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

\UseRawInputEncoding

On Secrecy Performance of RIS-Assisted MISO Systems over Rician Channels with
Spatially Random Eavesdroppers

Wei Shi, , Jindan Xu, , Wei Xu, , Chau Yuen, , A. Lee Swindlehurst, , and Chunming Zhao W. Shi, W. Xu, and C. Zhao are with the National Mobile Communications Research Laboratory, Southeast University, Nanjing 210096, China, and are also with the Purple Mountain Laboratories, Nanjing 211111, China (e-mail: {wshi, wxu, cmzhao}@seu.edu.cn).J. Xu and C. Yuen are with the School of Electrical and Electronics Engineering, Nanyang Technological University, Singapore 639798, Singapore (e-mail: [email protected], [email protected]).A. L. Swindlehurst is with the Center for Pervasive Communications and Computing, Henry Samueli School of Engineering, University of California at Irvine, Irvine, CA 92697 USA (e-mail: [email protected]).Part of this paper was accepted by IEEE GLOBECOM 2023 Workshops (GC Wkshps) [1].
Abstract

Reconfigurable intelligent surface (RIS) technology is emerging as a promising technique for performance enhancement for next-generation wireless networks. This paper investigates the physical layer security of an RIS-assisted multiple-antenna communication system in the presence of random spatially distributed eavesdroppers. The RIS-to-ground channels are assumed to experience Rician fading. Using stochastic geometry, exact distributions of the received signal-to-noise-ratios (SNRs) at the legitimate user and the eavesdroppers located according to a Poisson point process (PPP) are derived, and closed-form expressions for the secrecy outage probability (SOP) and the ergodic secrecy capacity (ESC) are obtained to provide insightful guidelines for system design. First, the secrecy diversity order is obtained as 2α2\frac{2}{\alpha_{2}}, where α2\alpha_{2} denotes the path loss exponent of the RIS-to-ground links. Then, it is revealed that the secrecy performance is mainly affected by the number of RIS reflecting elements, NN, and the impact of the number of transmit antennas and transmit power at the base station is marginal. In addition, when the locations of the randomly located eavesdroppers are unknown, deploying the RIS closer to the legitimate user rather than to the base station is shown to be more efficient. Moreover, it is also found that the density of randomly located eavesdroppers, λe\lambda_{e}, has an additive effect on the asymptotic ESC performance given by log2(1/λe)\log_{2}{\left({1}/{\lambda_{e}}\right)}. Finally, numerical simulations are conducted to verify the accuracy of these theoretical observations.

Index Terms:
Reconfigurable intelligent surface (RIS), physical layer security, secrecy outage probability (SOP), ergodic secrecy capacity (ESC), stochastic geometry.

I Introduction

Reconfigurable intelligent surface (RIS) technology has recently been recognized as a promising approach for realizing both spectral and energy efficient communications in future wireless networks [2, 5, 4, 3]. An RIS comprises a large number of low-cost passive reflecting elements that are able to independently control the phase shifts and/or amplitudes of their reflection coefficients. In this way, the RIS can realize accurate beamforming for adjusting the propagation environments and thus improving the signal quality at desired receivers. In addition, unlike traditional relay transmission, an RIS with miniaturized circuits does not generate new signals or thermal noise. Hence, RISs can be flexibly installed on outdoor buildings, signage, street lamps, and indoor ceilings to help provide additional high-quality links [5]. Due to these advantages, RISs have been widely studied to support a broad range of communication requirements, including data rate enhancement [6, 7, 8], coverage extension [9, 10, 11], and interference mitigation [12, 13, 14, 15].

In recent years, with the fast-growing number of wireless devices, security for wireless communication has become a critical issue. As a complement to conventional complicated cryptographic methods, physical layer security (PLS) approach leverages the physical characteristics of the propagation environment for enhancing cellular network security against eavesdropping attacks. The capability of an RIS to create a smart controllable wireless propagation environment makes it a promising approach for providing PLS [16]. For example, the authors of [17] and [18] investigated the joint optimization of the active and nearly passive beamforming at the transmitter and the RIS to maximize the theoretical secrecy rate. Furthermore, the design of artificial noise (AN) was also considered in [19] for maximizing the system sum-rate while limiting information leakage to potential eavesdroppers.

On the other hand, there are multiple works that investigate the theoretical secrecy performance for RIS-enhanced PLS systems in terms of secrecy outage probability (SOP) and ergodic secrecy capacity (ESC) [20, 21, 22, 23, 24, 25, 26]. In particular, the SOP of an RIS-aided single-antenna system was first studied in the presence of an eavesdropper [20]. An SOP and ESC analysis that considered implementation issues was conducted in [21] and [22], respectively, under the assumption of discrete RIS phase shifts. RIS-aided secure communications were also studied in emerging applications, such as Device-to-Device (D2D) [23], vehicular networks [24], unmanned aerial vehicle (UAV) [25], and non-terrestrial networks [26]. However, for analytical simplicity and mathematical tractability, most works have considered single-antenna nodes and Rayleigh fading channels, and overlooked randomly distributed eavesdroppers. In order to investigate a more practical RIS-aided secure system, the randomness of the eavesdropper locations has to be taken into account when analyzing the system performance.

Stochastic geometry is an efficient mathematical tool for capturing the topological randomness of networks [27]. In this approach, the wireless network is conveniently abstracted to a point process that can capture the essential network properties. A homogeneous Poisson point process (PPP) is the most popular and tractable point process used to model the locations of the mobile devices in wireless networks [28]. However, there have been few works studying the secrecy performance of an RIS-aided communication system with spatially random eavesdroppers, which leaves the impact of key system parameters under the stochastic geometry framework still unclear.

I-A Motivation and Contribution

PLS has been studied in diverse RIS-assisted communication scenarios, but rarely considered in the general case of randomly located eavesdroppers. Although the authors of [29] and [30] considered random eavesdropper locations, there are still several research gaps left to be filled. In [29], a single-antenna setting was adopted at the base station to analyze the ESC performance, but the Rician fading assumption and RIS phase shifts optimization were not taken into consideration for the multiple-antenna setting. In [30], a study was developed based on a simplified transmit beamforming design and the analyses of secrecy diversity order and capacity were not conducted. In addition, the works in [29] and [30] both need to be further explored to quantify the impact of the key parameters, e.g., the number of RIS reflecting elements, on the attainable secrecy performance in order to provide insightful guidelines for system design. Table I provides a summary of current studies related to the RIS-assisted secure communication systems and compares our work with them. In this paper, we investigate the SOP and ESC performance of an RIS-assisted multiple-input single-output (MISO) system over Rician fading channels with spatially random eavesdroppers. The main contributions of our work are summarized as follows.

TABLE I: Comparison between our work with the state-of-the-art.
[20] [21] [22] [23] [24] [25] [26] [29] [30] Our work
MISO \checkmark \checkmark \checkmark
Rician channel \checkmark \checkmark
Eves with PPP distribution \checkmark \checkmark \checkmark
SOP analysis \checkmark \checkmark \checkmark \checkmark \checkmark \checkmark \checkmark
ESC analysis \checkmark \checkmark \checkmark \checkmark
High-SNR analysis \checkmark \checkmark \checkmark \checkmark
  • Assuming maximum ratio transmission (MRT) for the transmit beamforming, the optimal reflect beamforming at the RIS is designed. Using tools from stochastic geometry, we derive accurate closed-form expressions for the distributions of the received signal-to-noise-ratios (SNRs) at the legitimate user and randomly located eavesdroppers located according to a homogeneous PPP.

  • We propose a new analytical PLS framework for the RIS-aided MISO secure communication system over Rician fading channels. Novel closed-form expressions are derived to characterize the SOP and ESC of the system. The derived results show that both the SOP and ESC are mainly affected by the number of RIS reflecting elements, and are not strong functions of the number of transmit antennas nor the transmit power at the base station, which means that increasing either of these latter resources will not significantly improve the secrecy performance.

  • To obtain more insightful observations, the asymptotic secrecy performance in the high SNR regime is also analyzed to characterize the SOP and ESC. The asymptotic SOP demonstrates that the secrecy diversity order of RIS-aided MISO secure communication systems depends on the path loss exponent of the RIS-to-ground links. In addition, when the locations of spatially random eavesdroppers are unknown, it is more efficient to deploy the RIS closer to the legitimate user than to the base station. Moreover, the impact of the randomly located eavesdropper density, λe\lambda_{e}, on the asymptotic ESC performance is quantitatively evaluated to be additive and proportional to log2(1/λe)\log_{2}{\left({1}/{\lambda_{e}}\right)}.

I-B Organization

The rest of this paper is organized as follows. We introduce the system model in Section II, and in Section III, we derive the exact distributions of the received SNRs for the legitimate user and randomly located eavesdroppers. In Section IV, we provide a closed-form expression for the SOP, and we study the secrecy diversity order at high SNR. The ergodic secrecy capacity is investigated in Section V and simulation results are discussed in Section VI. Finally, we draw our conclusions in Section VII.

Notation: Boldface lowercase (uppercase) letters represent vectors (matrices). The set of all complex numbers is denoted by \mathbb{C}. The set of all positive real numbers is denoted by +\mathbb{Z}^{+}. The superscripts ()T(\cdot)^{T}, ()(\cdot)^{\ast}, and ()H(\cdot)^{H} stand for the transpose, conjugate, and conjugate-transpose operations, respectively. A circularly symmetric complex Gaussian distribution is denoted by 𝒞𝒩{\cal{CN}}. A Rician distribution is denoted by RiceRice. 𝔼{}\mathbb{E}\{\cdot\} denotes the expectation of a random variable (RV). diag{}{\rm diag}\left\{\cdot\right\} indicates a diagonal matrix. The operators |||\cdot| and \left\|\cdot\right\| take the norm of a complex number and a vector, respectively. \angle returns the phase of a complex value. The symbol Γ()\Gamma\left(\cdot\right) is the Gamma function. The symbols γ(,)\gamma\left(\cdot,\cdot\right) and Γ(,)\Gamma\left(\cdot,\cdot\right) denote the lower and upper incomplete Gamma functions, respectively. [x]+=max{0,x}\left[x\right]^{+}={\rm max}\left\{0,x\right\} returns the maximum between 0 and xx.

II System Model

As illustrated in Fig. 1, an RIS-assisted secure communication system is considered, where a base station (SS) is equipped with KK antennas and an RIS is composed of NN reflecting elements. The reflection matrix of the RIS is denoted by 𝚯diag{η1ejθ1,,ηnejθn,,ηNejθN}\mathbf{\Theta}\triangleq{\rm diag}\left\{{\eta_{1}}{\rm e}^{j\theta_{1}},\ldots,{\eta_{n}}{\rm e}^{j\theta_{n}},\ldots,{\eta_{N}}{\rm e}^{j\theta_{N}}\right\}, where ηn[0,1]\eta_{n}\in[0,1] and θn[0,2π)\theta_{n}\in[0,2\pi) for n=1,2,,Nn=1,2,\ldots,N are, respectively, the amplitude coefficient and the phase shift introduced by the nn-th reflecting element. In order to exploit the maximum reflection capability of the RIS, the amplitude coefficients in this paper are set to 1, i.e., ηn=1\eta_{n}\!=\!1 for all nn. The eavesdroppers (EE) are randomly located within a disk of radius rer_{e} centered at the RIS, and their spatial distribution is modeled using a homogeneous PPP Φe\Phi_{e} with a density λe\lambda_{e} [28, 29, 30]. In contrast, the legitimate user is located without any spatial restrictions.

Refer to caption
Figure 1: System model of an RIS-assisted MISO communication system in the presence of spatially random eavesdroppers.

We assume that the direct link between SS and the legitimate user (DD) is blocked by obstacles, which is a common occurrence in high frequency bands. To address this issue, the data transmission between SS and DD is facilitated by the RIS. Since the base station and the RIS are typically deployed at an elevated height with few scatterers, the channel between SS and the RIS is assumed to obey a line-of-sight (LoS) model [31, 32, 33], denoted by 𝐇SRN×K\mathbf{H}_{S\!R}\in\mathbb{C}^{N\times K}. While the legitimate user and eavesdroppers are usually located on the ground, the RIS-related channels with these terminals undergo both direct LoS and rich scattering, which can be modeled using Rician fading. Here, 𝐡RiN×1\mathbf{h}_{Ri}\in\mathbb{C}^{N\times 1} is the channel vector of the RIS-ii links, where i{D,Em}i\in\left\{D,E_{m}\right\} and EmE_{m} represents the mm-th eavesdropper. Specifically, the expressions for 𝐇SR\mathbf{H}_{S\!R} and 𝐡Ri\mathbf{h}_{Ri} are given by

𝐇SR=ν𝐇¯SR,\displaystyle\mathbf{H}_{S\!R}=\sqrt{\nu}{\overline{\mathbf{H}}}_{S\!R}, (1)
𝐡Ri=μi(ϵϵ+1𝐡¯Ri+1ϵ+1𝐡~Ri),\displaystyle\mathbf{h}_{Ri}=\sqrt{\mu_{i}}\left(\sqrt{\frac{\epsilon}{\epsilon+1}}{\overline{\mathbf{h}}}_{Ri}+\sqrt{\frac{1}{\epsilon+1}}{\widetilde{\mathbf{h}}}_{Ri}\right), (2)

where ν=β0dSRα1\nu=\beta_{0}d_{S\!R}^{-\alpha_{1}} and μi=β0dRiα2\mu_{i}=\beta_{0}d_{Ri}^{-\alpha_{2}} denote the large-scale fading coefficients, β0\beta_{0} is the path loss at a reference distance of 11m, dSRd_{S\!R} and dRid_{Ri} are the distances of the SS-RIS and RIS-ii links respectively, α12\alpha_{1}\geq 2 and α22\alpha_{2}\geq 2 are the path loss exponents for the SS-RIS and RIS-ii links respectively, and ϵ\epsilon denotes the Rician factor. 111The Rician factors/path loss exponents for the RIS-DD and RIS-EE channels are assumed to be identical, as shown in [30][34][35], since the propagation environments around the legitimate user and eavesdroppers are similar. The vector 𝐡~Ri{\widetilde{\mathbf{h}}}_{Ri} represents the non-line-of-sight (NLoS) component, whose entries are standard independent and identically distributed (i.i.d.) Gaussian RVs, i.e., 𝒞𝒩(0,1)\mathcal{CN}(0,1). The LoS components 𝐇¯SR{\overline{\mathbf{H}}}_{S\!R} and 𝐡¯Ri{\overline{\mathbf{h}}}_{Ri} are expressed as

𝐇¯SR=𝐚N(ϕSRa,ϕSRe)𝐚KH(ψSRa,ψSRe)=𝐚N,SR𝐚K,SRH,\displaystyle{\overline{\mathbf{H}}}_{S\!R}=\mathbf{a}_{N}\left(\phi_{S\!R}^{a},\phi_{S\!R}^{e}\right)\mathbf{a}_{K}^{H}\left(\psi_{S\!R}^{a},\psi_{S\!R}^{e}\right)=\mathbf{a}_{N,S\!R}\mathbf{a}_{K,S\!R}^{H}, (3)
𝐡¯Ri=𝐚N(ψRia,ψRie)=𝐚N,Ri,\displaystyle{\overline{\mathbf{h}}}_{Ri}=\mathbf{a}_{N}\left(\psi_{Ri}^{a},\psi_{Ri}^{e}\right)=\mathbf{a}_{N,Ri}, (4)

where ϕSRa\phi_{S\!R}^{a} (ϕSRe\phi_{S\!R}^{e}) is the azimuth (elevation) angle of arrival (AoA) at the RIS, ψSRa\psi_{S\!R}^{a} (ψSRe\psi_{S\!R}^{e}) and ψRia\psi_{Ri}^{a} (ψRie\psi_{Ri}^{e}) are the azimuth (elevation) angles of departure (AoD) at the base station and RIS, respectively, and 𝐚Z(ϑa,ϑe)\mathbf{a}_{Z}\left(\vartheta^{a},\vartheta^{e}\right) is the array response vector. We consider a uniform square planar array (USPA) deployed at both the base station and the RIS. Thus, the array response vector can be written as [36]

𝐚Z(ϑa,ϑe)\displaystyle\mathbf{a}_{Z}\left(\vartheta^{a},\vartheta^{e}\right) =[1,,ej2πdλ(xsinϑasinϑe+ycosϑe),,\displaystyle\!=\!\left[1,\ldots,{\rm e}^{j2\pi\frac{d}{\lambda}\left(x\sin{\vartheta^{a}}\sin{\vartheta^{e}}+y\cos{\vartheta^{e}}\right)},\ldots,\right.
ej2πdλ((Z1)sinϑasinϑe+(Z1)cosϑe)]T,\displaystyle\left.{\rm e}^{j2\pi\frac{d}{\lambda}\left(\left(\sqrt{Z}\!-\!1\right)\sin{\vartheta^{a}}\sin{\vartheta^{e}}+\left(\sqrt{Z}\!-\!1\right)\cos{\vartheta^{e}}\right)}\right]^{T}, (5)

where dd and λ\lambda are the element spacing and signal wavelength, respectively, and 0x,y<Z0\leq x,y<\sqrt{Z} are the element indices in the plane. We assume that the channel coefficients of 𝐇SR\mathbf{H}_{S\!R} and 𝐡RD\mathbf{h}_{R\!D} are perfectly known to SS and extensive approaches have been proposed in literature for the channel estimation of RIS-aided links [37, 38, 39]. However, the channel coefficient 𝐡REm\mathbf{h}_{R\!E_{m}} is typically not available to SS, since eavesdroppers are usually passive devices that do not emit signals.

