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

Physical-Layer Security in the Finite Blocklength Regime over Fading Channels

Tong-Xing Zheng,  Hui-Ming Wang, 
Derrick Wing Kwan Ng,  and Jinhong Yuan
T.-X. Zheng is with the School of Information and Communications Engineering, Xi’an Jiaotong University, Xi’an 710049, China, also with the State Key Laboratory of Integrated Services Networks, Xidian University, Xi’an 710071, China, and also with the Ministry of Education Key Laboratory for Intelligent Networks and Network Security, Xi’an Jiaotong University, Xi’an 710049, China (e-mail: [email protected]).H.-M. Wang is with the School of Information and Communications Engineering, Xi’an Jiaotong University, Xi’an 710049, China, and also with the Ministry of Education Key Laboratory for Intelligent Networks and Network Security, Xi’an Jiaotong University, Xi’an 710049, China (e-mail: [email protected]).D. W. K. Ng and J. Yuan are with the School of Electrical Engineering and Telecommunications, University of New South Wales, Sydney, NSW 2052, Australia (e-mail: [email protected]; [email protected]).
Abstract

This paper studies physical-layer secure transmissions from a transmitter to a legitimate receiver against an eavesdropper over slow fading channels, taking into account the impact of finite blocklength secrecy coding. A comprehensive analysis and optimization framework is established to investigate secrecy throughput for both single- and multi-antenna transmitter scenarios. Both adaptive and non-adaptive design schemes are devised, in which the secrecy throughput is maximized by exploiting the instantaneous and statistical channel state information of the legitimate receiver, respectively. Specifically, optimal transmission policy, blocklength, and code rates are jointly designed to maximize the secrecy throughput. Additionally, null-space artificial noise is employed to improve the secrecy throughput for the multi-antenna setup with the optimal power allocation derived. Various important insights are developed. In particular, 1) increasing blocklength benefits both reliability and secrecy under the proposed transmission policy; 2) secrecy throughput monotonically increases with blocklength; 3) secrecy throughput initially increases but then decreases as secrecy rate increases, and the optimal secrecy rate maximizing the secrecy throughput should be carefully chosen in order to strike a good balance between rate and decoding correctness. Numerical results are eventually presented to verify theoretical findings.

Index Terms:
Physical-layer security, wiretap code, secrecy throughput, finite blocklength, optimization.

I Introduction

In the past decade, pursuing communication security at the physical layer has received a considerable interest, e.g., [2]-[8]. In particular, physical-layer security exploits the inherent randomness of noise and wireless channels to protect wireless secure transmissions [9]-[13], which can provide an additional mechanism for security guarantee and can coexist with those security techniques already employed at the upper layers, such as key-based encipherment. Most recent progress in developing physical-layer security is motivated by Wyner’s pioneering work. Specifically, the concept of secrecy capacity was first established which is defined as the supremum of secrecy rates at which both reliability and secrecy are achieved over a wiretap channel [14]. Wyner showed that the error probability and information leakage can be made arbitrarily low concurrently with an appropriate secrecy coding, provided that a data rate below the secrecy capacity is chosen and meanwhile the data is mapped to asymptotically long codewords, i.e., the coding blocklength tends to infinity. However, the upcoming 5G wireless communication systems are required to support various novel traffic types adopting short packets to reduce the end-to-end communication latency, e.g., smart-traffic safety and machine-to-machine communications [15, 16]. For the short-packet applications, conventional physical-layer security schemes originated from infinite blocklength are generally suboptimal and the impact of finite blocklength could be destructive for secure communications. Therefore, it is necessary to rethink the analysis and design of physical-layer security for the finite blocklength regime.

I-A Previous Works and Motivations

Decoding with finite blocklength will inevitably reduce the secrecy capacity and some preliminary works have been devoted to analyzing the impact of finite blocklength on secrecy for the wiretap channel. For example, the authors in [17] derived an upper bound for the information leakage probability for a given target decoding error probability demonstrating the inherent trade-off between secrecy and reliability. The authors in [18] provided both upper and lower bounds for the maximal secrecy rate capturing the impact of finite blocklength, error probability, and information leakage in both degraded discrete-memoryless wiretap channels and Gaussian wiretap channels. The obtained bounds were shown to be tighter than existing ones from [19, 20]. The work in [18] was further extended by [21], in which the optimal second-order secrecy rate was derived for a semi-deterministic wiretap channel, and the optimal tradeoff between secrecy and reliability with finite blocklength was analytically characterized. It should be noted that, all the above works were aimed to uncover the fundamental limits of secrecy performance from the information theory point of view, whereas the design of practical signaling and transmission schemes were not investigated.

In practice, due to finite blocklength penalty for practical coding schemes, even a secrecy rate below the secrecy capacity cannot guarantee a perfectly successful and secure communication. In this sense, in addition to exploring and/or improving the fundamental limits of the maximal secrecy rate, optimizing secrecy throughput seems more important from the perspective of transmission efficiency, particularly for fading channels where the code rates can be adapted to the fading status. Herein, the secrecy throughput denotes the amount of successfully delivered secret information subject to certain reliability and secrecy constraints. In fact, the secrecy throughput has extensively been taken as an optimization objective for the design of secure transmissions in slow fading channels in the context of infinite blocklength [22]-[26]. Nevertheless, to optimize the secrecy throughput under the constraint of finite blocklength is difficult, and the results derived for infinite blocklength, e.g., [22]-[26], cannot be directly applied. Indeed, the blocklength itself is an optimization variable, and it couples with other variables in a sophisticated manner which makes the optimization problem intractable. For instance, the authors in a recent work [27] investigated the secrecy throughput of a relay-aided secure transmission with finite blocklength, where neither the instantaneous channel state information (CSI) with respect to (w.r.t.) the legitimate receiver nor the eavesdropper is available at the transmitter side. Numerical results were presented therein to show that there exists a critical value of the blocklength that maximizes the secrecy throughput.

Despite the above endeavors, there are some fundamental questions regarding the design of physical-layer security schemes with finite blocklength that have not been thoroughly addressed. First of all, a theoretical proof of the optimal blocklength and the corresponding secrecy rate for maximizing the secrecy throughput is of great significance for the practical design of secure transmissions, which however has not yet been reported by existing literature. Also, in many applications, the transmitter is capable to acquire the instantaneous CSI of the legitimate receiver in slow fading channels via training or feedback. Yet, the potential of exploiting the instantaneous CSI to alleviate the negative impact of finite blocklength on the performance of secure communications has not been exploited. Furthermore, only the single-antenna transmitter scenario has been considered, e.g., [17]-[21], [27], and the design of the optimal signaling and code rates for multi-antenna systems with finite blocklength is still an open issue. This research work aims to provide an analytical framework and design schemes to address the abovementioned problems.

I-B Contributions

This paper investigates the security issue between a pair of legitimate communicating parties in the presence of an eavesdropper, considering the impact of finite blocklength in secrecy coding. The secrecy throughput is thoroughly analyzed and optimized for both single- and multi-antenna transmitter scenarios. In particular, both adaptive and non-adaptive parameter design schemes are proposed for each scenario. The main contributions of this work are summarized as follows:

  • For the single-antenna transmitter scenario, the secrecy throughput is maximized by jointly optimizing the transmission policy, blocklength, as well as code rates. Closed-form bounds and approximations for the secrecy rate are provided to facilitate the practical design of code rates for achieving a close-to-optimal performance.

  • For the multi-antenna transmitter configuration, the optimality of the null-space artificial noise (AN) scheme in terms of secrecy throughput maximization is first investigated. Afterwards, the optimal transmission policy, blocklength, code rates, and power allocation between the information-bearing signal and the AN are derived. Particularly, the power allocation and the secrecy rate are designed via the alternating optimization method, and their impacts on the system performance are further revealed.

  • Numerous useful insights into the design of secure transmissions are provided with finite blocklength. For example, 1) increasing the blocklength can improve both reliability and secrecy, with properly exploiting the instantaneous CSI of the main channel and the statistical CSI of the wiretap channel, which has not been revealed by existing literature, e.g., [17]-[21]; 2) using the maximal blocklength is profitable for boosting the secrecy throughput, which is distinguished from the observation in [27]; 3) due to the finite blocklength penalty, there is a critical secrecy rate that can maximize the secrecy throughput even for the adaptive scheme, rather than always employing the maximal available secrecy rate, which is fundamentally different from the phenomenon with infinite blocklength, e.g., [22, 23].

I-C Organization and Notations

The remainder of this paper is organized as below. Section II describes the system model and the underlying optimization problem. Sections III and IV detail the secrecy throughput maximization for both single- and multi-antenna transmitter scenarios. Section V draws a conclusion.

Notations: Bold lowercase letters denote column vectors. |||\cdot|, \|\cdot\|, ()(\cdot)^{\dagger}, ()T(\cdot)^{\rm T}, ln()\ln(\cdot), {}\mathbb{P}\{\cdot\}, 𝔼v[]\mathbb{E}_{v}[\cdot] denote the absolute value, Euclidean norm, conjugate, transpose, natural logarithm, probability, and the expectation over a random variable vv, respectively. fv()f_{v}(\cdot) and v()\mathcal{F}_{v}(\cdot) denote the probability density function (PDF) and cumulative distribution function (CDF) of vv, respectively. F1()F^{-1}(\cdot) denotes the inverse function of a function F()F(\cdot). 𝒞𝒩(μ,σ2)\mathcal{CN}(\mu,\sigma^{2}) denotes the circularly symmetric complex Gaussian distribution with mean μ\mu and variance σ2\sigma^{2}.

II System Model and Problem Description

II-A Channel Model

Refer to caption
Figure 1: Secure transmission from Alice (multi-antenna) to Bob (single-antenna) overheard by Eve (single-antenna). Alice adopts a wiretap code with a binning structure, where 2nRs2^{nR_{s}} messages each are mapped to a bin of 2nRe2^{nR_{e}} codewords with a finite blocklength nn. A codeword among a set of codewords representing the same message is randomly chosen for transmission [14].

Consider a secure transmission from a transmitter (Alice) to a legitimate receiver (Bob) coexisting with an eavesdropper (Eve), as depicted in Fig. 1. Alice is equipped with M1M\geq 1 transmit antennas, whereas Bob and Eve are single-antenna devices. Quasi-static Rayleigh fading channels are considered, where the channel coherence time is on the order of the blocklength. More specifically, the fading coefficients are assumed to remain constant during the transmission of an entire codeword, but change independently and randomly between two codewords [2]. Denote the coefficients of the main and wiretap channels by 𝒉b\bm{h}_{b} and 𝒉e\bm{h}_{e}, and each entry of 𝒉b\bm{h}_{b} and 𝒉e\bm{h}_{e} follow the Gaussian distribution 𝒞𝒩(0,σb2)\mathcal{CN}(0,\sigma_{b}^{2}) and 𝒞𝒩(0,σe2)\mathcal{CN}(0,\sigma_{e}^{2}), respectively.111The subscripts bb and ee are used to refer to Bob and Eve, respectively. A common hypothesis is adopted [22, 23], i.e., Bob and Eve know perfectly the instantaneous CSI of their individual channels 𝒉b\bm{h}_{b} and 𝒉e\bm{h}_{e}, and Alice has the instantaneous CSI of Bob’s channel 𝒉b\bm{h}_{b} but does not has the instantaneous CSI of Eve’s channel 𝒉e\bm{h}_{e}. Besides, the statistics of both channels 𝒉b\bm{h}_{b} and 𝒉e\bm{h}_{e} are available at Alice. Assume that 𝒉b\bm{h}_{b}, 𝒉e\bm{h}_{e}, and the receiver noise are mutually independent, where noise variances at Bob and Eve are denoted by wb2w_{b}^{2} and we2w_{e}^{2}, respectively. Alice adopts a constant transmit power PP. For notational simplicity, define PbPwb2P_{b}\triangleq\frac{P}{w_{b}^{2}} and PePwe2P_{e}\triangleq\frac{P}{w_{e}^{2}} as the normalized power for Bob and Eve, respectively.

II-B Finite Blocklength Secrecy Coding

To safeguard information confidentiality, secrecy coding should be employed to encode the secret information bits. Instead of investigating any explicit practical constructions of secrecy codes, the Wyner’s wiretap code [14], as a generic code structure, is employed in this paper. A synopsis of the state-of-the-art coding schemes for wiretap channels can be found in [28].

It is reported in [2] that the Wyner’s wiretap code possesses a binning structure, as illustrated in Fig. 1, where 2nRs2^{nR_{s}} messages are encoded to 2nRt2^{nR_{t}} codewords, and each message is mapped to a bin of 2nRe2^{nR_{e}} codewords. Here, nn denotes the blocklength (i.e., the codeword length or the number of channel uses), RsR_{s} and RtR_{t} (bits/s/Hz/channel) denote the secrecy rate and codeword rate, respectively. The binning codeword rate, i.e., the rate redundancy Re=RtRsR_{e}=R_{t}-R_{s}, reflects the cost of providing secrecy.

It is well-known that, for an infinite blocklength with nn\rightarrow\infty, as long as the codeword rate RtR_{t} is not larger than Bob’s channel capacity, Bob can recover messages with an arbitrarily low decoding error probability. On the other hand, perfect secrecy cannot always be guaranteed due to the absence of Eve’s instantaneous CSI: once the rate redundancy ReR_{e} falls below Eve’s channel capacity, perfect secrecy is compromised, and a secrecy outage event is said to have occurred. Nevertheless, in the finite blocklength regime which is restricted to a finite number of channel uses, no practical protocols can achieve perfectly reliable communications [29]. Hence, to capture the impact of finite blocklength, the maximal channel coding rate for sustaining a desired decoding error probability ϵ\epsilon at a finite blocklength nn (e.g., n100n\geq 100) for a given signal-to-noise ratio (SNR) γ\gamma was studied in [30] and can be approximated by

R(γ,n,ϵ)C(γ)V(γ)nQ1(ϵ),R(\gamma,n,\epsilon)\approx C(\gamma)-\sqrt{\frac{V(\gamma)}{n}}Q^{-1}(\epsilon), (1)

where C(γ)log2(1+γ)C(\gamma)\triangleq\log_{2}(1+\gamma) denotes the Shannon channel capacity, V(γ)(1(1+γ)2)log22eV(\gamma)\triangleq\left(1-(1+\gamma)^{-2}\right)\log_{2}^{2}e denotes the channel dispersion [30], and Q(x)Q(x) is the QQ-function defined as Q(x)12πxet22𝑑tQ(x)\triangleq\frac{1}{\sqrt{2\pi}}\int_{x}^{\infty}e^{-\frac{t^{2}}{2}}dt. Equivalently, the decoding error probability for a given coding rate RR can be expressed as

ϵ(γ,n,R)=Q(C(γ)RV(γ)/n).\epsilon(\gamma,n,R)=Q\left(\frac{C(\gamma)-R}{\sqrt{V(\gamma)/n}}\right). (2)

For ease of notation, let CiC(γi)C_{i}\triangleq C(\gamma_{i}) and ViV(γi)V_{i}\triangleq V(\gamma_{i}) for i{b,e}i\in\{b,e\}, where γi\gamma_{i} denotes the corresponding SNR. Define the successful decoding probability of Bob as the complement of its decoding error probability with the codeword rate RtR_{t}. Then, the successful decoding probability conditioned on the power gain of the main channel, i.e., η𝒉b2\eta\triangleq\|\bm{h}_{b}\|^{2}, can be expressed as

ps(η)1ϵ(γb,n,Rt)=1Q(CbRtVb/n).p_{s}(\eta)\triangleq 1-\epsilon(\gamma_{b},n,R_{t})=1-Q\left(\frac{C_{b}-R_{t}}{\sqrt{V_{b}/n}}\right). (3)

The secrecy performance is characterized by the information leakage probability defined below:

𝒪e𝔼γe[1ϵ(γe,n,Re)].\mathcal{O}_{e}\triangleq\mathbb{E}_{\gamma_{e}}\left[1-\epsilon(\gamma_{e},n,R_{e})\right]. (4)
Remark 1

Due to the finite blocklength, the secrecy metric information leakage probability in (4) appears to be distinguished from the widely used secrecy outage probability, defined as {ReCe}\mathbb{P}\{R_{e}\leq C_{e}\} [23], for the infinite blocklength regime with nn\rightarrow\infty.

II-C Optimization Problem

Since Alice knows Bob’s instantaneous CSI perfectly, she is able to adapt the code rates to the instantaneous channel gain η\eta, which implies that the code rates can be functions of η\eta. This paper focuses on the metric named secrecy throughput (bits/s/Hz/channel), which measures the average successfully transmitted information bits per second per Hertz per channel use subject to a secrecy constraint 𝒪eδ\mathcal{O}_{e}\leq\delta, where δ[0,1]\delta\in[0,1] is a pre-established threshold for the information leakage probability. Formally, the secrecy throughput is defined as

𝒯𝔼η[Rs(η)ps(η)]s.t.𝒪eδ,\mathcal{T}\triangleq\mathbb{E}_{\eta}\left[R_{s}(\eta)p_{s}(\eta)\right]~{}{\rm s.t.}~{}\mathcal{O}_{e}\leq\delta, (5)

which is averaged over η\eta. Note that the introduction of finite blocklength leads to a different definition of secrecy throughput compared to the case of infinite blocklength which is 𝒯𝔼η[Rs(η)]\mathcal{T}\triangleq\mathbb{E}_{\eta}\left[R_{s}(\eta)\right] [22, 23]. In addition, as will be shown later, in order to meet certain secrecy and reliability requirements during the transmission period, an on-off transmission policy is required;222The on-off policy was initially proposed for ergodic-fading channels [31], where a codeword experiences many channel realizations. It was later introduced to slow fading channels and well characterized the condition for secure transmissions [22]. i.e., the transmission should take place only when the channel gain η\eta exceeds some threshold μ>0\mu>0. With the on-off policy, Rs(η)R_{s}(\eta) is set to zero for η<μ\eta<\mu.

