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

Joint User Scheduling and Computing Resource Allocation Optimization in Asynchronous Mobile Edge Computing Networks

Yihan Cang, Ming Chen, Yijin Pan, Zhaohui Yang,
Ye Hu, Haijian Sun, and Mingzhe Chen
The work of Ming Chen was supported by the National Natural Science Foundation of China (NSFC) under grant 61871128, 61960206005 and 61960206006, by the National Key Research and Development Program under grant 2018YFB1801905, and by Fundamental Research on Foreword Leading Technology of Jiangsu Province under grant BK20192002. The work of Y. Pan was supported by Chongqing Natural Science Joint Fund Project under Grant No. CSTB2023NSCQ-LZX0121. (Corresponding author: Ming Chen). Y. Cang, Ming Chen and Y. Pan are with the National Mobile Communications Research Laboratory, Southeast University, Nanjing 210096, China (e-mails: yhcang@seu.edu.cn, chenming@seu.edu.cn, panyj@seu.edu.cn). Ming Chen is also with the Purple Mountain Laboratories, Nanjing 211100, China. Z. Yang is with College of Information Science and Electronic Engineering, Zhejiang University, Hangzhou 310027, China, and with International Joint Innovation Center, Zhejiang University, Haining 314400, China, and also with Zhejiang Provincial Key Laboratory of Info. Proc., Commun. & Netw. (IPCAN), Hangzhou 310027, China (e-mail: yang_zhaohui@zju.edu.cn). Y. Hu is with the Department of Industrial and System Engineering, University of Miami, Coral Gables, FL, 33146, (e-mail: yehu@miami.edu). H. Sun is with the School of Electrical and Computer Engineering, The University of Georgia, Athens, GA 30602 USA (e-mail: hsun@uga.edu). M. Chen is with the Department of Electrical and Computer Engineering and Institute for Data Science and Computing, University of Miami, Coral Gables, FL, 33146, USA (e-mail: mingzhe.chen@miami.edu).
Abstract

In this paper, the problem of joint user scheduling and computing resource allocation in asynchronous mobile edge computing (MEC) networks is studied. In such networks, edge devices will offload their computational tasks to an MEC server, using the energy they harvest from this server. To get their tasks processed on time using the harvested energy, edge devices will strategically schedule their task offloading, and compete for the computational resource at the MEC server. Then, the MEC server will execute these tasks asynchronously based on the arrival of the tasks. This joint user scheduling, time and computation resource allocation problem is posed as an optimization framework whose goal is to find the optimal scheduling and allocation strategy that minimizes the energy consumption of these mobile computing tasks. To solve this mixed-integer non-linear programming problem, the general benders decomposition method is adopted which decomposes the original problem into a primal problem and a master problem. Specifically, the primal problem is related to computation resource and time slot allocation, of which the optimal closed-form solution is obtained. The master problem regarding discrete user scheduling variables is constructed by adding optimality cuts or feasibility cuts according to whether the primal problem is feasible, which is a standard mixed-integer linear programming problem and can be efficiently solved. By iteratively solving the primal problem and master problem, the optimal scheduling and resource allocation scheme is obtained. Simulation results demonstrate that the proposed asynchronous computing framework reduces 87.17%87.17\% energy consumption compared with conventional synchronous computing counterpart.

Index Terms:
Mobile edge computing, asynchronous computing, user scheduling, wireless power transfer.

I Introduction

Mobile edge computing (MEC) provides powerful computing ability to edge devices [1, 2]. Numerous works have investigated MEC systems from the perspective of resource allocation. In [3], computing offloading and service caching are jointly optimized in MEC-enabled smart grid to minimize system cost. Work [4] proposed a reverse auction-based offloading and resource allocation scheme in MEC. With the aid of machine learning, a multi-agent deep deterministic policy gradient (MADDPG) algorithm is designed to maximize energy efficiency in [5]. However, deploying computing resources at edge servers of a wireless network faces several challenges. First, due to limited energy of edge servers, they may not be able to provide sufficient computation resource according to devices’ requirements [6, 7]. Second, executing all offloading tasks synchronously requires edge servers to wait the arrival of the task with maximum transmission delay which may not be efficient. Meanwhile, task scheduling sequence is nonnegligible in synchronous task offloading, which will also impact the network loads and task completion [8].

To address the first challenge, wireless power transfer (WPT) technology that exploits energy carried by radio frequency (RF) signals emerges [9]. Instead of using solar and wind sources, ambient RF signals can be a viable new source for energy scavenging. Harvesting energy from the environment provides perpetual energy supplies to wireless devices for tasks offloading [10]. Thus, WPT has been regarded as a promising paradigm for MEC scenarios. Combining WPT with MEC, the authors in [11] proposed a multi-user wireless-powered MEC framework aiming at minimizing the total energy consumption under latency constraints. In [12], considering binary computation offloading, the weighted sum computation rate of all wireless devices was maximized by optimizing computation mode selection and transmission time allocation. The work in [13] proposed a multiple intelligent reflecting surfaces (IRSs) assisted wireless powered MEC system, where the IRSs are deployed to assist both the downlink WPT from the access point (AP) to the wireless devices and the uplink computation offloading. However, the above works [14, 15, 16, 17, 11, 12, 13] assumed that all computational tasks offloaded by users will arrive at the server at the same time and then the server starts to process all tasks simultaneously, which is not efficient and even impractical due to users’ dynamic computational task processing requests [18].

