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

\UseRawInputEncoding

Secure Outage Analysis of RIS-Assisted Communications with Discrete Phase Control

Wei Shi, Jindan Xu, Wei Xu, , Marco Di Renzo, , and Chunming Zhao 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 [email protected] Shi, Wei Xu, Jindan Xu, and Chunming Zhao are with the National Mobile Communications Research Laboratory, Southeast University, Nanjing 210096, China (e-mail:{wshi, wxu, jdxu, cmzhao}@seu.edu.cn).Marco Di Renzo is with Université Paris-Saclay, CNRS, CentraleSupélec, Laboratoire des Signaux et Systèmes, 3 Rue Joliot-Curie, 91192 Gif-sur-Yvette, France (e-mail: [email protected]).
Abstract

This correspondence investigates a reconfigurable intelligent surface (RIS)-assisted wireless communication system with security threats. The RIS is deployed to enhance the secrecy outage probability (SOP) of the data sent to a legitimate user. By deriving the distributions of the received signal-to-noise-ratios (SNRs) at the legitimate user and the eavesdropper, we formulate, in a closed-form expression, a tight bound for the SOP under the constraint of discrete phase control at the RIS. The SOP is characterized as a function of the number of antenna elements, NN, and the number of discrete phase choices, 2b2^{b}. It is revealed that the performance loss in terms of SOP due to the discrete phase control is ignorable for large NN when b3b\!\geq\!3. In addition, we explicitly quantify this SOP loss when binary phase shifts with b=1b\!=\!1 is utilized. It is identified that increasing the RIS antenna elements by 1.61.6 times can achieve the same SOP with binary phase shifts as that by the RIS with ideally continuous phase shifts. Numerical simulations are conducted to verify the accuracy of these theoretical observations.

Index Terms:
Reconfigurable intelligent surface (RIS), physical layer security, secrecy outage probability, discrete phase shifts.

I Introduction

Reconfigurable intelligent surface (RIS) is a metasurface that consists of a large number of passive reflecting elements with integrated low power electronics [1][2]. A main feature of an RIS is that the amplitude and phase of each reflecting element can be independently controlled through software, thereby realizing passive beamforming (BF) for improving the signal quality at intended receivers. Due to these merits, RISs have been considered for various wireless applications, e.g., in millimeter-wave (mmWave) [3] and Terahertz (THz) [4] communications, to enhance spectral and energy efficiencies [5].

In recent years, physical layer security (PLS) has gained considerable interest for securing wireless communications. As a complement to conventional cryptographic methods, PLS ensures secure communications by exploiting the dynamics of propagation channels. Since RISs have the ability of adjusting the propagation channels, their deployment empowers the design of PLS with an additional dimension by exploiting passive BF. In order to maximize the theoretical secrecy, literature [6, 7, 8] investigated joint optimization of the active and passive BFs at the transmitter and RIS, respectively. In [9], the average secrecy rate (SR) was characterized for an RIS-assisted two-way communication system through a lower bound. More recently in [10], the SR was further analyzed for a system where the RIS reflection is utilized as a multiplicative randomness against a wiretapper.

Besides the analysis of SR, secrecy outage probability (SOP) is another relevant performance measure to quantify the performance of PLS, especially for systems undergoing slow-varying channels. The SOP is defined as the probability that the instantaneous secrecy capacity falls below a target SR. The SOPs of RIS-aided communication systems have been investigated in [11, 12, 13]. Both analytical and asymptotic analyses have been provided to reveal the impacts of key system parameters on SOP. In particular, the work in [12] considered the SOP of an RIS-aided unmanned aerial vehicle (UAV) relay system. In [13], the authors analyzed the SOP as well as the probability of non-zero secrecy capacity of an RIS-aided device-to-device (D2D) communication system. However, most studies on SOP analysis considered RIS with continuous phase shifts, which leads to unaffordable high complexity in practice. Even though discrete phase shifts have been considered for RIS reflection optimizations, e.g., in [14][15], few studies have been conducted on theoretical performance analysis with discrete RIS phase shifts especially in terms of SOP. Quantitative insight on RIS design has only been discovered for some non-security scenarios [16]. This is because discrete phase shifts make the performance expressions of the cascaded RIS channels much less tractable. In particular for secure communications, it can be even challenging to directly derive the corresponding cascaded channel distributions of both the legitimate user and eavesdropper.

In this work, we investigate the SOP of an RIS-assisted secure communication system and quantitatively characterize the impacts of discrete phase shifts in closed-form expressions. Concretely, we first derive the exact distributions of the received signal-to-noise-ratios (SNRs) at the legitimate user and the eavesdropper. Then, we present a closed-form expression for a tight upper bound of the SOP. Based on the obtained expressions, the asymptotic scaling law of SOP is characterized in high-SNR regimes. In particular, the SOP decreases with the slope of e0.8N{\rm e}^{-0.8N} for large NN and b3b\!\geq\!3, where NN is the number of RIS elements and bb is the number of quantization bits. Compared with the ideally continuous phase control, we further obtain that the performance loss in terms of SOP caused by using binary phase shifts, i.e., b=1b\!=\!1, can be compensated by deploying a larger RIS with size 1.6N1.6N.

The remainder of this paper is organized as follows. The system model is introduced in Section II. In Section III, we derive the distribution of the received SNRs at the legitimate user and at the eavesdropper. In Section IV, we provide a closed-form expression for an upper bound of the SOP, and we study it in notable asymptotic regimes. Simulation results and conclusions are given in Section V and VI, respectively.

II System Model

We consider an RIS-assisted secure communication system consisting of a source (SS), an RIS with NN reflecting elements, a legitimate user (DD), and an eavesdropper (EE), as illustrated in Fig. 1. The direct link between SS and DD is assumed to be blocked by obstacles, such as buildings, which is likely to occur at high frequency bands. In this scenario, the data transmission between SS and DD is ensured by the RIS. The eavesdropper is at a location where it can overhear the information from both SS and the RIS.111The direct link between SS and EE exists when the eavesdropper is not blocked by obstacles [11][13]. Nodes SS, DD, and EE are equipped with a single antenna for transmission and reception and all links experience Rayleigh fading.222The RIS-related links can be modeled as Rayleigh fading, also as [11][13], when the RIS is not optimally deployed to ensure strong LoS links. The channel coefficients of the SS-RIS, RIS-DD, RIS-EE, and SS-EE links are respectively denoted by hih_{i}, gig_{i}, pip_{i}, and hSE𝒞𝒩(0,1)h_{S\!E}\sim\mathcal{CN}(0,1), where 𝒞𝒩{\cal{CN}} is the complex Gaussian distribution. By applying channel estimation methods in [3, 4] and the references therein, we assume that the channel coefficients of hih_{i} and gig_{i} are perfectly known to SS. However, the channel gains of pip_{i} and hSEh_{S\!E} are not available to SS, as the eavesdropper is usually a passive device that does not emit signals.

Refer to caption
Figure 1: The system model of an RIS-assisted secure communication.