This paper aims to maximize the secrecy throughput by designing the optimal on-off threshold, signaling, blocklength, as well as code rates. The following two sections will detail the optimization for single- and multi-antenna transmitter scenarios, respectively. For each scenario, both adaptive and non-adaptive design schemes are examined, where Alice adjusts the arguments based on the instantaneous and statistical CSI of the main channel, respectively.

III Single-Antenna Transmitter Scenario

For the single-antenna transmitter scenario, the SNRs of Bob and Eve are given by γb=Pbη\gamma_{b}=P_{b}\eta with η=|hb|2\eta=|h_{b}|^{2} and γe=Pe|he|2\gamma_{e}=P_{e}|h_{e}|^{2}, respectively. Clearly, γi\gamma_{i} is exponentially distributed with mean Γi=Piσi2\Gamma_{i}=P_{i}\sigma_{i}^{2} for i{b,e}i\in\{b,e\}. The subsequent two subsections aim to maximize the secrecy throughput 𝒯\mathcal{T} defined in (5) by jointly designing the on-off threshold μ(η)\mu(\eta), the wiretap code rates Rs(η)R_{s}(\eta) and Re(η)R_{e}(\eta), and the blocklength n(η)n(\eta), via adaptive and non-adaptive ways, respectively. For notational convenience, these parameters are treated as functions of η\eta by default for the adaptive scheme, with the notation η\eta being dropped, and 𝒯A\mathcal{T}_{\rm A} and 𝒯N\mathcal{T}_{\rm N} are used to differentiate the adaptive scheme to its non-adaptive counterpart. The optimization problem then can be formulated as below:

maxμ>0,Re>0,Rs>0,n\displaystyle\max_{\mu>0,R_{e}>0,R_{s}>0,n}~{} 𝒯=𝔼η[Rsps]\displaystyle\mathcal{T}=\mathbb{E}_{\eta}\left[R_{s}p_{s}\right] (6a)
s.t.\displaystyle{\rm s.t.}~{}~{} CbRt=Rs+Re,η>μ,\displaystyle C_{b}\geq R_{t}=R_{s}+R_{e},~{}\forall\eta>\mu, (6b)
𝒪eδ,\displaystyle\mathcal{O}_{e}\leq\delta, (6c)
1nN,n,N+.\displaystyle 1\leq n\leq N,~{}n,N\in\mathbb{Z}^{+}. (6d)

Note that (6b) is interpreted as a reliability requirement since otherwise the successful decoding probability psp_{s} in (3) falls below 0.50.5 and it is no better than random guessing, which is definitely not acceptable; (6c) describes the secrecy constraint; (6d) is related to a latency constraint, where the integer NN denotes the maximal available blocklength imposed by a maximal tolerable delay.

III-A Adaptive Optimization Scheme

In the adaptive scheme, the parameters μ\mu, RsR_{s}, ReR_{e}, and nn are designed based on η\eta, i.e., they are adjusted in real time. A detailed optimization procedure is provided as follows.

III-A1 Solving ReR_{e}

Since QQ-function Q(x)Q(x) is a monotonically decreasing function of xx, it is known that psp_{s} defined in (3) decreases with ReR_{e} for a fixed RsR_{s}. This suggests that, the optimal ReR_{e} maximizing 𝒯A\mathcal{T}_{\rm A} should be the minimal ReR_{e} that satisfies the secrecy constraint 𝒪eδ\mathcal{O}_{e}\leq\delta. Now that 𝒪e\mathcal{O}_{e} in (4) decreases with ReR_{e}, the optimal ReR_{e} is given as the inverse of 𝒪e\mathcal{O}_{e} at δ\delta, i.e.:

Re=𝒪e1(δ).R_{e}^{*}=\mathcal{O}_{e}^{-1}(\delta). (7)

Obviously, ReR_{e}^{*} is independent of η\eta, but monotonically decreases with δ\delta. This is intuitive that a larger rate redundancy is required to combat the eavesdropper in order to meet a more rigorous secrecy constraint. Although it is difficult to derive a closed-form expression for ReR_{e}^{*} due to the complicated QQ-function, the value of ReR_{e}^{*} can be efficiently acquired via a bisection method with 𝒪e(Re)=δ\mathcal{O}_{e}(R_{e})=\delta, requiring only the computation of Q(x)Q(x) or a lookup table.

III-A2 Solving μ\mu

The secrecy throughput 𝒯A\mathcal{T}_{\rm A} given in (6a) can be calculated as

𝒯A=PbμRspsfγb(γ)𝑑γ,\mathcal{T}_{\rm A}=\int_{P_{b}\mu}^{\infty}R_{s}p_{s}f_{\gamma_{b}}(\gamma)d\gamma, (8)

where fγb(γ)=1Γbeγ/Γbf_{\gamma_{b}}(\gamma)=\frac{1}{\Gamma_{b}}e^{-{\gamma}/{\Gamma_{b}}} is the PDF of γb=Pbη\gamma_{b}=P_{b}\eta. It appears that choosing μ\mu as small as possible is beneficial for increasing 𝒯A\mathcal{T}_{\rm A}, on the premise of satisfying the reliability constraint (6b). In addition, constraint (6b) suggests that Cb>Reη=γbPb>2Re1PbC_{b}>R_{e}^{*}\Rightarrow\eta=\frac{\gamma_{b}}{P_{b}}>\frac{2^{R_{e}^{*}}-1}{P_{b}} must be ensured to achieve a positive RsR_{s}. Hence, the optimal on-off threshold is given by

μ=2Re1Pb.\mu^{*}=\frac{2^{R_{e}^{*}}-1}{P_{b}}. (9)

This result indicates that the transmission condition for the adaptive scheme is determined by the secrecy constraint. Apparently, μ\mu^{*} is monotonically decreasing with δ\delta since ReR_{e}^{*} decreases with δ\delta. This implies, a weaker channel is still allowed for transmission for a looser secrecy constraint.

Once μ\mu is obtained, to maximize 𝒯A\mathcal{T}_{\rm A} in (8) only calls for maximizing 𝒯A(η)Rsps\mathcal{T}_{\rm A}(\eta)\triangleq R_{s}p_{s} which is conditioned on η\eta. The subproblem is described as below:

maxRs,n𝒯A(η)=Rspss.t.(6d),0RsCbRe.\max_{R_{s},n}~{}\mathcal{T}_{\rm A}(\eta)=R_{s}p_{s}~{}~{}{\rm s.t.}~{}~{}\eqref{st_max_c3},~{}0\leq R_{s}\leq C_{b}-R_{e}^{*}. (10)

The basic idea to tackle the above problem is first to maximize psp_{s} over nn for a fixed RsR_{s} and then to design the optimal RsR_{s} that maximizes RspsR_{s}p_{s} with the optimal nn.

III-A3 Solving nn

For any fixed RtCbR_{t}\leq C_{b}, there is no doubt that psp_{s} increases with nn. However, as shown in (4), ϵ(γe,n,Re)\epsilon(\gamma_{e},n,R_{e}) decreases with nn for ReCeR_{e}\leq C_{e} but increases with nn otherwise. Then, it remains unclear how 𝒪e\mathcal{O}_{e} defined in (4), as well as ReR_{e}^{*} in (7), varies with nn. More importantly, it is less obvious if the monotonicity of psp_{s} w.r.t. nn can still hold, since Rt=Rs+ReR_{t}=R_{s}+R_{e}^{*} becomes independent of nn. Therefore, in order to derive the optimal nn^{*} maximizing 𝒯A(η)\mathcal{T}_{\rm A}(\eta) in (10), the monotonicity of 𝒪e\mathcal{O}_{e} or ReR_{e}^{*} w.r.t. nn should be first identified.

Lemma 1

𝒪e\mathcal{O}_{e} in (4) and ReR_{e}^{*} in (7) decrease with nn.

Proof 1

Please refer to Appendix -A.

Lemma 1 shows that increasing the blocklength is beneficial for decreasing the information leakage probability such that the required rate redundancy of the wiretap code can be lowered. This result is perhaps counter-intuitive, which makes sense when one realizes that a larger blocklength will yield a larger decoding error probability for Eve if Eve’s channel capacity falls below the rate redundancy. With Lemma 1, the monotonicity of 𝒯A(η)\mathcal{T}_{\rm A}(\eta) w.r.t. nn is uncovered, followed by the optimal nn^{*} that maximizes 𝒯A(η)\mathcal{T}_{\rm A}(\eta).

Theorem 1

𝒯A(η)\mathcal{T}_{\rm A}(\eta) in (10) increases with nn and is maximized at n=Nn^{*}=N.

Proof 2

Please refer to Appendix -B.

Theorem 1 reveals that exploiting a larger blocklength is beneficial for improving the secrecy throughput under given channel gains. This result is nontrivial in light of [27] where there exists a critical value of the blocklength, instead of the maximal one, that can achieve the maximal secrecy throughput. The main reason behind the two different results lies in that, Bob’s instantaneous CSI is available here and is adequately exploited, and the codeword rate will not exceed Bob’s channel capacity under the on-off policy such that using a larger blocklength can always lower the decoding error probability for Bob. Combined with Lemma 1, it can be seen that increasing the blocklength improves reliability and secrecy simultaneously, thus making the secrecy throughput higher. However, this can no longer be promised in [27] where the instantaneous CSI of the main channel is unknown, and using a larger blocklength might degrade the reliability once the codeword rate exceeds Bob’s channel capacity, just as implied in Lemma 1. Revisiting (8), since μ\mu^{*} in the lower limit of the integral decreases with nn (see (9) where ReR_{e}^{*} decreases with nn), it is clear that the global optimal blocklength that maximizes 𝒯A\mathcal{T}_{\rm A} is also n=Nn^{*}=N.

III-A4 Solving RsR_{s}

Substituting the derived optimal ReR_{e}^{*}, μ\mu^{*}, and nn^{*} into (3) yields the maximal psp_{s}, and then the optimal RsR_{s}^{*} can be determined by solving the following problem:

maxRs\displaystyle\max_{R_{s}} 𝒯A(η)=Rs[1Q(CbRsReVb/N)]\displaystyle~{}\mathcal{T}_{\rm A}(\eta)=R_{s}\left[1-Q\left(\frac{C_{b}-R_{s}-R_{e}^{*}}{\sqrt{V_{b}/N}}\right)\right] (11a)
s.t.\displaystyle~{}~{}{\rm s.t.} 0<RsCbRe.\displaystyle~{}~{}0<R_{s}\leq C_{b}-R_{e}^{*}. (11b)
Theorem 2

𝒯A(η)\mathcal{T}_{\rm A}(\eta) in (11) is a concave function of RsR_{s}, and its maximal value is achieved at

Rs={CbRe,ηγbPb,Rs,otherwise,\displaystyle R_{s}^{*}=\begin{cases}C_{b}-R_{e}^{*},&\eta\leq\frac{\gamma_{b}^{\circ}}{P_{b}},\\ R_{s}^{\circ},&\rm otherwise,\end{cases} (12)

where γb(12+14+π2N1,eπ2N+Reln21)\gamma_{b}^{\circ}\in\left(\sqrt{\frac{1}{2}+\sqrt{\frac{1}{4}+\frac{\pi}{2N}}}-1,e^{\sqrt{\frac{\pi}{2N}}+R_{e}^{*}\ln 2}-1\right) is the unique root γb>0\gamma_{b}>0 that satisfies CbπVb2N=ReC_{b}-\sqrt{\frac{\pi V_{b}}{2N}}=R_{e}^{*}, and RsR_{s}^{\circ} is the unique zero-crossing Rs<CbReR_{s}<C_{b}-R_{e}^{*} of the derivative

d𝒯A(η)dRs=1Q(CbRsReVb/N)RsN2πVbe(CbRsRe)22Vb/N.\frac{d\mathcal{T}_{\rm A}(\eta)}{dR_{s}}=1-Q\left(\frac{C_{b}-R_{s}-R_{e}^{*}}{\sqrt{V_{b}/N}}\right)-\frac{R_{s}\sqrt{N}}{\sqrt{2\pi V_{b}}}e^{-\frac{\left(C_{b}-R_{s}-R_{e}^{*}\right)^{2}}{2V_{b}/N}}. (13)
Proof 3

Please refer to Appendix -C.

Theorem 2 presents an optimal secrecy rate RsR_{s}^{*} that differs from the one for infinite blocklength with NN\rightarrow\infty, where in the latter employing the maximal achievable secrecy rate Rs=CbReR_{s}^{*}=C_{b}-R_{e}^{*} is always optimal for secrecy throughput improvement. The fundamental reason behind such difference lies in the decoding failure caused by finite blocklength. Specifically, when the quality of the main channel is poor (i.e., a small η\eta) or when a large rate redundancy ReR_{e}^{*} is required, e.g., due to a high average SNR of Eve or a stringent secrecy requirement, the successful decoding probability psp_{s} is initially small and decreases slowly with RsR_{s}. In this case, the secrecy throughput improvement is mainly bottlenecked by RsR_{s}, and hence it is necessary to choose the maximal secrecy rate Rs=CbReR_{s}^{*}=C_{b}-R_{e}^{*}. Otherwise, psp_{s} is initially large but drops rapidly with RsR_{s}, thus dramatically degrading the secrecy throughput. Therefore, a relatively small RsR_{s} is supposed to be chosen to strike a good balance between the decoding and throughput performance.

The optimal secrecy rate RsCbReR_{s}^{*}\leq C_{b}-R_{e}^{*} in (12) can be obtained efficiently using the Newton’s method, despite its implicit form. The following corollaries further give a closed-form asymptotically tight lower bound RsLR_{s}^{L} on RsR_{s}^{*} and provide useful insights into the behavior of RsR_{s}^{*}.

Corollary 1

The optimal secrecy rate RsR_{s}^{*} in (12) satisfies

RsRsLCbRe2VbNln(12+CbRe2πVb/N).R_{s}^{*}\geq R_{s}^{L}\triangleq C_{b}-R_{e}^{*}-\sqrt{\frac{2V_{b}}{N}\ln\left(\frac{1}{2}+\frac{C_{b}-R_{e}^{*}}{\sqrt{2\pi V_{b}/N}}\right)}. (14)
Proof 4

The result follows by finding a lower bound on d𝒯A(η)dRs\frac{d\mathcal{T}_{\rm A}(\eta)}{dR_{s}} in (13) applying the inequalities Q(x)12ex2/2Q(x)\leq\frac{1}{2}e^{-x^{2}/2} and RsCbReR_{s}\leq C_{b}-R_{e}^{*} and then setting the resultant lower bound to zero.

The term 2VbNln(12+CbRe2πVb/N)\sqrt{\frac{2V_{b}}{N}\ln\left(\frac{1}{2}+\frac{C_{b}-R_{e}^{*}}{\sqrt{2\pi V_{b}/N}}\right)} in (14) is interpreted as the secrecy rate loss arisen from finite blocklength. This term vanishes as NN\rightarrow\infty or ReCbπVb2NR_{e}^{*}\rightarrow C_{b}-\sqrt{\frac{\pi V_{b}}{2N}}, and accordingly RsR_{s}^{*} approaches CeReC_{e}-R_{e}^{*}. In this sense, the lower bound RsLR_{s}^{L} can be employed as a computational convenient alternative to the optimal RsR_{s}^{*}, particularly for the large blocklength scenarios.

Corollary 2

The optimal secrecy rate RsR_{s}^{*} monotonically increases with the channel gain η\eta.

Proof 5

It is proved that CbRsReVb\frac{C_{b}-R_{s}-R_{e}^{*}}{\sqrt{V_{b}}} in (13) increases with η\eta such that d𝒯A(η)dRs\frac{d\mathcal{T}_{\rm A}(\eta)}{dR_{s}} increases with η\eta. Then, using the derivative rule for implicit functions with d𝒯A(η)dRs=0\frac{d\mathcal{T}_{\rm A}(\eta)}{dR_{s}^{*}}=0 reaches dRsdη>0\frac{dR_{s}^{*}}{d\eta}>0.

Fig. 2 depicts secrecy throughput 𝒯A(η)\mathcal{T}_{\rm A}(\eta) versus secrecy rate RsR_{s} for different blocklength NN and channel gain η\eta. The concavity of 𝒯A(η)\mathcal{T}_{\rm A}(\eta) on RsR_{s} given by Theorem 2 is well verified. Specifically, 𝒯A(η)\mathcal{T}_{\rm A}(\eta) first increases and then decreases with RsR_{s}, and there exists an optimal RsR_{s}^{*} that maximizes 𝒯A(η)\mathcal{T}_{\rm A}(\eta). It is also found that 𝒯A(η)\mathcal{T}_{\rm A}(\eta) almost linearly increases with RsR_{s} at first, since the throughput loss due to decoding error is negligible. Note that the curves in the figure are cut in different points which represent different values of the maximal achievable secrecy rate RsmaxR_{s}^{\max} for different NN and η\eta, and it is obvious that RsmaxR_{s}^{\max} increases with NN and η\eta. As η\eta grows, 𝒯A(η)\mathcal{T}_{\rm A}(\eta) improves significantly and the corresponding optimal RsR_{s}^{*} increases, which validates Corollary 2. The underlying reason is that, when the main channel quality improves, choosing a larger RsR_{s} contributes more to improving 𝒯A(η)\mathcal{T}_{\rm A}(\eta) compared with increasing the successful decoding probability psp_{s} (by lowering RsR_{s}). In addition, as proved in Theorem 1, 𝒯A(η)\mathcal{T}_{\rm A}(\eta) increases with NN. It is also proved that the optimal RsR_{s}^{*} increases with NN as η\eta\rightarrow\infty. However, it is no longer true when η\eta is too small, e.g., η=3\eta=3 dB. This is because, for a low channel quality, the decoding performance becomes a key restricting factor on throughput improvement, and hence RsR_{s} should be decreased to ensure a large psp_{s} as NN increases. Moreover, the secrecy throughput obtained with the lower bound RsLR_{s}^{L} in Corollary 1 approaches closely the optimal one particularly when NN is sufficiently large, which demonstrates the usefulness of the lower bound.