Currently, only a few existing works [18, 19, 20, 21, 22, 23] optimized MEC networks under dynamic computation requests. The work in [18] designed a Whittle index based offloading algorithm to maximize the long-term reward for asynchronous MEC networks where computational tasks arrive randomly. In [19], the authors studied the co-channel interference caused by asynchronous task uploading in NOMA MEC systems. The work in [20] investigated the energy efficient asynchronous task offloading for a MEC system where computational tasks with various latency requirements arrive at different time slots. Task scheduling problem for MEC systems with task interruptions and insertions was studied in [21]. However, the above works [18, 19, 20, 21] that focused on the asynchronous task offloading neglected how the asynchronous task arrival affects the computation at the MEC server. The authors in [22] used a sequential computation method to solve the energy consumption minimization problem under asynchronous task arrivals. The work in [23] designed a computation strategy that only allows a task to be executed after the completion of the previous tasks. Yet, works in [22] and [23] are still constrained by their limited usage of the server computation capacity, and cannot act as resource efficient asynchronous task offloading solutions.

The sequential computation strategy [24] has shown to have the potential to improve the computation resource efficiency and task execution punctuality in an asynchronous MEC network. However, since the computation resource allocation at the server depends on the arrival of the offloaded tasks, the sequential scheduling of the tasks will inevitably affect system performance, which is a fact that has been wildly ignored [22, 25, 26].

The main contribution of this paper is a novel asynchronous MEC framework that jointly schedules tasks and allocates computation resource with optimized system energy efficiency. In brief, our key contributions include:

  • We develop a novel framework to manage computation resource for the sequential computation in asynchronous MEC networks. In particular, we consider a MEC network in which the edge devices sequentially harvest energy for transmission, offload their computational tasks to a MEC server, and then compete for computation source at the server to get their tasks accomplished. To achieve the high energy efficient task execution, a policy needs to be designed for determining the optimal task scheduling sequence, time and computational resource allocation. We pose this joint scheduling and resource allocation problem in an optimization framework and seek to find the strategy which minimizes the energy consumption of the tasks.

  • Then, a general benders decomposition (GBD) based algorithm is proposed to solve the formulated mixed-integer non-linear programming (MINLP) problem which is decomposed into a primal problem that allocates computation resource and time, and a master problem that schedules user tasks. By iteratively solving the primal problem and master problem, the optimal scheduling and resource allocation scheme is obtained.

  • To show the effectiveness of the proposed algorithm, we prove that the optimal energy efficient scheduling and resource allocation scheme also optimizes the task punctuality. Our analytical results also show that the optimal allocation scheme for a given offloading task follows a specific pattern: the computation frequency allocated to each task remains constant initially, then gradually decreases before eventually reaching zero. Notably, all tasks experience a simultaneous decrease, the time of which is given in a closed form, in terms of their required central processing unit (CPU) cycles. Leveraging these identified properties, we introduce a computation resource allocation algorithm that offers a low-complexity solution.

Simulation results demonstrate that the proposed asynchronous computing framework reduces 87.87%87.87\% energy consumption compared with conventional synchronous computing counterpart. Moreover, computational complexity of the proposed computation resource allocation algorithm is reduced by 100100 times compared with conventional interior point method.

The rest of this paper is structured as follows. Section II elaborates system model and problem formulation. In Section III, we investigate the properties of asynchronous frequency allocation with given time allocation and user scheduling. The joint optimization of user scheduling, time allocation, and computation resource allocation is rendered in Section IV. Simulation results are presented in Section V. Finally, Section VI draws the conclusions.

II System Model and Problem Formulation

Refer to caption
Figure 1: An illustration about the flow chart of the ordered TDMA system with asynchronous computing.

Consider a MEC network consisting of one MEC server, and a set 𝒦={1,2,,K}\mathcal{K}=\{1,2,\cdots,K\} of energy harvesting enabled edge devices. Within this network, each device k𝒦k\in\mathcal{K} needs to execute an AkA_{k} bits computational task, and will offload its computational task to the MEC server. As shown in Fig. 1, the devices need to first harvest energy from the server to enable such offloading. Then, using the time division multi-access (TDMA) technique, the devices need to schedule their offloading toward the MEC server. In other words, the computational tasks offloaded by devices will arrive at the MEC server asynchronously. To this end, the MEC server will process each device’s computational task in an asynchronous manner. In particular, the server will process devices’ computational tasks according to the time that it receives each of these computational tasks.