III Distributions of the Received SNRs

In order to analyze the secrecy performance of the system, we need to first characterize the distributions of the received SNRs at the legitimate user and the eavesdroppers.

III-A Distribution of the Received SNR at DD

Assuming quasi-static flat fading channels, the signal received at DD is expressed as

rD=𝐠DH𝐟s+nD,r_{D}=\mathbf{g}_{D}^{H}\mathbf{f}s+n_{D}, (6)

where the cascaded channel 𝐠DH𝐡RDH𝚯𝐇SR1×K\mathbf{g}_{D}^{H}\triangleq\mathbf{h}_{R\!D}^{H}\mathbf{\Theta}\mathbf{H}_{S\!R}\in\mathbb{C}^{1\times K}, 𝐟K×1\mathbf{f}\in\mathbb{C}^{K\times 1} is the normalized beamforming vector, ss denotes the transmit signal that satisfies the power constraint 𝔼{|s|2}=PT\mathbb{E}\{\left|s\right|^{2}\}=P_{T}, and nD𝒞𝒩(0,σD2)n_{D}\sim\mathcal{CN}(0,\sigma_{D}^{2}) is additive white Gaussian noise (AWGN) at DD with variance σD2\sigma_{D}^{2}. Therefore, the received SNR at DD is calculated as

γD=PT|𝐠DH𝐟|2σD2=ρd|A|2,\gamma_{D}=\frac{P_{T}\left|\mathbf{g}_{D}^{H}\mathbf{f}\right|^{2}}{\sigma_{D}^{2}}=\rho_{d}\left|{\rm A}\right|^{2}, (7)

where the RV |A||𝐠DH𝐟|\left|{\rm A}\right|\triangleq\left|\mathbf{g}_{D}^{H}\mathbf{f}\right|, and ρdPTσD2\rho_{d}\triangleq\frac{P_{T}}{\sigma_{D}^{2}} denotes the transmit SNR from SS to DD.

Due to passive eavesdropping, the channel state information of the eavesdropper is not known to the RIS, thus we determine the transmit beamforming at SS and the reflect beamforming at the RIS by maximizing the received signal power at DD.

Theorem 1: When MRT beamforming is adopted, i.e., 𝐟=𝐠D𝐠DH\mathbf{f}=\frac{\mathbf{g}_{D}}{\left\|\mathbf{g}_{D}^{H}\right\|}, the optimal reflection matrix of the RIS is given as

𝚯opt=diag{ej(diag{𝐡RDH}𝐚N,SR)}.\mathbf{\Theta}^{\rm opt}={\rm diag}\left\{{\rm e}^{-j\angle\left({\rm diag}\left\{\mathbf{h}_{R\!D}^{H}\right\}\mathbf{a}_{N\!,S\!R}\right)}\right\}. (8)

Proof: See Appendix A.\hfill\blacksquare

With the optimized RIS phase shifts in Theorem 1, the RV |A|\left|{\rm A}\right| is expressed as |A|=Kνn=1N|hRD(n)|\left|{\rm A}\right|\!=\!\sqrt{K\nu}\sum_{n=1}^{N}\left|h_{R\!D}\left(n\right)\right| which follows the distribution characterized in the following lemma.

Lemma 1: The cumulative distribution function (CDF) of |A|\left|{\rm A}\right| is well approximated by

F|A|(x)=1Γ(k)γ(k,xθ),F_{\left|{\rm A}\right|}\left(x\right)=\frac{1}{\Gamma\left(k\right)}\gamma\left(k,\frac{x}{\theta}\right), (9)

with shape parameter k=Nπ4(L12(ϵ))21+ϵπ4(L12(ϵ))2k=N\frac{\frac{\pi}{4}\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)^{2}}{1+\epsilon-\frac{\pi}{4}\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)^{2}} and scale parameter θ=KμDνϵ+11+ϵπ4(L12(ϵ))2π2L12(ϵ)\theta=\sqrt{K}\sqrt{\frac{\mu_{D}\nu}{\epsilon+1}}\frac{1+\epsilon-\frac{\pi}{4}\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)^{2}}{\frac{\sqrt{\pi}}{2}L_{\frac{1}{2}}\left(-\epsilon\right)}, in which Lq(x)L_{q}\left(x\right) is the Laguerre polynomial defined in [40, Eq. (2.66)].

Proof: See Appendix B.\hfill\blacksquare

By applying Lemma 1, we can obtain the CDF and probability density function (PDF) of γD\gamma_{D} in (7), respectively, as

FγD(x)=F|A|(x/ρd)=1Γ(k)γ(k,x/ρdθ),F_{\gamma_{D}}\left(x\right)=F_{\left|{\rm A}\right|}\left(\sqrt{{x}/{\rho_{d}}}\right)=\frac{1}{\Gamma\left(k\right)}\gamma\left(k,\frac{\sqrt{{x}/{\rho_{d}}}}{\theta}\right), (10)
fγD(x)=dFγD(x)dx=ex/ρdθ(x/ρdθ)k2Γ(k)x.f_{\gamma_{D}}\left(x\right)=\frac{dF_{\gamma_{D}}\left(x\right)}{dx}=\frac{{\rm e}^{-\frac{\sqrt{{x}/{\rho_{d}}}}{\theta}}\left(\frac{\sqrt{{x}/{\rho_{d}}}}{\theta}\right)^{k}}{2\Gamma\left(k\right)x}. (11)

Corollary 1 (Asymptotic Analysis): For large NN, the average received SNR at the legitimate user DD is obtained as

𝔼{γD}=π(L1/2(ϵ))24(ϵ+1)ρdμDνKN2.\mathbb{E}\{\gamma_{D}\}=\frac{\pi\left(L_{1/2}\left(-\epsilon\right)\right)^{2}}{4\left(\epsilon+1\right)}{\rho_{d}\mu_{D}\nu}KN^{2}. (12)

Proof: We can infer from (7) and Lemma 1 that 𝔼{γD}=ρd𝔼{|A|2}=ρdk(1+k)θ2\mathbb{E}\{\gamma_{D}\}={\rho_{d}}\mathbb{E}\{|{\rm A}|^{2}\}={\rho_{d}}k(1+k){\theta}^{2} [41]. By substituting the shape and scale parameters, the proof is completed. \hfill\blacksquare

Remark 1: From Corollary 1, we see that 𝔼{γD}\mathbb{E}\{\gamma_{D}\} scales with KK and N2N^{2}, which implies that deploying more transmit antennas and RIS reflecting elements both increase the average received SNR at the legitimate user, while the impact of NN is more dominant. Such a “squared improvement” in terms of NN is due to the fact that the optimal reflect beamforming attained in Theorem 1 not only enables the system to achieve a beamforming gain of KNKN in the SS-RIS link, but also acquires an additional gain of NN by coherently collecting signals in the RIS-DD link.

Remark 2: From Corollary 1, it is also found that 𝔼{γD}\mathbb{E}\{\gamma_{D}\} is increasing with respect to (w.r.t.) the Rician factor ϵ>0\epsilon>0. In other words, the average received SNR at the legitimate user is greater when the proportion of the LoS component in the RIS-DD link is higher. It is worth noting that the average SNR in (12) can be upper bounded as 𝔼{γD}𝔼{γD}U=ρdμDνKN2\mathbb{E}\{\gamma_{D}\}\leq\mathbb{E}\{\gamma_{D}\}^{\rm U}={\rho_{d}}\mu_{D}\nu KN^{2} which holds for large NN and ϵ\epsilon since limϵ(L1/2(ϵ))2ϵ+1=4π\lim_{\epsilon\to\infty}{\frac{\left(L_{1/2}\left(-\epsilon\right)\right)^{2}}{\epsilon+1}}=\frac{4}{\pi}.

III-B Distribution of the Received SNR at EE

Before calculating the effective SNR of the independent and homogeneous PPP distributed eavesdroppers, we first derive the SNR of the mm-th eavesdropper EmE_{m}. The signal received at EmE_{m} is formulated as

rEm=𝐡REmH𝚯𝐇SR𝐟s+nEm,r_{E_{m}}=\mathbf{h}_{R\!E_{m}}^{H}\mathbf{\Theta}\mathbf{H}_{S\!R}\mathbf{f}s+n_{E_{m}}, (13)

where nEm𝒞𝒩(0,σE2)n_{E_{m}}\sim\mathcal{CN}(0,\sigma_{E}^{2}) is AWGN at EmE_{m} with variance σE2\sigma_{E}^{2}. The received SNR at EmE_{m} is given in the following proposition.

Proposition 1: The received SNR at EmE_{m} is expressed as

γEm=ρeKν|ZEm|2,\gamma_{E_{m}}=\rho_{e}K\nu\left|Z_{E_{m}}\right|^{2}, (14)

where ρePTσE2\rho_{e}\triangleq\frac{P_{T}}{\sigma_{E}^{2}} denotes the transmit SNR and we define the RV ZEmn=1NhREm(n)ejhRD(n)Z_{E_{m}}\!\triangleq\!\sum_{n=1}^{N}{h_{R\!E_{m}}^{\ast}\left(n\right){\rm e}^{-j\angle h_{R\!D}^{\ast}\left(n\right)}}.

Proof: See Appendix C.\hfill\blacksquare

According to Proposition 1, we present Lemma 2 before deriving the distribution of γEm\gamma_{E_{m}}.

Lemma 2: The RV ZEmZ_{E_{m}} follows a complex Gaussian distribution with mean MEmM_{E_{m}} and variance VEmV_{E_{m}}, given by

MEm=\displaystyle~{}M_{E_{m}}= μEmϵ2π4(ϵ+1)(L12(ϵ))2ejπdλ(N1)(δ1+δ2)\displaystyle\sqrt{\frac{\mu_{E_{m}}\epsilon^{2}}{{\frac{\pi}{4}\left(\epsilon+1\right)\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}}{\rm e}^{j\pi\frac{d}{\lambda}\left(\sqrt{N}-1\right)\left(\delta_{1}+\delta_{2}\right)}
×sin(πdλNδ1)sin(πdλNδ2)sin(πdλδ1)sin(πdλδ2),\displaystyle\times\frac{\sin{\left(\pi\frac{d}{\lambda}\sqrt{N}\delta_{1}\right)}\sin{\left(\pi\frac{d}{\lambda}\sqrt{N}\delta_{2}\right)}}{\sin{\left(\pi\frac{d}{\lambda}\delta_{1}\right)}\sin{\left(\pi\frac{d}{\lambda}\delta_{2}\right)}}, (15)
VEm=NμEm[1ϵ2π4(ϵ+1)(L12(ϵ))2],V_{E_{m}}=N\mu_{E_{m}}\left[1-\frac{\epsilon^{2}}{\frac{\pi}{4}\left(\epsilon+1\right){\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}\right], (16)

where δ1=sinψRDasinψRDesinψREmasinψREme\delta_{1}\!=\!\sin{\psi_{R\!D}^{a}}\sin{\psi_{R\!D}^{e}}\!-\!\sin{\psi_{R\!E_{m}}^{a}}\sin{\psi_{R\!E_{m}}^{e}} and δ2=cosψRDecosψREme\delta_{2}\!=\!\cos{\psi_{R\!D}^{e}}\!-\!\cos{\psi_{R\!E_{m}}^{e}}.

