, M. Jiang, Q. Zhang, Q. Li, and J. Qin, “Secure beamforming in downlink MISO nonorthogonal multiple access systems,” IEEE Trans. Veh. Technol., vol. 66, no. 8, pp. 7563–7567, Aug. 2017.
  • [13] B. M. ElHalawany and K. Wu, “Physical-layer security of NOMA systems under untrusted users,” in Proc. IEEE GLOBECOM, United Arab Emirates, Dec. 2018, pp. 1–6.
  • [14] R. Saini, D. Mishra, and S. De, “Subcarrier pairing as channel gain tailoring: Joint resource allocation for relay-assisted secure OFDMA with untrusted users,” Physical Communication, vol. 32, pp. 217–230, 2019.
  • [15] R. Saini, D. Mishra, and S. De, “Utility regions for DF relay in OFDMA-based secure communication with untrusted users,” IEEE Commun. Lett., vol. 21, no. 11, pp. 2512–2515, Nov. 2017.
  • [16] D. Xu, P. Ren, Q. Du, L. Sun, and Y. Wang, “Combat eavesdropping by full-duplex technology and signal transformation in non-orthogonal multiple access transmission,” in Proc. IEEE ICC, Paris, France, May 2017, pp. 1–6.
  • [17] C. Yan, A. Harada, A. Benjebbour, Y. Lan, A. Li, and H. Jiang, “Receiver design for downlink non-orthogonal multiple access (NOMA),” in Proc. IEEE VTC Spring, Glasgow, United Kingdom, May 2015, pp. 1–6.
  • [18] M. Vaezi, R. Schober, Z. Ding, and H. V. Poor, “Non-orthogonal multiple access: Common myths and critical questions,” IEEE Wireless Commun., vol. 26, no. 5, pp. 174–180, 2019.
  • [19] X. Yue, Z. Qin, Y. Liu, S. Kang, and Y. Chen, “A unified framework for non-orthogonal multiple access,” IEEE Trans. Commun., vol. 66, no. 11, pp. 5346–5359, Nov. 2018.
  • [20] Z. Ding, M. Peng, and H. V. Poor, “Cooperative non-orthogonal multiple access in 5G systems,” IEEE Commun. Lett., vol. 19, no. 8, pp. 1462–1465, Aug. 2015.
  • [21] Z. Ding, P. Fan, and H. V. Poor, “Impact of user pairing on 5G nonorthogonal multiple-access downlink transmissions,” IEEE Trans. Veh. Technol., vol. 65, no. 8, pp. 6010–6023, Aug. 2016.
  • [22] T. M. Cover and J. A. Thomas, Elements of information theory.   John Wiley & Sons, 2012.
  • [23] Y. Liu, Z. Ding, M. Elkashlan, and H. V. Poor, “Cooperative non-orthogonal multiple access with simultaneous wireless information and power transfer,” IEEE J. Sel. Areas Commun., vol. 34, no. 4, pp. 938–953, Apr. 2016.
  • [24] N. T. Do, D. B. Da Costa, T. Q. Duong, and B. An, “A BNBF user selection scheme for NOMA-based cooperative relaying systems with SWIPT,” IEEE Commun. Lett., vol. 21, no. 3, pp. 664–667, Mar. 2017.
  • [25] X. Chen, A. Beiijebbour, A. Li, H. Jiang, and H. Kayama, “Consideration on successive interference canceller (SIC) receiver at cell-edge users for non-orthogonal multiple access (NOMA) with SU-MIMO,” in Proc. PIMRC, Hong Kong, China, Aug. 2015, pp. 522–526.
  • [26] S. Thapar, D. Mishra, and R. Saini, “Decoding orders and power allocation for untrusted NOMA: A secrecy perspective,” in Proc. IEEE WCNC, Seoul, South Korea, May 2020, pp. 1–6.
  • [27] D. Tse and P. Viswanath, Fundamentals of wireless communication.   Cambridge university press, 2005.
  • [28] A. Papoulis and S. U. Pillai, Probability, random variables, and stochastic processes.   Tata McGraw-Hill Education, 2002.
  • [29] S. Boyd and L. Vandenberghe, Convex optimization.   Cambridge university press, 2004.
  • [30] D. Mishra, S. De, and C.-F. Chiasserini, “Joint optimization schemes for cooperative wireless information and power transfer over rician channels,” IEEE Trans. Commun., vol. 64, no. 2, pp. 554–571, Feb. 2016.
  • [31] M. S. Bazaraa, H. D. Sherali, and C. M. Shetty, Nonlinear programming: theory and applications.   New York: John Wiley and Sons, 2006.
  • [32] Y.-C. Chang, “N-dimension golden section search: Its variants and limitations,” in Proc. 2nd Int. Conf. on Biomedical Engineering and Informatics (BMEI), Tianjin, China, Oct. 2009, pp. 1–6.
  • [33] A. Ravindran, G. V. Reklaitis, and K. M. Ragsdell, Engineering optimization: methods and applications.   John Wiley & Sons, 2006.
  • [34] D. Mishra and S. De, “Optimal relay placement in two-hop RF energy transfer,” IEEE Trans. Commun., vol. 63, no. 5, pp. 1635–1647, May 2015.