The server and edge devices must complete their computational tasks within a time period TT which is divided into (K+2)\left(K+2\right) time slots. The duration of each time slot n{0,,N+1}n\in\{0,\cdots,N+1\} is represented by Δtn\Delta t_{n}, with N=KN=K. Each device uses one time slot to offload its computational task. Let ak,na_{k,n} be the index to indicate whether device kk offloads its task to the server at time slot tnt_{n}. In particular, if device kk uses time slot tnt_{n} to offload its computational task, we have ak,n=1a_{k,n}=1; otherwise, ak,n=0a_{k,n}=0. Since each device uses only one time slot and each time slot can only be allocated to one device, we have k=1Kak,n=1,n{1,,N}\sum_{k=1}^{K}a_{k,n}=1,\ \forall n\in\{1,\cdots,N\}, and n=1Nak,n=1,k𝒦\sum_{n=1}^{N}a_{k,n}=1,\ \forall k\in\mathcal{K}. Meanwhile, when ak,n=1a_{k,n}=1, device kk will harvest energy from time slot t0t_{0} to tn1t_{n-1}. Once task kk arrives at the MEC server, i.e., at time slot slot tn+1t_{n+1}, the server will process this computational task.

The task computation process of the server and a device jointly completing a computational task kk consists of three stages: 1) energy harvesting, 2) task offloading, and 3) remote computing. Next, we first introduce the process of the energy harvesting, task offloading, and remote computing stages. Then, the problem formulation is given.

II-A Energy Harvesting Model

The path loss model is given by h¯k=A(c4πfcdk)\bar{h}_{k}=A\left(\frac{c}{4\pi f_{c}d_{k}}\right)^{\ell}, where AA represents antenna gain, cc denotes the speed of light, fcf_{c} is the carrier frequency, \ell denotes the path-loss factor, and dkd_{k} represents the distance between device kk and the server [27]. The instant channel gain between device kk and server denoted by hkh_{k}, follows an i.i.d. Rician distribution with line-of-sight (LoS) link gain equal to γh¯k\gamma\bar{h}_{k}, where γ\gamma is Rician factor. If device kk offloads its task at time slot tnt_{n} (i.e., ak,n=1a_{k,n}=1), the harvested energy of device kk is EkH=i=0n1ΔtihkηP0E_{k}^{H}=\sum_{i=0}^{n-1}\Delta{t_{i}}h_{k}\eta P_{0}, where η\eta is the energy harvesting efficiency of each device, which is assumed to be equal for all devices [28]. P0P_{0} denotes the transmit power of the server. Since each device has only a single time slot for task offloading (i.e., there exists only one n𝒩n\in\mathcal{N} such that ak,n=1a_{k,n}=1 for a certain device kk), the energy harvested by device kk can be reformulated by EkH=n=1Ki=0n1ak,nΔtihkηP0E_{k}^{H}=\sum_{n=1}^{K}\sum_{i=0}^{n-1}a_{k,n}\Delta{t_{i}}h_{k}\eta P_{0} (k𝒦)(\forall k\in\mathcal{K}).

II-B Tasks Offloading Model

Based on the monomial offloading power model [20, 29], the transmit power of device kk at its offloading time slot tnt_{n} is

pk,n=λ(rk,n)3hk=λ(Ak)3hk(Δtn)3,k𝒦,n{1,,N},\displaystyle p_{k,n}=\frac{\lambda(r_{k,n})^{3}}{h_{k}}=\frac{\lambda(A_{k})^{3}}{h_{k}(\Delta t_{n})^{3}},\forall k\in\mathcal{K},\forall n\in\{1,\cdots,N\}, (1)

where rk,n=Ak/Δtnr_{k,n}=A_{k}/\Delta t_{n} is the transmission rate, λ>0\lambda>0 is the energy coefficient related to the bandwidth and the noise power, and the order 33 is the monomial order associated with coding scheme. Since the transmit power of devices comes from harvested energy, we have n=1Nak,nΔtnpk,nEkH\sum_{n=1}^{N}a_{k,n}\Delta t_{n}p_{k,n}\leq E_{k}^{H} (k𝒦)(\forall k\in\mathcal{K}).

II-C Computing Model

Refer to caption
Figure 2: Asynchronous computation resource allocation.

The MEC server is equipped with multiple CPUs such that the computational tasks offloaded from different devices can be executed in parallel [30]. Let IkI_{k} be the computation intensity of task kk in terms of CPU cycles per bit. As shown in Fig. 2, to sufficiently utilize asynchronous computing, the computation resource of the server will be reallocated to the computational tasks offloaded from devices at each time slot from t2t_{2} to tN+1t_{N+1}. Intuitively, the first uploading task can occupy the whole computation capacity of the server before the second offloading task arrives, while all the tasks compete for computation resource at time slot tN+1t_{N+1}. At an arbitrary time slot m{2,,N+1}m\in\{2,\cdots,N+1\}, fn,mf_{n,m}, n{1,,m1}\forall n\in\{1,\cdots,m-1\} is set to be the computation resource allocated to the task that arrives at the server at time slot tnt_{n}. Given these definitions, we have n=1m1fn,mFmax\sum_{n=1}^{m-1}f_{n,m}\leq F_{\max} (m=2,,N+1)(\forall m=2,\cdots,N+1), where FmaxF_{\max} represents the maximum computation capacity of the MEC server. To complete the task computation for each device kk, we have n=1Nm=n+1N+1ak,nfn,mΔtmFk\sum_{n=1}^{N}\sum_{m=n+1}^{N+1}a_{k,n}f_{n,m}\Delta t_{m}\geq F_{k} (k𝒦)(\forall k\in\mathcal{K}), where Fk=AkIkF_{k}=A_{k}I_{k} represents the computation cycles of device kk. Besides, the energy consumption of the MEC server for all tasks computation can be formulated by EMEC=n=1Nm=n+1N+1κfn,m3ΔtmE_{MEC}=\sum_{n=1}^{N}\sum_{m=n+1}^{N+1}\kappa f_{n,m}^{3}\Delta t_{m}, where κ\kappa denotes the energy coefficient of the MEC server.