Proof: See Appendix D.\hfill\blacksquare

As disclosed in Lemma 2, we conclude that γEm\gamma_{E_{m}} is a non-central Chi-squared RV with two degrees of freedom. Then, the CDF of γEm\gamma_{E_{m}} is given by

FγEm(x)=1Q1(sσ,xσ),F_{\gamma_{E_{m}}}\left(x\right)=1-Q_{1}\left(\frac{s}{\sigma},\frac{\sqrt{x}}{\sigma}\right), (17)

where Q1(,)Q_{1}(\cdot,\cdot) is the first-order Marcum QQ-function [42], and

s=ρeKνμEmϵ2π4(ϵ+1)(L12(ϵ))2|sin(πdλNδ1)sin(πdλNδ2)sin(πdλδ1)sin(πdλδ2)|,\displaystyle\!\!\!\!s\!=\!\!\sqrt{\frac{\rho_{e}K\nu\mu_{E_{m}}\epsilon^{2}}{\!\frac{\pi}{4}\!\left(\!\epsilon\!+\!1\!\right){\!\left(\!L_{\frac{1}{2}}\!\left(\!-\epsilon\right)\right)}^{2}}}\!\left|\!\frac{\sin\!{\left(\!\pi\frac{d}{\lambda}\sqrt{N}\delta_{1}\!\right)}\!\sin{\!\left(\!\pi\frac{d}{\lambda}\sqrt{N}\delta_{2}\!\right)}}{\sin\!{\left(\pi\frac{d}{\lambda}\delta_{1}\right)}\!\sin\!{\left(\pi\frac{d}{\lambda}\delta_{2}\right)}}\!\right|, (18)
σ2=12ρeKNνμEm[1ϵ2π4(ϵ+1)(L12(ϵ))2].\displaystyle\sigma^{2}\!=\!\frac{1}{2}\rho_{e}KN\nu\mu_{E_{m}}\left[1-\frac{\epsilon^{2}}{\frac{\pi}{4}\left(\epsilon+1\right){\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}\right]. (19)

In the case of non-colluding eavesdroppers, the eavesdropper with the strongest channel dominates the secrecy performance. Thus, the corresponding CDF of the eavesdropper SNR is derived as

FγE(x)\displaystyle F_{\gamma_{E}}\left(x\right) =Pr{maxmΦeγEmx}\displaystyle={\rm Pr}\left\{\mathop{\max}\limits_{m\in{\Phi_{e}}}{\gamma_{E_{m}}\leq x}\right\}
=(a)𝔼Φe{mΦe,rmreFγEm(x)}\displaystyle\mathop{=}^{\left(\rm a\right)}\mathbb{E}_{\Phi_{e}}\left\{\prod_{m\in\Phi_{e},r_{m}\leq r_{e}}{F_{\gamma_{E_{m}}}\left(x\right)}\right\}
=(b)exp[2πλe0re(1FγEm(x))rdr]\displaystyle\mathop{=}^{\left(\rm b\right)}{\rm exp}\left[-2\pi\lambda_{e}\int_{0}^{r_{e}}{\left(1-F_{\gamma_{E_{m}}}\left(x\right)\right)r\,{\rm{d}}r}\right]
=(c)exp[2πλe0reQ1(ϖ,Ξxrα22)rdr],\displaystyle\mathop{=}^{\left(\rm c\right)}{\rm exp}\left[-2\pi\lambda_{e}\int_{0}^{r_{e}}{Q_{1}\left(\varpi,\Xi\sqrt{x}r^{\frac{\alpha_{2}}{2}}\right)r\,{\rm{d}}r}\right], (20)

where (a)({\rm a}) follows from the i.i.d. characteristic of the eavesdroppers’ SNRs and their independence from the point process Φe{\Phi_{e}}, (b)({\rm b}) follows from the probability generating functional (PGFL) of the PPP [43, Eq. (4.55)], and (c)({\rm c}) is obtained by using μEm=β0rα2\mu_{E_{m}}=\beta_{0}r^{-\alpha_{2}} and defining the parameters

ϖ\displaystyle\varpi\triangleq 2|sin(πdλNδ1)sin(πdλNδ2)sin(πdλδ1)sin(πdλδ2)|\displaystyle\sqrt{2}\left|\frac{\sin{\left(\pi\frac{d}{\lambda}\sqrt{N}\delta_{1}\right)}\sin{\left(\pi\frac{d}{\lambda}\sqrt{N}\delta_{2}\right)}}{\sin{\left(\pi\frac{d}{\lambda}\delta_{1}\right)}\sin{\left(\pi\frac{d}{\lambda}\delta_{2}\right)}}\right|
×[N(π4(ϵ+1)(L12(ϵ))2ϵ21)]12,\displaystyle\times\left[N\left(\frac{\frac{\pi}{4}\left(\epsilon+1\right){\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}{\epsilon^{2}}-1\right)\right]^{-\frac{1}{2}}, (21)
Ξ2[NKνβ0ρe(1ϵ2π4(ϵ+1)(L12(ϵ))2)]12.\displaystyle\Xi\!\triangleq\!{\sqrt{2}}\!\left[NK\nu\beta_{0}\rho_{e}\!\left(\!1\!-\!\frac{\epsilon^{2}}{\frac{\pi}{4}\!\left(\epsilon\!+\!1\right)\!{\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}\right)\right]^{-\frac{1}{2}}. (22)

From the characterization in [42, Eq. (2)], we have the following approximation for the Marcum QQ-function in (20): Q1(ϖ,Ξxrα22)exp[ev(ϖ)(Ξxrα22)μ(ϖ)]Q_{1}\left(\varpi,\Xi\sqrt{x}r^{\frac{\alpha_{2}}{2}}\right)\simeq{\rm exp}\left[-{\rm e}^{v\left(\varpi\right)}\left(\Xi\sqrt{x}r^{\frac{\alpha_{2}}{2}}\right)^{\mu\left(\varpi\right)}\right], where v(ϖ)v\left(\varpi\right) and μ(ϖ)\mu\left(\varpi\right) are polynomial functions of ϖ\varpi defined as v(ϖ)=0.840+0.327ϖ0.740ϖ2+0.083ϖ30.004ϖ4v\left(\varpi\right)\!=\!-0.840\!+\!0.327\varpi\!-\!0.740\varpi^{2}\!+\!0.083\varpi^{3}\!-\!0.004\varpi^{4} and μ(ϖ)=2.1740.592ϖ+0.593ϖ20.092ϖ3+0.005ϖ4\mu\left(\varpi\right)\!=\!2.174\!-\!0.592\varpi\!+\!0.593\varpi^{2}\!-\!0.092\varpi^{3}\!+\!0.005\varpi^{4}. Then, (20) is further calculated as

FγE(x)\displaystyle F_{\gamma_{E}}\left(x\right) =exp[2πλe0reexp[ev(ϖ)(Ξxrα22)μ(ϖ)]rdr]\displaystyle\!=\!{\rm exp}\left[\!-\!2\pi\lambda_{e}\!\int_{0}^{r_{e}}\!{{\rm exp}\!\left[-{\rm e}^{v\left(\varpi\right)}\left(\Xi\sqrt{x}r^{\frac{\alpha_{2}}{2}}\right)^{\mu\left(\varpi\right)}\right]\!r{\rm{d}}r}\right]
=exp[t0Γ(t1)Γ(t1,t2xt3)xt4],\displaystyle\!=\!{\rm exp}\left[\!-t_{0}\frac{\Gamma\left(t_{1}\right)-\Gamma\left(t_{1},t_{2}x^{t_{3}}\right)}{x^{t_{4}}}\right], (23)

where the last equality is obtained from [46, Eq. (3.326)] with the definitions t0=2πλeα22μ(ϖ)e4v(ϖ)α2μ(ϖ)Ξ4α2t_{0}=\frac{2\pi\lambda_{e}}{\frac{\alpha_{2}}{2}\mu\left(\varpi\right){\rm e}^{\frac{4v\left(\varpi\right)}{\alpha_{2}\mu\left(\varpi\right)}}\Xi^{\frac{4}{\alpha_{2}}}}, t1=2α22μ(ϖ)t_{1}=\frac{2}{\frac{\alpha_{2}}{2}\mu\left(\varpi\right)}, t2=ev(ϖ)Ξμ(ϖ)reα22μ(ϖ)t_{2}={\rm e}^{v\left(\varpi\right)}\Xi^{\mu\left(\varpi\right)}{r_{e}}^{\frac{\alpha_{2}}{2}\mu\left(\varpi\right)}, t3=μ(ϖ)2t_{3}=\frac{\mu\left(\varpi\right)}{2}, and t4=2α2t_{4}=\frac{2}{\alpha_{2}}.

Therefore, the PDF of the overall eavesdropper SNR can be further derived from (23) as

fγE(x)=\displaystyle f_{\gamma_{E}}\left(x\right)= dFγE(x)dx=t0xt41(t4γ(t1,t2xt3)\displaystyle\frac{dF_{\gamma_{E}}\left(x\right)}{dx}=t_{0}x^{-t_{4}-1}\left(t_{4}\gamma\left(t_{1},t_{2}x^{t_{3}}\right)\right.
t3(t2xt3)t1et2xt3)et0xt4γ(t1,t2xt3).\displaystyle\left.-t_{3}\left(t_{2}x^{t_{3}}\right)^{t_{1}}{\rm e}^{-t_{2}x^{t_{3}}}\right){\rm e}^{-t_{0}x^{-t_{4}}\gamma\left(t_{1},t_{2}x^{t_{3}}\right)}. (24)

IV Secrecy Outage Analysis

In this section, we apply the derived statistical properties of γD\gamma_{D} and γE\gamma_{E} in the above section to conduct the secrecy outage analysis of the RIS-aided MISO system.

IV-A Theoretical SOP Analysis

A popular metric for quantifying PLS is the SOP, which is defined as the probability that the instantaneous secrecy capacity falls below a target secrecy rate CthC_{\rm th}. Mathematically, the SOP is evaluated by

SOP\displaystyle{\rm SOP} =Pr(ln(1+γD)ln(1+γE)<Cth)\displaystyle={\rm Pr}\left(\ln{\left(1+\gamma_{D}\right)}-\ln{\left(1+\gamma_{E}\right)}<C_{\rm th}\right)
=0FγD((1+x)φ1)fγE(x)dx,\displaystyle=\int_{0}^{\infty}{F_{\gamma_{D}}\left(\left(1+x\right)\varphi-1\right)}f_{\gamma_{E}}\left(x\right){\rm{d}}x, (25)

where φeCth\varphi\triangleq{\rm e}^{C_{\rm th}}. By substituting (10) and (24) into (25), the SOP can be easily expressed as follows

SOP=1Γ(k)t0(t4I1t3t2t1I2),\displaystyle{\rm SOP}=\frac{1}{\Gamma\left(k\right)}t_{0}\left(t_{4}I_{1}-t_{3}{t_{2}}^{t_{1}}I_{2}\right), (26)

where I1I_{1} and I2I_{2} are defined as

I1=0+\displaystyle~{}~{}~{}~{}~{}I_{1}=\int_{0}^{+\infty} γ(k,((1+x)φ1)ρdθ)xt41\displaystyle\gamma\left(k,\frac{\sqrt{\frac{\left(\left(1+x\right)\varphi-1\right)}{\rho_{d}}}}{\theta}\right)x^{-t_{4}-1}
×\displaystyle\times~{} γ(t1,t2xt3)et0xt4γ(t1,t2xt3)dx,\displaystyle\gamma\left(t_{1},t_{2}x^{t_{3}}\right){\rm e}^{-t_{0}x^{-t_{4}}\gamma\left(t_{1},t_{2}x^{t_{3}}\right)}{\rm{d}}x, (27)
I2=0+\displaystyle I_{2}=\int_{0}^{+\infty} γ(k,((1+x)φ1)ρdθ)xt41xt1t3\displaystyle\gamma\left(k,\frac{\sqrt{\frac{\left(\left(1+x\right)\varphi-1\right)}{\rho_{d}}}}{\theta}\right)x^{-t_{4}-1}x^{t_{1}t_{3}}
×\displaystyle\times~{} et2xt3et0xt4γ(t1,t2xt3)dx.\displaystyle{\rm e}^{-t_{2}x^{t_{3}}}{\rm e}^{-t_{0}x^{-t_{4}}\gamma\left(t_{1},t_{2}x^{t_{3}}\right)}{\rm{d}}x. (28)

However, it is difficult to directly compute an accurate closed-form expression for (26), because (27) and (28) both involve an intractable integral. In order to analyze the secrecy performance, an approximate closed-form expression for the SOP is presented in Proposition 2.

Proposition 2: When rer_{e}\rightarrow\infty, the SOP can be approximated as222The assumption of a large rer_{e}, similar to [44][45], only denotes an upper limit of distance and does not mean that the eavesdroppers must be far away from the RIS. Also, we will illustrate that the insights we obtained under this assumption are still applicable when rer_{e} is relatively small in the simulations.

SOP\displaystyle{\rm SOP}\simeq~{} 11Γ(k)p12qk122p+4q22kπp+4q21\displaystyle 1-\frac{1}{\Gamma\left(k\right)}\frac{p^{\frac{1}{2}}q^{k-\frac{1}{2}}}{2^{\frac{p+4q}{2}-2k}\pi^{\frac{p+4q}{2}-1}}
×G0,p+4qp+4q,0((t0Γ(t1)φt4)ppp(4qρdθ)4q|Δ),\displaystyle\times G_{0,p+4q}^{p+4q,0}\left(\frac{\left(t_{0}\Gamma\left(t_{1}\right)\varphi^{t_{4}}\right)^{p}}{p^{p}\left(4q\sqrt{\rho_{d}}\theta\right)^{4q}}\middle|\begin{matrix}-\\ \Delta\\ \end{matrix}\right), (29)

where Gs,tm,n(z)G_{s,t}^{m,n}\left(z\right) is Meijer’s GG function [46], p,q+p,q\in\mathbb{Z}^{+}, p/q=α2{p}/{q}=\alpha_{2}, and Δ=[0,1p,,p1p,k4q,k+14q,,k+4q14q]\Delta=\left[0,\frac{1}{p},\ldots,\frac{p-1}{p},\frac{k}{4q},\frac{k+1}{4q},\ldots,\frac{k+4q-1}{4q}\right].

Proof: See Appendix E.\hfill\blacksquare

Proposition 2 provides an explicit relation between the secrecy outage probability and various system parameters. A number of interesting points can be made based on this expression.

Remark 3: The SOP in (29) is not a function of the transmit power PTP_{T} at the base station, which implies that increasing PTP_{T} does not enhance the system’s secrecy performance. This is intuitive since an increase in PTP_{T} would yield a proportional increase in the transmit SNRs at both the legitimate user and eavesdroppers.

Proof: According to Proposition 2, we see that the transmit power PTP_{T} affects only the term t0pρd2q\frac{t_{0}^{p}}{\rho_{d}^{2q}} in (29), which is given by

t0pρd2q\displaystyle\frac{t_{0}^{p}}{\rho_{d}^{2q}} =(2πλeα22μ(ϖ)e4v(ϖ)α2μ(ϖ))p1(Ξ2ρd)2q(ρeρd)2q.\displaystyle=\left(\frac{2\pi\lambda_{e}}{\frac{\alpha_{2}}{2}\mu\left(\varpi\right)e^{\frac{4v\left(\varpi\right)}{\alpha_{2}\mu\left(\varpi\right)}}}\right)^{p}\frac{1}{\left(\Xi^{2}\rho_{d}\right)^{2q}}\propto\left(\frac{\rho_{e}}{\rho_{d}}\right)^{2q}. (30)

This equality shows that the SOP depends on the ratio of the transmit SNRs at the legitimate user and the eavesdroppers, i.e., ρeρd\frac{\rho_{e}}{\rho_{d}}, regardless of the specific values of PTP_{T}. \hfill\blacksquare

Remark 4: The SOP in (29) is mainly affected by the number of RIS reflecting elements, NN, while the impact of the number of transmit antennas, KK, is marginal. This is readily checked by calculating t0pθ4q\frac{t_{0}^{p}}{\theta^{4q}} in (29) because the other terms are obviously irrelevant to KK. We obtain t0pθ4q=(2πλeα22μ(ϖ)e4v(ϖ)α2μ(ϖ))p1(Ξθ)4q\frac{t_{0}^{p}}{\theta^{4q}}\!\!=\!\!\left(\frac{2\pi\lambda_{e}}{\frac{\alpha_{2}}{2}\mu\left(\varpi\right){\rm e}^{\frac{4v\left(\varpi\right)}{\alpha_{2}\mu\left(\varpi\right)}}}\right)^{p}\!\!\!\!\frac{1}{\left(\Xi\theta\right)^{4q}}, where Ξθ=χN\Xi\theta\!=\!\frac{\chi}{\sqrt{N}} depends only on NN, since the coefficient χ\chi is independent of NN and KK.

In addition, Proposition 2 is a general analysis of the SOP for any path loss exponent α2\alpha_{2}. Some specific case studies are reported below. Note that other values of α2\alpha_{2} also admit closed-form expressions using (29).