Refer to caption
Figure 2: 𝒯A(η)\mathcal{T}_{\rm A}(\eta) vs. RsR_{s} for different NN and η\eta, with Pb=0P_{b}=0 dB, Γe=0\Gamma_{e}=0 dB, and δ=0.2\delta=0.2.

III-B Non-Adaptive Optimization Scheme

This section devises a non-adaptive optimization scheme where the parameters μ\mu, RsR_{s}, ReR_{e}, and nn are designed based on the statistical CSI of the main channel and remain unchanged during the transmission period. Such a non-adaptive scheme can be computed off-line, which significantly lowers the complexity compared with an adaptive one.

Since all the parameters are independent of the channel gain η\eta, the problem of maximizing the secrecy throughput in (5) can be recast as follows:

maxμ,Re,Rs,n𝒯N=Rsp¯ss.t.(6b)(6d),\max_{\mu,R_{e},R_{s},n}~{}\mathcal{T}_{\rm N}=R_{s}\bar{p}_{s}~{}~{}{\rm s.t.}~{}~{}\eqref{st_max_c1}-\eqref{st_max_c3}, (15)

where p¯s=Pbμpsfγb(γ)𝑑γ\bar{p}_{s}=\int_{P_{b}\mu}^{\infty}p_{s}f_{\gamma_{b}}(\gamma)d\gamma denotes the average successful decoding probability.

The above problem can be handled via similar steps for its adaptive counterpart in Sec. III-A. To begin with, in order to increase psp_{s} for a given RsR_{s}, a minimal rate redundancy ReR_{e} should be chosen while satisfying the secrecy constraint 𝒪eδ\mathcal{O}_{e}\leq\delta. Hence, the optimal ReR_{e}^{*} is given in (7). It can be inferred from (15) that a smaller transmission threshold μ\mu can produce a larger 𝒯N\mathcal{T}_{\rm N}. Nonetheless, Cblog2(1+Pbμ)Rs+ReC_{b}\geq\log_{2}(1+P_{b}\mu)\geq R_{s}+R_{e}^{*} must be ensured, since otherwise there would always exist a transmission initiated when η>μ\eta>\mu while violating the reliability constraint (6b). Consequently, the optimal μ\mu^{*} for a fixed RsR_{s} is given by

μ=2Rs+Re1Pb.\mu^{*}=\frac{2^{R_{s}+R_{e}^{*}}-1}{P_{b}}. (16)

Note that in order to support a constant secrecy rate RsR_{s}, the optimal on-off threshold μ\mu^{*} for the non-adaptive scheme is generally larger than that of the adaptive one as given in (9). On the other hand, the optimal μ\mu^{*} monotonically decreases with δ\delta and nn, which is similar to the adaptive case. That is to say, the transmission condition can be relaxed when facing a looser secrecy requirement or using a larger blocklength.

Substituting ReR_{e}^{*} and μ\mu^{*} into p¯s\bar{p}_{s} and invoking the approximation of QQ-function in (49) yields

p¯s\displaystyle\bar{p}_{s} =Pbμ[1Ξ(γb,n,Rs+Re)]fγb(γ)𝑑γ\displaystyle=\int_{P_{b}\mu^{*}}^{\infty}\left[1-\Xi(\gamma_{b},n,R_{s}+R_{e}^{*})\right]f_{\gamma_{b}}(\gamma)d\gamma
=(a)1γb(θb2)θb2τbu(12βθb(γθb2))fγb(γ)𝑑γ\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{=}}1-\mathcal{F}_{\gamma_{b}}(\theta_{b}^{2})\int_{\theta_{b}^{2}}^{\tau^{u}_{b}}\left(\frac{1}{2}-\frac{\beta}{\theta_{b}}(\gamma-\theta_{b}^{2})\right)f_{\gamma_{b}}(\gamma)d\gamma
=(b)112γb(θb2)βθbθb2τbuγb(γ)𝑑γ,\displaystyle\stackrel{{\scriptstyle\mathrm{(b)}}}{{=}}1-\frac{1}{2}\mathcal{F}_{\gamma_{b}}(\theta_{b}^{2})-\frac{\beta}{\theta_{b}}\int_{\theta_{b}^{2}}^{\tau^{u}_{b}}\mathcal{F}_{\gamma_{b}}(\gamma)d\gamma, (17)

where (a)\mathrm{(a)} is due to θb=Pbμ=2Rs+Re1\theta_{b}=\sqrt{P_{b}\mu^{*}}=\sqrt{2^{R_{s}+R_{e}^{*}}-1}, β=n2π\beta=\frac{\sqrt{n}}{2\pi}, and τbu=θb2+θb2β\tau^{u}_{b}=\theta_{b}^{2}+\frac{\theta_{b}}{2\beta}, and (b)\mathrm{(b)} stems from the use of partial integration. With (III-B), the problem of maximizing 𝒯N\mathcal{T}_{\rm N} over nn and RsR_{s} can be equivalently transformed as below:

maxβ,θb\displaystyle\max_{\beta,\theta_{b}} 𝒯N=[log2(1+θb2)Re]p¯s\displaystyle~{}~{}\mathcal{T}_{\rm N}=\left[\log_{2}(1+\theta_{b}^{2})-R_{e}^{*}\right]{\bar{p}_{s}} (18a)
s.t.\displaystyle~{}~{}{\rm s.t.} 12πβN2π,θb>2Re1.\displaystyle~{}~{}\frac{1}{2\pi}\leq\beta\leq\frac{\sqrt{N}}{2\pi},~{}\theta_{b}>\sqrt{2^{R_{e}^{*}}-1}. (18b)
Theorem 3

𝒯N\mathcal{T}_{\rm N} in (18) is a monotonically increasing function of β\beta or nn.

Proof 6

The result follows by proving that

d𝒯Ndβ\displaystyle\frac{d\mathcal{T}_{\rm N}}{d\beta} =dRedβp¯s+[log2(1+θb2)Re]dp¯sdβ\displaystyle=-\frac{dR_{e}^{*}}{d\beta}\bar{p}_{s}+\left[\log_{2}(1+\theta_{b}^{2})-R_{e}^{*}\right]\frac{d\bar{p}_{s}}{d\beta}
>(a)[log2(1+θb2)Re]dp¯sdβ>(b)0,\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{>}}\left[\log_{2}(1+\theta_{b}^{2})-R_{e}^{*}\right]\frac{d\bar{p}_{s}}{d\beta}\stackrel{{\scriptstyle\mathrm{(b)}}}{{>}}0, (19)

where (a)\mathrm{(a)} is due to dRedn<0\frac{dR_{e}^{*}}{dn}<0 from (51), and (b)\mathrm{(b)} follows from dp¯sdβ=1θbθb2τbu[γb(τbu)γb(γ)]𝑑γ>0\frac{d\bar{p}_{s}}{d\beta}=\frac{1}{\theta_{b}}\int_{\theta_{b}^{2}}^{\tau^{u}_{b}}\left[\mathcal{F}_{\gamma_{b}}(\tau^{u}_{b})-\mathcal{F}_{\gamma_{b}}(\gamma)\right]d\gamma>0 as γb(γ)\mathcal{F}_{\gamma_{b}}(\gamma) is an increasing function of γ\gamma.

Theorem 3 suggests that Alice should use the maximal blocklength to maximize the secrecy throughput for the non-adaptive scheme, regardless of other parameters, i.e., the globally optimal blocklength is n=Nn^{*}=N. More importantly, this conclusion holds for any distribution of γb\gamma_{b}.

Substituting the CDF γb(γ)=1eγ/Γb\mathcal{F}_{\gamma_{b}}(\gamma)=1-e^{-{\gamma}/{\Gamma_{b}}} into (III-B) yields

𝒯N=12[log2(1+θb2)Re][1+Y(θb)]eθb2Γb,\mathcal{T}_{\rm N}=\frac{1}{2}\left[\log_{2}(1+\theta_{b}^{2})-R_{e}^{*}\right]\left[1+Y(\theta_{b})\right]e^{-\frac{\theta_{b}^{2}}{\Gamma_{b}}}, (20)

where Y(θb)=2βΓbθb(1eθb2βΓb)>0Y(\theta_{b})=\frac{2\beta\Gamma_{b}}{\theta_{b}}(1-e^{-\frac{\theta_{b}}{2\beta\Gamma_{b}}})>0. The optimal θb\theta_{b}^{*} that maximizes 𝒯N\mathcal{T}_{\rm N} is provided below.

Theorem 4

𝒯N\mathcal{T}_{\rm N} in (20) is first-increasing-then-decreasing w.r.t. θb\theta_{b}; the optimal θb\theta_{b}^{*} maximizing 𝒯N\mathcal{T}_{\rm N} is the unique root θb>2Re1\theta_{b}>\sqrt{2^{R_{e}^{*}}-1} of G(θb)=0G(\theta_{b})=0, where G(θb)G(\theta_{b}) is a decreasing function of θb\theta_{b}:

G(θb)=1+Y(θb)ln2[log2(1+θb2)Re]1+θb2θbg(θb),G(\theta_{b})=\frac{1+Y(\theta_{b})}{\ln 2}-\left[\log_{2}(1+\theta_{b}^{2})-R_{e}^{*}\right]\frac{1+\theta_{b}^{2}}{\theta_{b}}g(\theta_{b}), (21)

with g(θb)=(12θb+14βΓb+θbΓb)Y(θb)+θbΓb12θbg(\theta_{b})=\left(\frac{1}{2\theta_{b}}+\frac{1}{4\beta\Gamma_{b}}+\frac{\theta_{b}}{\Gamma_{b}}\right)Y(\theta_{b})+\frac{\theta_{b}}{\Gamma_{b}}-\frac{1}{2\theta_{b}}.

Proof 7

Please refer to Appendix -D.

Based on Theorem 4, the optimal θb\theta_{b}^{*} or secrecy rate Rs=log2(1+(θb)2)ReR_{s}^{*}=\log_{2}(1+(\theta_{b}^{*})^{2})-R_{e}^{*} can be efficiently calculated using a bisection search with G(θb)=0G(\theta_{b})=0, and thus the maximal 𝒯N\mathcal{T}_{\rm N}^{*} can be obtained from (18). The following corollaries demonstrate the behavior of RsR_{s}^{*} w.r.t. to the average channel power gain σb2=ΓbPb\sigma_{b}^{2}=\frac{\Gamma_{b}}{P_{b}} and provide a closed-form approximation of RsR_{s}^{*} at the large σb2\sigma_{b}^{2} regime.

Corollary 3

The optimal RsR_{s}^{*} monotonically increases with σb2\sigma_{b}^{2}.

Proof 8

Following similar steps as the proof of Theorem 4, it can be verified that G(θb)G(\theta_{b}) in (21) increases with σb2\sigma_{b}^{2} such that dRsdσb2=G(θb)/σb2G(θb)/Rs>0\frac{dR_{s}^{*}}{d\sigma_{b}^{2}}=-\frac{\partial G(\theta_{b})/\partial\sigma_{b}^{2}}{\partial G(\theta_{b})/\partial R_{s}^{*}}>0, which completes the proof.

Corollary 3 suggests that a larger secrecy rate should be employed to boost the secrecy throughput when the quality of the main channel improves, despite the fact that it might deteriorate the decoding correctness at Bob.

Corollary 4

At the regime of σb2\sigma_{b}^{2}\rightarrow\infty, the optimal secrecy rate RsR_{s}^{*} is approximated by

RsRsA\displaystyle R_{s}^{*}\approx R_{s}^{A} =log2(e)𝒲0(σb22Re)\displaystyle={\log_{2}(e)}{\mathcal{W}_{0}\left(\sigma_{b}^{2}2^{-R_{e}^{*}}\right)}
log2(σb2)Relog2[ln(σb2)Reln2],\displaystyle\approx\log_{2}(\sigma_{b}^{2})-R_{e}^{*}-\log_{2}\left[\ln(\sigma_{b}^{2})-R_{e}^{*}\ln 2\right], (22)

where 𝒲0(x)\mathcal{W}_{0}(x) is the Lambert’s WW function [39, Sec. 4.13] that satisfies x=𝒲0(x)e𝒲0(x)x=\mathcal{W}_{0}(x)e^{\mathcal{W}_{0}(x)}.

Proof 9

It is clear that Y(θb)1Y(\theta_{b})\rightarrow 1 and g(θb)2θbΓbg(\theta_{b})\rightarrow\frac{2\theta_{b}}{\Gamma_{b}} as σb2\sigma_{b}^{2}\rightarrow\infty. Substituting the results into (21) with θb2=2Rs+Re1\theta_{b}^{2}=2^{R_{s}+R_{e}^{*}}-1 and letting G(θb)=0G(\theta_{b})=0 produce the first approximation. The second approximation comes from the expansion of 𝒲0(x)\mathcal{W}_{0}(x) as xx\rightarrow\infty that 𝒲0(x)lnxln(lnx)\mathcal{W}_{0}(x)\approx\ln x-\ln(\ln x).

Fig. 3 plots the secrecy throughput 𝒯N\mathcal{T}_{\rm N} versus the secrecy rate RsR_{s} for different values of the blocklength NN and the average channel gain σb2\sigma_{b}^{2}. It can be seen that 𝒯N\mathcal{T}_{\rm N} first increases and then decreases with RsR_{s}, which validates Theorem 4. The optimal RsR_{s}^{*} maximizing 𝒯N\mathcal{T}_{\rm N} increases with σb2\sigma_{b}^{2}, which verifies Corollary 3 well, and the reason behind is similar to that for Corollary 2. It can also be observed that the optimal RsR_{s}^{*} is almost impervious to different NN. This is because, the optimal secrecy rate for the non-adaptive scheme only depends on the average successful decoding probability, and the averaging process softens the impact of the blocklength. Theorem 3 is also confirmed, where it is found that 𝒯N\mathcal{T}_{\rm N} increases with NN. In addition, the secrecy throughput with the approximate RsAR_{s}^{A} obtained in Corollary 4 is almost coincided with that of the optimal RsR_{s}^{*}, which demonstrates the practicability of the low-complexity approximation.

Refer to caption
Figure 3: 𝒯N\mathcal{T}_{\rm N} vs. RsR_{s} for different NN and σb2\sigma_{b}^{2}, with Pb=0P_{b}=0 dB, Γe=0\Gamma_{e}=0 dB, and δ=0.2\delta=0.2.

Fig. 4 compares the secrecy throughput for adaptive and non-adaptive schemes with different blocklength NN. The left-hand-side figure depicts the maximal secrecy throughput 𝒯\mathcal{T}^{*}, where 𝒯A\mathcal{T}_{\rm A}^{*} for the adaptive case improves as NN increases whereas 𝒯N\mathcal{T}_{\rm N}^{*} for the non-adaptive case almost remains unchanged. When the average channel gain σb2\sigma_{b}^{2} increases or the secrecy constraint becomes relaxed (i.e., a larger δ\delta), the maximal 𝒯\mathcal{T}^{*} for both schemes improves significantly, and the gap 𝒯A𝒯N\mathcal{T}_{\rm A}^{*}-\mathcal{T}_{\rm N}^{*} increases. The right-hand-side figure illustrates the relative throughput gain Δ𝒯𝒯A𝒯N𝒯N\Delta\mathcal{T}\triangleq\frac{\mathcal{T}_{\rm A}^{*}-\mathcal{T}_{\rm N}^{*}}{\mathcal{T}_{\rm N}^{*}} which reflects the superiority of the adaptive scheme over its non-adaptive counterpart. It is shown that Δ𝒯\Delta\mathcal{T} grows dramatically with NN but decreases with σb2\sigma_{b}^{2} and δ\delta. This suggests that the adaptive scheme is more preferred for some unfavorable scenarios, e.g., with a large blocklengh (large delay), a poor channel quality, or a stringent secrecy requirement; otherwise, the non-adaptive scheme could be an alternative choice owing to its low implementation complexity.

Refer to caption
Figure 4: 𝒯\mathcal{T}^{*} and Δ𝒯\Delta\mathcal{T} vs. NN for different σb2\sigma_{b}^{2} and δ\delta, with Pb=0P_{b}=0 dB and Γe=0\Gamma_{e}=0 dB.

IV Multi-Antenna Transmitter Scenario

When Alice is equipped with multiple antennas, she can intentionally transmit AN together with the information-bearing signal to degrade Eve’s channel quality. Generally, the null-space AN scheme, in which the AN is injected uniformly in directions orthogonal to the main channel, is heuristically employed in the context of infinite blocklength [34]. The near-optimality of AN in terms of improving secrecy capacity for the multi-input single-output wiretap channel was first proved in [35] from a rigorous information-theoretic perspective, and its degraded performance was later observed for the multi-input multi-output wiretap channel [36]. On the other hand, it was argued in [37] that distributing a certain proportion of AN in the direction of main channel can surprisingly gain a larger ergodic secrecy rate. When it comes to finite blocklength, since decoding failure might occur even when the codeword rate lies below the channel capacity, which is quite different from the infinite blocklength case, it is still unclear whether the null-space AN is optimal and how the optimal power allocation of the AN scheme should be determined for maximizing the secrecy throughput. To this end, this section focuses on the optimization of secrecy throughput with finite blocklength for the multi-antenna scenario, where the optimality of the null-space AN scheme will be identified first.