II-D Problem Formulation

Our goal is to minimize the MEC server’s energy consumption of completing the tasks offloaded by all devices, which is formulated as an optimization problem as

minΔ𝒕,𝑨,𝒇\displaystyle\min_{\Delta\boldsymbol{t},\boldsymbol{A},\boldsymbol{f}}\ n=1Nm=n+1N+1κfn,m3Δtm,\displaystyle\sum_{n=1}^{N}\sum_{m=n+1}^{N+1}\kappa f_{n,m}^{3}\Delta t_{m}, (2)
s.t. n=1Nak,nλ(Ak)3hk(Δtn)2n=1Ni=0n1ak,nΔtihkηP0,k𝒦,\displaystyle\sum_{n=1}^{N}a_{k,n}\frac{\lambda(A_{k})^{3}}{h_{k}(\Delta t_{n})^{2}}\leq\sum_{n=1}^{N}\sum_{i=0}^{n-1}a_{k,n}\Delta t_{i}h_{k}\eta P_{0},\forall k\in\mathcal{K}, (2a)
n=1m1fn,mFmax,m=2,,N+1,\displaystyle\sum_{n=1}^{m-1}f_{n,m}\leq F_{\max},\quad\forall m=2,\cdots,N+1, (2b)
n=1Nm=n+1N+1ak,nfn,mΔtmAkIk,k𝒦,\displaystyle\sum_{n=1}^{N}\sum_{m=n+1}^{N+1}a_{k,n}f_{n,m}\Delta t_{m}\geq A_{k}I_{k},\quad\forall k\in\mathcal{K}, (2c)
i=0N+1ΔtiT,\displaystyle\sum_{i=0}^{N+1}\Delta t_{i}\leq T, (2d)
k=1Kak,n=1,n{1,,N},\displaystyle\sum_{k=1}^{K}a_{k,n}=1,\quad\forall n\in\{1,\cdots,N\}, (2e)
n=1Nak,n=1,k𝒦,\displaystyle\sum_{n=1}^{N}a_{k,n}=1,\quad\forall k\in\mathcal{K}, (2f)
ak,n{0,1},k𝒦,n{1,,N},\displaystyle a_{k,n}\in\{0,1\},\quad\forall k\in\mathcal{K},\forall n\in\{1,\cdots,N\}, (2g)

where Δ𝒕=[Δt0,,ΔtN+1]T\Delta\boldsymbol{t}=[\Delta t_{0},\cdots,\Delta t_{N+1}]^{T}, 𝒇=(fn,m)n{1,,N},m{n+1,,N+1}\boldsymbol{f}=(f_{n,m})_{\forall n\in\{1,\cdots,N\},m\in\{n+1,\cdots,N+1\}}, and 𝑨=(ak,n)K×N\boldsymbol{A}=(a_{k,n})_{K\times N}. In (2), (2a) is energy consumption causality constraint; (2b) represents a computational resource allocation constraint; (2c) ensures the completion of task computing; (2d) implies that the execution time of all devices should be less than TT; (2e)-(2g) are user scheduling constraints. Since the discrete user scheduling variables ak,na_{k,n} and continuous resource allocation variables Δti\Delta t_{i}, fn,mf_{n,m} are highly coupled, problem (2) is a standard MINLP problem which is difficult to solve. To handle this issue, we first analyze the optimal computation resource allocation with given user scheduling and time allocation in Section III, based on which an efficient low-complexity computation frequency optimization algorithm is proposed. Finally, in Section IV, we propose a GBD-based algorithm to jointly optimize user scheduling and resource allocation so as to solve problem (2). 111 For multi-server edge computing systems, new indicator variables can be introduced to denote the association between tasks and servers. Then the energy minimization problem can be formulated as a MINLP problem containing two kinds of binary optimization variables for task-server association and scheduling sequence, respectively. Despite being more complex, the problem can be solved efficiently using conventional MINLP methods such as convex relaxation and branch-and-bound, or latest approach using machine learning (see e.g., [27]). It is worth noting that with given task-server association, the proposed algorithm in this work is still applicable to scheduling and resource allocation optimization for each server. The detailed transmission protocol and algorithm procedure are left for future works.