Corollary 2: For the special case of α2=2\alpha_{2}=2, i.e., p=2p=2 and q=1q=1, which corresponds to free space propagation [47], the SOP in (29) reduces to

SOP12k1πΓ(k)G0,33,0(t0Γ(t1)φ4ρdθ2|0,k2,k+12).\displaystyle{\rm SOP}\simeq 1-\frac{2^{k-1}}{\sqrt{\pi}\Gamma\left(k\right)}\ G_{0,3}^{3,0}\left(\frac{t_{0}\Gamma\left(t_{1}\right)\varphi}{4\rho_{d}\theta^{2}}\middle|\begin{matrix}-\\ 0,\frac{k}{2},\frac{k+1}{2}\\ \end{matrix}\right). (31)

Corollary 3: For the special case of α2=4\alpha_{2}=4, i.e., p=4p=4 and q=1q=1, which is a common practical value for the path-loss exponent in outdoor urban environments [47], the SOP in (29) simplifies to the following expression

SOP12Γ(k)(t0Γ(t1)φρdθ)k2Kk(2(t0Γ(t1)φρdθ)12),\displaystyle{\rm SOP}\simeq 1\!-\!\frac{2}{\Gamma\left(k\right)}\!\left(\frac{t_{0}\Gamma\left(t_{1}\right)\sqrt{\varphi}}{\sqrt{\rho_{d}}\theta}\right)^{\frac{k}{2}}\!\!K_{k}\!\!\left(\!2\!\left(\frac{t_{0}\Gamma\left(t_{1}\right)\sqrt{\varphi}}{\sqrt{\rho_{d}}\theta}\right)^{\frac{1}{2}}\!\right), (32)

where Kν()K_{\nu}\left(\cdot\right) denotes the ν\nu-th-order modified Bessel function of the second kind [46, Eq. (8.407)].

From (32), it is obvious that the SOP is a monotonically increasing function w.r.t. t0ρdθ=ρeρdλedRD2β(N,ϵ)\frac{t_{0}}{\sqrt{\rho_{d}}\theta}=\sqrt{\frac{\rho_{e}}{\rho_{d}}}\lambda_{e}d_{R\!D}^{2}\beta\left(N,\epsilon\right) with fixed kk, where β(N,ϵ)\beta\left(N,\epsilon\right) consists of the parameters NN, ϵ\epsilon, and constant terms. Some new observations can thus be obtained in addition to the results presented in Remark 3 and Remark 4. We evince that the SOP increases with the density parameter λe\lambda_{e}, which implies that a larger density of randomly located eavesdroppers leads to a negative effect on the secrecy performance. Moreover, we can also see that the SOP is only related to the distance of the RIS-DD link, i.e., dRDd_{R\!D}. Therefore, when the locations of the eavesdroppers are unknown, this suggests that the RIS should be deployed closer to the legitimate user than to the base station.

IV-B Secrecy Diversity Order Analysis

In order to derive the secrecy diversity order and gain further insights, we adopt the analytical framework proposed in [48] where the secrecy diversity order is defined as follows

ds=limρdlogSOPlogρd,\displaystyle d_{s}=-\lim_{\rho_{d}\rightarrow\infty}{\frac{\log{{\rm SOP}^{\infty}}}{\log{\rho_{d}}}}, (33)

where SOP{\rm SOP}^{\infty} represents the asymptotic value of the SOP in (29) for ρd\rho_{d}\rightarrow\infty, and the transmit SNR ρe\rho_{e} is fixed.

According to [49, Eq. (07.34.06.0006.01)], the SOP in (29) can be expanded as

SOP\displaystyle{\rm SOP}\simeq 11Γ(k)p12qk122p+4q22kπp+4q21×G0,p+4qp+4q,0(x|Δ)\displaystyle 1-\frac{1}{\Gamma\left(k\right)}\frac{p^{\frac{1}{2}}q^{k-\frac{1}{2}}}{2^{\frac{p+4q}{2}-2k}\pi^{\frac{p+4q}{2}-1}}\times G_{0,p+4q}^{p+4q,0}\left(x\middle|\begin{matrix}-\\ \Delta\\ \end{matrix}\right)
=\displaystyle= 11Γ(k)p12qk122p+4q22kπp+4q21×\displaystyle 1-\frac{1}{\Gamma\left(k\right)}\frac{p^{\frac{1}{2}}q^{k-\frac{1}{2}}}{2^{\frac{p+4q}{2}-2k}\pi^{\frac{p+4q}{2}-1}}\times
l=1p+4qj=1,jlp+4qΓ(Δ(j)Δ(l))xΔ(l)(1+𝒪(x)),\displaystyle\sum_{l=1}^{p+4q}{\prod_{j=1,j\neq l}^{p+4q}\!\Gamma\left(\Delta\left(j\right)\!-\!\Delta\left(l\right)\right)x^{\Delta\left(l\right)}}\left(1\!+\!\mathcal{O}\left(x\right)\right), (34)

where x=(t0Γ(t1)φt4)ppp(4qρdθ)4q0x\!=\!\!\frac{\left(t_{0}\Gamma\left(t_{1}\right)\varphi^{t_{4}}\right)^{p}}{p^{p}\left(4q\sqrt{\rho_{d}}\theta\right)^{4q}}\!\rightarrow\!0, and 𝒪\mathcal{O} denotes higher order terms.

When the transmit SNR from SS to DD is sufficiently large, i.e., ρd\rho_{d}\rightarrow\infty, only the dominant terms l=0l=0 and l=1l=1 in the summation of (34) are retained, which yields the asymptotic SOP as expressed in (35),

SOP=\displaystyle{\rm SOP}^{\infty}= 11Γ(k)p12qk122p+4q22kπp+4q21[j=2p+4qΓ(Δ(j))x0+j=1,j2p+4qΓ(Δ(j)1p)x1p]\displaystyle 1-\frac{1}{\Gamma\left(k\right)}\frac{p^{\frac{1}{2}}q^{k-\frac{1}{2}}}{2^{\frac{p+4q}{2}-2k}\pi^{\frac{p+4q}{2}-1}}\left[\prod_{j=2}^{p+4q}\Gamma\left(\Delta\left(j\right)\right)x^{0}+\prod_{j=1,j\neq 2}^{p+4q}\Gamma\left(\Delta\left(j\right)-\frac{1}{p}\right)x^{\frac{1}{p}}\right]
=\displaystyle= 11Γ(k)p12qk122p+4q22kπp+4q21[j=0p1Γ(1p+jp)j=04q1Γ(k4q+j4q)pj=0p1Γ(1p+jp)\displaystyle 1-\frac{1}{\Gamma\left(k\right)}\frac{p^{\frac{1}{2}}q^{k-\frac{1}{2}}}{2^{\frac{p+4q}{2}-2k}\pi^{\frac{p+4q}{2}-1}}\left[\prod_{j=0}^{p-1}\Gamma\left(\frac{1}{p}+\frac{j}{p}\right)\prod_{j=0}^{4q-1}\Gamma\left(\frac{k}{4q}+\frac{j}{4q}\right)-p\prod_{j=0}^{p-1}\Gamma\left(\frac{1}{p}+\frac{j}{p}\right)\right.
×j=04q1Γ(k4q1p+j4q)x1p]=t0Γ(t1)φ2α2Γ(k4α2)θ4α2Γ(k)(ρd)2α2,\displaystyle\left.\times\prod_{j=0}^{4q-1}\Gamma\left(\frac{k}{4q}-\frac{1}{p}+\frac{j}{4q}\right)x^{\frac{1}{p}}\right]=\frac{t_{0}\mathrm{\Gamma}\left(t_{1}\right)\varphi^{\frac{2}{\alpha_{2}}}\Gamma\left(k-\frac{4}{\alpha_{2}}\right)}{\theta^{\frac{4}{\alpha_{2}}}\Gamma\left(k\right)}\left(\rho_{d}\right)^{-\frac{2}{\alpha_{2}}}, (35)

where the last step is calculated by applying Gauss’ multiplication formula [50, Eq. (6.1.20)].

Remark 5: By substituting (35) into (33), the secrecy diversity order is obtained as 2α2\frac{2}{\alpha_{2}}, which only depends on the path loss exponent of the RIS-to-ground links. This implies that the secrecy diversity order of this system improves when the RIS is deployed to provide better LoS links to the terminals.

V Ergodic Secrecy Capacity Analysis

In this section, we obtain closed-form expressions for both the theoretical and asymptotic ESC. We also characterize the impact of various parameters, including NN, KK, PTP_{T}, and λe\lambda_{e}, on the ESC performance of the system.

V-A Theoretical ESC Analysis

The ESC is an alternative fundamental metric that denotes the statistical average of the secrecy rate over fading channels, which is mathematically expressed as

Cs=𝔼{[log2(1+γD)log2(1+γE)]+}.C_{s}=\mathbb{E}\left\{\left[\log_{2}{\left(1+\gamma_{D}\right)}-\log_{2}{\left(1+\gamma_{E}\right)}\right]^{+}\right\}. (36)

Given the received SNRs at both the legitimate user and eavesdroppers in Section III, we first derive an approximate expression for the ESC in the following proposition.

Proposition 3: The ESC for the RIS-assisted system is evaluated as

C¯s=[RDRE]+,\displaystyle{\bar{C}}_{s}=\left[R_{D}-R_{E}\right]^{+}, (37)

where RDR_{D} and RER_{E} are the ergodic rates of the legitimate user DD and the eavesdroppers EE, respectively, and are expressed as

RD=1ln21Γ(k)2k122πG2,44,1(14ρdθ2|0,10,0,k2,k+12),\displaystyle R_{D}=\frac{1}{\ln{2}}\frac{1}{\Gamma\left(k\right)}\frac{2^{k-\frac{1}{2}}}{\sqrt{2\pi}}G_{2,4}^{4,1}\left(\frac{1}{4\rho_{d}\theta^{2}}\middle|\begin{matrix}0,1\\ 0,0,\frac{k}{2},\frac{k+1}{2}\\ \end{matrix}\right), (38)
RE=1ln20+1exp[t0Γ(t1)Γ(t1,t2xt3)xt4]1+xdx.\displaystyle~{}~{}R_{E}=\frac{1}{\ln{2}}\int_{0}^{+\infty}\frac{1-{\rm exp}\left[-t_{0}\frac{\Gamma\left(t_{1}\right)-\Gamma\left(t_{1},t_{2}x^{t_{3}}\right)}{x^{t_{4}}}\right]}{1+x}{\rm{d}}x. (39)

Proof: Using Jensen’s inequality, an effective approximation of the ESC can be written as [51][52]

C¯s=[𝔼{log2(1+γD)log2(1+γE)}]+=[RDRE]+,\displaystyle\!\!{\bar{C}}_{s}\!=\!\left[\mathbb{E}\!\left\{\log_{2}\!{\left(1\!+\!\gamma_{D}\right)}\!-\!\log_{2}\!{\left(1\!+\!\gamma_{E}\right)}\right\}\right]^{+}\!=\!\left[R_{D}\!-\!R_{E}\right]^{+}, (40)

where RDR_{D} and RER_{E} are respectively calculated as follows

RD\displaystyle R_{D} =𝔼{log2(1+γD)}=1ln20+F¯γD(x)1+x𝑑x\displaystyle=\mathbb{E}\!\left\{\log_{2}{\left(1+\gamma_{D}\right)}\right\}=\frac{1}{\ln{2}}\int_{0}^{+\infty}\frac{{\bar{F}}_{\gamma_{D}}\left(x\right)}{1+x}dx~{}~{}~{}~{}~{}~{}~{}~{}
=1ln20+Γ(k,x/ρdθ)Γ(k)(1+x)dx,\displaystyle=\frac{1}{\ln{2}}\int_{0}^{+\infty}\frac{\Gamma\left(k,\frac{\sqrt{{x}/{\rho_{d}}}}{\theta}\right)}{\Gamma\left(k\right)\left(1+x\right)}{\rm{d}}x, (41)
RE\displaystyle R_{E} =𝔼{log2(1+γE)}=1ln20+F¯γE(x)1+x𝑑x\displaystyle=\mathbb{E}\left\{\log_{2}{\left(1+\gamma_{E}\right)}\right\}=\frac{1}{\ln{2}}\int_{0}^{+\infty}\frac{{\bar{F}}_{\gamma_{E}}\left(x\right)}{1+x}dx
=1ln20+1exp[t0Γ(t1)Γ(t1,t2xt3)xt4]1+xdx.\displaystyle=\frac{1}{\ln{2}}\int_{0}^{+\infty}\frac{1-{\rm exp}\left[-t_{0}\frac{\Gamma\left(t_{1}\right)-\Gamma\left(t_{1},t_{2}x^{t_{3}}\right)}{x^{t_{4}}}\right]}{1+x}{\rm{d}}x. (42)

Using [49, Eq. (07.34.03.0613.01)] and [46, Eq. (9.31.5)], (41) is equivalently given by

RD\displaystyle R_{D} =1ln21Γ(k)0+11+xG1,22,0(xρdθ|1k,0)dx\displaystyle=\frac{1}{\ln{2}}\frac{1}{\Gamma\left(k\right)}\int_{0}^{+\infty}{\frac{1}{1+x}G_{1,2}^{2,0}\left(\frac{\sqrt{x}}{\sqrt{\rho_{d}}\theta}\middle|\begin{matrix}1\\ k,0\\ \end{matrix}\right)}{\rm{d}}x
=1ln21Γ(k)2k122πG3,55,1(14ρdθ2|0,12,10,0,12,k2,k+12),\displaystyle=\frac{1}{\ln{2}}\frac{1}{\Gamma\left(k\right)}\frac{2^{k-\frac{1}{2}}}{\sqrt{2\pi}}G_{3,5}^{5,1}\left(\frac{1}{4\rho_{d}\theta^{2}}\middle|\begin{matrix}0,\frac{1}{2},1\\ 0,0,\frac{1}{2},\frac{k}{2},\frac{k+1}{2}\\ \end{matrix}\right), (43)

where the integral of (43) is evaluated through [49, Eq. (07.34.21.0086.01)].

By further applying the identity given in [46, Eq. (9.31.1)], a simplified expression for the ergodic rate of the legitimate user DD is obtained in (38). The proof is thus complete.\hfill\blacksquare

As a general analysis, the ergodic rate RER_{E} given by (39) involves an intractable integral that is difficult to compute. So in the sequel we focus on a few relevant cases for widely-used values of path loss exponents where significant simplification is possible and intuitive ergodic secrecy rate expressions can be obtained.

Corollary 4: For rer_{e}\rightarrow\infty and α2=2\alpha_{2}=2, which corresponds to the free space model [47], the ESC in (37) reduces to the closed-form expression as

C¯s=[RDRE,1]+,\displaystyle{\bar{C}}_{s}=\left[R_{D}-R_{E,1}\right]^{+}, (44)

where RDR_{D} is given in closed-form by (38) and

RE,1=\displaystyle R_{E,1}~{}=~{} 1ln20+1exp[t0Γ(t1)x1]1+xdx\displaystyle\frac{1}{\ln{2}}\int_{0}^{+\infty}\frac{1-{\rm exp}\left[-t_{0}\Gamma\left(t_{1}\right)x^{-1}\right]}{1+x}{\rm{d}}x
=\displaystyle=~{} 1ln2{γ+ln(t0Γ(t1))+exp[t0Γ(t1)]\displaystyle\frac{1}{\ln{2}}\left\{\gamma+\ln{\left(t_{0}\Gamma\left(t_{1}\right)\right)}+{\rm exp}\left[t_{0}\Gamma\left(t_{1}\right)\right]\right.
×(Shi(t0Γ(t1))Chi(t0Γ(t1)))},\displaystyle\left.\times\left({\rm Shi}\left(t_{0}\Gamma\left(t_{1}\right)\right)-{\rm Chi}\left(t_{0}\Gamma\left(t_{1}\right)\right)\right)\right\}, (45)

where γ\gamma denotes Euler’s constant, and Shi(){\rm Shi}\left(\cdot\right) and Chi(){\rm Chi}\left(\cdot\right) are the hyperbolic sine and cosine integral functions, respectively [46, Eq. (8.221)].