Considering a general scenario where the AN is not restricted to be orthogonal to the main channel, Alice’s transmitted signal can be constructed in the form of

𝒙=ϕP𝒘(αs+1αv)+(1ϕ)PM1𝑾𝒛,\bm{x}=\sqrt{\phi P}{\bm{w}}\left(\sqrt{\alpha}s+\sqrt{1-\alpha}v\right)+\sqrt{\frac{(1-\phi)P}{M-1}}{\bm{W}}_{\bot}{\bm{z}}, (23)

where 𝒘=𝒉b𝒉b{\bm{w}}=\frac{\bm{h}_{b}^{{\dagger}}}{\|\bm{h}_{b}\|} denotes the beamforming vector for the main channel, 𝑾{\bm{W}}_{\bot} denotes the M×(M1)M\times(M-1) projection matrix onto the null space of 𝒉b\bm{h}_{b} such that 𝒉bT𝑾=𝟎{\bm{h}_{b}^{\mathrm{T}}}{\bm{W}}_{\bot}=\bm{0}, and the columns of [𝒘𝑾][{\bm{w}}~{}{\bm{W}}_{\bot}] constitute an orthogonal basis; ss, vv, and 𝒛{\bm{z}} denote the information signal, the AN in the direction of 𝒘{\bm{w}}, and the AN in the null space 𝑾{\bm{W}}_{\bot}, with each element obeying 𝒞𝒩(0,1)\mathcal{CN}(0,1); ϕ[0,1]\phi\in[0,1] represents the fraction of the total transmit power PP allocated to the direction of 𝒘{\bm{w}}, and α[0,1]\alpha\in[0,1] represents the power allocation ratio of the information signal to ϕP\phi P. With (23), the received signal-to-interference-plus-noise ratios (SINRs) at Bob and Eve are respectively

γb\displaystyle\gamma_{b} =αϕPbη(1α)ϕPbη+1,\displaystyle=\frac{\alpha\phi P_{b}\eta}{(1-\alpha)\phi P_{b}\eta+1}, (24)
γe\displaystyle\gamma_{e} =αϕPe|𝒉eT𝒘|2(1α)ϕPe|𝒉eT𝒘|2+(1ϕ)Pe𝒉eT𝑾2M1+1,\displaystyle=\frac{\alpha\phi P_{e}|\bm{h}_{e}^{\mathrm{T}}\bm{w}|^{2}}{(1-\alpha)\phi P_{e}|\bm{h}_{e}^{\mathrm{T}}\bm{w}|^{2}+\frac{(1-\phi)P_{e}\|\bm{h}_{e}^{\mathrm{T}}\bm{W}_{\bot}\|^{2}}{M-1}+1}, (25)

where η=𝒉b2\eta=\|\bm{h}_{b}\|^{2}. The successful decoding probability psp_{s} and the information leakage probability 𝒪e\mathcal{O}_{e} for the multi-antenna case are still given by (2) and (4), respectively. The corresponding secrecy throughput optimization problem can be formulated as below:

maxμ,Re,Rs,n,α,ϕ𝒯=𝔼η[Rsps]s.t.(6b)(6d),0α,ϕ1.\displaystyle\max_{\mu,R_{e},R_{s},n,\alpha,\phi}~{}\mathcal{T}=\mathbb{E}_{\eta}\left[R_{s}p_{s}\right]~{}~{}{\rm s.t.}~{}~{}\eqref{st_max_c1}-\eqref{st_max_c3},~{}0\leq\alpha,\phi\leq 1. (26)

The following subsections will first detail the optimization procedure for both adaptive and non-adaptive schemes, and then briefly discuss the scenario of a multi-antenna Eve.

IV-A Adaptive Optimization Scheme

This subsection optimizes the secrecy throughput 𝒯A\mathcal{T}_{\rm A} by designing the parameters involved in problem (26) adaptively according to the instantaneous channel realization 𝒉b\bm{h}_{b}.

IV-A1 Solving ReR_{e}

Similar to the single-antenna case, the optimal rate redundancy is given by Re=𝒪e1(δ)R_{e}^{*}=\mathcal{O}_{e}^{-1}(\delta) with 𝒪e\mathcal{O}_{e} in (4). Note that ReR_{e}^{*} herein is a function of ϕ\phi and α\alpha.

IV-A2 Solving α\alpha

Resort to a function κ(x,α)xαx(1α)+1\kappa(x,\alpha)\triangleq\frac{x\alpha}{x(1-\alpha)+1} defined in [38], which increases with xx for α>0\alpha>0. Then, the SINRs γb\gamma_{b} in (24) and γe\gamma_{e} in (25) can be reformulated as γb(ϕ,α)=κ(γb(ϕ,1),α)\gamma_{b}(\phi,\alpha)=\kappa\left(\gamma_{b}(\phi,1),\alpha\right) and γe(ϕ,α)=κ(γe(ϕ,1),α)\gamma_{e}(\phi,\alpha)=\kappa\left(\gamma_{e}(\phi,1),\alpha\right). Define Φe(ϕ,α)2Re1\Phi_{e}(\phi,\alpha)\triangleq 2^{R_{e}^{*}}-1 as the SINR threshold for γe(ϕ,α)\gamma_{e}(\phi,\alpha) such that Re=log2(1+Φe(ϕ,α))R_{e}^{*}=\log_{2}(1+\Phi_{e}(\phi,\alpha)). Recalling the secrecy constraint 𝒪e(Φe;θ,α)=δ\mathcal{O}_{e}(\Phi_{e};\theta,\alpha)=\delta, Φe(ϕ,α)\Phi_{e}(\phi,\alpha) is the δ\delta-upper quantile of γe(ϕ,α)\gamma_{e}(\phi,\alpha) such that it also follows the form Φe(ϕ,α)=κ(Φe(ϕ,1),α)\Phi_{e}(\phi,\alpha)=\kappa(\Phi_{e}(\phi,1),\alpha) [38]. Hence, the condition for guaranteeing a positive secrecy rate is described as

γb(ϕ,α)>Φe(ϕ,α)\displaystyle\gamma_{b}(\phi,\alpha)>\Phi_{e}(\phi,\alpha) κ(γb(ϕ,1),α)>κ(Φe(ϕ,1),α)\displaystyle\Rightarrow\kappa(\gamma_{b}(\phi,1),\alpha)>\kappa(\Phi_{e}(\phi,1),\alpha)
γb(ϕ,1)>Φe(ϕ,1)\displaystyle\Rightarrow\gamma_{b}(\phi,1)>\Phi_{e}(\phi,1)
(a)ρb>ρe(ϕ),\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{\Rightarrow}}\rho_{b}>\rho_{e}(\phi), (27)

where ρbPbη\rho_{b}\triangleq P_{b}\eta, ρe(ϕ)Φe(ϕ,1)ϕ\rho_{e}(\phi)\triangleq\frac{\Phi_{e}(\phi,1)}{\phi}, and (a)\mathrm{(a)} is due to γb(ϕ,1)=ϕPη\gamma_{b}(\phi,1)=\phi P\eta. Then, the threshold μ\mu can be simply set as μ(ϕ)=ρe(ϕ)Pb\mu(\phi)=\frac{\rho_{e}(\phi)}{P_{b}} for any fixed ϕ\phi. Revisiting (8), since μ(ϕ)\mu(\phi) is independent of α\alpha, the optimal α\alpha^{*} that maximizes 𝒯A\mathcal{T}_{\rm A} can be obtained by maximizing 𝒯A(η)=Rsps\mathcal{T}_{\rm A}(\eta)=R_{s}p_{s}, where psp_{s} is defined in (3) and can be rewritten as

ps=1Q(nλblnλblnλeRsln2λb21),p_{s}=1-Q\left(\sqrt{n}\lambda_{b}\frac{\ln\lambda_{b}-\ln\lambda_{e}-R_{s}\ln 2}{\sqrt{\lambda_{b}^{2}-1}}\right), (28)

with λb1+κ(γb(ϕ,1),α)>λe1+κ(Φe(ϕ,1),α)>1\lambda_{b}\triangleq 1+\kappa(\gamma_{b}(\phi,1),\alpha)>\lambda_{e}\triangleq 1+\kappa(\Phi_{e}(\phi,1),\alpha)>1. Although it is difficult to see how psp_{s} varies with α\alpha for a fixed Rs<log2λbλeR_{s}<\log_{2}\frac{\lambda_{b}}{\lambda_{e}} as both λb\lambda_{b} and λe\lambda_{e} increase with α\alpha, the following theorem provides the optimal α\alpha^{*} that maximizes 𝒯A\mathcal{T}_{\rm A}.

Theorem 5

α=1\alpha^{*}=1 is optimal for maximizing the secrecy throughput 𝒯A\mathcal{T}_{\rm A}.

Proof 10

Please refer to Appendix -E.

Theorem 5 suggests that there is no need to inject the AN in the main channel direction for secrecy throughput improvement with finite blocklength. The reason is that, once the main channel quality suffices to guarantee λb>λe\lambda_{b}>\lambda_{e}, a larger α\alpha can improve the term lnλblnλe1λb2\frac{\ln\lambda_{b}-\ln\lambda_{e}}{\sqrt{1-\lambda_{b}^{-2}}} in (28) which reflects the channel superiority of the main channel over the wiretap channel.

Define ξϕ11M1\xi\triangleq\frac{\phi^{-1}-1}{M-1}. Substituting α=1\alpha^{*}=1 into (24) and (25) yields the CDFs of γb\gamma_{b} and γe\gamma_{e}:

γb(γ)\displaystyle\mathcal{F}_{\gamma_{b}}(\gamma) =1eγϕΓbk=0M11k!(γϕΓb)k,\displaystyle=1-e^{-\frac{\gamma}{\phi\Gamma_{b}}}\sum_{k=0}^{M-1}\frac{1}{k!}\left(\frac{\gamma}{\phi\Gamma_{b}}\right)^{k}, (29)
γe(γ)\displaystyle\mathcal{F}_{\gamma_{e}}(\gamma) =1eγϕΓe(1+ξγ)1M,\displaystyle=1-e^{-\frac{\gamma}{\phi\Gamma_{e}}}\left(1+\xi\gamma\right)^{1-M}, (30)

IV-A3 Solving μ\mu

The threshold μ(ϕ)=ρe(ϕ)Pb\mu(\phi)=\frac{\rho_{e}(\phi)}{P_{b}} mentioned in the last step is related to ϕ\phi. This step further determines the optimal μ\mu^{*} which is independent of ϕ\phi and η\eta. For tractability, consider an asymptotically large blocklength and exploit the tail property of the QQ-function, then the information leakage probability 𝒪e\mathcal{O}_{e} is approximated as [33]

𝒪e(Φe)eΦeϕΓe(1+ξΦe)1M.\mathcal{O}_{e}(\Phi_{e})\approx e^{-\frac{\Phi_{e}}{\phi\Gamma_{e}}}\left(1+\xi\Phi_{e}\right)^{1-M}. (31)

Fig. 5 shows that the approximate 𝒪e(Φe)\mathcal{O}_{e}(\Phi_{e}) is extremely close to the exact value for quite a wide range of ϕ\phi, MM, nn, and Γe\Gamma_{e}, and it then can be adopted to facilitate the subsequent analysis and optimization. Revisiting ρe(ϕ)=Φe(ϕ)ϕ\rho_{e}(\phi)=\frac{\Phi_{e}(\phi)}{\phi} with 𝒪e(Φe(ϕ))=δ\mathcal{O}_{e}(\Phi_{e}(\phi))=\delta, the following lemma is obtained.

Refer to caption
Figure 5: 𝒪e\mathcal{O}_{e} vs. ϕ\phi for different MM, nn, and Γe\Gamma_{e}.
Lemma 2 ([38])

ρe(ϕ)>0\rho_{e}(\phi)>0, dρe(ϕ)dϕ=ρe(ϕ)[1+ϕρe(ϕ)ξ]/Γe+1ϕ>0\frac{d\rho_{e}(\phi)}{d\phi}=\frac{\rho_{e}(\phi)}{\left[1+\phi\rho_{e}(\phi)\xi\right]/\Gamma_{e}+1-\phi}>0, and d2ρe(ϕ)dϕ2>2ρe(ϕ)[dρe(ϕ)dϕ]2>0\frac{d^{2}\rho_{e}(\phi)}{d\phi^{2}}>\frac{2}{\rho_{e}(\phi)}\left[\frac{d\rho_{e}(\phi)}{d\phi}\right]^{2}>0.

Lemma 2 indicates that ρe(ϕ)\rho_{e}(\phi) increases with ϕ\phi. It is observed from (IV-A2) that no positive RsR_{s} can be achieved if Pbηρe(0)P_{b}\eta\leq\rho_{e}(0). To avoid this, the optimal on-off threshold should be chosen as

μ=ρe(0)Pb.\mu^{*}=\frac{\rho_{e}(0)}{P_{b}}. (32)

IV-A4 Solving nn

This step gives the optimal blocklength nn^{*} that maximizes secrecy throughput.

Theorem 6

n=Nn^{*}=N is optimal for maximizing 𝒯A\mathcal{T}_{\rm A} in (26).

Proof 11

The proof is similar to that of Theorem 1. According to (53), one only needs to prove that fγe(τel)>fγe(τeu)f_{\gamma_{e}}(\tau^{l}_{e})>f_{\gamma_{e}}(\tau^{u}_{e}). The PDF of γe\gamma_{e} is calculated from (30) and is given by

fγe(γ)=(1ϕΓe(1+ξγ)M1+ξ(M1)(1+ξγ)M)eγϕΓe.f_{\gamma_{e}}(\gamma)=\left(\frac{1}{\phi\Gamma_{e}\left(1+\xi\gamma\right)^{M-1}}+\frac{\xi(M-1)}{\left(1+\xi\gamma\right)^{M}}\right)e^{-\frac{\gamma}{\phi\Gamma_{e}}}. (33)

Apparently, fγe(γ)f_{\gamma_{e}}(\gamma) decreases with γ\gamma such that fγe(τel)>fγe(τeu)f_{\gamma_{e}}(\tau^{l}_{e})>f_{\gamma_{e}}(\tau^{u}_{e}), which completes the proof.

Theorem 6 suggests that a multi-antenna transmitter also should adopt the maximal blocklength to maximize the secrecy throughput, regardless of the power allocation and code rates. This is validated by Fig. 6, and the reason behind is similar to that of the single-antenna case.

IV-A5 Solving ϕ\phi

By now, the secrecy throughput 𝒯A(η)\mathcal{T}_{\rm A}(\eta) conditioned on η\eta is given by

𝒯A(η)=Rs(1Q[NλblnλbλeRsln2λb21]),\mathcal{T}_{\rm A}(\eta)=R_{s}\left(1-Q\left[\sqrt{N}\lambda_{b}\frac{\ln\frac{\lambda_{b}}{\lambda_{e}}-R_{s}\ln 2}{\sqrt{\lambda_{b}^{2}-1}}\right]\right), (34)

where λb=1+ϕρb\lambda_{b}=1+\phi\rho_{b} and λe=1+ϕρe\lambda_{e}=1+\phi\rho_{e} with ρb>ρe\rho_{b}>\rho_{e}. For notational simplicity, ϕ\phi has been dropped from ρe(ϕ)\rho_{e}(\phi). Obviously, maximizing 𝒯A(η)\mathcal{T}_{\rm A}(\eta) is equivalent to maximizing the following function:

L(ϕ)=λbλb21(lnλbλeRsln2).L(\phi)=\frac{\lambda_{b}}{\sqrt{\lambda_{b}^{2}-1}}\left({\ln\frac{\lambda_{b}}{\lambda_{e}}-R_{s}\ln 2}\right). (35)
Theorem 7

L(ϕ)L(\phi) in (35) is a concave function of ϕ\phi, and the optimal ϕ\phi^{*} maximizing L(ϕ)L(\phi) is

ϕ={1,ηρbPbandρe(1)1+ρe(1)<11+Γe,ϕ,otherwise.f\displaystyle\phi^{*}=\begin{cases}1,&\eta\geq\frac{\rho_{b}^{\circ}}{P_{b}}~{}{\rm and}~{}\frac{\rho_{e}(1)}{1+\rho_{e}(1)}<\frac{1}{1+\Gamma_{e}},\\ \phi^{\circ},&\rm otherwise.\end{cases}f (36)

Here ϕ\phi^{\circ} is the unique zero-crossing ϕ[0,1)\phi\in[0,1) of the following derivative:

dL(ϕ)dϕ=(1Aϕ)λb1ϕλb21(λb1)(lnλbBϕ)ϕ(λb21)3/2,\frac{dL(\phi)}{d\phi}=\frac{(1-A_{\phi}){\lambda_{b}}-1}{\phi\sqrt{\lambda_{b}^{2}-1}}-\frac{({\lambda_{b}-1})\left({\ln{\lambda_{b}}-B_{\phi}}\right)}{\phi\left(\lambda_{b}^{2}-1\right)^{3/2}}, (37)

where Aϕϕλe(ρe+ϕdρedϕ)A_{\phi}\triangleq\frac{\phi}{\lambda_{e}}\left(\rho_{e}+\phi\frac{d\rho_{e}}{d\phi}\right) and Bϕlnλe+Rsln2B_{\phi}\triangleq\ln\lambda_{e}+R_{s}\ln 2 with dρedϕ\frac{d\rho_{e}}{d\phi} given in Lemma 2, and ρb\rho_{b}^{\circ} is the unique root ρb\rho_{b} of the equation X(ρb)=0X(\rho_{b})=0 with X(ρb)X(\rho_{b}) given below:

X(ρb)=(1A1)(1+ρb)1ln(1+ρb)B12+ρb.X(\rho_{b})=(1-A_{1})(1+\rho_{b})-1-\frac{\ln(1+\rho_{b})-B_{1}}{2+\rho_{b}}. (38)
Proof 12

Please refer to Appendix -F.