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

rD=P[η(dSRdRD)υ/2i=1Nhigiejϕi]x+nD,r_{D}=\sqrt{P}\left[{\eta(d_{S\!R}d_{R\!D})}^{-\upsilon/2}\sum_{i=1}^{N}{h_{i}g_{i}{\rm e}^{j\phi_{i}}}\right]x+n_{D}, (1)

where PP denotes the transmit power at SS, xx is the transmit signal, η(0,1]\eta\in(0,1] is the RIS amplitude reflection coefficient with η=1\eta=1 corresponding to lossless reflection, {ϕi}i=1N\left\{\phi_{i}\right\}_{i=1}^{N} represents the phase shift of the iith reflecting element of the RIS, and nD𝒞𝒩(0,N0)n_{D}\sim\mathcal{CN}(0,N_{0}) is the additive white Gaussian noise (AWGN) with zero mean and variance N0N_{0}. Without loss of generality, the signal power is normalized, i.e., 𝔼[|x|2]=1\mathbb{E}[\left|x\right|^{2}]=1 where 𝔼[]\mathbb{E}\left[\cdot\right] denotes the expectation of a random variable (RV). In addition, dSRd_{S\!R} and dRDd_{R\!D} are the distances of the SS-RIS and RIS-DD links, respectively, and υ\upsilon is the path loss exponent.

From (1), the received SNR at DD is calculated as

γD=η2P|i=1Nhigiejϕi|2N0dSRυdRDυ.\gamma_{D}=\frac{{\eta}^{2}P\left|\sum_{i=1}^{N}{h_{i}g_{i}{\rm e}^{j\phi_{i}}}\right|^{2}}{N_{0}d_{S\!R}^{\upsilon}d_{R\!D}^{\upsilon}}. (2)

In the case of ideally continuous phase shifts, γD\gamma_{D} is maximized by setting the phases of the RIS elements equal to ϕi=ϕiopt(higi)\phi_{i}\!=\!\phi_{i}^{\rm opt}\!\triangleq\!-\angle\left({h_{i}}{g_{i}}\right), where \angle returns the phase of the complex number. This optimized phase compensates the phase shift introduced by fading channels. Due to hardware limitations, however, the phase shifts {ϕi}i=1N\left\{\phi_{i}\right\}_{i=1}^{N} of the RIS elements are usually limited to a finite number of controllable discrete values. In particular, the set of discrete phase shifts is denoted by 𝒜{0,2π2b,,(2b1)2π2b}\mathcal{A}\triangleq\left\{0,\frac{2\pi}{2^{b}},...,\frac{(2^{b}-1)2\pi}{2^{b}}\right\}, where bb is the number of quantization bits. In this case, usually, the phase shift of the iith RIS element, ϕisub𝒜\phi_{i}^{\rm sub}\in\mathcal{A}, is chosen as

ϕisubargminϕ𝒜{|ϕioptϕ|}.\phi_{i}^{\rm sub}\triangleq\arg\mathop{\min}\limits_{{\phi\in\mathcal{A}}}\left\{\left|\phi_{i}^{\rm opt}-\phi\right|\right\}. (3)

Then, the received SNR in the presence of the discrete phase shifts is rewritten as

γD=γ¯SRD|i=1Nhigiejϕisub|2,\displaystyle\gamma_{D}\!=\!{\bar{\gamma}_{S\!R\!D}}\left|\sum_{i=1}^{N}{h_{i}g_{i}{\rm e}^{j\phi_{i}^{\rm sub}}}\right|^{2}, (4)

where γ¯SRDη2PN0dSRυdRDυ{\bar{\gamma}}_{S\!R\!D}\triangleq\frac{{\eta}^{2}P}{N_{0}d_{S\!R}^{\upsilon}d_{R\!D}^{\upsilon}} denotes the average SNR.

The eavesdropper receives signals from the direct link from SS and the reflected link from the RIS. Then, the received signal at EE is written as

rE=P[η(dSRdRE)υ/2i=1Nhipiejϕisub+dSEυ/2hSE]x+nE,r_{E}\!=\!\sqrt{P}\!\left[\eta{(d_{S\!R}d_{R\!E})}^{-\upsilon/2}\sum_{i=1}^{N}{h_{i}p_{i}{\rm e}^{j\phi_{i}^{\rm sub}}}\!+\!d_{S\!E}^{-\upsilon/2}h_{S\!E}\right]\!x+n_{E}, (5)

where dREd_{R\!E} and dSEd_{S\!E} denote the distances of the RIS-EE and SS-EE links, respectively, and nE𝒞𝒩(0,N0)n_{E}\sim\mathcal{CN}(0,N_{0}) is the AWGN at EE. Then, the received SNR at EE is

γE=|γ¯SREi=1Nhipiejϕisub+γ¯SEhSE|2,\displaystyle\gamma_{E}\!=\!\left|\sqrt{{\bar{\gamma}}_{S\!R\!E}}\sum_{i=1}^{N}{h_{i}p_{i}{\rm e}^{j\phi_{i}^{\rm sub}}}+\sqrt{{\bar{\gamma}}_{S\!E}}h_{S\!E}\right|^{2}, (6)

where γ¯SREη2PN0dSRυdREυ{\bar{\gamma}}_{S\!R\!E}\triangleq\frac{{\eta}^{2}P}{N_{0}d_{S\!R}^{\upsilon}d_{R\!E}^{\upsilon}} and γ¯SEη2PN0dSEυ{\bar{\gamma}}_{S\!E}\triangleq\frac{{\eta}^{2}P}{N_{0}d_{S\!E}^{\upsilon}} represent the average SNRs of the SS-RIS-EE and SS-EE links, respectively.

III Distributions of the Received SNRs

In order to analyze the SOP of the system, we need to first characterize the distributions of γD\gamma_{D} and γE\gamma_{E}.

III-A Distribution of γD\gamma_{D}

Let us denote the quantization error of phase shifts by Θiϕisubϕiopt\Theta_{i}\!\triangleq\!\phi_{i}^{\rm sub}\!-\!\phi_{i}^{\rm opt}, which is uniformly distributed [14][15][17], i.e., Θi𝒰(2bπ,2bπ)\Theta_{i}\!\sim\!\mathcal{U}\left(-2^{-b}\pi,2^{-b}\pi\right). Then, γD\gamma_{D} in (4) is rewritten as