Corollary 5: For rer_{e}\rightarrow\infty and α2=4\alpha_{2}=4, which is a common value for the path-loss exponent in outdoor urban environments [47], the ESC in (37) reduces to the closed-form expression

C¯s=[RDRE,2]+,\displaystyle{\bar{C}}_{s}=\left[R_{D}-R_{E,2}\right]^{+}, (46)

where RDR_{D} is the same as (38), and

RE,2\displaystyle R_{E,2} =1ln20+1exp[t0Γ(t1)x12]1+xdx\displaystyle=\frac{1}{\ln{2}}\int_{0}^{+\infty}\frac{1-{\rm exp}\left[-t_{0}\Gamma\left(t_{1}\right)x^{-\frac{1}{2}}\right]}{1+x}{\rm{d}}x
=1ln21πG2,43,2(t02Γ(t1)24|1,112,1,1,0).\displaystyle=\frac{1}{\ln{2}}\frac{1}{\sqrt{\pi}}\ G_{2,4}^{3,2}\left(\frac{t_{0}^{2}{\Gamma\left(t_{1}\right)}^{2}}{4}\middle|\begin{matrix}1,1\\ \frac{1}{2},1,1,0\\ \end{matrix}\right). (47)

V-B Asymptotic ESC Analysis

In order to obtain useful insights for system design, we analyze the asymptotic ESC at high SNR in this subsection. For the sake of tractability, we first derive new expressions for bounding the ergodic rate RDR_{D} in the following lemma.

Lemma 3: The ergodic rate RDR_{D} can be upper bounded by

RDU=log2(1+ρdKNνμD(1+π4(N1)(L12(ϵ))2ϵ+1)).\displaystyle\!\!\!\!R_{D}^{\rm U}\!=\!\log_{2}\!{\!\left(\!1\!+\!\rho_{d}KN\nu\mu_{D}\!\!\left(\!1\!+\!\!\frac{\pi}{4}\!\left(N\!-\!1\right)\frac{\left(\!L_{\frac{1}{2}}\!\left(\!-\epsilon\right)\!\right)^{2}}{\epsilon\!+\!1}\!\right)\!\right)}. (48)

Proof: See Appendix F.\hfill\blacksquare

By applying Lemma 3, we are ready to give the following corollary quantitatively analyzing the asymptotic ESC in the high SNR region.

Corollary 6: In the high SNR regime, the ESC in (44) is further simplified as follows

C¯s\displaystyle{\bar{C}}_{s}\rightarrow {log2(σE2σD2)+log2(dRD2πλe)γln2+C\displaystyle\left\{\log_{2}{\left(\frac{\sigma_{E}^{2}}{\sigma_{D}^{2}}\right)}+\log_{2}{\left(\frac{d_{RD}^{-2}}{\pi\lambda_{e}}\right)}-\frac{\gamma}{\ln{2}}+C\right.
+log2(1+(N1)π41ϵ+1(L12(ϵ))21ϵ2π4(ϵ+1)(L12(ϵ))2)}+,\displaystyle+\left.\log_{2}{\left(\frac{1\!+\!\left(N-1\right)\frac{\pi}{4}\frac{1}{\epsilon+1}\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)^{2}}{1-\frac{\epsilon^{2}}{{\frac{\pi}{4}\left(\epsilon+1\right)\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}}\right)}\right\}^{+}, (49)

where C=log2μ(ϖ)e2v(ϖ)μ(ϖ)Γ(2μ(ϖ))C=\log_{2}\frac{\mu\left(\varpi\right){\rm e}^{\frac{2v\left(\varpi\right)}{\mu\left(\varpi\right)}}}{\Gamma\left(\frac{2}{\mu\left(\varpi\right)}\right)} is a constant.

Proof: See Appendix G.\hfill\blacksquare

From Corollary 6, we have the following remarks on the impact of key system parameters, including NN, KK, PTP_{T}, and λe\lambda_{e}, on the secrecy performance.

Remark 6: By direct inspection of (49), it is evident that the asymptotic ESC does not depend on the number of transmit antennas, KK, but it increases with the number of RIS reflecting elements, NN. This suggests that deploying more RIS reflecting elements rather than transmit antennas achieves better secrecy performance. Moreover, for large NN, the asymptotic ESC scales logarithmically with NN.

Remark 7: Corollary 6 also indicates that the asymptotic ESC is not a function of dSRd_{S\!R} whereas it increases with decreasing dRDd_{R\!D}. This result might seem a bit counterintuitive at first, but it actually makes sense and can be explained as follows. When the distance from the base station to the RIS decreases, the received SNRs at both the legitimate user and the eavesdroppers increase by the same amount, and they offset each other. Therefore, if the locations of the eavesdroppers are unknown, this suggests that the RIS should be deployed closer to the legitimate user than to the base station.

Remark 8: In agreement with Remark 3, the asymptotic ESC in (49) is only related to the ratio of the noise power at the legitimate user and the eavesdroppers σD2σE2\frac{\sigma_{D}^{2}}{\sigma_{E}^{2}}, and is independent of the transmit power PTP_{T}. In addition, in line with intuition, the asymptotic ESC decreases when the density of the homogeneous PPP λe\lambda_{e} increases. We see that the effect of λe\lambda_{e} on the ESC is additive, and is proportional to log2(1/λe)\log_{2}{\left({1}/{\lambda_{e}}\right)}.

VI Numerical Results

In this section, Monte-Carlo simulations are presented to validate the analytical results. All the simulation results are obtained by averaging over 10510^{5} independent channel realizations. We first verify the approximations of the received SNR distributions at the legitimate user and the eavesdroppers in Fig. 2. The simulation parameters are set to K=16K=16, N=36N=36, α1=α2=2\alpha_{1}=\alpha_{2}=2, ϵ=2\epsilon=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, and λe=103\lambda_{e}=10^{-3}. We see from Fig. 2 that both the analytical CDF of the received SNRs at the legitimate user DD and the eavesdroppers EE characterized by (10) and (23), respectively, match well with the numerical curves. In addition, the asymptotic CDF of the received SNR at the eavesdroppers EE calculated from (61) for large rer_{e} is also verified to be quite accurate.

Refer to caption
Figure 2: The CDF of the received SNRs at the legitimate user DD and the eavesdroppers EE, respectively.

VI-A Secrecy Outage Probability

In this subsection, we compare the SOP obtained from Monte-Carlo simulations and the analytical results calculated from (29). Fig. 3 and Fig. 4 plot the SOP versus the transmit SNR ρd\rho_{d} for different values of NN and KK, respectively. Again, the analytical expressions in (29) match very well with the numerical results, and the SOP always decreases as the transmit SNR ρd\rho_{d} increases. Furthermore, as expected from Remark 4, the SOP obviously decreases as NN increases. However, the SOP remains almost the same when KK increases with fixed NN, which means that the impact of the number of transmit antennas on the secrecy outage performance is negligible.

Fig. 5 depicts the SOP as a function of the Rician factor ϵ\epsilon and the eavesdropper density λe\lambda_{e}. It is observed that the SOP improves as ϵ\epsilon increases. This is because with a large Rician factor, the channels are dominated by the LoS component with better link quality than NLoS. Additionally, it is noteworthy that the SOP increases with λe\lambda_{e}, since a larger λe\lambda_{e} leads to more eavesdroppers which increases the likelihood that the worst-case eavesdropper obtains higher quality information.

Refer to caption
Figure 3: The SOP versus ρd\rho_{d}, with K=16K=16, α1=α2=2\alpha_{1}=\alpha_{2}=2, ϵ=2\epsilon=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, λe=103\lambda_{e}=10^{-3}, Cth=0.05C_{\rm th}=0.05, and ρe=30dB\rho_{e}=30~{}{\rm dB}.
Refer to caption
Figure 4: The SOP versus ρd\rho_{d}, with N=16N=16, α1=α2=2\alpha_{1}=\alpha_{2}=2, ϵ=2\epsilon=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, λe=103\lambda_{e}=10^{-3}, Cth=0.05C_{\rm th}=0.05, and ρe=30dB\rho_{e}=30~{}{\rm dB}.

Fig. 6 illustrates the SOP as a function of the transmit SNRs ρd\rho_{d} and ρe\rho_{e}. It is clearly shown that increasing ρd\rho_{d} and decreasing ρe\rho_{e} improve the SOP performance. We can further observe that the SOP remains the same when the ratio of the transmit SNRs at the legitimate user and the eavesdroppers, i.e., ρdρe\frac{\rho_{d}}{\rho_{e}}, is kept fixed. This implies that increasing the transmit power PTP_{T} will not help improve the secrecy outage performance of the system, which is consistent with Remark 3.

Fig. 7 shows the SOP versus the transmit SNR ρd\rho_{d} for various values of the path loss exponents α1\alpha_{1} and α2\alpha_{2}. It can be seen that the slope of the secrecy outage curves is determined by α2\alpha_{2}, which becomes less steep as α2\alpha_{2} increases. Besides, according to the definition in (33), the secrecy diversity order presented in Remark 5 can be demonstrated by calculating the negative slope of the SOP curves on a log-log scale.

Refer to caption
Figure 5: The SOP versus ϵ\epsilon and λe\lambda_{e}, with K=16K=16, N=36N=36, α1=α2=2\alpha_{1}=\alpha_{2}=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, Cth=0.05C_{\rm th}=0.05, ρd=20dB\rho_{d}=20~{}{\rm dB}, and ρe=30dB\rho_{e}=30~{}{\rm dB}.
Refer to caption
Figure 6: The SOP versus ρd\rho_{d} and ρe\rho_{e}, with K=16K=16, N=16N=16, α1=α2=2\alpha_{1}=\alpha_{2}=2, ϵ=2\epsilon=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, λe=103\lambda_{e}=10^{-3}, and Cth=0.05C_{\rm th}=0.05.
Refer to caption
Figure 7: The SOP versus ρd\rho_{d}, with K=16K=16, N=16N=16, ϵ=2\epsilon=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, λe=103\lambda_{e}=10^{-3}, Cth=0.05C_{\rm th}=0.05, and ρe=60dB\rho_{e}=60~{}{\rm dB}.

VI-B Ergodic Secrecy Capacity

In this subsection, we compare the ESC obtained from Monte-Carlo simulations with the analytical and asymptotic results calculated from (44) and (49), respectively. Fig. 8 illustrates the ESC versus the transmit SNR ρd\rho_{d} for different values of NN and KK. We can see that both the analytical and asymptotic expressions match well with the numerical results, and the ESC increases with ρd\rho_{d}. As expected from Remark 6, we can observe that the ESC at high SNR is independent of the number of transmit antennas, KK, but obviously increases with the number of RIS reflecting elements, NN. The ESC increases by approximately 22 bps/Hz when the number of RIS reflecting elements increases by a factor of 44, validating the conclusions presented in Remark 6, i.e., the ESC increases logarithmically with the number of RIS reflecting elements.

Fig. 9 plots the ESC as a function of the transmit SNRs ρd\rho_{d} and ρe\rho_{e}. First, similar to the results shown in Fig. 6, we see that increasing ρd\rho_{d} and decreasing ρe\rho_{e} both improve the secrecy performance. Also, as predicted in Remark 8, the ESC remains constant for a fixed ratio ρdρe\frac{\rho_{d}}{\rho_{e}}, regardless of the specific value of the transmit power PTP_{T}. In addition, in Fig. 9, when the ratio ρdρe\frac{\rho_{d}}{\rho_{e}} increases by a factor of 44, the ESC increases by 22 bps/Hz, which confirms the results derived in (49).

Refer to caption
Figure 8: The ESC versus ρd\rho_{d}, with α1=α2=2\alpha_{1}=\alpha_{2}=2, ϵ=2\epsilon=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, λe=103\lambda_{e}=10^{-3}, and ρe=50dB\rho_{e}=50~{}{\rm dB}.
Refer to caption
Figure 9: The ESC versus ρd\rho_{d} and ρe\rho_{e}, with K=16K=16, N=16N=16, α1=α2=2\alpha_{1}=\alpha_{2}=2, ϵ=2\epsilon=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, and λe=103\lambda_{e}=10^{-3}.

Fig. 10 shows the ESC as a function of the distances dSRd_{S\!R} and dRDd_{R\!D}. It is clear that the ESC improves as dRDd_{R\!D} decreases, and it keeps unchanged with arbitrary variation of dSRd_{S\!R}, which has been validated by Remark 7. This phenomenon can be explained by the fact that decreasing the distance from the base station to the RIS will simultaneously increase the received SNRs for both the legitimate user and the eavesdroppers.

Fig. 11 presents the ESC versus the density λe\lambda_{e} of the randomly located eavesdroppers in the logarithmic domain. It is observed that the ESC decreases linearly with log2(λe)\log_{2}{\left({\lambda_{e}}\right)}. This behavior is due to the fact that a larger λe\lambda_{e} results in the presence of more eavesdroppers within a fixed range, which degrades the secrecy performance. In addition, it can also be seen from Fig. 11 that the ESC increases with the Rician factor ϵ\epsilon. This is because the channels are mainly influenced by the LoS component when the Rician factor is large.

Refer to caption
Figure 10: The ESC versus dSRd_{S\!R} and dRDd_{R\!D}, with K=16K=16, N=16N=16, α1=α2=2\alpha_{1}=\alpha_{2}=2, ϵ=2\epsilon=2, re=200mr_{e}=200~{}{\rm m}, λe=103\lambda_{e}=10^{-3}, ρd=50dB\rho_{d}=50~{}{\rm dB}, and ρe=50dB\rho_{e}=50~{}{\rm dB}.
Refer to caption
Figure 11: The ESC versus log2(λe)\log_{2}{\left({\lambda_{e}}\right)}, with K=16K=16, N=16N=16, α1=α2=2\alpha_{1}=\alpha_{2}=2, dSR=30md_{S\!R}=30~{}{\rm m}, dRD=40md_{R\!D}=40~{}{\rm m}, re=200mr_{e}=200~{}{\rm m}, ρd=50dB\rho_{d}=50~{}{\rm dB}, and ρe=50dB\rho_{e}=50~{}{\rm dB}.

VI-C Cases for A Relatively Small rer_{e} and Rician Fading Model

In this subsection, we verify the generality of the analytical results of the SOP and ESC with a relatively small value of rer_{e}, i.e., re=50mr_{e}=50~{}{\rm m}, and under the assumption that the S-RIS and RIS-ii channels both follow Rician fading, expressed as

𝐇SR=ν(ϵ1ϵ1+1𝐇¯SR+1ϵ1+1𝐇~SR),\displaystyle\mathbf{H}_{S\!R}=\sqrt{\nu}\left(\sqrt{\frac{\epsilon_{1}}{\epsilon_{1}+1}}{\overline{\mathbf{H}}}_{S\!R}+\sqrt{\frac{1}{\epsilon_{1}+1}}{\widetilde{\mathbf{H}}}_{S\!R}\right), (50)
𝐡Ri=μi(ϵ2ϵ2+1𝐡¯Ri+1ϵ2+1𝐡~Ri),\displaystyle\mathbf{h}_{Ri}=\sqrt{\mu_{i}}\left(\sqrt{\frac{\epsilon_{2}}{\epsilon_{2}+1}}{\overline{\mathbf{h}}}_{Ri}+\sqrt{\frac{1}{\epsilon_{2}+1}}{\widetilde{\mathbf{h}}}_{Ri}\right), (51)

where ϵ1\epsilon_{1} and ϵ2\epsilon_{2} are the Rician factors for the SS-RIS and RIS-ii channels, respectively. Note that when the Rician factor ϵ1\epsilon_{1} is large, (50) simplifies to the LoS model we consider in (1). Therefore, the Rician factor ϵ1\epsilon_{1} is set to be small, e.g., ϵ1=2\epsilon_{1}=2.