Theorem 7 shows that, the naive beamforming scheme without injecting any AN is optimal for maximizing the secrecy throughput only when the quality of the main channel is good enough and meanwhile the quality of the wiretap channel is poor or a high information leakage probability is acceptable. Using the derivative rule for implicit functions with (37) proves that dϕdRs>0\frac{d\phi^{*}}{dR_{s}}>0, which suggests that in order to support a higher secrecy rate, a larger fraction of power should be allocated to the information signal although at the cost of a larger required rate redundancy.

For a robust design perspective, a worst-case scenario is considered by ignoring Eve’s thermal noise, i.e., Γe\Gamma_{e}\rightarrow\infty in (31), such that ρe=Λ1ϕ\rho_{e}=\frac{\Lambda}{1-\phi} with Λ=(M1)(δ11M1)\Lambda=(M-1)(\delta^{\frac{1}{1-M}}-1). It is seen from (37) that ϕ\phi^{*} is a function of η\eta and δ\delta, and the monotonicity of ϕ\phi^{*} is revealed as below.

Corollary 5

For the worst case Γe\Gamma_{e}\rightarrow\infty, the optimal power allocation ϕ\phi^{*} is non-decreasing w.r.t. η\eta and δ\delta. Moreover, limηϕ=1Λ+1\lim_{\eta\rightarrow\infty}\phi^{*}=\frac{1}{\sqrt{\Lambda}+1} and limδ1ϕ=1\lim_{\delta\rightarrow 1}\phi^{*}=1.

Proof 13

Please refer to Appendix -G.

Corollary 5 suggests that when the quality of the main channel improves (i.e., a larger η\eta) or the secrecy requirement is relaxed (i.e., a larger δ\delta), it would be more appealing to use a higher signal power to promote the main channel than to increase the AN power to degrade the wiretap channel. This is because that the main channel becomes the dominate factor to the improvement of secrecy throughput. Different from Theorem 7 where ϕ=1\phi^{*}=1 can be achieved, the optimal ϕ\phi^{*} here only can be increased up to 1Λ+1\frac{1}{\sqrt{\Lambda}+1} as η\eta\rightarrow\infty due to Eve’s background noise being ignored. Besides, it is unsurprising that ϕ=1\phi^{*}=1 for δ=1\delta=1 since there is no secrecy requirement.

IV-A6 Solving RsR_{s}

For any given power allocation ϕ\phi^{*}, it can be proved that the secrecy throughput 𝒯A(η)\mathcal{T}_{\rm A}(\eta) is a concave function of the secrecy rate RsR_{s} as done in Theorem 2. Hence, the optimal RsR_{s}^{*} maximizing 𝒯A(η)\mathcal{T}_{\rm A}(\eta) is given by (12) and a closed-form lower bound on RsR_{s}^{*} can be found in (14). Eventually, problem (26) can be addressed via an alternating optimization (AO) method, which is summarized in Algorithm 1. In addition, at the high η\eta regime, the optimal ϕ\phi^{*} is independent of RsR_{s}, and hence a global optimal pair (ϕ,Rs)(\phi^{*},R_{s}^{*}) is obtained for maximizing 𝒯A(η)\mathcal{T}_{\rm A}(\eta).

Algorithm 1 AO Algorithm for Solving Problem (26)
1:  Initialize k=1k=1, ϕ(0)[0,1]\phi^{(0)}\in[0,1], Rs(0)0R_{s}^{(0)}\geq 0, and assign ϵ\epsilon a sufficiently small positive value, e.g., ϵ=1010\epsilon=10^{-10};
2:  Input δ[0,1]\delta\in[0,1], N1N\geq 1, and Pb,Γe,η=𝒉b2>0P_{b},\Gamma_{e},\eta=\|\bm{h}_{b}\|^{2}>0;
3:  Calculate μ\mu from (32) and 𝒯A(0)(η)=Rs(0)ps(0)\mathcal{T}_{\rm A}^{(0)}(\eta)=R_{s}^{(0)}p_{s}^{(0)};
4:  if η<μ\eta<\mu then
5:     𝒯A(k)(η)0\mathcal{T}_{\rm A}^{(k)}(\eta)\leftarrow 0;
6:  else
7:      Update ϕ(k)ϕ(k1)\phi^{(k)}\leftarrow\phi^{(k-1)}, Rs(k)Rs(k1)R_{s}^{(k)}\leftarrow R_{s}^{(k-1)};
8:     Calculate ρb\rho_{b}^{\circ} from (38);
9:     if PbηρbP_{b}\eta\geq\rho_{b}^{\circ} then
10:        ϕ(k)1\phi^{(k)}\leftarrow 1;
11:     else
12:        Calculate ϕ(k)\phi^{(k)} from (37);
13:     end if
14:     Calculate Rs(k)R_{s}^{(k)} from (12);
15:      Update 𝒯A(k)(η)Rs(k)ps(k)\mathcal{T}_{\rm A}^{(k)}(\eta)\leftarrow R_{s}^{(k)}p_{s}^{(k)};
16:     while |[𝒯A(k)(η)𝒯A(k1)(η)]/𝒯A(k1)(η)|ϵ\left|\left[\mathcal{T}_{\rm A}^{(k)}(\eta)-\mathcal{T}_{\rm A}^{(k-1)}(\eta)\right]/{\mathcal{T}_{\rm A}^{(k-1)}(\eta)}\right|\geq\epsilon do
17:        Update kk+1k\leftarrow k+1;
18:        Repeat step 7 to step 15;
19:     end while
20:  end if
21:  Output 𝒯A(k)(η)\mathcal{T}_{\rm A}^{(k)}(\eta)

Fig. 6 illustrates the optimal power allocation ϕ\phi^{*} and the corresponding maximal secrecy throughput 𝒯A(η)\mathcal{T}_{\rm A}(\eta) for varying secrecy rate RsR_{s}. The maximal 𝒯A(η)\mathcal{T}_{\rm A}(\eta) is concave on RsR_{s}, which guarantees the global optimality of the solution and the convergence of the proposed AO algorithm. The optimal ϕ\phi^{*} increases with η\eta and RsR_{s}, which verifies Corollary 5. In addition, the curves of ϕ\phi^{*} and 𝒯A(η)\mathcal{T}_{\rm A}(\eta) are truncated after RsR_{s} exceeds some critical values. This can be explained similarly as that of Fig. 2. It is shown that ϕ\phi^{*} increases with the blocklength NN, although slightly. This is because, increasing NN will mildly decrease the information leakage probability 𝒪e\mathcal{O}_{e}, thus allowing a larger portion of power to be devoted to transmitting the information-bearing signal.

Refer to caption
Figure 6: Optimal ϕ\phi^{*} and 𝒯A(η)\mathcal{T}_{\rm A}(\eta) vs. RsR_{s} for different NN and η\eta, with M=4M=4, Pb=0P_{b}=0 dB, Γe=0\Gamma_{e}=0 dB, and δ=0.2\delta=0.2.

IV-B Non-Adaptive Optimization Scheme

This subsection examines the secrecy throughput maximization through a non-adaptive design manner for the multi-antenna transmitter case. The problem can be formulated as

maxμ,Re,Rs,n,α,ϕ\displaystyle\max_{\mu,R_{e},R_{s},n,\alpha,\phi} 𝒯N=Rsp¯s\displaystyle~{}\mathcal{T}_{\rm N}=R_{s}\bar{p}_{s} (39a)
s.t.\displaystyle~{}~{}{\rm s.t.}~{} (6b)(6d),0α,ϕ1,\displaystyle~{}\eqref{st_max_c1}-\eqref{st_max_c3},~{}0\leq\alpha,\phi\leq 1, (39b)

where p¯s\bar{p}_{s} is the average successful decoding probability.

The basic idea to solve problem (39) is similar to that of problem (15). Again, the optimal rate redundancy is Re=𝒪e1(δ)R_{e}^{*}=\mathcal{O}_{e}^{-1}(\delta) with 𝒪e\mathcal{O}_{e} given in (4). For the adaptive case, it is known from (IV-A2) that κ(γb(ϕ,1),α)>κ(Φe(ϕ,1),α)η>μ(ϕ)=ρe(ϕ)Pb\kappa(\gamma_{b}(\phi,1),\alpha)>\kappa(\Phi_{e}(\phi,1),\alpha)\Rightarrow\eta>\mu(\phi)=\frac{\rho_{e}(\phi)}{P_{b}} suffices to guarantee a positive secrecy rate RsR_{s} with the threshold μ(ϕ)\mu(\phi) independent of α\alpha, and then α=1\alpha^{*}=1 is optimal for secrecy throughput maximization. As for the non-adaptive one, supporting a certain secrecy rate RsR_{s} requires that 1+κ(γb(ϕ,1),α)>2Rs(1+κ(Φe(ϕ,1),α))1+\kappa(\gamma_{b}(\phi,1),\alpha)>2^{R_{s}}(1+\kappa(\Phi_{e}(\phi,1),\alpha)) which is further transformed to

η>μ(ϕ)=1ϕPb1α2Rs(1+κ(Φe(ϕ,1),α))11+α.\eta>\mu(\phi)=\frac{1}{\phi P_{b}}\frac{1}{\frac{\alpha}{2^{R_{s}}(1+\kappa(\Phi_{e}^{*}(\phi,1),\alpha))-1}-1+\alpha}. (40)

Although μ(ϕ)\mu(\phi) herein depends on α\alpha, it is proved that μ(ϕ)\mu(\phi) monotonically decreases with α\alpha. Hence, α=1\alpha^{*}=1 is still throughput-optimal for the non-adaptive case. Accordingly, the optimal threshold is μ=2Rs(1+ϕρe)ϕPb\mu^{*}=\frac{2^{R_{s}}(1+\phi\rho_{e})}{\phi P_{b}}. Similar to the proof of Theorem 3, using the maximal blocklength is optimal for maximizing secrecy throughput, regardless of the distribution of γb\gamma_{b}. Hence, the optimal blocklength is n=Nn^{*}=N. Afterwards, the secrecy throughput is calculated from (III-B):

𝒯N\displaystyle\mathcal{T}_{\rm N} =Rsp¯s=Rs[112γb(θb2)βθbθb2τbuγb(γ)𝑑γ]\displaystyle=R_{s}\bar{p}_{s}=R_{s}\left[1-\frac{1}{2}\mathcal{F}_{\gamma_{b}}(\theta_{b}^{2})-\frac{\beta}{\theta_{b}}\int_{\theta_{b}^{2}}^{\tau^{u}_{b}}\mathcal{F}_{\gamma_{b}}(\gamma)d\gamma\right]
=(a)Rs[Γ¯(M,ϱ1)2+ϕΓbβθbΔΓ],\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{=}}R_{s}\left[\frac{\bar{\Gamma}(M,\varrho_{1})}{2}+\frac{\phi\Gamma_{b}\beta}{\theta_{b}}\Delta\Gamma\right], (41)

where (a)\mathrm{(a)} holds by invoking the CDF γb(γ)\mathcal{F}_{\gamma_{b}}(\gamma) of γb\gamma_{b} in (29) and computing the integral, with ΔΓk=0M1[Γ¯(k+1,ϱ1)Γ¯(k+1,ϱ2])\Delta\Gamma\triangleq\sum_{k=0}^{M-1}\left[\bar{\Gamma}(k+1,\varrho_{1})-\bar{\Gamma}(k+1,\varrho_{2}\right]) and Γ¯(m+1,x)k=0mxkexk!\bar{\Gamma}(m+1,x)\triangleq\sum_{k=0}^{m}\frac{x^{k}e^{-x}}{k!} being the regularized upper incomplete gamma function, with ϱ1=θb2ϕΓb\varrho_{1}=\frac{\theta_{b}^{2}}{\phi\Gamma_{b}}, ϱ2=θb2ϕΓb+θb2βϕΓb\varrho_{2}=\frac{\theta_{b}^{2}}{\phi\Gamma_{b}}+\frac{{\theta_{b}}}{2\beta\phi\Gamma_{b}}, θb=2Rs+Re1\theta_{b}=\sqrt{2^{R_{s}+R_{e}^{*}}-1}, and β=N2π\beta=\frac{\sqrt{N}}{2\pi}. Differentiating 𝒯N\mathcal{T}_{\rm N} w.r.t. ϕ\phi yields

d𝒯Ndϕ=\displaystyle\frac{d\mathcal{T}_{\rm N}}{d\phi}= Rs[ϖ1ϱ1Meϱ12(M1)!+ϕΓbβϖ2ΔΓθb+βθbϖ1Γ(M,ϱ1)\displaystyle R_{s}\bigg{[}\frac{\varpi_{1}\varrho_{1}^{M}e^{-\varrho_{1}}}{2(M-1)!}+\frac{\phi\Gamma_{b}\beta\varpi_{2}\Delta\Gamma}{\theta_{b}}+\beta\theta_{b}\varpi_{1}\Gamma(M,\varrho_{1})
(βθbϖ1+ϖ22)Γ¯(M,ϱ2)],\displaystyle-\left(\beta\theta_{b}\varpi_{1}+\frac{\varpi_{2}}{2}\right)\bar{\Gamma}(M,\varrho_{2})\bigg{]}, (42)

where ϖ1=1ϕ2Rsθb2dλedϕ\varpi_{1}=\frac{1}{\phi}-\frac{2^{R_{s}}}{\theta_{b}^{2}}\frac{d\lambda_{e}}{d\phi} and ϖ2=1ϕ2Rs2θb2dλedϕ\varpi_{2}=\frac{1}{\phi}-\frac{2^{R_{s}}}{2\theta_{b}^{2}}\frac{d\lambda_{e}}{d\phi} with λe\lambda_{e} given in (34). It is verified that the derivative d𝒯Ndϕ\frac{d\mathcal{T}_{\rm N}}{d\phi} is monotonically decreasing with ϕ\phi. In other words, for a fixed RsR_{s}, the optimal ϕ\phi^{*} that maximizes 𝒯N\mathcal{T}_{\rm N} is unique, which is ϕ=1\phi^{*}=1 if d𝒯Ndϕ|ϕ=1>0\frac{d\mathcal{T}_{\rm N}}{d\phi}|_{\phi=1}>0 or otherwise satisfies d𝒯Ndϕ=0\frac{d\mathcal{T}_{\rm N}}{d\phi}=0. Likewise, it is confirmed that the derivative

d𝒯NdRs=\displaystyle\frac{d\mathcal{T}_{\rm N}}{dR_{s}}= Γ¯(M,ϱ1)2+ϕΓbβθbΔΓλeRs2Rsln2θb[βΓ¯(M,ϱ1)\displaystyle\frac{\bar{\Gamma}(M,\varrho_{1})}{2}+\frac{\phi\Gamma_{b}\beta}{\theta_{b}}\Delta\Gamma-\frac{\lambda_{e}R_{s}2^{R_{s}}\ln 2}{\theta_{b}}\bigg{[}{\beta}\bar{\Gamma}(M,\varrho_{1})
+ϕΓbβΔΓ2θb2+ϱ1Meϱ12θb(M1)!(β+14θb)Γ¯(M,ϱ2)]\displaystyle+\frac{\phi\Gamma_{b}\beta\Delta\Gamma}{2\theta_{b}^{2}}+\frac{\varrho_{1}^{M}e^{-\varrho_{1}}}{2\theta_{b}(M-1)!}-\left({\beta}+\frac{1}{4\theta_{b}}\right)\bar{\Gamma}(M,\varrho_{2})\bigg{]} (43)

is first positive and then negative with increasing RsR_{s}, and the unique optimal RsR_{s}^{*} maximizing 𝒯N\mathcal{T}_{\rm N} can be calculated via a bisection method with the equation d𝒯NdRs=0\frac{d\mathcal{T}_{\rm N}}{dR_{s}}=0.

Refer to caption
Figure 7: Above: RsR_{s}^{*} and 𝒯N\mathcal{T}_{\rm N} vs. ϕ\phi; Bottom: ϕ\phi^{*} and 𝒯N\mathcal{T}_{\rm N} vs. RsR_{s}; for different NN and δ\delta, with M=4M=4, Γb=3\Gamma_{b}=3 dB and Γe=0\Gamma_{e}=0 dB.

The monotonicity of 𝒯N\mathcal{T}_{\rm N} w.r.t. ϕ\phi and RsR_{s} is verified in Fig. 7, where, similar to Fig. 6, 𝒯N\mathcal{T}_{\rm N} is given with the optimal ϕ\phi^{*} or RsR_{s}^{*}. This implies that the global maximal 𝒯N\mathcal{T}_{\rm N} is practically achieved even by alternatively solving the optimal ϕ\phi and RsR_{s}. As expected, 𝒯N\mathcal{T}_{\rm N} improves with a larger blocklength NN and a looser secrecy constraint (a larger δ\delta). It is found that RsR_{s}^{*} first increases and then might decrease with ϕ\phi, which means that a moderate RsR_{s} is desired to balance the decoding and throughput performance. On the other hand, a larger ϕ\phi^{*} is required to support an increasing RsR_{s}. It is also shown that RsR_{s}^{*} for a fixed ϕ\phi increases with NN, since a larger NN improves the decoding performance which then affords a larger RsR_{s}. Nevertheless, ϕ\phi^{*} decreases with NN in the low RsR_{s} regime whereas increases with NN in the high RsR_{s} regime. It can be explained as follows: for a low RsR_{s}, the rate redundancy ReR_{e} has a great impact on the decoding performance, and hence the AN power should be increased as NN increases to better combat the eavesdropper; in contrast, for a large RsR_{s}, the decoding correctness is more affected by the main channel quality, which requires a larger signal power to maintain a high decoding probability.