γD\displaystyle\gamma_{D} =(a)γ¯SRD|i=1Nhigiej(Θi+ϕiopt)|2=(b)γ¯SRD|i=1N|hi||gi|ejΘi|2\displaystyle\!\overset{({\rm a})}{=}\!{\bar{\gamma}_{S\!R\!D}}\left|\sum_{i=1}^{N}{h_{i}g_{i}{\rm e}^{j(\Theta_{i}+\phi_{i}^{\rm opt})}}\right|^{2}\!\overset{({\rm b})}{=}\!{\bar{\gamma}_{S\!R\!D}}\left|\sum_{i=1}^{N}{\left|h_{i}\right|\left|g_{i}\right|{\rm e}^{j\Theta_{i}}}\right|^{2}
=(c)γ¯SRD(X2+Y2)=(d)γD1+γD2,\displaystyle\!\overset{({\rm c})}{=}\!\bar{\gamma}_{S\!R\!D}\left(X^{2}\!+\!Y^{2}\right)\!\overset{({\rm d})}{=}\!\gamma_{D_{1}}\!+\!\gamma_{D_{2}}, (7)

where (a)({\rm a}) follows by the identity ϕisub=Θi+ϕiopt\phi_{i}^{\rm sub}\!=\!\Theta_{i}+\phi_{i}^{\rm opt}, (b)({\rm b}) is obtained by using ϕiopt=(higi)\phi_{i}^{\rm opt}\!=\!-\angle\left({h_{i}}{g_{i}}\right), (c)({\rm c}) follows by the definitions Xi=1N|hi||gi|cosΘiX\!\triangleq\!\sum_{i=1}^{N}{\left|h_{i}\right|\left|g_{i}\right|\cos{\Theta_{i}}} and Yi=1N|hi||gi|sinΘiY\!\triangleq\!\sum_{i=1}^{N}{\left|h_{i}\right|\left|g_{i}\right|\sin{\Theta_{i}}}, and (d)({\rm d}) is obtained by defining γD1γ¯SRDX2\gamma_{D_{1}}\!\triangleq\!{\bar{\gamma}}_{S\!R\!D}X^{2} and γD2γ¯SRDY2\gamma_{D_{2}}\!\triangleq\!{\bar{\gamma}}_{S\!R\!D}Y^{2}. Before deriving the distribution of γD\gamma_{D}, we introduce the following lemma.

Lemma 1: If NN is large, γD1\gamma_{D_{1}} and γD2\gamma_{D_{2}} are statistically independent. Also, the cumulative distribution function (CDF) of γD1\gamma_{D_{1}} and the probability density function (PDF) of γD2\gamma_{D_{2}} are, respectively,

FγD1(x)=112erfc(α+xβ)12erfc(xαβ),F_{\gamma_{D_{1}}}\left(x\right)=1-\frac{1}{2}{\rm erfc}\left(\frac{\alpha\!+\!\sqrt{x}}{\beta}\right)\!-\!\frac{1}{2}{\rm erfc}\left(\frac{\sqrt{x}\!-\!\alpha}{\beta}\right), (8)
fγD2(y)=λμyμ1Γ(μ)eλy,f_{\gamma_{D_{2}}}\left(y\right)=\frac{\lambda^{\mu}y^{\mu-1}}{\Gamma\left(\mu\right)}{\rm e}^{-\lambda y}, (9)

where α=m1γ¯SRD\alpha\!=\!m_{1}\sqrt{{\bar{\gamma}}_{S\!R\!D}}, β=2σ1γ¯SRD\beta\!=\!\sqrt{2}\sigma_{1}\sqrt{{\bar{\gamma}}_{S\!R\!D}}, m1=Nπ4sinc(2b)m_{1}\!=\!\frac{N\pi}{4}{\rm sinc}\left(2^{-b}\right), σ12=N2[1+sinc(21b)]Nπ216sinc2(2b)\sigma_{1}^{2}\!=\!\frac{N}{2}[1+{\rm sinc}(2^{1-b})]-\frac{N\pi^{2}}{16}{\rm sinc}^{2}\left(2^{-b}\right), λ=12σ22γ¯SRD\lambda\!=\!\frac{1}{2\sigma_{2}^{2}{\bar{\gamma}}_{S\!R\!D}}, μ=12\mu\!=\!\frac{1}{2}, σ22=N2[1sinc(21b)]\sigma_{2}^{2}\!=\!\frac{N}{2}[1-{\rm sinc}(2^{1-b})], where sinc(x)sinπxπx{\rm sinc}(x)\!\triangleq\!\frac{\sin{\pi x}}{\pi x} and Γ()\Gamma\left(\cdot\right) is the Gamma function [18, Eq. (8.310)].

Proof: The proof of the independence of γD1\gamma_{D_{1}} and γD2\gamma_{D_{2}} for large values of NN is provided in Appendix A. Specifically, by applying the central limit theorem (CLT) [19], XX and YY converge in distribution to Gaussian RVs for large NN. Since |hi|\left|h_{i}\right| and |gi|\left|g_{i}\right| are independently distributed Rayleigh RVs with mean π/2\sqrt{\pi}/2 and variance (4π)/4(4-\pi)/4, we obtain 𝔼[X]=m1\mathbb{E}\left[X\right]=m_{1}, Var[X]=σ12{\rm Var}[X]=\sigma_{1}^{2}, 𝔼[Y]=0\mathbb{E}\left[Y\right]=0, and Var[Y]=σ22{\rm Var}[Y]=\sigma_{2}^{2}. It follows

Xd𝒩(m1,σ12),Yd𝒩(0,σ22),X\xrightarrow{\rm{d}}\mathcal{N}(m_{1},\sigma_{1}^{2}),~{}Y\xrightarrow{\rm{d}}\mathcal{N}(0,\sigma_{2}^{2}), (10)

where d\xrightarrow{\rm{d}} denotes the convergence in distribution by virtue of the CLT. γD1\gamma_{D_{1}} is a non-central χ2\chi^{2} RV and γD2\gamma_{D_{2}} is a central χ2\chi^{2} RV with one degree of freedom, where χ2\chi^{2} denotes the Chi-square distribution. Then, by using [20, Eq. (2.3-35)] and [21, Eq. (27)], FγD1()F_{\gamma_{D_{1}}}\left(\cdot\right) is derived. The PDF fγD2()f_{\gamma_{D_{2}}}\left(\cdot\right) is obtained from [20, Eq. (2.3-28)]. The proof completes.\hfill\blacksquare

By applying Lemma 1 and (7), the CDF of γD\gamma_{D} equals

FγD(z)=\displaystyle F_{\gamma_{D}}\left(z\right)= DfγD1,γD2(x,y)dxdy\displaystyle\iint_{D}{f_{\gamma_{D_{1}},\gamma_{D_{2}}}\left(x,y\right){\rm{d}}x{\rm{d}}y}
=(d)\displaystyle\overset{({\rm d})}{=} 0zfγD2(y)FγD1(zy)dy\displaystyle\int_{0}^{z}{f_{\gamma_{D_{2}}}\left(y\right)F_{\gamma_{D_{1}}}\left(z-y\right){\rm{d}}y}
=(e)\displaystyle\overset{({\rm e})}{=} 0zλμyμ1Γ(μ)eλy×[112erfc(zy+αβ)\displaystyle\int_{0}^{z}{\frac{\lambda^{\mu}y^{\mu-1}}{\Gamma\left(\mu\right)}e^{-\lambda y}}\times\left[1\!-\!\frac{1}{2}{\rm erfc}\left(\frac{\sqrt{z\!-\!y}\!+\!\alpha}{\beta}\right)\right.
12erfc(zyαβ)]dy,\displaystyle\left.-\frac{1}{2}{\rm erfc}\left(\frac{\sqrt{z\!-\!y}\!-\!\alpha}{\beta}\right)\right]{\rm{d}}y, (11)

where D{(x,y):x+yz,x>0,y>0}D\triangleq\left\{\left(x,y\right):~{}x+y\leq z,~{}x>0,~{}y>0\right\}, (d)({\rm d}) utilizes the independence of γD1\gamma_{D_{1}} and γD2\gamma_{D_{2}} in Lemma 1, and (e)({\rm e}) is obtained by using (8) and (9).