As shown in Fig. 12 and Fig. 13, it can be seen that the numerical results with small rer_{e} and Rician fading model still match well with the analytical results under the assumption of large rer_{e} and LoS model. Also, the observations that the SOP and ESC are mainly affected by the number of RIS reflecting elements rather than the number of transmit antennas, which is obtained with large rer_{e} and LoS model, are also applicable.

Refer to caption
Figure 12: The SOP versus ρd\rho_{d}, with α1=α2=2\alpha_{1}\!=\!\alpha_{2}\!=\!2, ϵ1=ϵ2=2\epsilon_{1}\!=\!\epsilon_{2}\!=\!2, dSR=30md_{S\!R}\!=\!30~{}{\rm m}, dRD=40md_{R\!D}\!=\!40~{}{\rm m}, re=50mr_{e}\!=\!50~{}{\rm m}, λe=102\lambda_{e}\!=\!10^{-2}, Cth=0.05C_{\rm th}=0.05, and ρe=30dB\rho_{e}\!=\!30~{}{\rm dB}.
Refer to caption
Figure 13: The ESC versus ρd\rho_{d}, with α1=α2=2\alpha_{1}\!=\!\alpha_{2}\!=\!2, ϵ1=ϵ2=2\epsilon_{1}\!=\!\epsilon_{2}\!=\!2, dSR=30md_{S\!R}\!=\!30~{}{\rm m}, dRD=40md_{R\!D}\!=\!40~{}{\rm m}, re=50mr_{e}\!=\!50~{}{\rm m}, λe=102\lambda_{e}\!=\!10^{-2}, and ρe=50dB\rho_{e}\!=\!50~{}{\rm dB}.

VII Conclusion

In this paper, the secrecy performance of an RIS-assisted communication system with spatially random eavesdroppers was studied. Specifically, the exact distributions of the received SNRs at the legitimate user and the eavesdroppers were first derived. Then, based on these distributions, the closed-form SOP and ESC expressions were derived. Furthermore, a high-SNR secrecy diversity order and the asymptotic ESC analysis have also been conducted. The obtained results quantify the impact of key parameters on the secrecy performance and provide insightful guidelines for system design. Several simulations were provided to demonstrate the obtained results.

Appendix A Proof of Theorem 1

For the transmit beamformer 𝐟=𝐠D𝐠DH\mathbf{f}=\frac{\mathbf{g}_{D}}{\left\|\mathbf{g}_{D}^{H}\right\|}, we compute the optimal reflecting phase shifts at the RIS by maximizing the received signal power as follows

𝚯opt=argmax𝚯\displaystyle\mathbf{\Theta}^{\rm opt}=\arg\mathop{\max}\limits_{\mathbf{\Theta}} |𝐠DH𝐟|2=argmax𝚯𝐡RDH𝚯𝐇SR2\displaystyle{\left|\mathbf{g}_{D}^{H}\mathbf{f}\right|^{2}}=\arg\mathop{\max}\limits_{\mathbf{\Theta}}{\left\|\mathbf{h}_{R\!D}^{H}\mathbf{\Theta}\mathbf{H}_{S\!R}\right\|^{2}}
=argmax𝚯\displaystyle=\arg\mathop{\max}\limits_{\mathbf{\Theta}} |𝐡RDH𝚯𝐚N,SR|2𝐚K,SRH2\displaystyle{\left|\mathbf{h}_{R\!D}^{H}\mathbf{\Theta}\mathbf{a}_{N,S\!R}\right|^{2}\left\|\mathbf{a}_{K,S\!R}^{H}\right\|^{2}}
=(d)argmax𝚯\displaystyle\mathop{=}^{\left(\rm d\right)}\arg\mathop{\max}\limits_{\mathbf{\Theta}} |𝜽Hdiag{𝐡RDH}𝐚N,SR|2\displaystyle{\left|{\bm{\theta}}^{H}{\rm diag}\left\{\mathbf{h}_{R\!D}^{H}\right\}\mathbf{a}_{N,S\!R}\right|^{2}}
=argmax𝚯\displaystyle=\arg\mathop{\max}\limits_{\mathbf{\Theta}} |n=1NhRD(n)aN,SR(n)ejθn|2,\displaystyle\left|\sum_{n=1}^{N}h_{R\!D}^{\ast}\left(n\right){\rm a}_{N,S\!R}\left(n\right){\rm e}^{j\theta_{n}}\right|^{2}, (52)

where (d)({\rm d}) follows by defining 𝜽H=[ejθ1,,ejθn,,ejθN]\bm{\theta}^{H}\!=\!\!\left[{\rm e}^{j\theta_{1}},\ldots,{\rm e}^{j\theta_{n}},\ldots,{\rm e}^{j\theta_{N}}\right], and using the identity 𝐚K,SRH2=K\left\|\mathbf{a}_{K,S\!R}^{H}\right\|^{2}=K. From (52), we see that maximizing |𝐠DH𝐟|2\left|\mathbf{g}_{D}^{H}\mathbf{f}\right|^{2} is equivalent to ensuring the phases of NN complex RVs hRD(n)aN,SR(n)ejθnh_{R\!D}^{\ast}\left(n\right){\rm a}_{N,S\!R}\left(n\right){\rm e}^{j\theta_{n}} being identical. Therefore, the optimal RIS phase shifts are given by

θnopt=(hRD(n)aN,SR(n)),\displaystyle\theta_{n}^{\rm opt}=-\angle\left(h_{R\!D}^{\ast}\left(n\right){\rm a}_{N,S\!R}\left(n\right)\right), (53)

and the corresponding reflection matrix can be easily obtained as (8). This completes the proof.

Appendix B Proof of Lemma 1

Since |hRD(1)|,|hRD(2)|,,|hRD(N)|\left|h_{R\!D}\left(1\right)\right|,\left|h_{R\!D}\left(2\right)\right|,\ldots,\left|h_{R\!D}\left(N\right)\right| are i.i.d. RVs, the mean and variance of the RV |A|=Kνn=1N|hRD(n)|\left|{\rm A}\right|=\sqrt{K\nu}\sum_{n=1}^{N}\left|h_{R\!D}\left(n\right)\right| can be respectively calculated as

𝔼{|A|}=KνN𝔼{|hRD(n)|},\mathbb{E}\{\left|{\rm A}\right|\}=\sqrt{K\nu}N\cdot\mathbb{E}\{\left|h_{R\!D}\left(n\right)\right|\}, (54)

and

Var{|A|}=KνNVar{|hRD(n)|},{\rm Var}\{\left|{\rm A}\right|\}=K\nu N\cdot{\rm Var}\{\left|h_{R\!D}\left(n\right)\right|\}, (55)

where |hRD(n)|Rice(μDϵϵ+1,12μDϵ+1)\left|h_{R\!D}\left(n\right)\right|\!\sim\!Rice\left(\sqrt{\frac{\mu_{D}\epsilon}{\epsilon+1}},\sqrt{\frac{1}{2}\frac{\mu_{D}}{\epsilon+1}}\right), whose mean and variance are given as 𝔼{|hRD(n)|}=μDϵ+1π2L12(ϵ)\mathbb{E}\{\left|h_{R\!D}\left(n\right)\right|\}\!\!=\!\!\sqrt{\frac{\mu_{D}}{\epsilon+1}}\frac{\sqrt{\pi}}{2}L_{\frac{1}{2}}\left(-\epsilon\right) and Var{|hRD(n)|}=μDϵ+1[1+ϵπ4(L12(ϵ))2]{\rm Var}\{\left|h_{R\!D}\left(n\right)\right|\}=\frac{\mu_{D}}{\epsilon+1}\left[1+\epsilon-\frac{\pi}{4}\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)^{2}\right], respectively. Therefore, according to [41], the RV |A|\left|{\rm A}\right| can be approximated by a Gamma distributed RV with shape parameter k=𝔼{|A|}2Var{|A|}k=\frac{\mathbb{E}\{\left|{\rm A}\right|\}^{2}}{{\rm Var}\{\left|{\rm A}\right|\}} and scale parameter θ=Var{|A|}𝔼{|A|}\theta=\frac{{\rm Var}\{\left|{\rm A}\right|\}}{\mathbb{E}\{\left|{\rm A}\right|\}}, which yields the desired result in (9).

Appendix C Proof of Proposition 1

The received SNR at EmE_{m} can be written as

γEm\displaystyle\gamma_{E_{m}} =ρe|𝐡REmH𝚯𝐇SR𝐟|2\displaystyle=\rho_{e}\left|\mathbf{h}_{R\!E_{m}}^{H}\mathbf{\Theta}\mathbf{H}_{S\!R}\mathbf{f}\right|^{2}
=(e)ρe|𝐡REmH𝚯𝐇SR𝐇SRH𝚯H𝐡RD𝐡RDH𝚯𝐇SR|2\displaystyle\mathop{=}^{\left(\rm e\right)}\rho_{e}\left|\mathbf{h}_{R\!E_{m}}^{H}\mathbf{\Theta}\mathbf{H}_{S\!R}\frac{\mathbf{H}_{S\!R}^{H}\mathbf{\Theta}^{H}\mathbf{h}_{R\!D}}{\left\|\mathbf{h}_{R\!D}^{H}\mathbf{\Theta}\mathbf{H}_{S\!R}\right\|}\right|^{2}
=ρeν|𝐡REmH𝚯𝐚N,SR|2|𝐚N,SRH𝚯H𝐡RD|2𝐚K,SRH4|𝐡RDH𝚯𝐚N,SR|2𝐚K,SRH2\displaystyle=\rho_{e}\nu\frac{\left|\mathbf{h}_{RE_{m}}^{H}\mathbf{\Theta}\mathbf{a}_{N,SR}\right|^{2}\left|\mathbf{a}_{N,SR}^{H}\mathbf{\Theta}^{H}\mathbf{h}_{RD}\right|^{2}\left\|\mathbf{a}_{K,SR}^{H}\right\|^{4}}{\left|\mathbf{h}_{RD}^{H}\mathbf{\Theta}\mathbf{a}_{N,SR}\right|^{2}\left\|\mathbf{a}_{K,SR}^{H}\right\|^{2}}
=ρeKν|𝐡REmH𝚯𝐚N,SR|2\displaystyle=\rho_{e}K\nu\left|\mathbf{h}_{RE_{m}}^{H}\mathbf{\Theta}\mathbf{a}_{N,SR}\right|^{2}
=ρeKν|n=1NhREm(n)aN,SR(n)ejθn|2\displaystyle=\rho_{e}K\nu\left|\sum_{n=1}^{N}{h_{RE_{m}}^{\ast}\left(n\right){\rm a}_{N,SR}\left(n\right){\rm e}^{j\theta_{n}}}\right|^{2}
=(f)ρeKν|n=1NhREm(n)ejhRD(n)|2,\displaystyle\mathop{=}^{\left(\rm f\right)}\rho_{e}K\nu\left|\sum_{n=1}^{N}{h_{RE_{m}}^{\ast}\left(n\right){\rm e}^{-j\angle h_{RD}^{\ast}\left(n\right)}}\right|^{2}, (56)

where (e)({\rm e}) is obtained by substituting the expressions of 𝐟\mathbf{f} and 𝐠DH\mathbf{g}_{D}^{H}, and (f)({\rm f}) comes from (53) with |aN,SR(n)|=1\left|{\rm a}_{N,SR}\left(n\right)\right|=1.

Appendix D Proof of Lemma 2

From (2), we see that hRi(n)𝒞𝒩(μiϵϵ+1h¯Ri(n),μiϵ+1)h_{Ri}\left(n\right)\!\sim\!\mathcal{CN}\left(\sqrt{\frac{\mu_{i}\epsilon}{\epsilon+1}}{\bar{h}}_{Ri}\left(n\right),\frac{\mu_{i}}{\epsilon+1}\right), and |hRi(n)|Rice(μiϵϵ+1,12μiϵ+1)\left|h_{Ri}\left(n\right)\right|\!\sim\!Rice\left(\sqrt{\frac{\mu_{i}\epsilon}{\epsilon+1}},\sqrt{\frac{1}{2}\frac{\mu_{i}}{\epsilon+1}}\right). Then, it can be easily obtained that 𝔼{hRi(n)}=μiϵϵ+1aN,Ri(n)\mathbb{E}\{h_{Ri}\!\left(n\right)\}\!=\!\sqrt{\frac{\mu_{i}\epsilon}{\epsilon+1}}{\rm a}_{N,Ri}\!\left(n\right), 𝔼{|hRi(n)|}=μiϵ+1π2L12(ϵ)\mathbb{E}\{\left|h_{Ri}\!\left(n\right)\right|\}\!=\!\sqrt{\frac{\mu_{i}}{\epsilon+1}}\frac{\sqrt{\pi}}{2}L_{\frac{1}{2}}\!\left(-\epsilon\right), and 𝔼{|hRi(n)|2}=μi\mathbb{E}\{\left|h_{Ri}\left(n\right)\right|^{2}\}\!=\!\mu_{i}. It follows that

𝔼{ejhRD(n)}=(𝔼{hRD(n)}𝔼{|hRD(n)|})=ϵaN,RD(n)π2L12(ϵ).\displaystyle\mathbb{E}\{{\rm e}^{-j\angle h_{RD}^{\ast}\left(n\right)}\}\!=\!\left(\frac{\mathbb{E}\{h_{RD}^{\ast}\left(n\right)\}}{\mathbb{E}\{\left|h_{RD}^{\ast}\left(n\right)\right|\}}\right)^{\ast}\!=\!\frac{\sqrt{\epsilon}{\rm a}_{N,RD}\left(n\right)}{\frac{\sqrt{\pi}}{2}L_{\frac{1}{2}}\left(-\epsilon\right)}. (57)

Therefore, the mean and variance of the RV xn=hREm(n)ejhRD(n)x_{n}=h_{RE_{m}}^{\ast}\left(n\right){\rm e}^{-j\angle h_{RD}^{\ast}\left(n\right)} can be calculated, respectively, as

𝔼{xn}=μEmϵ2ϵ+1aN,REm(n)aN,RD(n)π2L12(ϵ),\displaystyle\mathbb{E}\{x_{n}\}=\sqrt{\frac{\mu_{E_{m}}\epsilon^{2}}{\epsilon+1}}\frac{{\rm a}_{N,RE_{m}}^{\ast}\left(n\right){\rm a}_{N,RD}\left(n\right)}{\frac{\sqrt{\pi}}{2}L_{\frac{1}{2}}\left(-\epsilon\right)}, (58)

and

Var{xn}=μEm[1ϵ2π4(ϵ+1)(L12(ϵ))2].\displaystyle{\rm Var}\{x_{n}\}=\mu_{E_{m}}\left[1-\frac{\epsilon^{2}}{{\frac{\pi}{4}\left(\epsilon+1\right)\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}\right]. (59)