Proposition 1

For the high average channel gain Γb\Gamma_{b}\rightarrow\infty, 𝒯N\mathcal{T}_{\rm N} in (IV-B) is approximated as

limΓb𝒯N=Rs(1ϱ1M2M!).\lim_{\Gamma_{b}\rightarrow\infty}\mathcal{T}_{\rm N}=R_{s}\left(1-\frac{\varrho_{1}^{M}}{2M!}\right). (44)
Proof 14

Please refer to Appendix -H.

Proposition 1 shows that for a high average channel gain, the secrecy throughput becomes independent of the blocklength. In consequence, the optimal ϕ\phi^{*} and RsR_{s}^{*} maximizing 𝒯N\mathcal{T}_{\rm N} in (44) admit the following closed-form approximations [22, (19), (20)]

limΓbϕ\displaystyle\lim_{\Gamma_{b}\rightarrow\infty}\phi^{*} =1Λ+1,\displaystyle=\frac{1}{\sqrt{\Lambda}+1},~{}~{}~{} (45)
limΓbRs\displaystyle\lim_{\Gamma_{b}\rightarrow\infty}R_{s}^{*} =1Mln2[𝒲0(2exp(1)M!ΓbM(Λ+1)2M)1].\displaystyle=\frac{1}{M\ln 2}\left[\mathcal{W}_{0}\left(\frac{2\exp(1)M!\Gamma_{b}^{M}}{(\sqrt{\Lambda}+1)^{2M}}\right)-1\right]. (46)

Fig. 8 illustrates the influence of the number of transmit antennas MM on the maximal secrecy throughput 𝒯\mathcal{T}^{*} for both adaptive and non-adaptive schemes and the relative secrecy throughput gain Δ𝒯=𝒯A𝒯N𝒯N\Delta\mathcal{T}=\frac{\mathcal{T}_{\rm A}^{*}-\mathcal{T}_{\rm N}^{*}}{\mathcal{T}_{\rm N}^{*}}. It is not surprising that deploying more transmit antennas can significantly improve the secrecy throughput for both schemes. Similar to the observation in Fig. 4, both 𝒯A\mathcal{T}_{\rm A}^{*} and 𝒯N\mathcal{T}_{\rm N}^{*} increase with δ\delta and NN, but the benefit to 𝒯N\mathcal{T}_{\rm N}^{*} brought by a larger NN is nearly negligible. The right-hand-side subgraph shows that Δ𝒯\Delta\mathcal{T} drops sharply as MM increases but grows for a larger NN and a smaller δ\delta. This indicates that the superiority of the adaptive scheme over its non-adaptive counterpart is more pronounced for the scenarios requiring a large blocklengh, having few transmit antennas, suffering from a stringent secrecy constraint, etc; otherwise, the non-adaptive scheme might be appealing because of the low-complexity off-line design.

Refer to caption
Figure 8: 𝒯\mathcal{T}^{*} and Δ𝒯\Delta\mathcal{T} vs. MM for different NN and δ\delta, with Γb=3\Gamma_{b}=3 dB and Γe=0\Gamma_{e}=0 dB.

IV-C A Note on Multi-Antenna Eve

This subsection examines the secure transmission in the presence of an Eve with MeM_{e} antennas. Assume that Eve employs the minimum mean-squared error (MMSE) receiver, and then the CDF of Eve’s SINR under the null-space AN scheme can be given as [24]:

γe(x)=1exϕPen=1MeAn(x)(n1)!(xϕPe)n1,\mathcal{F}_{\gamma_{e}}(x)=1-e^{-\frac{x}{\phi P_{e}}}\sum_{n=1}^{M_{e}}\frac{A_{n}(x)}{(n-1)!}\left(\frac{x}{\phi P_{e}}\right)^{n-1}, (47)

where

An(x)={1,MeM1+n,m=0Men(M1m)(ξx)m(1+ξx)M1,Me<M1+n.\displaystyle A_{n}(x)=\begin{cases}1,&M_{e}\geq M-1+n,\\ \frac{\sum_{m=0}^{M_{e}-n}\binom{M-1}{m}(\xi x)^{m}}{(1+\xi x)^{M-1}},&M_{e}<M-1+n.\\ \end{cases} (48)

The information leakage probability 𝒪e\mathcal{O}_{e} is obtained by substituting (47) into (50), and the secrecy throughput can be optimized similarly as described in the above two subsections.

By ignoring the receiver noise at Eve, i.e., considering Eve’s transmit power PeP_{e}\rightarrow\infty, one can obtain γe(x)=1A1(x)\mathcal{F}_{\gamma_{e}}(x)=1-A_{1}(x). Furthermore, if Eve has more antennas than Alice, i.e., MeMM_{e}\geq M, one have A1(x)=1A_{1}(x)=1, γe(x)=0\mathcal{F}_{\gamma_{e}}(x)=0, and accordingly 𝒪e=1\mathcal{O}_{e}=1. This means, when PeP_{e}\rightarrow\infty, Eve with enough antennas can completely eliminate all the AN signals with an MMSE receiver such that her SNR will approach infinity. As a consequence, the SOP constraint can no longer be satisfied for any chosen rate redundancy, and no positive secrecy rate can be achieved from the perspective of secrecy outage. In other words, the null-space AN scheme can safeguard secure transmissions well for the finite blocklength regime only when the eavesdropper has fewer antennas than the transmitter, and this conclusion is the same as that for the infinite blocklength case.

V Conclusions

This paper investigated the design of secure transmissions in slow fading channels, where secrecy encoding with finite blocklength was employed to confront the eavesdropper. Both adaptive and non-adaptive schemes were devised to maximize the secrecy throughput, providing the optimal threshold of the on-off transmission policy, blocklength, code rates, and power allocation of the AN scheme. Theoretical and numerical results showed that, under the on-off policy, increasing the blocklength can simultaneously enhance the reliability and secrecy, and thus the secrecy throughput is maximized when using the maximal blocklength. In addition, since an overly large secrecy rate will significantly decrease the successful decoding probability thus lowering the secrecy throughput, there exists a critical secrecy rate, but not as large as possible, that can achieve the maximal secrecy throughput.

-A Proof of Lemma 1

For tractability, a piece-wise linear approximation approach is leveraged to approximate the QQ-function given in (2), i.e., Q(CiRiVi/n)Ξ(γi,n,Ri)Q\left(\frac{C_{i}-R_{i}}{\sqrt{V_{i}/n}}\right)\approx\Xi(\gamma_{i},n,R_{i}) for i{b,e}i\in\{b,e\} [32, 33],333 This approximation has been extensively applied to the finite-blocklength scenarios, and its accuracy has been well validated. with

Ξ(γi,n,Ri)={0,γi>τiu,12βθi(γiθi2),τilγiτiu,1,γi<τil,\displaystyle\Xi(\gamma_{i},n,R_{i})=\begin{cases}0,&\gamma_{i}>\tau^{u}_{i},\\ \frac{1}{2}-\frac{\beta}{\theta_{i}}(\gamma_{i}-\theta_{i}^{2}),&\tau^{l}_{i}\leq\gamma_{i}\leq\tau^{u}_{i},\\ 1,&\gamma_{i}<\tau^{l}_{i},\\ \end{cases} (49)

where βn2π\beta\triangleq\frac{\sqrt{n}}{2\pi}, θi2Ri1\theta_{i}\triangleq\sqrt{2^{R_{i}}-1}, τiuθi2+θi2β\tau^{u}_{i}\triangleq\theta_{i}^{2}+\frac{\theta_{i}}{2\beta}, and τilθi2θi2β\tau^{l}_{i}\triangleq\theta_{i}^{2}-\frac{\theta_{i}}{2\beta}.444Generally, θi>12β\theta_{i}>\frac{1}{2\beta} or Ri>log2(1+π2/n)R_{i}>\log_{2}\left(1+{\pi^{2}}/{n}\right) should be satisfied to ensure a positive τil\tau_{i}^{l}. With (49), the information leakage probability 𝒪e\mathcal{O}_{e} defined in (4) is calculated as

𝒪e=1𝔼γe[Ξ(γe,n,Re)]=1βθeτelτeuγe(γ)𝑑γ,\displaystyle\mathcal{O}_{e}=1-\mathbb{E}_{\gamma_{e}}\left[\Xi(\gamma_{e},n,R_{e})\right]=1-\frac{\beta}{\theta_{e}}\int_{\tau^{l}_{e}}^{\tau^{u}_{e}}\mathcal{F}_{\gamma_{e}}(\gamma)d\gamma, (50)

where γi(γ)=1eγ/Γi2\mathcal{F}_{\gamma_{i}}(\gamma)=1-e^{-\gamma/\Gamma_{i}^{2}} is the CDF of γi\gamma_{i} for i{b,e}i\in\{b,e\}, and the last equality in (50) follows from invoking (49) and using partial integration. Next, treat nn as a continuous variable. As ReR_{e}^{*} satisfies 𝒪e(Re)=δ\mathcal{O}_{e}(R_{e}^{*})=\delta, the derivative dRedn\frac{dR_{e}^{*}}{dn} is obtained by using the derivative rule for implicit functions [23] with 𝒪e(Re)=δ\mathcal{O}_{e}(R_{e}^{*})=\delta, i.e.:

dRedn=𝒪e/n𝒪e/Re.\frac{dR_{e}^{*}}{dn}=-\frac{\partial\mathcal{O}_{e}/\partial n}{\partial\mathcal{O}_{e}/\partial R_{e}^{*}}. (51)

First, it can be proved that 𝒪eRe=𝒪eθeθeRe<0\frac{\partial\mathcal{O}_{e}}{\partial R_{e}^{*}}=\frac{\partial\mathcal{O}_{e}}{\partial\theta_{e}}\frac{\partial\theta_{e}}{\partial R_{e}^{*}}<0 by noting that θeRe>0\frac{\partial\theta_{e}}{\partial R_{e}^{*}}>0 and

𝒪eθe=βθe2τelτeuγe(γ)𝑑γβθe[dτeudθeγe(τeu)dτeldθeγe(τel)]\displaystyle\frac{\partial\mathcal{O}_{e}}{\partial\theta_{e}}=\frac{\beta}{\theta_{e}^{2}}\int_{\tau^{l}_{e}}^{\tau^{u}_{e}}\mathcal{F}_{\gamma_{e}}(\gamma)d\gamma-\frac{\beta}{\theta_{e}}\left[\frac{d\tau^{u}_{e}}{d\theta_{e}}\mathcal{F}_{\gamma_{e}}(\tau^{u}_{e})-\frac{d\tau^{l}_{e}}{d\theta_{e}}\mathcal{F}_{\gamma_{e}}(\tau^{l}_{e})\right]
(a)βθe2[γγe(γ)]|τelτeu4βθe+12θeγe(τeu)+4βθe12θeγe(τel)\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{\leq}}\frac{\beta}{\theta_{e}^{2}}\left[\gamma\mathcal{F}_{\gamma_{e}}(\gamma)\right]|^{\tau^{u}_{e}}_{\tau^{l}_{e}}-\frac{4\beta\theta_{e}+1}{2\theta_{e}}\mathcal{F}_{\gamma_{e}}(\tau^{u}_{e})+\frac{4\beta\theta_{e}-1}{2\theta_{e}}\mathcal{F}_{\gamma_{e}}(\tau^{l}_{e})
=β[γe(τel)γe(τeu)]<0,\displaystyle=\beta\left[\mathcal{F}_{\gamma_{e}}(\tau^{l}_{e})-\mathcal{F}_{\gamma_{e}}(\tau^{u}_{e})\right]<0,

where (a)\mathrm{(a)} follows from the partial integration. The next step is to determine the sign of 𝒪en=𝒪eββn\frac{\partial\mathcal{O}_{e}}{\partial n}=\frac{\partial\mathcal{O}_{e}}{\partial\beta}\frac{\partial\beta}{\partial n}. The first and second derivatives of 𝒪e\mathcal{O}_{e} w.r.t. β\beta are respectively given by

𝒪eβ\displaystyle\frac{\partial\mathcal{O}_{e}}{\partial\beta} =12β[γe(τeu)+γe(τel)]1θeτelτeuγe(γ)𝑑γ,\displaystyle=\frac{1}{2\beta}\left[\mathcal{F}_{\gamma_{e}}(\tau^{u}_{e})+\mathcal{F}_{\gamma_{e}}(\tau^{l}_{e})\right]-\frac{1}{\theta_{e}}\int_{\tau^{l}_{e}}^{\tau^{u}_{e}}\mathcal{F}_{\gamma_{e}}(\gamma)d\gamma, (52)
2𝒪eβ2\displaystyle\frac{\partial^{2}\mathcal{O}_{e}}{\partial\beta^{2}} =θe4β3[fγe(τel)fγe(τeu)].\displaystyle=\frac{\theta_{e}}{4\beta^{3}}\left[f_{\gamma_{e}}(\tau^{l}_{e})-f_{\gamma_{e}}(\tau^{u}_{e})\right]. (53)

It is easy to see 2𝒪eβ2>0\frac{\partial^{2}\mathcal{O}_{e}}{\partial\beta^{2}}>0 as fγe(γ)=1Γeeγ/Γef_{\gamma_{e}}(\gamma)=\frac{1}{\Gamma_{e}}e^{-{\gamma}/{\Gamma_{e}}} decreases with γ\gamma and τeu>τel\tau^{u}_{e}>\tau^{l}_{e}. This indicates that 𝒪eβ\frac{\partial\mathcal{O}_{e}}{\partial\beta} increases with β\beta such that 𝒪eβ<𝒪eβ|β=0\frac{\partial\mathcal{O}_{e}}{\partial\beta}<\frac{\partial\mathcal{O}_{e}}{\partial\beta}|_{\beta\rightarrow\infty}=0. Combining 𝒪eβ<0\frac{\partial\mathcal{O}_{e}}{\partial\beta}<0 and βn>0\frac{\partial\beta}{\partial n}>0 yields 𝒪en<0\frac{\partial\mathcal{O}_{e}}{\partial n}<0. With 𝒪eRe<0\frac{\partial\mathcal{O}_{e}}{\partial R_{e}^{*}}<0 and 𝒪en<0\frac{\partial\mathcal{O}_{e}}{\partial n}<0 in (51), dRedn<0\frac{dR_{e}^{*}}{dn}<0 is obtained, which completes the proof.

-B Proof of Theorem 1

The derivative of psp_{s} w.r.t. nn is given by

dpsdn=12πen(CbRt)22Vb[CbRt2nVbnVbdRedn],\frac{dp_{s}}{dn}=\frac{1}{\sqrt{2\pi}}e^{-\frac{n(C_{b}-R_{t})^{2}}{2V_{b}}}\left[\frac{C_{b}-R_{t}}{2\sqrt{nV_{b}}}-\sqrt{\frac{n}{V_{b}}}\frac{dR_{e}^{*}}{dn}\right], (54)

which follows from the derivative dQ(x)dx=12πex22\frac{dQ(x)}{dx}=\frac{-1}{\sqrt{2\pi}}e^{-\frac{x^{2}}{2}}. Plugging dRedn<0\frac{dR_{e}^{*}}{dn}<0, as shown in Lemma 1, into (54) yields dpsdn>0\frac{dp_{s}}{dn}>0. For a fixed RsR_{s} in (10), it is directly concluded that d𝒯A(η)dn>0\frac{d\mathcal{T}_{\rm A}(\eta)}{dn}>0, which means that 𝒯A(η)\mathcal{T}_{\rm A}(\eta) monotonically increases with nn. Since nn is an integer, 𝒯A(η)\mathcal{T}_{\rm A}(\eta) is maximized at the maximal integer of nn, i.e., n=Nn=N, which completes the proof.

-C Proof of Theorem 2

From (13), it is easy to prove that d2𝒯A(η)dRs2<0\frac{d^{2}\mathcal{T}_{\rm A}(\eta)}{dR_{s}^{2}}<0, i.e., 𝒯A(η)\mathcal{T}_{\rm A}(\eta) is concave on RsR_{s}. It is verified that d𝒯A(η)dRs|Rs=0=1Q(CbReVb/N)>0\frac{d\mathcal{T}_{\rm A}(\eta)}{dR_{s}}|_{R_{s}=0}=1-Q\left(\frac{C_{b}-R_{e}^{*}}{\sqrt{V_{b}/N}}\right)>0. As a result, 𝒯A(η)\mathcal{T}_{\rm A}(\eta) is maximized at the boundary Rs=CbReR_{s}=C_{b}-R_{e}^{*} if d𝒯A(η)dRs|Rs=CbRe=12CbRe2πVb/N0ReCbπVb2N\frac{d\mathcal{T}_{\rm A}(\eta)}{dR_{s}}|_{R_{s}=C_{b}-R_{e}^{*}}=\frac{1}{2}-\frac{C_{b}-R_{e}^{*}}{\sqrt{2\pi V_{b}/N}}\geq 0\Rightarrow R_{e}^{*}\geq C_{b}-\sqrt{\frac{\pi V_{b}}{2N}} or otherwise at the unique zero-crossing of d𝒯A(η)dRs\frac{d\mathcal{T}_{\rm A}(\eta)}{dR_{s}}, i.e., RsR_{s}^{\circ}. Next, the condition ReCbπVb2NR_{e}^{*}\geq C_{b}-\sqrt{\frac{\pi V_{b}}{2N}} is equivalently transformed to that γb\gamma_{b} does not exceed a critical value γb\gamma_{b}^{\circ}. Let ψ(γb)=CbπVb2N\psi(\gamma_{b})=C_{b}-\sqrt{\frac{\pi V_{b}}{2N}}. It can be readily confirmed that ψ(γb)<0\psi(\gamma_{b})<0, and ψ(γb)\psi(\gamma_{b}) decreases with γb\gamma_{b} if 0<γb<γbL12+14+π2N10<\gamma_{b}<\gamma_{b}^{L}\triangleq\sqrt{\frac{1}{2}+\sqrt{\frac{1}{4}+\frac{\pi}{2N}}}-1 or otherwise increases with γb\gamma_{b}. This leads to Reψ(γb)γbγbψ1(Re)R_{e}^{*}\geq\psi(\gamma_{b})\Rightarrow\gamma_{b}\leq\gamma_{b}^{\circ}\triangleq\psi^{-1}(R_{e}^{*}). An upper bound for γb\gamma_{b}^{\circ} is further provided by realizing that ψ(γb)=Re>log2(1+γb)π2Nlog2eγb<γbUeπ2N+Reln21\psi(\gamma_{b}^{\circ})=R_{e}^{*}>\log_{2}(1+\gamma_{b}^{\circ})-\sqrt{\frac{\pi}{2N}}\log_{2}e\Rightarrow\gamma_{b}^{\circ}<\gamma_{b}^{U}\triangleq e^{\sqrt{\frac{\pi}{2N}}+R_{e}^{*}\ln 2}-1. Then, γb\gamma_{b}^{\circ} can be quickly calculated using the bisection method with ψ(γb)=Re\psi(\gamma_{b})=R_{e}^{*} in the range (γbL,γbU)(\gamma_{b}^{L},\gamma_{b}^{U}). This completes the proof.