III Analysis and Algorithm of the Optimal Computation Resource Allocation

In this section, we first analyze the properties of the optimal computation resource allocation, and then a low-complexity computation resource allocation algorithm is accordingly proposed. For ease of notation, we use FnF_{n} (n{1,,N})(\forall n\in\{1,\cdots,N\}) to represent the computation cycles to complete the task that arrives at the server with order nn. With given time slot allocation vector Δ𝒕\Delta\boldsymbol{t} and user scheduling matrix 𝑨\boldsymbol{A}, problem (2) is simplified as follows:

min{fn,m}\displaystyle\min_{\{f_{n,m}\}}\ n=1Nm=n+1N+1κfn,m3Δtm,\displaystyle\sum_{n=1}^{N}\sum_{m=n+1}^{N+1}\kappa f_{n,m}^{3}\Delta t_{m}, (3)
s.t.\displaystyle s.t.\ \ n=1m1fn,mFmax,m=2,,N+1,\displaystyle\sum_{n=1}^{m-1}f_{n,m}\leq F_{\max},\quad\forall m=2,\cdots,N+1, (3a)
m=n+1N+1fn,mΔtmFn,n{1,,N},\displaystyle\sum_{m=n+1}^{N+1}f_{n,m}\Delta t_{m}\geq F_{n},\quad\forall n\in\{1,\cdots,N\}, (3b)
fn,m0,n{1,,N},m=n+1,,N+1.\displaystyle f_{n,m}\geq 0,\forall n\in\{1,\cdots,N\},\forall m=n+1,\cdots,N+1. (3c)

Before solving problem (3), we provide the feasibility condition as follows.

Proposition 1.

Problem (3) is feasible if and only if Fmaxmaxn{1,,N}i=nNFii=n+1N+1Δti+1F_{\max}\geq\max_{n\in\{1,\cdots,N\}}\frac{\sum_{i=n}^{N}F_{i}}{\sum_{i=n+1}^{N+1}\Delta t_{i+1}}.

Proof.

Please refer to Appendix A. ∎

Denote {αm}\{\alpha_{m}\}, {βn}\{\beta_{n}\}, and {γn,m}\{\gamma_{n,m}\} as the non-negative Lagrangian multipliers associated with the maximum frequency constraints (3a), task computation completion constraints (3b) and non-negative frequency constraints (3c), respectively. The optimal computation resource allocation is given by the following proposition.

Proposition 2.

Given the optimal {αm}\{\alpha^{*}_{m}\}, {βn}\{\beta^{*}_{n}\}, the optimal solution of problem (3) is given by

fn,m=[βn3καm3κΔtm]+,n𝒦,m{2,,K+1}.\displaystyle f^{*}_{n,m}=\sqrt{\left[\frac{\beta^{*}_{n}}{3\kappa}-\frac{\alpha^{*}_{m}}{3\kappa\Delta t_{m}}\right]^{+}},\forall n\in\mathcal{K},\forall m\in\{2,\cdots,K+1\}. (4)
Proof.

Since (4) can be effectively obtained by solving Karush-Kuhn-Tucker (KKT) and Slater conditions, the proofs is omitted here. ∎

According to Propostion 2, we can use the sub-gradient method to obtain the optimal {αm}\{\alpha^{*}_{m}\} and {βn}\{\beta^{*}_{n}\} so as to acquire the optimal computation resource allocation. To further reduce the computational complexity and provide some design insights, the properties of the optimal solution of problem (3) are summarized in the following theorem.

Theorem 3.

Denote ϝ(i)=n=1i2Fnm=n+1K+1Δtm+n=i1KFnm=iK+1Δtm\digamma(i)=\sum_{n=1}^{i-2}\frac{F_{n}}{\sum_{m=n+1}^{K+1}\Delta t_{m}}+\frac{\sum_{n=i-1}^{K}F_{n}}{\sum_{m=i}^{K+1}\Delta t_{m}} (2iK+1)(2\leq i\leq K+1). The optimal computation resource has the following properties:

  • 1)

    The optimal solution of problem (3) satisfies fn,n+1==fn,i>>fn,j==fn,K+1=0,f_{n,n+1}^{*}=\cdots=f_{n,i}^{*}>\cdots>f^{*}_{n,j}=\cdots=f_{n,K+1}^{*}=0, (n+1i<jK+1)(n+1\leq i<j\leq K+1), where tit_{i} is referred as “transition point”.

  • 2)

    The optimal {αm}\{\alpha^{*}_{m}\} satisfies 0=α2Δt2==αiΔti<αi+1Δti+1<αK+1ΔtK+10=\frac{\alpha^{*}_{2}}{\Delta t_{2}}=\cdots=\frac{\alpha^{*}_{i}}{\Delta t_{i}}<\frac{\alpha^{*}_{i+1}}{\Delta t_{i+1}}\cdots<\frac{\alpha_{K+1}^{*}}{\Delta t_{K+1}}.

  • 3)

    The transition point is tit_{i} (3iK+1)(3\leq i\leq K+1) if and only if ϝ(i1)Fmax<ϝ(i)\digamma(i-1)\leq F_{\max}<\digamma(i).

Proof.

The proofs of 1), 2) and 3) are provided in Appendix B, C, D, respectively. ∎

According to property 1) in Theorem 3, the optimal frequency allocation scheme for a certain offloading task always follows a specific pattern: the frequency allocated to each device remains constant initially, then gradually decreases and eventually reaches zero. This property motivates us to deduce the condition fn,i>fn,i+1f^{*}_{n,i}>f^{*}_{n,i+1}. The property 2) in Theorem 3 implies that the computation resource of the server is redundant at time slots from t2t_{2} to tit_{i}, while the maximum computation resource FmaxF_{\max} is utilized at slots from ti+1t_{i+1} to tK+1t_{K+1}.