It can be seen from (58) that 𝔼{xn}\mathbb{E}\{x_{n}\} depends on nn, which means that xnx_{n} is not identically distributed. Therefore, the distribution of the RV ZEm=n=1NxnZ_{E_{m}}=\sum_{n=1}^{N}x_{n} cannot be directly approximated as Gaussian by applying the central limit theorem (CLT). In order to characterize the distribution of ZEmZ_{E_{m}}, we first define a new RV xn𝔼{xn}x_{n}-\mathbb{E}\{x_{n}\}, and it can be easily verified that x1𝔼{x1},x2𝔼{x2},,xN𝔼{xN}x_{1}-\mathbb{E}\{x_{1}\},x_{2}-\mathbb{E}\{x_{2}\},\ldots,x_{N}-\mathbb{E}\{x_{N}\} are i.i.d. RVs with zero mean and variance Var{xn}{\rm Var}\{x_{n}\}. By virtue of the CLT, n=1N(xn𝔼{xn})\sum_{n=1}^{N}\left(x_{n}-\mathbb{E}\{x_{n}\}\right) converges in distribution to a complex Gaussian RV with zero mean and variance NVar{xn}N\!\cdot\!{\rm Var}\{x_{n}\}. Then, we can obtain that ZEm=n=1Nxn=n=1N(xn𝔼{xn})+n=1N𝔼{xn}𝒞𝒩(n=1N𝔼{xn},NVar{xn})Z_{E_{m}}=\sum_{n=1}^{N}x_{n}\!=\!\sum_{n=1}^{N}\left(x_{n}-\mathbb{E}\{x_{n}\}\right)+\sum_{n=1}^{N}\mathbb{E}\{x_{n}\}\sim\mathcal{CN}\left(\sum_{n=1}^{N}\mathbb{E}\{x_{n}\},N\!\cdot\!{\rm Var}\{x_{n}\}\right), where

n=1N𝔼{xn}=\displaystyle\sum_{n=1}^{N}\mathbb{E}\{x_{n}\}\!=\! μEmϵ2π4(ϵ+1)(L12(ϵ))2n=1NaN,REm(n)aN,RD(n)\displaystyle\sqrt{\frac{\mu_{E_{m}}\epsilon^{2}}{{\frac{\pi}{4}\!\left(\epsilon\!+\!1\right)\!\left(\!L_{\frac{1}{2}}\!\left(-\epsilon\right)\right)}^{2}}}\sum_{n=1}^{N}{{\rm a}_{N,RE_{m}}^{\ast}\!\!\left(n\right){\rm a}_{N,RD}\!\left(n\right)}
=(g)\displaystyle\mathop{=}^{\left(\rm g\right)} μEmϵ2π4(ϵ+1)(L12(ϵ))20x,yN1ej2πdλ(xδ1+yδ2)\displaystyle\sqrt{\frac{\mu_{E_{m}}\epsilon^{2}}{{\frac{\pi}{4}\!\left(\epsilon\!+\!1\right)\!\left(\!L_{\frac{1}{2}}\!\left(-\epsilon\right)\right)}^{2}}}\sum_{0\leq x,y\leq\sqrt{N}-1}\!\!\!\!\!\!\!\!{\rm e}^{j2\pi\frac{d}{\lambda}\left(x\delta_{1}+y\delta_{2}\right)}
=\displaystyle\!=\! μEmϵ2π4(ϵ+1)(L12(ϵ))2ejπdλ(N1)(δ1+δ2)\displaystyle\sqrt{\frac{\mu_{E_{m}}\epsilon^{2}}{{\frac{\pi}{4}\!\left(\epsilon\!+\!1\right)\!\left(\!L_{\frac{1}{2}}\!\left(-\epsilon\right)\right)}^{2}}}{\rm e}^{j\pi\frac{d}{\lambda}\left(\sqrt{N}-1\right)\left(\delta_{1}+\delta_{2}\right)}
×sin(πdλNδ1)sin(πdλNδ2)sin(πdλδ1)sin(πdλδ2),\displaystyle\times\frac{\sin{\left(\pi\frac{d}{\lambda}\sqrt{N}\delta_{1}\right)}\sin{\left(\pi\frac{d}{\lambda}\sqrt{N}\delta_{2}\right)}}{\sin{\left(\pi\frac{d}{\lambda}\delta_{1}\right)}\sin{\left(\pi\frac{d}{\lambda}\delta_{2}\right)}}, (60)

and (g)({\rm g}) follows by making use of a mapping from the index nn to the two-dimensional index (x,y)\left(x,y\right) and substituting aN,Ri(n)=ej2πdλ(xsinψRiasinψRie+ycosψRie){\rm a}_{N,Ri}\left(n\right)={\rm e}^{j2\pi\frac{d}{\lambda}\left(x\sin{\psi_{Ri}^{a}}\sin{\psi_{Ri}^{e}}+y\cos{\psi_{Ri}^{e}}\right)} from (5), δ1=sinψRDasinψRDesinψREmasinψREme\delta_{1}=\sin{\psi_{RD}^{a}}\sin{\psi_{RD}^{e}}-\sin{\psi_{RE_{m}}^{a}}\sin{\psi_{RE_{m}}^{e}}, and δ2=cosψRDecosψREme\delta_{2}=\cos{\psi_{RD}^{e}}-\cos{\psi_{RE_{m}}^{e}}.

Appendix E Proof of Proposition 2

When rer_{e}\rightarrow\infty, the CDF of the overall eavesdropper SNR in (23) is given as

FγE(x)=\displaystyle F_{\gamma_{E}}\left(x\right)= exp[t0xt4Γ(t1)]exp[t0xt4Γ(t1,t2xt3)]\displaystyle{\rm exp}\left[-t_{0}x^{-t_{4}}\Gamma\left(t_{1}\right)\right]{\rm exp}\left[t_{0}x^{-t_{4}}\Gamma\left(t_{1},t_{2}x^{t_{3}}\right)\right]
=(h)\displaystyle\mathop{=}^{\left(\rm h\right)} exp[t0xt4Γ(t1)]\displaystyle{\rm exp}\left[-t_{0}x^{-t_{4}}\Gamma\left(t_{1}\right)\right]
×(1+t0xt4𝒪((t2xt3)t11et2xt3)),\displaystyle\times\left(1+t_{0}x^{-t_{4}}\mathcal{O}\left(\left(t_{2}x^{t_{3}}\right)^{t_{1}-1}{\rm e}^{-t_{2}x^{t_{3}}}\right)\right), (61)

where (h)({\rm h}) comes from the asymptotic expansion of the upper incomplete Gamma function when t2=ev(ϖ)Ξμ(ϖ)reα22μ(ϖ)t_{2}={\rm e}^{v\left(\varpi\right)}\Xi^{\mu\left(\varpi\right)}{r_{e}}^{\frac{\alpha_{2}}{2}\mu\left(\varpi\right)}\rightarrow\infty [50, Eq. (6.5.32)]. Therefore, for large rer_{e}, the SOP in (25) is further rewritten as

SOP\displaystyle{\rm SOP} =10+FγE(1φ(1+x)1)fγD(x)dx\displaystyle=1-\int_{0}^{+\infty}{F_{\gamma_{E}}\left(\frac{1}{\varphi}\left(1+x\right)-1\right)f_{\gamma_{D}}\left(x\right)}{\rm{d}}x
(i)10+exp[t0Γ(t1)(xφ)t4]\displaystyle\mathop{\simeq}^{\left(\rm i\right)}1-\int_{0}^{+\infty}{\rm exp}\left[-t_{0}\Gamma\left(t_{1}\right)\left(\frac{x}{\varphi}\right)^{-t_{4}}\right]
×1Γ(k)ex/ρdθ(x/ρdθ)k2xdx\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\times\frac{1}{\Gamma\left(k\right)}\frac{{\rm e}^{-\frac{\sqrt{{x}/{\rho_{d}}}}{\theta}}\left(\frac{\sqrt{{x}/{\rho_{d}}}}{\theta}\right)^{k}}{2x}{\rm{d}}x
=112Γ(k)(ρdθ)kI,\displaystyle=1-\frac{1}{2\Gamma\left(k\right)}\left(\sqrt{\rho_{d}}\theta\right)^{-k}\ I, (62)

where (i)({\rm i}) is obtained by substituting (11) and (61), I=0+xaexp[bxcυx]dxI=\int_{0}^{+\infty}{x^{a}{\rm exp}\left[-bx^{-c}-\upsilon\sqrt{x}\right]}{\rm{d}}x, a=k21a=\frac{k}{2}-1, b=t0Γ(t1)φt4b=t_{0}\Gamma\left(t_{1}\right)\varphi^{t_{4}}, c=t4=2qpc=t_{4}=\frac{2q}{p}, and υ=1ρdθ\upsilon=\frac{1}{\sqrt{\rho_{d}}\theta}.

By applying the Mellin convolution theorem [44], the Mellin transform of II is given by

[I;s]=2p2qυ2s+2a+2Γ(ps2q)Γ(2s+2a+2).\displaystyle\mathcal{M}\left[I;s\right]=\frac{2p}{2q\upsilon^{2s+2a+2}}\Gamma\left(\frac{ps}{2q}\right)\Gamma\left(2s+2a+2\right). (63)

Therefore, we can calculate II using the inverse transform as follows

I=\displaystyle I= pπiυ2a+2uiu+iΓ(ps)Γ(4q(s+a+12q))(υ4qbp)sds\displaystyle\frac{p}{\pi i\upsilon^{2a+2}}\int_{u-i\infty}^{u+i\infty}{\!\Gamma\left(ps\right)\!\Gamma\left(4q\left(s\!+\!\frac{a\!+\!1}{2q}\right)\right)\!\left(\upsilon^{4q}b^{p}\right)^{-s}}{\rm{d}}s
=(j)\displaystyle\mathop{=}^{\left(\rm j\right)} p12q2a+32υ2a+22p+4q24a5πp+4q2112πiuiu+i(υ4qbppp256qq4q)s\displaystyle\frac{p^{\frac{1}{2}}q^{2a+\frac{3}{2}}}{\upsilon^{2a+2}2^{\frac{p+4q}{2}-4a-5}\pi^{\frac{p+4q}{2}-1}}\frac{1}{2\pi i}\int_{u-i\infty}^{u+i\infty}\left(\frac{\upsilon^{4q}b^{p}}{p^{p}{256}^{q}q^{4q}}\right)^{-s}
×n=0p1Γ(s+np)n=04q1Γ(s+n+2a+24q)ds\displaystyle\times\prod_{n=0}^{p-1}\Gamma\left(s+\frac{n}{p}\right)\prod_{n=0}^{4q-1}\Gamma\left(s+\frac{n+2a+2}{4q}\right){\rm{d}}s
=\displaystyle= p12q2a+32υ2a+22p+4q24a5πp+4q21G0,p+4qp+4q,0(υ4qbppp256qq4q|Δ),\displaystyle\frac{p^{\frac{1}{2}}q^{2a+\frac{3}{2}}}{\upsilon^{2a+2}2^{\frac{p+4q}{2}-4a-5}\pi^{\frac{p+4q}{2}-1}}G_{0,p+4q}^{p+4q,0}\left(\frac{\upsilon^{4q}b^{p}}{p^{p}{256}^{q}q^{4q}}\middle|\begin{matrix}-\\ \Delta\\ \end{matrix}\right), (64)

where (j)({\rm j}) follows from Gauss’ multiplication formula [50, Eq. (6.1.20)], and the last equality is derived by applying the definition of Meijer’s GG function. Subsequently, by substituting (64) into (62), the SOP is obtained as shown in (29).

Appendix F Proof of Lemma 4

First, by using Jensen’s inequality, we have

𝔼{log2(1+γD)}\displaystyle\!\!\!\mathbb{E}\!\left\{\log_{2}\!{\left(1\!+\!\gamma_{D}\right)}\right\} log2(1+𝔼{γD})=log2(1+ρdKνJ),\displaystyle\!\leq\!\log_{2}\!{\left(1\!+\!\mathbb{E}\!\left\{\gamma_{D}\right\}\right)}\!=\!\log_{2}\!{\left(1\!+\!\rho_{d}K\nu J\right)}, (65)
C¯s\displaystyle{\bar{C}}_{s} ={log2(1+ρdKNνμD(1+π4(N1)(L12(ϵ))2ϵ+1))\displaystyle=\left\{\log_{2}{\left(1+\rho_{d}KN\nu\mu_{D}\!\left(1+\frac{\pi}{4}\left(N-1\right)\frac{\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)^{2}}{\epsilon+1}\!\right)\right)}\right.
1ln2{γ+lnt0Γ(t1)+exp(t0Γ(t1))×(Shi(t0Γ(t1))Chi(t0Γ(t1)))}}+\displaystyle\left.-\frac{1}{\ln{2}}\!\left\{\gamma+\ln{t_{0}\Gamma\left(t_{1}\right)}+{\rm exp}\left(t_{0}\Gamma\left(t_{1}\right)\right)\!\times\!\left({\rm Shi}\left(t_{0}\Gamma\left(t_{1}\right)\right)-{\rm Chi}\left(t_{0}\Gamma\left(t_{1}\right)\right)\right)\right\}\right\}^{+}
(k){log2ρdKNνμD[1+(N1)1ϵ+1π4(L12(ϵ))2]1ln2{γ+lnt0Γ(t1)}}+\displaystyle\mathop{\to}\limits^{\left(\rm k\right)}\left\{\log_{2}{\rho_{d}KN\nu\mu_{D}\left[1+\left(N-1\right)\frac{1}{\epsilon+1}\frac{\pi}{4}\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)^{2}\right]}-\frac{1}{\ln{2}}\left\{\gamma+\ln{t_{0}\Gamma\left(t_{1}\right)}\right\}\right\}^{+}
={log2ρdρe+log2μDπλeβ0+log21+(N1)1ϵ+1π4(L12(ϵ))21ϵ2π4(ϵ+1)(L12(ϵ))2γln2+C}+,\displaystyle=\left\{\log_{2}\frac{\rho_{d}}{\rho_{e}}+\log_{2}{\frac{\mu_{D}}{\pi\lambda_{e}\beta_{0}}}+\log_{2}\frac{1+\left(N-1\right)\frac{1}{\epsilon+1}\frac{\pi}{4}\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)^{2}}{1-\frac{\epsilon^{2}}{{\frac{\pi}{4}\left(\epsilon+1\right)\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}}-\frac{\gamma}{\ln{2}}+C\right\}^{+}, (65)

where JJ is expressed as

J\displaystyle J =𝔼{(n=1N|hRD(n)|)2}\displaystyle=\mathbb{E}\!\left\{\left(\sum_{n=1}^{N}\left|h_{R\!D}\left(n\right)\right|\right)^{2}\right\}
=𝔼{n=1N|hRD(n)|2+21i<jN|hRD(i)||hRD(j)|}\displaystyle=\mathbb{E}\!\left\{\sum_{n=1}^{N}\left|h_{R\!D}\left(n\right)\right|^{2}+2\sum_{1\leq i<j\leq N}\left|h_{R\!D}\left(i\right)\right|\left|h_{R\!D}\left(j\right)\right|\right\}
=N𝔼{|hRD(n)|2}+N(N1)(𝔼{|hRD(n)|})2\displaystyle=N\cdot\mathbb{E}\!\left\{\left|h_{R\!D}\left(n\right)\right|^{2}\right\}\!+N\left(N-1\right)\cdot\left(\mathbb{E}\!\left\{\left|h_{R\!D}\left(n\right)\right|\right\}\right)^{2}
=NμD[1+(N1)1ϵ+1π4(L1/2(ϵ))2].\displaystyle=N\mu_{D}\left[1+\left(N-1\right)\frac{1}{\epsilon+1}\frac{\pi}{4}\left(L_{1/2}\left(-\epsilon\right)\right)^{2}\right]. (64)

Then, substituting (64) into (65), the upper bound for RDR_{D} is obtained as shown in (48).