III-B Distribution of γE\gamma_{E}

Let us first consider the distribution of Zγ¯SREi=1Nhipiejϕisub+γ¯SEhSEZ\!\triangleq\!\sqrt{{\bar{\gamma}}_{S\!R\!E}}\sum_{i=1}^{N}{h_{i}p_{i}{\rm e}^{j\phi_{i}^{\rm sub}}}\!+\!\sqrt{{\bar{\gamma}}_{S\!E}}h_{S\!E}. Then, γE\gamma_{E} in (6) can be calculated according to the relationship of γE=|Z|2\gamma_{E}=\left|Z\right|^{2}. The distribution of ZZ is provided in the following lemma.

Lemma 2: For large NN, Zd𝒞𝒩(0,Nγ¯SRE+γ¯SE)Z\xrightarrow{\rm{d}}\mathcal{CN}(0,N{\bar{\gamma}}_{S\!R\!E}+{\bar{\gamma}}_{S\!E}), and the real and imaginary parts of ZZ are independent RVs with equal variance.

Proof: See Appendix B.\hfill\blacksquare

As disclosed in Lemma 2, γE\gamma_{E} has an exponential distribution with mean Nγ¯SRE+γ¯SEN{\bar{\gamma}}_{S\!R\!E}+{\bar{\gamma}}_{S\!E}. Therefore, the PDF of γE\gamma_{E} is

fγE(x)=ϵeϵx,x0f_{\gamma_{E}}\left(x\right)=\epsilon{\rm e}^{-\epsilon x},~{}x\geq 0 (12)

where ϵ=1/(Nγ¯SRE+γ¯SE)\epsilon={1}/{(N{\bar{\gamma}}_{S\!R\!E}+{\bar{\gamma}}_{S\!E})}.

IV Theoretical Analysis of the SOP

IV-A SOP Analysis

The SOP is an essential performance metric to quantify the performance of PLS, which is defined as the probability that the instantaneous secrecy capacity falls below a target positive SR CthC_{\rm th}. From [11, 12, 13][22], the SOP is calculated as

SOP\displaystyle 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, (13)

where φeCth\varphi\triangleq{\rm e}^{C_{\rm th}}. It is still difficult to compute (13) because the CDF of γD\gamma_{D} in (11) involves an intractable integral. Thus, instead of seeking for a closed-form expression for the SOP, we derive an upper bound as follows

SOP<0FγD1((1+x)φ1)fγE(x)dx=SOP¯.SOP<\int_{0}^{\infty}{F_{\gamma_{D_{1}}}\left(\left(1+x\right)\varphi-1\right)}f_{\gamma_{E}}\left(x\right){\rm{d}}x=\overline{SOP}. (14)

Remark 1: Note that SOP¯\overline{SOP} in (14) is tight when NN is large, because γD1γD2\gamma_{D_{1}}\!\!\gg\!\!\gamma_{D_{2}} holds with high probability, whose proof is provided in Appendix C.

Lemma 3: The upper bound in (14) can be expressed as

SOP¯=112(I1+I2),\displaystyle\overline{SOP}=1-\frac{1}{2}\left(I_{1}+I_{2}\right), (15)

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

I1=\displaystyle I_{1}=~{} erfc(φ1+αβ)2AβeAB2+(φ1)ϵφα2β2\displaystyle{\rm erfc}\left(\frac{\sqrt{\varphi\!-\!1}\!+\!\alpha}{\beta}\right)\!-\!\frac{2\sqrt{A}}{\beta}{\rm e}^{AB^{2}\!+\!\frac{\left(\varphi\!-\!1\right)\epsilon}{\varphi}\!-\!\frac{\alpha^{2}}{\beta^{2}}}
×[1erf(BA+φ12A)],\displaystyle\times\left[1\!-\!{\rm erf}\left(B\sqrt{A}\!+\!\frac{\sqrt{\varphi\!-\!1}}{2\sqrt{A}}\right)\right], (16)
I2=\displaystyle I_{2}=~{} erfc(φ1αβ)2AβeAB2+(φ1)ϵφα2β2\displaystyle{\rm erfc}\left(\frac{\sqrt{\varphi\!-\!1}\!-\!\alpha}{\beta}\right)\!-\!\frac{2\sqrt{A}}{\beta}{\rm e}^{AB^{2}\!+\!\frac{\left(\varphi\!-\!1\right)\epsilon}{\varphi}\!-\!\frac{\alpha^{2}}{\beta^{2}}}
×[1erf(BA+φ12A)],\displaystyle\times\left[1\!-\!{\rm erf}\left(-B\sqrt{A}\!+\!\frac{\sqrt{\varphi\!-\!1}}{2\sqrt{A}}\right)\right], (17)

where A=β2φ4(β2ϵ+φ)A=\frac{\beta^{2}\varphi}{4(\beta^{2}\epsilon+\varphi)} and B=2αβ2B=\frac{2\alpha}{\beta^{2}}.

Proof: See Appendix D.\hfill\blacksquare

IV-B Asymptotic SOP Analysis

We consider application scenarios characterized by a low transmission rate but high security requirements, such as for the Internet of Things [23]. The target SR CthC_{\rm th} can be so small that φ1\varphi\rightarrow 1. In this case, we obtain

SOP¯(f)12σ12γ¯SRDϵ+1em12γ¯SRDϵ2σ12γ¯SRDϵ+1(g)12σ12γ¯SRDϵem122σ12\displaystyle\overline{SOP}\mathop{\to}\limits^{\left(\rm f\right)}\sqrt{\frac{1}{2\sigma_{1}^{2}{\bar{\gamma}}_{S\!R\!D}\epsilon\!+\!1}}{\rm e}^{-\frac{m_{1}^{2}{\bar{\gamma}}_{S\!R\!D}\epsilon}{2\sigma_{1}^{2}{{\bar{\gamma}}_{S\!R\!D}}\epsilon+1}}\mathop{\to}\limits^{\left(\rm g\right)}\sqrt{\frac{1}{2\sigma_{1}^{2}{\bar{\gamma}}_{S\!R\!D}\epsilon}}{\rm e}^{-\frac{m_{1}^{2}}{2\sigma_{1}^{2}}}
=\displaystyle= 1k[1+sinc(2x)π28sinc2(x)]e116[1+sinc(2x)]π2sinc2(x)2N,\displaystyle\sqrt{\frac{1}{k\!\left[1\!+\!{\rm sinc}\left(2x\right)\!-\!\frac{\pi^{2}}{8}{\rm sinc}^{2}\left(x\right)\!\right]}}{\rm e}^{-\frac{1}{\frac{16\left[1+{\rm sinc}\left(2x\right)\right]}{\pi^{2}{\rm sinc}^{2}\left(x\right)}-2}N}, (18)

where (f) is obtained from (15) by setting φ1\varphi\rightarrow 1, and (g) holds true in the high-SNR regime when γ¯SRD{γ¯SRE,γ¯SE}{\bar{\gamma}}_{S\!R\!D}\gg\{{\bar{\gamma}}_{S\!R\!E},{\bar{\gamma}}_{S\!E}\} and by using the following inequalities