According to property 1) in Theorem 3, unless fn,n+1==fn,K+1f^{*}_{n,n+1}=\cdots=f^{*}_{n,K+1} (n𝒦)(\forall n\in\mathcal{K}), there always exists a special time slot tϰt_{\varkappa} we called “transition point” such that fn,n+1==fn,ϰ1>fn,ϰfn,K+1f^{*}_{n,n+1}=\cdots=f^{*}_{n,\varkappa-1}>f^{*}_{n,\varkappa}\geq\cdots\geq f^{*}_{n,K+1} (3ϰK+1)(3\leq\varkappa\leq K+1). The transition point indicates the number of time slots that the computation resource remains the same. The computation resource decreases for all tasks at the transition point. The method to find out the transition point when it exists is given by property 3) in Theorem 3.

Property 3) in Theorem 3 also shows that the transition point is impacted by the computation ability of the server. We can directly determine the transition point tϰt_{\varkappa} utilizing property 3) in Theorem 3 without the need of solving problem (3). After determining the transition point, we have αm=0\alpha^{*}_{m}=0 (2mϰ1)(2\leq m\leq\varkappa-1) according to property 2) in Theorem 3.

Refer to caption
Figure 3: Properties of the optimal computation resource allocation.

Fig. 3 depicts an illustration of properties in the optimal computation resource allocation. As can be seen, before the transition point tϰt_{\varkappa}, the optimal αm=0\alpha^{*}_{m}=0 and fn,mf^{*}_{n,m} keeps unchanged as m=2,,ϰ1m=2,\cdots,\varkappa-1. Based on Theorem 3, a low-complexity algorithm is proposed in Algorithm 1. First, we check the feasibility of problem (3) according to Proposition 1. Then, we determine the transition point tϰt_{\varkappa} based on property 3) in Theorem 3. If there is no transition points, which means that the computation resource of server is abundant, we can directly obtain the optimal solution fn,n+1==fn,K+1=Fnm=n+1K+1Δtmf^{*}_{n,n+1}=\cdots=f^{*}_{n,K+1}=\frac{F_{n}}{\sum_{m=n+1}^{K+1}\Delta t_{m}} for n=1,,Kn=1,\cdots,K; otherwise, we obtain the transition point tϰt_{\varkappa} and have αm=0\alpha^{*}_{m}=0 for m=2,,ϰ1m=2,\cdots,\varkappa-1. Hence, we only need to find out the optimal αm\alpha^{*}_{m} (m=ϰ,,K+1)(m=\varkappa,\cdots,K+1) and βn\beta^{*}_{n} (n=1,,K)(n=1,\cdots,K). Note that with given βn\beta^{*}_{n} (n=1,,K)(n=1,\cdots,K), we can obtain the optimal αm\alpha^{*}_{m} by solving the following (Kϰ+2)(K-\varkappa+2) equalities

G(m,αm)n=1m1\displaystyle G(m,\alpha_{m})\triangleq\sum_{n=1}^{m-1} [βn3καm3κΔtm]+=Fmax,\displaystyle\sqrt{\left[\frac{\beta_{n}}{3\kappa}-\frac{\alpha_{m}}{3\kappa\Delta t_{m}}\right]^{+}}=F_{\max},
(m=ϰ,,K+1),\displaystyle\quad\quad\quad(m=\varkappa,\cdots,K+1), (5)

since the maximum frequency is utilized at time slots from tϰt_{\varkappa} to tK+1t_{K+1}. Since G(m,αm)G(m,\alpha_{m}) decreases with respect to αm\alpha_{m}, the bisection method is adopted. It should be noticed that G(m,αm)G(m,\alpha_{m}) achieves the maximum value of n=1m1βn3κ\sum_{n=1}^{m-1}\sqrt{\frac{\beta_{n}}{3\kappa}} when αm=0\alpha_{m}=0 and the minimum value of 0 when αmΔtmmaxn=1,,m1βn\frac{\alpha_{m}}{\Delta t_{m}}\geq\max_{n=1,\cdots,m-1}\beta_{n}. Therefore, the upper bound of αmΔtm\frac{\alpha_{m}}{\Delta t_{m}} is set as αmubΔtm=maxn=1,,m1βn\frac{\alpha_{m}^{ub}}{\Delta t_{m}}=\max_{n=1,\cdots,m-1}\beta_{n}. For the lower bound, we set αmlbΔtm=αm1Δtm1\frac{\alpha_{m}^{lb}}{\Delta t_{m}}=\frac{\alpha_{m-1}}{\Delta t_{m-1}} according to property 2) in Theorem 3. After obtaining αm\alpha_{m} for m=ϰ,,K+1m=\varkappa,\cdots,K+1, βn\beta_{n} is updated by a sub-gradient method [31], where ϕn\phi_{n} is the dynamically chosen step-size. Through repeating Steps 5 to 13 until the objective of (3) converges, we can obtain the optimal αm\alpha^{*}_{m} for m=2,,K+1m=2,\cdots,K+1 and βn\beta^{*}_{n} for n=1,,Kn=1,\cdots,K.