Appendix G Proof of Corollary 7

By substituting (48) into (44), the approximate ESC can be represented as (65), where t0=2πλeμ(ϖ)e2v(ϖ)μ(ϖ)Ξ2=1μ(ϖ)e2v(ϖ)μ(ϖ)πλeNKνβ0ρe[1ϵ2π4(ϵ+1)(L12(ϵ))2]t_{0}=\frac{2\pi\lambda_{e}}{\mu\left(\varpi\right){\rm e}^{\frac{2v\left(\varpi\right)}{\mu\left(\varpi\right)}}\Xi^{2}}=\frac{1}{\mu\left(\varpi\right){\rm e}^{\frac{2v\left(\varpi\right)}{\mu\left(\varpi\right)}}}\pi\lambda_{e}NK\nu\beta_{0}\rho_{e}\left[1-\frac{\epsilon^{2}}{{\frac{\pi}{4}\left(\epsilon+1\right)\left(L_{\frac{1}{2}}\left(-\epsilon\right)\right)}^{2}}\right], and (k) comes from the fact that when ρd,ρe\rho_{d},\rho_{e}\rightarrow\infty, the formula limt0Γ(t1){Shi(t0Γ(t1))Chi(t0Γ(t1))}=0\lim_{t_{0}\Gamma\left(t_{1}\right)\to\infty}{\left\{{\rm Shi}\left(t_{0}\Gamma\left(t_{1}\right)\right)-{\rm Chi}\left(t_{0}\Gamma\left(t_{1}\right)\right)\right\}}=0 holds. We complete the proof by substituting the expressions of ρd\rho_{d}, ρe\rho_{e}, and μD\mu_{D} into (65).

References

  • [1] W. Shi et al., “Secure outage analysis for RIS-aided MISO systems with randomly located eavesdroppers,” in Proc. IEEE Globecom Workshops (GC Wkshps), Kuala Lumpur, Malaysia, 2023, pp. 1445–1450. doi: 10.1109/GCWkshps58843.2023.10465162.
  • [2] M. Di Renzo et al., “Smart radio environments empowered by reconfigurable intelligent surfaces: How it works, state of research, and the road ahead,” IEEE J. Sel. Areas Commun., vol. 38, no. 11, pp. 2450–2525, Nov. 2020.
  • [3] X. Lei, M. Wu, F. Zhou, X. Tang, R. Q. Hu, and P. Fan, “Reconfigurable intelligent surface-based symbiotic radio for 6G: Design, challenges, and opportunities,” IEEE Wireless Commun., vol. 28, no. 5, pp. 210–216, Oct. 2021.
  • [4] F. Naeem, M. Ali, G. Kaddoum, C. Huang, and C. Yuen, “Security and privacy for reconfigurable intelligent surface in 6G: A review of prospective applications and challenges”, IEEE Open J. Commun. Soc., vol. 4, pp. 1196–1217, May 2023.
  • [5] W. Shi, W. Xu, X. You, C. Zhao, and K. Wei, “Intelligent reflection enabling technologies for integrated and green Internet-of-Everything beyond 5G: Communication, sensing, and security,” IEEE Wireless Commun., vol. 30, no. 2, pp. 147–154, Apr. 2023.
  • [6] S. Li, B. Duo, X. Yuan, Y.-C. Liang, and M. Di Renzo, “Reconfigurable intelligent surface assisted UAV communication: Joint trajectory design and passive beamforming,” IEEE Wireless Commun. Lett., vol. 9, no. 5, pp. 716–720, May 2020.
  • [7] H. Zhang, B. Di, L. Song, and Z. Han, “Reconfigurable intelligent surfaces assisted communications with limited phase shifts: How many phase shifts are enough?,” IEEE Trans. Veh. Technol., vol. 69, no. 4, pp. 4498–4502, Apr. 2020.
  • [8] S. Lin, B. Zheng, G. C. Alexandropoulos, M. Wen, F. Chen, and S. Mumtaz, “Adaptive transmission for reconfigurable intelligent surface-assisted OFDM wireless communications,” IEEE J. Sel. Areas Commun., vol. 38, no. 11, pp. 2653–2665, Nov. 2020.
  • [9] C. Huang, A. Zappone, G. C. Alexandropoulos, M. Debbah, and C. Yuen, “Reconfigurable intelligent surfaces for energy efficiency in wireless communication,” IEEE Trans. Wireless Commun., vol. 18, no. 8, pp. 4157–4170, Aug. 2019.
  • [10] S. Liu, Z. Gao, J. Zhang, M. Di Renzo, and M.-S. Alouini, “Deep denoising neural network assisted compressive channel estimation for mmWave intelligent reflecting surfaces,” IEEE Trans. Veh. Technol., vol. 69, no. 8, pp. 9223–9228, Aug. 2020.
  • [11] Z. Wan, Z. Gao, F. Gao, M. Di Renzo, and M.-S. Alouini, “Terahertz massive MIMO with holographic reconfigurable intelligent surfaces,” IEEE Trans. Commun., vol. 69, no. 7, pp. 4732–4750, Jul. 2021.
  • [12] J. Yao, J. Xu, W. Xu, D. W. K. Ng, C. Yuen, and X. You, “Robust beamforming design for RIS-aided cell-free systems with CSI uncertainties and capacity-limited backhaul,” IEEE Trans. Commun., vol. 71, no. 8, pp. 4636–4649, Aug. 2023.
  • [13] J. Yao, W. Xu, X. You, D. W. K. Ng, and J. Fu, “Robust beamforming design for reconfigurable intelligent surface-aided cell-free systems,” in Proc. IEEE Int. Symp. Wireless Commun. Syst. (ISWCS), Hangzhou, China, 2022, pp. 1–6.
  • [14] W. Ni, Y. Liu, Y. C. Eldar, Z. Yang, and H. Tian, “STAR-RIS integrated nonorthogonal multiple access and over-the-air federated learning: Framework, analysis, and optimization,” IEEE Internet Things J., vol. 9, no. 18, pp. 17136–17156, Sept. 2022.
  • [15] W. Xu et al., “Edge learning for B5G networks with distributed signal processing: Semantic communication, edge computing, and wireless sensing,” IEEE J. Sel. Topics Signal Process., vol. 17, no. 1, pp. 9–39, Jan. 2023.
  • [16] J. Xu et al., “Reconfiguring wireless environment via intelligent surfaces for 6G: Reflection, modulation, and security,” Sci China Inf Sci, vol. 66, no. 3, pp. 130304:1–-20, Mar. 2023.
  • [17] H. Shen, W. Xu, S. Gong, Z. He, and C. Zhao, “Secrecy rate maximization for intelligent reflecting surface assisted multi-antenna communications,” IEEE Commun. Lett., vol. 23, no. 9, pp. 1488–1492, Sept. 2019.
  • [18] G. Zhou, C. Pan, H. Ren, K. Wang, and Z. Peng, “Secure wireless communication in RIS-aided MISO system with hardware impairments,” IEEE Wireless Commun. Lett., vol. 10, no. 6, pp. 1309–1313, Jun. 2021.
  • [19] X. Yu, D. Xu, Y. Sun, D. W. K. Ng, and R. Schober, “Robust and secure wireless communications via intelligent reflecting surfaces,” IEEE J. Sel. Areas Commun., vol. 38, no. 11, pp. 2637–2652, Nov. 2020.
  • [20] L. Yang, J. Yang, W. Xie, M. O. Hasna, T. Tsiftsis, and M. Di Renzo, “Secrecy performance analysis of RIS-aided wireless communication systems,” IEEE Trans. Veh. Technol., vol. 69, no. 10, pp. 12296–12300, Oct. 2020.
  • [21] W. Shi, J. Xu, W. Xu, M. Di Renzo, and C. Zhao, “Secure outage analysis of RIS-assisted communications with discrete phase control,” IEEE Trans. Veh. Technol., vol. 72, no. 4, pp. 5435–5440, Apr. 2023.
  • [22] P. Xu, G. Chen, G. Pan, and M. Di Renzo, “Ergodic secrecy rate of RIS-assisted communication systems in the presence of discrete phase shifts and multiple eavesdroppers,” IEEE Wireless Commun. Lett., vol. 10, no. 3, pp. 629–633, Mar. 2021.
  • [23] M. H. Khoshafa, T. M. N. Ngatched, and M. H. Ahmed, “Reconfigurable intelligent surfaces-aided physical layer security enhancement in D2D underlay communications,” IEEE Commun. Lett., vol. 25, no. 5, pp. 1443–1447, May 2021.
  • [24] Y. Ai, F. A. P. deFigueiredo, L. Kong, M. Cheffena, S. Chatzinotas, and B. Ottersten, “Secure vehicular communications through reconfigurable intelligent surfaces,” IEEE Trans. Veh. Technol., vol. 70, no. 7, pp. 7272–7276, Jul. 2021.
  • [25] D. Wang, M. Wu, Z. Wei, K. Yu, L. Min, and S. Mumtaz, “Uplink secrecy performance of RIS-based RF/FSO three-dimension heterogeneous networks,” IEEE Trans. Wireless Commun., early access. doi: 10.1109/TWC.2023.3292073.
  • [26] J. Yuan, G. Chen, M. Wen, R. Tafazolli, and E. Panayirci, “Secure transmission for THz-empowered RIS-assisted non-terrestrial networks,” IEEE Trans. Veh. Technol., vol. 72, no. 5, pp. 5989–6000, May 2023.
  • [27] S. N. Chiu, D. Stoyan, W. S. Kendall, and J. Mecke, Stochastic Geometry and Its Applications. Hoboken, NJ, USA: Wiley, 2013.
  • [28] J. G. Andrews, F. Baccelli, and R. K. Ganti, “A tractable approach to coverage and rate in cellular networks,” IEEE Trans. Commun., vol. 59, no. 11, pp. 3122–3134, Nov. 2011.
  • [29] L. Wei, K. Wang, C. Pan, and M. Elkashlan, “Secrecy performance analysis of RIS-aided communication system with randomly flying eavesdroppers,” IEEE Wireless Commun. Lett., vol. 11, no. 10, pp. 2240–2244, Oct. 2022.
  • [30] W. Wang, H. Tian, and W. Ni, “Secrecy performance analysis of IRS-aided UAV relay system,” IEEE Wireless Commun. Lett., vol. 10, no. 12, pp. 2693–2697, Dec. 2021.
  • [31] J. Yao, J. Xu, W. Xu, C. Yuen, and X. You, “A universal framework of superimposed RIS-phase modulation for MISO communication,” IEEE Trans. Veh. Technol., vol. 72, no. 4, pp. 5413–5418, Apr. 2023.
  • [32] J. Yao, J. Xu, W. Xu, C. Yuen, and X. You, “Superimposed RIS-phase modulation for MIMO communications: A novel paradigm of information transfer,” IEEE Trans. Wireless Commun., early access. doi: 10.1109/TWC.2023.3304695.
  • [33] A. Papazafeiropoulos, C. Pan, P. Kourtessis, S. Chatzinotas, and J. M. Senior, “Intelligent reflecting surface-assisted MU-MISO systems with imperfect hardware: Channel estimation and beamforming design,” IEEE Trans. Wireless Commun., vol. 21, no. 3, pp. 2077–2092, Mar. 2022.
  • [34] S. Fang, G. Chen, Z. Abdullah, and Y. Li, “Intelligent omni surface-assisted secure MIMO communication networks with artificial noise,” IEEE Commun. Lett., vol. 26, no. 6, pp. 1231–1235, Jun. 2022.
  • [35] J. Bai, H.-M. Wang, and P. Liu, “Robust IRS-aided secrecy transmission with location optimization,” IEEE Trans. Commun., vol. 70, no. 9, pp. 6149–6163, Sept. 2022.
  • [36] S. Zhou, W. Xu, K. Wang, M. Di Renzo, and M.-S. Alouini, “Spectral and energy efficiency of IRS-assisted MISO communication with hardware impairments,” IEEE Wireless Commun. Lett., vol. 9, no. 9, pp. 1366–1369, Sept. 2020.
  • [37] Z.-Q. He and X. Yuan, “Cascaded channel estimation for large intelligent metasurface assisted massive MIMO,” IEEE Wireless Commun. Lett., vol. 9, no. 2, pp. 210–214, Feb. 2020.
  • [38] L. Wei, C. Huang, G. C. Alexandropoulos, C. Yuen, Z. Zhang, and M. Debbah, “Channel estimation for RIS-empowered multi-user MISO wireless communications,” IEEE Trans. Commun., vol. 69, no. 6, pp. 4144–4157, Jun. 2021.
  • [39] Z. Wang, L. Liu, and S. Cui, “Channel estimation for intelligent reflecting surface assisted multiuser communications: Framework, algorithms, and analysis,” IEEE Trans. Wireless Commun., vol. 19, no. 10, pp. 6607–6620, Oct. 2020.
  • [40] S. Primak, V. Kontorovich, and V. Lyandres, Stochastic Methods and their Applications to Communications: Stochastic Differential Equations Approach. West Sussex, U.K.: Wiley, 2004.
  • [41] R. W. Heath, M. Kountouris, and T. Bai, “Modeling heterogeneous network interference using Poisson point processes,” IEEE Trans. Signal Process., vol. 61, no. 16, pp. 4114–4126, Aug. 2013.
  • [42] M. Z. Bocus, C. P. Dettmann, and J. P. Coon, “An approximation of the first order Marcum QQ-function with application to network connectivity analysis,” IEEE Commun. Lett., vol. 17, no. 3, pp. 499–502, Mar. 2013.
  • [43] S. N. Chiu, D. Stoyan, W. S. Kendall, and J. Mecke, Stochastic Geometry and Its Applications. Hoboken, NJ, USA: Wiley, 2013.
  • [44] G. Chen, J. P. Coon, and M. Di Renzo, “Secrecy outage analysis for downlink transmissions in the presence of randomly located eavesdroppers,” IEEE Trans. Inf. Forensics Security, vol. 12, no. 5, pp. 1195–1206, May 2017.
  • [45] H. Lei et al., “Safeguarding UAV IoT communication systems against randomly located eavesdroppers,” IEEE Internet Things J., vol. 7, no. 2, pp. 1230–1244, Feb. 2020.
  • [46] I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products, 7th ed. San Diego, CA, USA: Academic, 2007.
  • [47] F. Baccelli and B. Blaszczyszyn, Stochastic Geometry and Wireless Networks, Volume II: Applications, Hanover, MA, USA: Now, 2009.
  • [48] 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.
  • [49] Wolfram Research, “The Wolfram functions site: Meijer G-function,” 2001. [Online]. Available: https://functions.wolfram.com/PDF/MeijerG.pdf
  • [50] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions With Formulas, Graphs, and Mathematical Tables. New York, NY, USA: Dover, 1972.
  • [51] Y. Guo, R. Zhao, S. Lai, L. Fan, X. Lei, and G. K. Karagiannidis, “Distributed machine learning for multiuser mobile edge computing systems,” IEEE J. Sel. Topics Signal Process., vol. 16, no. 3, pp. 460–473, Apr. 2022.
  • [52] J. Xu, W. Xu, D. W. K. Ng, and A. L. Swindlehurst, “Secure communication for spatially sparse millimeter-wave massive MIMO channels via hybrid precoding,” IEEE Trans. Commun., vol. 68, no. 2, pp. 887–901, Feb. 2020.