2σ12γ¯SRDϵ=\displaystyle 2\sigma_{1}^{2}{\bar{\gamma}}_{S\!R\!D}\epsilon= 1+sinc(2x)π28sinc2(x)γ¯SRE+γ¯SE/Nγ¯SRD\displaystyle\frac{1\!+\!{\rm sinc}\left(2x\right)\!-\!\frac{\pi^{2}}{8}{\rm sinc}^{2}\left(x\right)}{{\bar{\gamma}}_{S\!R\!E}\!+\!{\bar{\gamma}}_{S\!E}/N}{\bar{\gamma}}_{S\!R\!D}
\displaystyle\geq γ¯SRD2(γ¯SRE+γ¯SE)1,\displaystyle\frac{{\bar{\gamma}}_{S\!R\!D}}{2\left({\bar{\gamma}}_{S\!R\!E}\!+\!{\bar{\gamma}}_{S\!E}\right)}\gg 1, (19)

where x2bx\triangleq 2^{-b} and the inequality in (19) follows by 1+sinc(2x)π28sinc2(x)121+{\rm sinc}\left(2x\right)\!-\!\frac{\pi^{2}}{8}{\rm sinc}^{2}\left(x\right)\geq\frac{1}{2} for x(0,12]x\in\left(0,\frac{1}{2}\right]. The last equality in (18) is obtained by defining kγ¯SRD/(γ¯SRE+1Nγ¯SE)k\!\triangleq\!{{\bar{\gamma}}_{S\!R\!D}}/\left({{\bar{\gamma}}_{S\!R\!E}}+\frac{1}{N}{\bar{\gamma}}_{S\!E}\right).

By direct inspection of (18), we evince that SOP¯\overline{SOP} decreases if γ¯SRD{\bar{\gamma}}_{S\!R\!D} increases, which means that enhancing the average SNR at the legitimate user always improves the secrecy performance even for a limited number of discrete phase shift status. In particular, we have SOP¯0\overline{SOP}\rightarrow 0 when γ¯SRD{\bar{\gamma}}_{S\!R\!D}\rightarrow\infty while keeping NN, bb, γ¯SRE{\bar{\gamma}}_{S\!R\!E}, and γ¯SE{\bar{\gamma}}_{S\!E} fixed.

Besides, inspired by [24], we have the following remark to show the impact of the RIS location on SOP.

Remark 2: The SOP improves when {dSR,dRD}\{d_{S\!R},d_{R\!D}\} decreases and dREd_{R\!E} increases. For large NN, it is further found that the asymptotic SOP hardly changes with a moderate variation of dSRd_{S\!R}. This behavior is also explained from the fact that reducing the distance from the source to RIS increases the received SNRs for both the legitimate user and the eavesdropper. Therefore, in the case that the location of the eavesdropper is unknown, we should prioritize deploying the RIS closer to the legitimate user than to the source.

Proof: From (18), we see that the location of the RIS is only reflected in the parameter k=dRDυdREυ+1N(dSEdSR)υk=\frac{d_{R\!D}^{-\upsilon}}{d_{R\!E}^{-\upsilon}+\frac{1}{N}\left(\frac{d_{S\!E}}{d_{S\!R}}\right)^{-\upsilon}}, and the asymptotic SOP in (18) decreases as kk grows. \hfill\blacksquare

Moreover, the high-SNR expression in (18) is also useful for understanding the asymptotic secrecy performance given the number of control bits of discrete phase shifts. Some relevant case studies are reported as follows.

Case 1: Under the assumption of continuous-valued phase shifts, i.e., b=+b=+\infty, the SOP in (18) reduces to

SOP¯|b=+8k(16π2)eπ2322π2N.{\overline{SOP}}\bigg{|}_{b=+\infty}\rightarrow\sqrt{\frac{8}{k\left(16-{\pi^{2}}\right)}}{\rm e}^{-\frac{\pi^{2}}{32-2\pi^{2}}N}. (20)

Case 2: Under the assumption of 11-bit binary phase shifts, i.e., b=1b=1, the SOP in (18) reduces to

SOP¯|b=12keN2.{\overline{SOP}}\bigg{|}_{b=1}\rightarrow\sqrt{\frac{2}{k}}{\rm e}^{-\frac{N}{2}}. (21)

Case 3: Under the assumption of b3b\!\geq\!3, we prove in Appendix E that the quantization noise, which is due to the use of discrete phase shifts, is one order-of-magnitude smaller than the SOP of the continuous phase shifts disclosed in Case 1.

Remark 3: According to Case 1 and Case 2, the SOP tends to SOP¯|b=+8γ¯SRE/[(16π2)γ¯SRD]e0.8N{\overline{SOP}}|_{b=+\infty}\!\rightarrow\!\sqrt{{8{{\bar{\gamma}}_{S\!R\!E}}}/\left[{\left(16-{\pi^{2}}\right){{\bar{\gamma}}_{S\!R\!D}}}\right]}{\rm e}^{-0.8N} and SOP¯|b=12γ¯SRE/γ¯SRDe0.5N{\overline{SOP}}|_{b=1}\!\rightarrow\!\sqrt{{2{{\bar{\gamma}}_{S\!R\!E}}}/{{{\bar{\gamma}}_{S\!R\!D}}}}{\rm e}^{-0.5N} for sufficiently large NN. We see that the SOP loss due to the 11-bit quantization, compared to the ideal continuous-valued phase shifts, can be asymptotically compensated by increasing the number of RIS elements by about 1.61.6 times.

Proof: Let N1N_{1} and N2N_{2} denote the numbers of RIS elements corresponding to b=+b=+\infty and b=1b=1, respectively. By solving SOP¯|b=1SOP¯|b=+{\overline{SOP}}|_{b=1}\leq{\overline{SOP}}|_{b=+\infty}, it follows that N2π216π2N1ln416π21.6N1N_{2}\geq\frac{\pi^{2}}{16-\pi^{2}}N_{1}-{\rm ln}\frac{4}{16-\pi^{2}}\approx 1.6N_{1}. This completes the proof. \hfill\blacksquare

V Numerical Results

In this section, Monte-Carlo simulations are illustrated to validate our analysis. The tested parameters are set to γ¯SE=5dB{\bar{\gamma}}_{S\!E}\!=\!-5~{}\rm{dB} and γ¯SRE=0dB{\bar{\gamma}}_{S\!R\!E}\!=\!0~{}\rm{dB}.

Refer to caption
Figure 2: SOP versus γ¯SRD{\bar{\gamma}}_{S\!R\!D} for different values of bb.
Refer to caption
Figure 3: SOP versus γ¯SRD{\bar{\gamma}}_{S\!R\!D} for different values of NN.

Fig. 2 shows the impact of bb on the SOP when N=30N=30 and Cth=0.05C_{\rm th}=0.05. We observe that the analytical results in (15) match well with the numerical curves. Monte-Carlo simulations are illustrated for values of the SOP no smaller than 10310^{-3} due to the limited number of channel realizations simulated. As stated in Case 3, the gap between b=3b=3 and b=+b=+\infty in Fig. 2 does appear negligible for high SNRs.