-D Proof of Theorem 4

First, display the derivative dY(θb)dθb\frac{dY(\theta_{b})}{d\theta_{b}} in a recursive form dY(θb)dθb=1θb(1θb+12βΓb)Y(θb)\frac{dY(\theta_{b})}{d\theta_{b}}=\frac{1}{\theta_{b}}-\left(\frac{1}{\theta_{b}}+\frac{1}{2\beta\Gamma_{b}}\right)Y(\theta_{b}) with Y(θb)Y(\theta_{b}) in (20). Then, the derivative d𝒯Ndθb\frac{d\mathcal{T}_{\rm N}}{d\theta_{b}} is given by d𝒯Ndθb=θb1+θb2eθb2ΓbG(θb)\frac{d\mathcal{T}_{\rm N}}{d\theta_{b}}=\frac{\theta_{b}}{1+\theta_{b}^{2}}e^{-\frac{\theta_{b}^{2}}{\Gamma_{b}}}G(\theta_{b}), with G(θb)G(\theta_{b}) presented in (21). It is easily proved that G(θb)>0G(\theta_{b})>0 when θb=2Re1\theta_{b}=\sqrt{2^{R_{e}^{*}}-1} and G(θb)<0G(\theta_{b})<0 as θb\theta_{b}\rightarrow\infty. The key step of the proof is to argue that G(θb)G(\theta_{b}) monotonically decreases with θb\theta_{b}, which guarantees a unique zero-crossing of G(θb)G(\theta_{b}) within θb(2Re1,)\theta_{b}\in(\sqrt{2^{R_{e}^{*}}-1},\infty). In other words, 𝒯N{\mathcal{T}_{\rm N}} initially increases and then decreases with θb\theta_{b} and reaches the maximum when θb\theta_{b} arrives at the unique zero-crossing of G(θb)G(\theta_{b}). To this end, it is necessary to calculate the derivative dG(θb)dθb\frac{dG(\theta_{b})}{d\theta_{b}}:

dG(θb)dθb=dY(θb)dθb2g(θb)ln2[log2(1+θb2)Re]h(θb),\frac{dG(\theta_{b})}{d\theta_{b}}=\frac{\frac{dY(\theta_{b})}{d\theta_{b}}-2g(\theta_{b})}{\ln 2}-\left[\log_{2}(1+\theta_{b}^{2})-R_{e}^{*}\right]h(\theta_{b}), (55)

where h(θb)=(11θb2)g(θb)+1+θb2θbdg(θb)dθbh(\theta_{b})=\left(1-\frac{1}{\theta_{b}^{2}}\right)g(\theta_{b})+\frac{1+\theta_{b}^{2}}{\theta_{b}}\frac{dg(\theta_{b})}{d\theta_{b}}. To proceed, the following lemma is introduced.

Lemma 3

Y(θb)Y(\theta_{b}) decreases with θb\theta_{b} and satisfies

2βΓbθb+2βΓb<Y(θb)<min{1,2βΓbθb}.\frac{2\beta\Gamma_{b}}{\theta_{b}+2\beta\Gamma_{b}}<Y(\theta_{b})<\min\left\{1,\frac{2\beta\Gamma_{b}}{\theta_{b}}\right\}. (56)
Proof 15

Define xθb2βΓbx\triangleq\frac{\theta_{b}}{2\beta\Gamma_{b}} such that Y(θb)=1exxY(\theta_{b})=\frac{1-e^{-x}}{x}. The monotonicity of Y(θb)Y(\theta_{b}) w.r.t. θb\theta_{b} is due to dY(θb)dθb=(1+x)ex1x<0\frac{dY(\theta_{b})}{d\theta_{b}}=\frac{(1+x)e^{-x}-1}{x}<0. The lower bound of Y(θb)Y(\theta_{b}) is obtained from dY(θb)dθb=1θb(1θb+12βΓb)Y(θb)<0\frac{dY(\theta_{b})}{d\theta_{b}}=\frac{1}{\theta_{b}}-\left(\frac{1}{\theta_{b}}+\frac{1}{2\beta\Gamma_{b}}\right)Y(\theta_{b})<0 and the upper bound follows from Y(θb)<1xY(\theta_{b})<\frac{1}{x} and 1ex<x1-e^{-x}<x.

With the lower bound of Y(θb)Y(\theta_{b}) given in (56), it can be readily proved that g(θb)>0g(\theta_{b})>0 such that the term dY(θb)dθb2g(θb){\frac{dY(\theta_{b})}{d\theta_{b}}-2g(\theta_{b})} in (55) is negative. Besides, since h(θb)0h(\theta_{b})\geq 0 directly yields dG(θb)dθb<0\frac{dG(\theta_{b})}{d\theta_{b}}<0, one only needs to discuss the situation h(θb)<0h(\theta_{b})<0 and prove that

dG(θb)dθbln2\displaystyle\frac{dG(\theta_{b})}{d\theta_{b}}\ln 2 dY(θb)dθb2g(θb)h(θb)ln(1+θb2)\displaystyle\leq{\frac{dY(\theta_{b})}{d\theta_{b}}-2g(\theta_{b})}-h(\theta_{b})\ln(1+\theta_{b}^{2})
<(a)dY(θb)dθb2g(θb)θb2h(θb)\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{<}}{\frac{dY(\theta_{b})}{d\theta_{b}}-2g(\theta_{b})}-\theta_{b}^{2}h(\theta_{b})
(b)1θb+2βΓb(θb2Γb+θb4Γb+8βθb+8βθb3θb2)\displaystyle\stackrel{{\scriptstyle\mathrm{(b)}}}{{\leq}}-\frac{1}{\theta_{b}+2\beta\Gamma_{b}}\left(\frac{\theta_{b}^{2}}{\Gamma_{b}}+\frac{\theta_{b}^{4}}{\Gamma_{b}}+8\beta\theta_{b}+8\beta\theta_{b}^{3}-\theta_{b}^{2}\right)
<(c)0,\displaystyle\stackrel{{\scriptstyle\mathrm{(c)}}}{{<}}0, (57)

where (a)\mathrm{(a)} is due to ln(1+x)x\ln(1+x)\leq x, (b)\mathrm{(b)} holds by invoking Lemma 3 along with algebraic manipulations, and (c)\mathrm{(c)} derives from 8βθb+8βθb316βθb2>θb28\beta\theta_{b}+8\beta\theta_{b}^{3}\geq 16\beta\theta_{b}^{2}>\theta_{b}^{2} as β=N2π>18\beta=\frac{\sqrt{N}}{2\pi}>\frac{1}{8}.

-E Proof of Theorem 5

First fix RsR_{s}, and it is clear that the term λbRsλb21\frac{-\lambda_{b}R_{s}}{\sqrt{\lambda_{b}^{2}-1}} in (28) increases with α\alpha as λb\lambda_{b} increases with α\alpha. It is also verified that the term Z(α)λb(lnλblnλe)λb21Z(\alpha)\triangleq\frac{\lambda_{b}(\ln\lambda_{b}-\ln\lambda_{e})}{\sqrt{\lambda_{b}^{2}-1}} in (28) increases with α\alpha by computing the derivative of Z(α)Z(\alpha) w.r.t. α\alpha:

dZ(α)dα\displaystyle\frac{dZ(\alpha)}{d\alpha} =dλbdα(λb21lnλbλe)λb(λb21)λedλedα(λb21)3/2\displaystyle=\frac{\frac{d\lambda_{b}}{d\alpha}\left(\lambda_{b}^{2}-1-\ln\frac{\lambda_{b}}{\lambda_{e}}\right)-\frac{\lambda_{b}(\lambda_{b}^{2}-1)}{\lambda_{e}}\frac{d\lambda_{e}}{d\alpha}}{(\lambda_{b}^{2}-1)^{3/2}}
=(a)λb(λb1)[(λbλe)(λb+1)lnλbλe]α(λb21)3/2\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{=}}\frac{\lambda_{b}(\lambda_{b}-1)\left[(\lambda_{b}-\lambda_{e})(\lambda_{b}+1)-\ln\frac{\lambda_{b}}{\lambda_{e}}\right]}{\alpha(\lambda_{b}^{2}-1)^{3/2}}
(b)λb(λb1)(λbλe)(λb+11λe)α(λb21)3/2>(c)0,\displaystyle\stackrel{{\scriptstyle\mathrm{(b)}}}{{\geq}}\frac{\lambda_{b}(\lambda_{b}-1)(\lambda_{b}-\lambda_{e})\left(\lambda_{b}+1-\frac{1}{\lambda_{e}}\right)}{\alpha(\lambda_{b}^{2}-1)^{3/2}}\stackrel{{\scriptstyle\mathrm{(c)}}}{{>}}0, (58)

where (a)\mathrm{(a)} holds by substituting dλidα=λi(λi1)α\frac{d\lambda_{i}}{d\alpha}=\frac{\lambda_{i}(\lambda_{i}-1)}{\alpha} for i{b,e}i\in\{b,e\}, (b)\mathrm{(b)} follows from the inequality lnλbλeλbλeλe\ln\frac{\lambda_{b}}{\lambda_{e}}\leq\frac{\lambda_{b}-\lambda_{e}}{\lambda_{e}} with λb>λe>0\lambda_{b}>\lambda_{e}>0, and (c)\mathrm{(c)} is due to λb>λe>1\lambda_{b}>\lambda_{e}>1. Hence, psp_{s} in (28) increases with α\alpha as Q(x)Q(x) decreases with xx. This indicates, α=1\alpha^{*}=1 is optimal for maximizing 𝒯A(η)=Rsps\mathcal{T}_{\rm A}(\eta)=R_{s}p_{s} for any given RsR_{s} and η\eta and is also optimal for maximizing 𝒯A\mathcal{T}_{\rm A}.

-F Proof of Theorem 7

Let L(ϕ)=L1L2L(\phi)=L_{1}L_{2}, where L1=λbλb21L_{1}=\frac{\lambda_{b}}{\sqrt{\lambda_{b}^{2}-1}} and L2=lnλbλeRsln2L_{2}={\ln\frac{\lambda_{b}}{\lambda_{e}}-R_{s}\ln 2} such that dL1dϕ=ρb(L121)3/2\frac{dL_{1}}{d\phi}=-\rho_{b}(L_{1}^{2}-1)^{3/2} and dL2dϕ=ρbλbρe+ϕdρedϕλe\frac{dL_{2}}{d\phi}=\frac{\rho_{b}}{\lambda_{b}}-\frac{\rho_{e}+\phi\frac{d\rho_{e}}{d\phi}}{\lambda_{e}}. Rewrite the second derivative as d2L(ϕ)dϕ2=L1(L121)2I(ϕ)\frac{d^{2}L(\phi)}{d\phi^{2}}=L_{1}(L_{1}^{2}-1)^{2}I(\phi) with I(ϕ)I(\phi) given by (-F) at the top of this page, where (a)\mathrm{(a)} holds by recalling the definition L2lnλbλeL_{2}\leq\ln\frac{\lambda_{b}}{\lambda_{e}} and invoking the result d2ρedϕ2>2ρe(dρedϕ)2>0\frac{d^{2}\rho_{e}}{d\phi^{2}}>\frac{2}{\rho_{e}}\left(\frac{d\rho_{e}}{d\phi}\right)^{2}>0 from Lemma 2, (b)\mathrm{(b)} follows from plugging L1=λbλb21L_{1}=\frac{\lambda_{b}}{\sqrt{\lambda_{b}^{2}-1}}, using the inequality lnλbλeλbλeλe\ln\frac{\lambda_{b}}{\lambda_{e}}\leq\frac{\lambda_{b}-\lambda_{e}}{\lambda_{e}}, and omitting the term (ϕ2+2ϕρe)(dρedϕ)2(\phi^{2}+\frac{2\phi}{\rho_{e}})(\frac{d\rho_{e}}{d\phi})^{2}, (c)\mathrm{(c)} holds by substituting dL2dϕ\frac{dL_{2}}{d\phi} into (b)\mathrm{(b)}, and (d)\mathrm{(d)} is established after some manipulation operations and by discarding the negative term 2(λb21)λbλe2[ϕρbλeλb(λb21)]\frac{2(\lambda_{b}^{2}-1)}{\lambda_{b}\lambda_{e}^{2}}\left[\phi\rho_{b}\lambda_{e}-\lambda_{b}\left(\lambda_{b}^{2}-1\right)\right] noting that λb=1+ϕρb>λe\lambda_{b}=1+\phi\rho_{b}>\lambda_{e}. As indicated by (-F) that L(ϕ)L(\phi) is concave on ϕ\phi, L(ϕ)L(\phi) is maximized at ϕ=1\phi=1 if dL(ϕ)dϕ|ϕ=10\frac{dL(\phi)}{d\phi}|_{\phi=1}\geq 0 or otherwise at the unique zero-crossing of dL(ϕ)dϕ\frac{dL(\phi)}{d\phi}. Besides, dL(ϕ)dϕ|ϕ=10\frac{dL(\phi)}{d\phi}|_{\phi=1}\geq 0 is equivalent to X(ρb)0X(\rho_{b})\geq 0 in (38). Clearly, A1=(1+Γe)ρe(1)1+ρe(1)<1A_{1}=\frac{(1+\Gamma_{e})\rho_{e}(1)}{1+\rho_{e}(1)}<1 must be ensured to yield a positive X(ρb)X(\rho_{b}), with which it can be verified that X(ρb)X(\rho_{b}) increases with ρb\rho_{b}. As a consequence, dL(ϕ)dϕ|ϕ=10\frac{dL(\phi)}{d\phi}|_{\phi=1}\geq 0 can be transformed to an explicit form with relation to ρb\rho_{b}, namely, ρbρb\rho_{b}\geq\rho_{b}^{\circ}. This completes the proof.