The complexity of Algorithm 1 is 𝒪((K+2ϰ)ϵ1log2(1ϵ0))\mathcal{O}\left(\frac{(K+2-\varkappa)}{\sqrt{\epsilon_{1}}}\log_{2}(\frac{1}{\epsilon_{0}})\right), where ϵ0\epsilon_{0} denotes the accuracy of the bisection method and ϵ1\epsilon_{1} is the accuracy of the objective of problem (3). Compared with the complexity of 𝒪((K2+K)3.5)\mathcal{O}\left(\left(K^{2}+K\right)^{3.5}\right) by the interior point method, the complexity of the proposed algorithm is significantly reduced. Moreover, when ϰ\varkappa is large, the complexity can be further reduced since more numbers of αm\alpha^{*}_{m} are zeros.

1
2If Fmaxmaxn{1,,K}i=nKFii=n+1K+1Δti+1F_{\max}\geq\max_{n\in\{1,\cdots,K\}}\frac{\sum_{i=n}^{K}F_{i}}{\sum_{i=n+1}^{K+1}\Delta t_{i+1}}, go to Step 2; otherwise, problem (3) is infeasible.
3According to Theorem 3, if there is no transition point, the optimal solution is given by fn,n+1==fn,K+1=Fnm=n+1K+1Δtmf^{*}_{n,n+1}=\cdots=f^{*}_{n,K+1}=\frac{F_{n}}{\sum_{m=n+1}^{K+1}\Delta t_{m}} for n=1,,Kn=1,\cdots,K; otherwise, obtain transition point tϰt_{\varkappa} and let αm=0\alpha^{*}_{m}=0 for m=2,,ϰ1m=2,\cdots,\varkappa-1.
4Initialize βn=(Fnm=n+1K+1Δtm)2\beta_{n}=\left(\frac{F_{n}}{\sum_{m=n+1}^{K+1}\Delta t_{m}}\right)^{2} for n=1,,Kn=1,\cdots,K and required precision ϵ0\epsilon_{0}.
5repeat
6      
7      for m=ϰ,,K+1m=\varkappa,\cdots,K+1 do
8             Let αmlbΔtm=αm1Δtm1\frac{\alpha_{m}^{lb}}{\Delta t_{m}}=\frac{\alpha_{m-1}}{\Delta t_{m-1}} and αmubΔtm=maxn=1,,m1βn\frac{\alpha_{m}^{ub}}{\Delta t_{m}}=\max_{n=1,\cdots,m-1}\beta_{n}.
9            while αmubΔtmαmlbΔtm>ϵ0\frac{\alpha_{m}^{ub}}{\Delta t_{m}}-\frac{\alpha_{m}^{lb}}{\Delta t_{m}}>\epsilon_{0} do
10                   Set αmΔtm(αmlbΔtm+αmubΔtm)/2\frac{\alpha_{m}}{\Delta t_{m}}\leftarrow(\frac{\alpha_{m}^{lb}}{\Delta t_{m}}+\frac{\alpha_{m}^{ub}}{\Delta t_{m}})/2.
11                  Calculate fn,m=[βn3καm3κΔtm]+f_{n,m}=\sqrt{\left[\frac{\beta_{n}}{3\kappa}-\frac{\alpha_{m}}{3\kappa\Delta t_{m}}\right]^{+}}.
12                  If n=1m1fn,m>Fmax\sum_{n=1}^{m-1}f_{n,m}>F_{\max}, let αmlbtmαmΔtm\frac{\alpha_{m}^{lb}}{t_{m}}\leftarrow\frac{\alpha_{m}}{\Delta t_{m}}; otherwise, let αmubtmαmΔtm\frac{\alpha_{m}^{ub}}{t_{m}}\leftarrow\frac{\alpha_{m}}{\Delta t_{m}}.
13             end while
14            
15       end for
16      
17      Update βn[βn+ϕn(Fnm=n+1K+1fn,mΔtm)]+\beta_{n}\leftarrow\left[\beta_{n}+\phi_{n}\left(F_{n}-\sum_{m=n+1}^{K+1}f_{n,m}\Delta t_{m}\right)\right]^{+} for n=1,,Kn=1,\cdots,K.
18until the objective of (3) converges;
Output the optimal {fn,m}\{f^{*}_{n,m}\}.
Algorithm 1 Optimal Computation Frequency Allocation Algorithm

IV Joint User Scheduling and Resource Allocation Algorithm

In this section, we employ the GBD method to solve problem (2). The core idea of GBD method is decomposing the original MINLP problem into a primal problem related to continuous variables and a master problem associated with integer variables, which are iteratively solved222Interested readers may refer to [29, 32, 33, 34] for details. . Specifically, for problem (2), the primal problem is a joint communication and computation resource optimization problem with fixed user scheduling. The master problem optimizes user scheduling by utilizing the optimal solutions and dual variables of the primal problem. Next, we describe the detailed procedures.