In Fig. 3, we consider a larger Cth=0.2C_{\rm th}=0.2 to verify the effectiveness of the asymptotic analysis. We plot the SOP for N=30N=30 by setting b=1b=1 and b=+b=+\infty. As expected, the asymptotic expressions in (20) and (21) are quite tight in the high-SNR regime. Then, we plot the SOP for b=1b=1 by setting N=48N=48 (i.e., which is equal to 1.6×301.6\times 30) and N=60N=60 (2×302\times 30). We see that the setup N=48N=48 with b=1b=1 provides, in the high-SNR regime, the same SOP as the setup N=30N=30 with b=+b=+\infty, which validates the obtained guideline in Remark 3.

VI Conclusion

This paper studied the SOP of an RIS-assisted communication system with discrete phase shifts. The main contribution is to unveil the achievable scaling law of SOP with respect to NN and bb. Specifically, the increased number of RIS elements was quantified to compensate the performance loss caused by binary phase shifts.

Appendix A Proof of The Independence of γD1\gamma_{D_{1}} and γD2\gamma_{D_{2}}

By taking into account that the distribution of Θi\Theta_{i} is symmetric around its mean value, which is equal to zero, we have 𝔼[XY]=0\mathbb{E}\left[XY\right]=0 [25]. Then, the covariance of XX and YY is

Cov[X,Y]=𝔼[XY]𝔼[X]E[Y]=𝔼[XY]=0,{\rm Cov}\left[X,Y\right]=\mathbb{E}\left[XY\right]-\mathbb{E}\left[X\right]E\left[Y\right]=\mathbb{E}\left[XY\right]=0, (22)

which indicates that XX and YY are uncorrelated RVs.

For large values of NN, i=1N|hi||gi|ejΘi\sum_{i=1}^{N}{\left|h_{i}\right|\left|g_{i}\right|{\rm e}^{j\Theta_{i}}} is approximately a complex Gaussian RV by virtue of the CLT, and its real part XX and imaginary part YY are jointly Gaussian RVs. Since two uncorrelated Gaussian RVs are independent as well, it follows that γD1\gamma_{D_{1}} and γD2\gamma_{D_{2}} are independent.

Appendix B Proof of Lemma 2

First, we note that hih_{i} and ϕisub\phi_{i}^{\rm sub} are dependent RVs since ϕisub\phi_{i}^{\rm sub} and ϕiopt=(higi)\phi_{i}^{\rm opt}=-\angle\left({h_{i}}{g_{i}}\right) are correlated RVs. Since ϕisub=Θi+ϕiopt\phi_{i}^{{\rm{sub}}}={\Theta_{i}}+\phi_{i}^{{\rm{opt}}}, the RV ZZ can be rewritten as

Z=γ¯SREi=1N|hi||pi|ej(pigi+Θi)+γ¯SEhSE,\displaystyle Z\!=\!\sqrt{{\bar{\gamma}}_{S\!R\!E}}\sum_{i=1}^{N}{\left|h_{i}\right|\left|p_{i}\right|{\rm e}^{j\left({\angle p_{i}}-{\angle g_{i}}+\Theta_{i}\right)}}\!+\!\sqrt{{\bar{\gamma}}_{S\!E}}h_{S\!E}, (23)

where pi{\angle p_{i}} and gi{\angle g_{i}} are uniformly distributed in [0,2π)\left[0,2\pi\right). The RV Θi\Theta_{i} depends on the phase error at the RIS, and pi{\angle p_{i}} and gi{\angle g_{i}} depend on the positions of the eavesdropper and the legitimate user, respectively. Since the RV pi=|pi|ejpi{p_{i}}=\left|{{p_{i}}}\right|{{\rm e}^{j\angle{p_{i}}}} is independent of the three RVs |hi|\left|{{h_{i}}}\right|, gi\angle{g_{i}}, and Θi{\Theta_{i}} and has zero mean, we obtain 𝔼[|hi||pi|ej(pigi+Θi)]=0\mathbb{E}\left[\left|h_{i}\right|\left|p_{i}\right|{\rm e}^{j\left({\angle p_{i}}-{\angle g_{i}}+\Theta_{i}\right)}\right]\!=\!0, and Var[|hi||pi|ej(pigi+Θi)]=1{\rm Var}\left[\left|h_{i}\right|\left|p_{i}\right|{\rm e}^{j\left({\angle p_{i}}-{\angle g_{i}}+\Theta_{i}\right)}\right]\!=\!1. For large values of NN, i=1N|hi||pi|ej(pigi+Θi)𝑑𝒞𝒩(0,N)\sum_{i=1}^{N}{\left|h_{i}\right|\left|p_{i}\right|{\rm e}^{j\left({\angle p_{i}}-{\angle g_{i}}+\Theta_{i}\right)}}\xrightarrow{d}\mathcal{CN}(0,N) by virtue of the CLT, and then Z𝑑𝒞𝒩(0,Nγ¯SRE+γ¯SE)Z\xrightarrow{d}\mathcal{CN}(0,N{\bar{\gamma}}_{S\!R\!E}+{\bar{\gamma}}_{S\!E}).

Since pip_{i} is a circularly-symmetric Gaussian RV, we have Pr(ejpi)=Pr(ejpiej(gi+Θi)){\rm Pr}({\rm e}^{j\angle p_{i}})\!=\!{\rm Pr}({\rm e}^{j\angle p_{i}}{\rm e}^{j(-\angle g_{i}+\Theta_{i})}). Thus, |pi|ej(pigi+Θi)\left|{{p_{i}}}\right|{\rm e}^{j\left(\angle p_{i}-\angle g_{i}+\Theta_{i}\right)} is a zero-mean circularly-symmetric complex Gaussian RV as well. This implies that the real and imaginary parts of ZZ are uncorrelated and hence independent since they are Gaussian distributed. Also, they have zero means and the same variance.

Appendix C Proof of Remark 1

Since the RV γD2/γD1{\gamma_{D_{2}}}/{\gamma_{D_{1}}} is nonnegative, we obtain

Pr(γD2γD1<0.1)(h)1𝔼[γD2γD1]0.1=(i)1𝔼[γD2]0.1𝔼[γD1],\displaystyle{\rm Pr}\!\left(\frac{\gamma_{D_{2}}}{\gamma_{D_{1}}}\!<\!0.1\right)\mathop{\geq}\limits^{\left(\rm h\right)}1-\frac{\mathbb{E}\left[\frac{\gamma_{D_{2}}}{\gamma_{D_{1}}}\right]}{0.1}\mathop{=}\limits^{\left(\rm i\right)}1-\frac{\mathbb{E}\left[{\gamma_{D_{2}}}\right]}{0.1\mathbb{E}\left[{\gamma_{D_{1}}}\right]}, (24)

where (h) is obtained by applying the Markov inequality [19], and (i) comes from the fact that γD1\gamma_{D_{1}} and γD2\gamma_{D_{2}} are independent.