I(ϕ)\displaystyle I(\phi) =3ρb2L22ρbdL2dϕL1(L121)1/21(L121)2[ρb2λb2+2dρedϕ+(ϕ+ϕ2ρe)d2ρedϕ2ρe2ϕ2(dρedϕ)2λe2]\displaystyle=3\rho_{b}^{2}L_{2}-\frac{2\rho_{b}\frac{dL_{2}}{d\phi}}{L_{1}(L_{1}^{2}-1)^{1/2}}-\frac{1}{(L_{1}^{2}-1)^{2}}\left[{\frac{\rho_{b}^{2}}{\lambda_{b}^{2}}+\frac{2\frac{d\rho_{e}}{d\phi}+(\phi+\phi^{2}\rho_{e})\frac{d^{2}\rho_{e}}{d\phi^{2}}-\rho_{e}^{2}-\phi^{2}(\frac{d\rho_{e}}{d\phi})^{2}}{\lambda_{e}^{2}}}\right]
(a)3ρb2lnλbλe2ρbL1L121dL2dϕ1(L121)2[ρb2λb2+2dρedϕ+(ϕ2+2ϕρe)(dρedϕ)2ρe2λe2]\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{\leq}}3\rho_{b}^{2}\ln\frac{\lambda_{b}}{\lambda_{e}}-\frac{2\rho_{b}}{L_{1}\sqrt{L_{1}^{2}-1}}\frac{dL_{2}}{d\phi}-\frac{1}{{(L_{1}^{2}-1)^{2}}}\left[{\frac{\rho_{b}^{2}}{\lambda_{b}^{2}}+\frac{2\frac{d\rho_{e}}{d\phi}+(\phi^{2}+\frac{2\phi}{\rho_{e}})(\frac{d\rho_{e}}{d\phi})^{2}-\rho_{e}^{2}}{\lambda_{e}^{2}}}\right]
(b)[3ρb2λbλeλe2ρbλb21λbdL2dϕ(λb21)2(ρb2λb2+2dρedϕρe2λe2)]\displaystyle\stackrel{{\scriptstyle\mathrm{(b)}}}{{\leq}}\left[3\rho_{b}^{2}\frac{\lambda_{b}-\lambda_{e}}{\lambda_{e}}-{2\rho_{b}}\frac{\lambda_{b}^{2}-1}{\lambda_{b}}\frac{dL_{2}}{d\phi}-\left(\lambda_{b}^{2}-1\right)^{2}\left({\frac{\rho_{b}^{2}}{\lambda_{b}^{2}}+\frac{2\frac{d\rho_{e}}{d\phi}-\rho_{e}^{2}}{\lambda_{e}^{2}}}\right)\right]
=(c)ρb2[3λb2(λbλe)+λe(1λb4)]λb2λe+2ρb(λb21)λbλe(ρe+ϕdρedϕ)+(λb21)2λe2(ρe22dρedϕ)\displaystyle\stackrel{{\scriptstyle\mathrm{(c)}}}{{=}}\frac{\rho_{b}^{2}\left[3\lambda_{b}^{2}(\lambda_{b}-\lambda_{e})+\lambda_{e}(1-\lambda_{b}^{4})\right]}{\lambda_{b}^{2}\lambda_{e}}+\frac{2\rho_{b}(\lambda_{b}^{2}-1)}{\lambda_{b}\lambda_{e}}\left(\rho_{e}+\phi\frac{d\rho_{e}}{d\phi}\right)+\frac{(\lambda_{b}^{2}-1)^{2}}{\lambda_{e}^{2}}\left(\rho_{e}^{2}-2\frac{d\rho_{e}}{d\phi}\right)
(d)ϕρb2(ρbρe)λb2λe2[2ϕ4ρb3ρe+ϕ3ρb2(ρb+6ρe)+ϕ2ρb(ρb+8ρe)+5ϕρe+1]<0,\displaystyle\stackrel{{\scriptstyle\mathrm{(d)}}}{{\leq}}-\frac{\phi\rho_{b}^{2}(\rho_{b}-\rho_{e})}{\lambda_{b}^{2}\lambda_{e}^{2}}\left[2\phi^{4}\rho_{b}^{3}\rho_{e}+\phi^{3}\rho_{b}^{2}(\rho_{b}+6\rho_{e})+\phi^{2}\rho_{b}(\rho_{b}+8\rho_{e})+5\phi\rho_{e}+1\right]<0, (59)

-G Proof of Corollary 5

Let K(ϕ)K(\phi) denote dL(ϕ)dϕ\frac{dL(\phi)}{d\phi} in (37). It is verified that dϕdη=K(ϕ)/ηK(ϕ)/ϕ|ϕ=ϕ>0\frac{d\phi^{*}}{d\eta}=-\frac{\partial K(\phi)/\partial\eta}{\partial K(\phi)/\partial\phi}|_{\phi=\phi^{*}}>0 by recalling that K(ϕ)ϕ|ϕ=ϕ<0\frac{\partial K(\phi)}{\partial\phi}|_{\phi=\phi^{*}}<0 from Theorem 7 and proving that

K(ϕ)ρb|ϕ=ϕ\displaystyle\frac{\partial K(\phi)}{\partial\rho_{b}}|_{\phi=\phi^{*}} =λb2λb+1λb(1Aϕ)+(2λb1)(lnλbBϕ)λb+1ϕ(λb21)5/2/Pb\displaystyle=\frac{\frac{\lambda_{b}^{2}-\lambda_{b}+1}{\lambda_{b}}-(1-A_{\phi^{*}})+\frac{(2\lambda_{b}-1)(\ln\lambda_{b}-B_{\phi^{*}})}{\lambda_{b}+1}}{\phi^{*}(\lambda_{b}^{2}-1)^{{5}/{2}}/P_{b}}
=(a)1λbλb+(2λb2λb1)(1Aϕ)ϕ(λb21)5/2/Pb\displaystyle\stackrel{{\scriptstyle\mathrm{(a)}}}{{=}}\frac{\frac{1}{\lambda_{b}}-\lambda_{b}+(2\lambda_{b}^{2}-\lambda_{b}-1)(1-A_{\phi^{*}})}{\phi^{*}(\lambda_{b}^{2}-1)^{{5}/{2}}/P_{b}}
(b)λb1ϕ(λb21)5/2/Pb>0,\displaystyle\stackrel{{\scriptstyle\mathrm{(b)}}}{{\geq}}\frac{\lambda_{b}-1}{\phi^{*}(\lambda_{b}^{2}-1)^{{5}/{2}}/P_{b}}>0, (60)

where (a)\mathrm{(a)} is due to lnλbBϕλb+1=(1Aϕ)λb1\frac{\ln\lambda_{b}-B_{\phi^{*}}}{\lambda_{b}+1}={(1-A_{\phi^{*}}){\lambda_{b}}-1} from K(ϕ)=0K(\phi^{*})=0, and (b)\mathrm{(b)} is because (1Aϕ)λb1>0(1-A_{\phi^{*}})\lambda_{b}-1>0. Moreover, limηK(ϕ)=1Aϕϕ\lim_{\eta\rightarrow\infty}K(\phi^{*})=\frac{1-A_{\phi^{*}}}{\phi^{*}}. Solving K(ϕ)=0K(\phi^{*})=0 with Aϕ=ϕΛ(1ϕ)(1ϕ+ϕΛ)A_{\phi^{*}}=\frac{\phi^{*}\Lambda}{(1-\phi^{*})(1-\phi^{*}+\phi^{*}\Lambda)} yields ϕ=1Λ+1\phi^{*}=\frac{1}{\sqrt{\Lambda}+1}. Similarly, one can prove that dϕdΛ<0dϕdδ>0\frac{d\phi^{*}}{d\Lambda}<0\Rightarrow\frac{d\phi^{*}}{d\delta}>0 and limδ1ϕ=1\lim_{\delta\rightarrow 1}\phi^{*}=1.

-H Proof of Proposition 1

Note that ϱ1,ϱ20\varrho_{1},\varrho_{2}\rightarrow 0 as Γb\Gamma_{b}\rightarrow\infty. Resorting to [22, Eqn. (44)] yields

Γ¯(M,ϱi)=eϱik=0M1ϱikk!1ϱiMM!,i{1,2},\bar{\Gamma}(M,\varrho_{i})=e^{-\varrho_{i}}\sum_{k=0}^{M-1}\frac{\varrho_{i}^{k}}{k!}\approx 1-\frac{\varrho_{i}^{M}}{M!},~{}i\in\{1,2\}, (61)

and the term ΔΓ\Delta\Gamma in (IV-B) is approximated as

ΔΓ=k=0M1[Γ¯(k+1,ϱ1)Γ¯(k+1,ϱ2)]\displaystyle\Delta\Gamma=\sum_{k=0}^{M-1}\left[\bar{\Gamma}(k+1,\varrho_{1})-\bar{\Gamma}(k+1,\varrho_{2})\right]
=k=0M1(eϱ1m=0kϱ1mm!eϱ2m=0kϱ2mm!)\displaystyle=\sum_{k=0}^{M-1}\left(e^{-\varrho_{1}}\sum_{m=0}^{k}\frac{\varrho_{1}^{m}}{m!}-e^{-\varrho_{2}}\sum_{m=0}^{k}\frac{\varrho_{2}^{m}}{m!}\right)
=k=0M1Mkk!(eϱ1ϱ1keϱ2ϱ2k)\displaystyle=\sum_{k=0}^{M-1}\frac{M-k}{k!}\left(e^{-\varrho_{1}}{\varrho_{1}^{k}}-e^{-\varrho_{2}}{\varrho_{2}^{k}}\right)
=MΔΓ(M1,ϱ1,ϱ2)ϱ1Γ¯(M1,ϱ1)+ϱ2Γ¯(M1,ϱ2)\displaystyle=M\Delta\Gamma(M-1,\varrho_{1},\varrho_{2})-\varrho_{1}\bar{\Gamma}(M-1,\varrho_{1})+\varrho_{2}\bar{\Gamma}(M-1,\varrho_{2})
M(ϱ1MM!ϱ2MM!)[ϱ1ϱ1M(M1)!]+[ϱ2ϱ2M(M1)!]\displaystyle\approx M\left(\frac{\varrho_{1}^{M}}{M!}-\frac{\varrho_{2}^{M}}{M!}\right)-\left[\varrho_{1}-\frac{\varrho_{1}^{M}}{(M-1)!}\right]+\left[\varrho_{2}-\frac{\varrho_{2}^{M}}{(M-1)!}\right]
=ϱ2ϱ1=θb2βϕΓb.\displaystyle=\varrho_{2}-\varrho_{1}=\frac{{\theta_{b}}}{2\beta\phi\Gamma_{b}}. (62)

Substituting (61) and (-H) into (IV-B) completes the proof.

References

  • [1]
  • [2] M. Bloch and J. Barros, Physical-Layer Security: From Information Theory to Security Engineering. Cambridge University Press, 2011.
  • [3] A. Mukherjee, S. A. A. Fakoorian, J. Huang, and A. L. Swindlehurst, “Principles of physical layer security in multiuser wireless networks: A survey,” IEEE Commun. Surveys Tuts., vol. 16, no. 3, pp. 1550–1573, Aug. 2014.
  • [4] N. Yang, L. Wang, G. Geraci, M. Elkashlan, J. Yuan, and M. D. Renzo, “Safeguarding 5G wireless communication networks using physical layer security,” IEEE Commun. Mag., vol. 53, no. 4, pp. 20–27, Apr. 2015.
  • [5] Y. Zou, X. Wang, and L. Hanzo, “A survey on wireless security: Technical challenges, recent advances and future trends,” Proc. IEEE, vol. 104, no. 9, pp. 1727–1765, Sep. 2016.
  • [6] H.-M. Wang and T.-X. Zheng, Physical Layer Security in Random Cellular Networks. Singapore: Springer, Oct. 2016.
  • [7] X. Chen, D. W. K. Ng, W. Gerstacker, and H.-H. Chen, “A survey on multiple-antenna techniques for physical layer security,” IEEE Commun. Surveys Tuts., vol. 19, no. 2, pp. 1027–1053, 2nd Quart., 2017.
  • [8] Y. Wu, A. Khisti, C. Xiao, G. Caire, K.-K. Wong, and Xiqi Gao, “A survey of physical layer security techniques for 5G wireless networks and challenges ahead,” IEEE J. Sel. Areas Commun., vol. 36, no. 4, pp. 679–695, Apr. 2018.
  • [9] Y. Deng, L. Wang, S. A. R. Zaidi, J. Yuan, and M. Elkashlan, “Artificial-noise aided secure transmission in large scale spectrum sharing networks,” IEEE Trans. Commun., vol. 64, no. 5, pp. 2116–2129, May 2016.
  • [10] 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.
  • [11] W. Wang, K. C. Teh, and K. H. Li, “Artificial noise aided physical layer security in multi-antenna small-cell networks,” IEEE Trans. Inf. Forensics Security, vol. 12, no. 6, pp. 1470–1482, Jun. 2017.
  • [12] Z. Ding, Z. Zhao, M. Peng, and H. V. Poor, “On the spectral efficiency and security enhancements of NOMA assisted multicast-unicast streaming,” IEEE Trans. Commun., vol. 65, no. 7, pp. 3151–3163, Jul. 2017.
  • [13] S. Yan, N. Yang, I. Land, R. Malaney, J. Yuan, “Three artificial-noise-aided secure transmission schemes in wiretap channels,” IEEE Trans. Veh. Technol., vol. 67, no. 4, pp. 3669–3673, Apr. 2018.
  • [14] A. D. Wyner, “The wire-tap channel,” Bell Syst. Tech. J., vol. 54, no. 8, pp. 1355–1387, Oct. 1975.
  • [15] G. Durisi, T. Koch, and P. Popovski, “Toward massive, ultrareliable, and low-latency wireless communication with short packets,” Proc. IEEE, vol. 104, no. 9, pp. 1711–1726, Sep. 2016.
  • [16] V. W. S. Wong, R. Schober, D. W. K. Ng, and L.-C. Wang, Key Technologies for 5G Wireless Systems. Cambridge, UK: Cambridge University Press, Apr. 2017.
  • [17] C. Cao, H. Li, Z. Hu, W. Liu, and X. Zhang, “Physical-layer secrecy performance in finite blocklength case,” in Proc. IEEE Global Commun. Conf. (GLOBECOM), San Diego, CA, USA, Dec 2015, pp. 1–6.
  • [18] W. Yang, R. F. Schaefer, and H. V. Poor, “Finite-blocklength bounds for wiretap channels,” in Proc. IEEE Int. Symp. Inf. Theory (ISIT), Barcelona, Spain, Jul. 2016, pp. 3087–3091.
  • [19] M. H. Yassaee, M. R. Aref, and A. Gohari, “Non-asymptotic output statistics of random binning and its applications,” in Proc. IEEE Int. Symp. Inf. Theory (ISIT), Istanbul, Turkey, Jul. 2013, pp. 1849–1853.
  • [20] V. Y. F. Tan, “Achievable second-order coding rates for the wiretap channel,” in Proc. IEEE Int. Conf. Comm. Syst. (ICCS), Singapore, Nov. 2012, pp. 65–69.
  • [21] W. Yang, R. F. Schaefer, and H. V. Poor, “Secrecy-reliability tradeoff for semi-deterministic wiretap channels at finite blocklength,” in Proc. IEEE Int. Symp. Inf. Theory (ISIT), Aachen, Germany, Jun. 2017, pp. 2133–2137.
  • [22] X. Zhang, X. Zhou, and M. R. McKay, “On the design of artificial-noise-aided secure multi-antenna transmission in slow fading channels,” IEEE Trans. Veh. Technol., vol. 62, no. 5, pp. 2170–2181, Jun. 2013.
  • [23] T.-X. Zheng, H.-M. Wang, J. Yuan, D. Towsley, and M. H. Lee, “Multi-antenna transmission with artificial noise against randomly distributed eavesdroppers,” IEEE Trans. Commun., vol. 63, no. 11, pp. 4347–4362, Nov. 2015.
  • [24] T.-X. Zheng, H.-M. Wang, Q. Yang, and M. H. Lee, “Safeguarding decentralized wireless networks using full-duplex jamming receivers,” IEEE Trans. Wireless Commun., vol. 16, no. 1, pp. 278–292, Jan. 2017.
  • [25] T.-X. Zheng, H.-M. Wang, J. Yuan, Z. Han, and M. H. Lee, “Physical layer security in wireless ad hoc networks under a hybrid full-/half-duplex receiver deployment strategy,” IEEE Trans. Wireless Commun., vol. 16, no. 6, pp. 3827–3839, Jun. 2017.
  • [26] T.-X. Zheng, H.-M. Wang, and J. Yuan, “Secure and energy-efficient transmissions in cache-enabled heterogeneous cellular networks: Performance analysis and optimization,” IEEE Trans. Commun., vol. 66, no. 11, pp. 5554–5567, Nov. 2018.
  • [27] J. Farhat, G. Brante, and R. D. Souza, “Secure throughput optimization of selective decode-and-forward with finite blocklength,” in Proc. IEEE Veh. Technol. Conf. (VTC Spring), Porto, Portugal, Jul. 2018, pp. 1–5.
  • [28] W. K. Harrison, J. Almeida, M.R. Bloch, S.W. McLaughlin, and J. Barros, “Coding for secrecy: An overview of error-control coding techniques for physical-layer security,” IEEE Signal Process. Mag., vol. 30, no. 5, pp. 41–50, Sep. 2013.
  • [29] D. P. Bertsekas and R. G. Gallager, Data Networks, 2nd ed. Englewood Cliffs, NJ, USA: Prentice–Hall, 1992.
  • [30] Y. Polyanskiy, H. Poor, and S. Verdu, “Channel coding rate in the finite blocklength regime,” IEEE Trans. Inf. Theory, vol. 56, no. 5, pp. 2307–2359, May 2010.
  • [31] P. Gopala, L. Lai, and H. El Gamal, “On the secrecy capacity of fading channels,” IEEE Trans. Inf. Theory, vol. 54, no. 10, pp. 4687–4698, Oct. 2008.
  • [32] B. Makki, T. Svensson and M. Zorzi, “Finite block-length analysis of the incremental redundancy HARQ,” IEEE Wireless Commun. Lett., vol. 3, no. 5, pp. 529–532, Oct. 2014.
  • [33] B. Makki, T. Svensson, and M. Zorzi, “Wireless energy and information transmission using feedback: Infinite and finite block-length analysis,” IEEE Trans. Commun., vol. 64, no. 12, pp. 5304–5318, Dec. 2016.
  • [34] S. Goel and R. Negi, “Guaranteeing secrecy using artificial noise,” IEEE Trans. Wireless Commun., vol. 7, no. 6, pp. 2180–2189, Jun. 2008.
  • [35] A. Khisti and G. W. Wornell, “Secure transmission with multiple antennas I: The MISOME wiretap channel,” IEEE Trans. Inf. Theory, vol. 56, no. 7, pp. 3088–3104, Jul. 2010.
  • [36] A. Khisti and G. W. Wornell, “Secure transmission with multiple antennas–part II: The MIMOME wiretap channel,” IEEE Trans. Inf. Theory, vol. 56, no. 11, pp. 5515–5532, Nov. 2010.
  • [37] P.-H. Lin, S.-H. Lai, S.-C. Lin, and H.-J. Su, “On secrecy rate of the generalized artificial-noise assisted secure beamforming for wiretap channels,” IEEE J. Sel. Areas Commun., vol. 31, no. 9, pp. 1728–1740, Sep. 2013.
  • [38] B. Wang, P. Mu, and Z. Li, “Secrecy rate maximization with artificial-noise-aided beamforming for MISO wiretap channels under secrecy outage constraint,” IEEE Commun. Lett., vol. 19, no. 1, pp. 18–21, Jan. 2015.
  • [39] F. W. J. Olver, D. W. Lozier, R. F. Boisvert, and C. W. Clark, NIST Handbook of Mathematical Functions, Cambridge, UK: Cambridge University Press, 2010.