Furthermore, 𝔼[γD2]𝔼[γD1]\frac{\mathbb{E}\left[{\gamma_{D_{2}}}\right]}{\mathbb{E}\left[{\gamma_{D_{1}}}\right]} is calculated as

𝔼[γD2]𝔼[γD1]=σ22m12+σ12=8[1sinc(2x)](N1)π2sinc2(x)+8[1+sinc(2x)],\displaystyle\frac{\mathbb{E}\left[{\gamma_{D_{2}}}\right]}{\mathbb{E}\left[{\gamma_{D_{1}}}\right]}\!=\!\frac{\sigma_{2}^{2}}{m_{1}^{2}\!+\!\sigma_{1}^{2}}\!=\!\frac{8[1-{\rm sinc}(2x)]}{{(N\!-\!1)\pi^{2}}{\rm sinc}^{2}\left(x\right)\!+\!8[1\!+\!{\rm sinc}(2x)]}, (25)

where x=2b(0,12]x=2^{-b}\in(0,\frac{1}{2}] and it is easy to check that 𝔼[γD2]𝔼[γD1]\frac{\mathbb{E}\left[{\gamma_{D_{2}}}\right]}{\mathbb{E}\left[{\gamma_{D_{1}}}\right]} is increasing as xx grows by calculating its first order derivative. It follows

𝔼[γD2]𝔼[γD1]𝔼[γD2]𝔼[γD1]|x=12=2N+1.\displaystyle\frac{\mathbb{E}\left[{\gamma_{D_{2}}}\right]}{\mathbb{E}\left[{\gamma_{D_{1}}}\right]}\leq\left.\frac{\mathbb{E}\left[{\gamma_{D_{2}}}\right]}{\mathbb{E}\left[{\gamma_{D_{1}}}\right]}\right|_{x=\frac{1}{2}}=\frac{2}{N+1}. (26)

Therefore, (24) is further expressed as

Pr(γD2γD1<0.1)1𝔼[γD2]0.1𝔼[γD1]=120N+1.\displaystyle{\rm Pr}\!\left(\frac{\gamma_{D_{2}}}{\gamma_{D_{1}}}\!<\!0.1\right)\geq 1-\frac{\mathbb{E}\left[{\gamma_{D_{2}}}\right]}{0.1\mathbb{E}\left[{\gamma_{D_{1}}}\right]}=1-\frac{20}{N+1}. (27)

When NN is large, we finally obtain Pr(γD2γD1<0.1)1{\rm Pr}\!\left(\frac{\gamma_{D_{2}}}{\gamma_{D_{1}}}\!<\!0.1\right)\!\mathop{\to}\!1, which means that γD1γD2\gamma_{D_{1}}\!\gg\!\gamma_{D_{2}} holds with high probability.

Appendix D Proof of Lemma 3

By inserting (8) and (12) in (14), the upper bound of the SOP can be rewritten as

SOP¯=\displaystyle\overline{SOP}= 0[112erfc(α+(1+x)φ1β)\displaystyle\int_{0}^{\infty}\left[1-\frac{1}{2}{\rm erfc}\left(\frac{\alpha+\sqrt{\left(1+x\right)\varphi-1}}{\beta}\right)\right.
12erfc((1+x)φ1αβ)]ϵeϵxdx\displaystyle-\left.\frac{1}{2}{\rm erfc}\left(\frac{\sqrt{\left(1+x\right)\varphi-1}-\alpha}{\beta}\right)\right]\epsilon{\rm e}^{-\epsilon x}{\rm{d}}x
=\displaystyle= 112(0erfc(α+(1+x)φ1β)ϵeϵxdx\displaystyle 1-\frac{1}{2}\left(\int_{0}^{\infty}{\rm erfc}\left(\frac{\alpha+\sqrt{\left(1+x\right)\varphi-1}}{\beta}\right)\epsilon{\rm e}^{-\epsilon x}{\rm{d}}x\right.
+0erfc((1+x)φ1αβ)ϵeϵxdx)\displaystyle+\left.\int_{0}^{\infty}{\rm erfc}\left(\frac{\sqrt{\left(1+x\right)\varphi-1}-\alpha}{\beta}\right)\epsilon{\rm e}^{-\epsilon x}{\rm{d}}x\right)
=\displaystyle= 112(I1+I2).\displaystyle 1-\frac{1}{2}\left(I_{1}+I_{2}\right). (28)

The integral I1I_{1} is calculated by using the integration by parts method. We have

I1=\displaystyle I_{1}~{}=~{} 0erfc((1+x)φ1+αβ)ϵeϵxdx\displaystyle\int_{0}^{\infty}{\rm erfc}\left(\frac{\sqrt{\left(1+x\right)\varphi-1}+\alpha}{\beta}\right)\epsilon{\rm e}^{-\epsilon x}{\rm{d}}x
=\displaystyle=~{} erfc((1+x)φ1+αβ)eϵx|x=0\displaystyle\left.-{\rm erfc}\left(\frac{\sqrt{\left(1+x\right)\varphi-1}+\alpha}{\beta}\right){\rm e}^{-\epsilon x}\right|_{x=0}^{\infty}
+0eϵxd(erfc((1+x)φ1+αβ))\displaystyle+\int_{0}^{\infty}{\rm e}^{-\epsilon x}{\rm{d}}\left({\rm erfc}\left(\frac{\sqrt{\left(1+x\right)\varphi-1}+\alpha}{\beta}\right)\right)
=\displaystyle=~{} erfc(φ1+αβ)1πβJ1,\displaystyle{\rm erfc}\left(\frac{\sqrt{\varphi\!-\!1}\!+\!\alpha}{\beta}\right)\!-\!\frac{1}{\sqrt{\pi}\beta}J_{1}, (29)

where J1J_{1} is expressed as

J1\displaystyle J_{1} =0φeϵxe((1+x)φ1+α)2/β2(1+x)φ1dx\displaystyle=\int_{0}^{\infty}{\varphi{\rm e}^{-\epsilon x}\frac{{\rm e}^{-{\left(\sqrt{\left(1+x\right)\varphi-1}+\alpha\right)^{2}}/{\beta^{2}}}\ }{\sqrt{\left(1+x\right)\varphi-1}}}{\rm{d}}x
=2e(φ1)ϵφα2β2φ1e(1β2+ϵφ)t22αβ2tdt\displaystyle=2{\rm e}^{\frac{\left(\varphi-1\right)\epsilon}{\varphi}-\frac{\alpha^{2}}{\beta^{2}}}\int_{\sqrt{\varphi-1}}^{\infty}{{\rm e}^{-\left(\frac{1}{\beta^{2}}+\frac{\epsilon}{\varphi}\right)t^{2}-\frac{2\alpha}{\beta^{2}}t}}{\rm{d}}t
=2e(φ1)ϵφα2β2πAeAB2[1erf(BA+φ12A)],\displaystyle=2{\rm e}^{\frac{\left(\varphi-1\right)\epsilon}{\varphi}-\frac{\alpha^{2}}{\beta^{2}}}\sqrt{\pi A}{\rm e}^{AB^{2}}\left[1\!-\!{\rm erf}\left(B\sqrt{A}\!+\!\frac{\sqrt{\varphi-1}}{2\sqrt{A}}\right)\right], (30)

t=(1+x)φ1t=\sqrt{\left(1+x\right)\varphi-1}, and the last equality is obtained by using [18, Eq. (3.322)], where A=β2φ4(β2ϵ+φ)A=\frac{\beta^{2}\varphi}{4(\beta^{2}\epsilon+\varphi)} and B=2αβ2B=\frac{2\alpha}{\beta^{2}}. Further, by substituting (30) into (29), I1I_{1} is obtained as shown in (16). Analogously, I2I_{2} is calculated by replacing α\alpha in (16) with α-\alpha, which yields the desired result in (17).

Appendix E Proof of Case 3

We apply the Taylor expansion to the SOP in (18). More precisely, around x=0x=0 and for small values of x=2b<1x=2^{-b}<1, we have

1k[1+sinc(2x)π28sinc2(x)]=c1+c2x2+O(x4),\sqrt{\frac{1}{k\!\left[1\!+\!{\rm sinc}\left(2x\right)\!-\!\frac{\pi^{2}}{8}{\rm sinc}^{2}\left(x\right)\!\right]}}={\rm c}_{1}+{\rm c}_{2}x^{2}+O\left(x^{4}\right), (31)
e116[1+sinc(2x)]π2sinc(x)2N=ec3N+O(x4),{\rm e}^{-\frac{1}{\frac{16\left[1+{\rm sinc}(2x)\right]}{\pi^{2}{\rm sinc}(x)}-2}N}={\rm e}^{-{\rm c}_{3}N}+O\left(x^{4}\right), (32)

where c1=8k(16π2){\rm c}_{1}=\sqrt{\frac{8}{k\left(16-\pi^{2}\right)}}, c2=π26c1{\rm c}_{2}=\frac{\pi^{2}}{6}{\rm c}_{1}, and c3=π2322π2{\rm c}_{3}=\frac{\pi^{2}}{32-2\pi^{2}}.

By substituting (31) and (32) in (18), the SOP is further rewritten as

SOP¯\displaystyle\overline{SOP} (c1+c2x2+O(x4))(ec3N+O(x4))\displaystyle\rightarrow\left({\rm c}_{1}+{\rm c}_{2}x^{2}+O\left(x^{4}\right)\right)\left({\rm e}^{-{\rm c}_{3}N}+O\left(x^{4}\right)\right)
=c1ec3N+c2ec3Nx2+O(x4).\displaystyle={\rm c}_{1}{\rm e}^{-{\rm c}_{3}N}+{\rm c}_{2}{\rm e}^{-{\rm c}_{3}N}x^{2}+O\left(x^{4}\right). (33)

Since the first term in (33) is the SOP for b=+b\!=\!+\infty as given in Case 1, the performance loss due to finite values of bb is dominated by the second term in (33), i.e., c2ec3Nx2{\rm c}_{2}{\rm e}^{-{\rm c}_{3}N}x^{2}. If we restrict that the performance loss is one order of magnitude smaller than Case 1 for b=+b\!=\!+\infty, it gives (c2ec3Nx2)/(c1ec3N)<1/10\left({{\rm c}_{2}{\rm e}^{-{\rm c}_{3}N}x^{2}}\right)/\left({{\rm c}_{1}{\rm e}^{-{\rm c}_{3}N}}\right)<{1}/{10}, which implies x<3/(5π2)x<\sqrt{3/\left(5{\pi}^{2}\right)} and equivalently b3b\geq 3.

References

  • [1] E. Basar, M. Di Renzo, J. De Rosny, M. Debbah, M.-S. Alouini, and R. Zhang, “Wireless communications through reconfigurable intelligent surfaces,” IEEE Access, vol. 7, pp. 116753–116773, Aug. 2019.
  • [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] 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.
  • [4] 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.
  • [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., early access. doi: 10.1109/MWC.018.2100717.
  • [6] 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.
  • [7] L. Dong, H.-M. Wang, J. Bai, and H. Xiao, “Double intelligent reflecting surface for secure transmission with inter-surface signal reflection,” IEEE Trans. Veh. Technol., vol. 70, no. 3, pp. 2912–2916, Mar. 2021.
  • [8] 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.
  • [9] L. Lv, Q. Wu, Z. Li, N. Al-Dhahir, and J. Chen, “Secure two-way communications via intelligent reflecting surfaces,” IEEE Commun. Lett., vol. 25, no. 3, pp. 744–748, Mar. 2021.
  • [10] J. Luo, F. Wang, S. Wang, H. Wang, and D. Wang, “Reconfigurable intelligent surface: Reflection design against passive eavesdropping,” IEEE Trans. Wireless Commun., vol. 20, no. 5, pp. 3350–3364, May 2021.
  • [11] 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.
  • [12] 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.
  • [13] 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.
  • [14] Q. Wu and R. Zhang, “Beamforming optimization for wireless network aided by intelligent reflecting surface with discrete phase shifts,” IEEE Trans. Commun., vol. 68, no. 3, pp. 1838–1851, Mar. 2020.
  • [15] 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.
  • [16] 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.
  • [17] P. Xu, G. Chen, Z. Yang, and M. Di Renzo, “Reconfigurable intelligent surfaces-assisted communications with discrete phase shifts: How many quantization levels are required to achieve full diversity?,” IEEE Wireless Commun. Lett., vol. 10, no. 2, pp. 358–362, Feb. 2021.
  • [18] I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products, 7th ed. San Diego, CA, USA: Academic, 2007.
  • [19] H. Pishro-Nik, Introduction to Probability, Statistics and Random Processes, Gaithersburg, MD, USA: Kappa Research, 2014.
  • [20] J. G. Proakis and M. Salehi, Digital Communications, 5th ed., New York, NY, USA: McGraw-Hill, 2008.
  • [21] V. M. Kapinas, S. K. Mihos, and G. K. Karagiannidis, “On the monotonicity of the generalized Marcum and Nuttall Q-functions,” IEEE Trans. Inf. Theory, vol. 55, no. 8, pp. 3701–3710, Aug. 2009.
  • [22] M. Elkashlan, L. Wang, T. Q. Duong, G. K. Karagiannidis, and A. Nallanathan, “On the security of cognitive radio networks,” IEEE Trans. Veh. Technol., vol. 64, no. 8, pp. 3790–3795, Aug. 2015.
  • [23] N. Wang, P. Wang, A. Alipour-Fanid, L. Jiao, and K. Zeng, “Physical-layer security of 5G wireless networks for IoT: Challenges and opportunities,” IEEE Internet Things J., vol. 6, no. 5, pp. 8169–8181, Oct. 2019.
  • [24] S. Zeng, H. Zhang, B. Di, Z. Han, and L. Song, “Reconfigurable intelligent surface (RIS) assisted wireless coverage extension: RIS orientation and location optimization,” IEEE Commun. Lett., vol. 25, no. 1, pp. 269–273, Jan. 2021.
  • [25] I. Trigui, W. Ajib, W.-P. Zhu, and M. Di Renzo, “Performance evaluation and diversity analysis of RIS-assisted communications over generalized fading channels in the presence of phase noise,” IEEE Open J. Commun. Soc., vol. 3, pp. 593–607, 2022.