IV-A Primal Problem

With given user scheduling 𝑨\boldsymbol{A}, problem (2) is reduced to the following optimization problem:

minΔ𝒕,𝒇\displaystyle\min_{\Delta\boldsymbol{t},\boldsymbol{f}}\quad n=1Km=n+1K+1κfn,m3Δtm,\displaystyle\sum_{n=1}^{K}\sum_{m=n+1}^{K+1}\kappa f_{n,m}^{3}\Delta t_{m}, (6)
s.t.\displaystyle s.t.\quad λ(Aπn)3hπn(Δtn)2i=0n1ΔtihπnηP0,n𝒦,\displaystyle\frac{\lambda(A_{\pi_{n}})^{3}}{h_{\pi_{n}}(\Delta t_{n})^{2}}\leq\sum_{i=0}^{n-1}\Delta t_{i}h_{\pi_{n}}\eta P_{0},\quad\forall n\in\mathcal{K}, (6a)
m=n+1K+1fn,mΔtmAπnIπn,n𝒦,\displaystyle\sum_{m=n+1}^{K+1}f_{n,m}\Delta t_{m}\geq A_{\pi_{n}}I_{\pi_{n}},\quad\forall n\in\mathcal{K}, (6b)
(2b),(2d),\displaystyle\text{(\ref{P1}b)},\text{(\ref{P1}d)}, (6c)

where πn\pi_{n} denotes the index of the nn-th offloading device, i.e., we have πn=k\pi_{n}=k if ak,n=1a_{k,n}=1. Since the user scheduling scheme 𝑨\boldsymbol{A} is known, the value of πn,(n𝒦)\pi_{n},(\forall n\in\mathcal{K}) can be deduced and substituted into problem (2). Since problem (6) is non-convex due to the constraints (6a), (6b) and the objective, we introduce xn,m=fn,mΔtm(n=1,,K,m=n+1,,K+1)x_{n,m}=f_{n,m}\Delta t_{m}(\forall n=1,\cdots,K,\forall m=n+1,\cdots,K+1) to represent computation amounts of the nn-th offloading task at time slot mm. Hence, problem (6) is equivalent to

minΔ𝒕,𝒙\displaystyle\min_{\Delta\boldsymbol{t},\boldsymbol{x}}\quad n=1Km=n+1K+1κ(xn,m)3(Δtm)2,\displaystyle\sum_{n=1}^{K}\sum_{m=n+1}^{K+1}\kappa\frac{(x_{n,m})^{3}}{(\Delta t_{m})^{2}}, (7)
s.t.\displaystyle s.t.\quad n=1m1xn,mFmaxΔtm,m=2,,K+1,\displaystyle\sum_{n=1}^{m-1}x_{n,m}\leq F_{\max}\Delta t_{m},\forall m=2,\cdots,K+1, (7a)
m=n+1K+1xn,mAπnIπn,n𝒦,\displaystyle\sum_{m=n+1}^{K+1}x_{n,m}\geq A_{\pi_{n}}I_{\pi_{n}},\quad\forall n\in\mathcal{K}, (7b)
(2d),(6a),\displaystyle\text{(\ref{P1}d)},\text{(\ref{P2}a)}, (7c)

where 𝒙=(xn,m)n𝒦,m{n+1,,K+1}\boldsymbol{x}=(x_{n,m})_{\forall n\in\mathcal{K},m\in\{n+1,\cdots,K+1\}} is the collections of xn,mx_{n,m}. It can be proved that problem (7) is convex utilizing the tricks of perspective function [35]. To further provide useful insights and reduce computation complexity, we utilize the block coordinate decent (BCD) method to iteratively optimize time allocation and computation resource. Since the low-complexity computation resource allocation algorithm with given time allocation has been provided in Algorithm 1, next we propose time allocation algorithm with fixed computation resource allocation.

The Lagrangian function of problem (7) with respect to Δ𝒕\Delta\boldsymbol{t} is given by

=\displaystyle\mathcal{L}= n=1Km=n+1K+1κ(xn,m)3(Δtm)2+n=1Kρn(λ(Aπn)3hπn(Δtn)2i=0n1ΔtihπnηP0)\displaystyle\sum_{n=1}^{K}\sum_{m=n+1}^{K+1}\!\!\kappa\frac{(x_{n,m})^{3}}{(\Delta t_{m})^{2}}\!+\!\sum_{n=1}^{K}\rho_{n}\!\left(\!\frac{\lambda(A_{\pi_{n}})^{3}}{h_{\pi_{n}}(\Delta t_{n})^{2}}\!-\!\sum_{i=0}^{n-1}\Delta t_{i}h_{\pi_{n}}\eta P_{0}\right)
+ξ(i=0K+1ΔtiT)+m=2K+1ωm(n=1m1xn,mFmaxΔtm),\displaystyle+\xi\left(\sum_{i=0}^{K+1}\Delta t_{i}-T\right)+\sum_{m=2}^{K+1}\omega_{m}\left(\sum_{n=1}^{m-1}x_{n,m}-F_{\max}\Delta t_{m}\right), (8)

where ρn\rho_{n}, ωm\omega_{m} and ξ\xi are dual variables related to constraints (6a), (7a) and (2d), respectively. Taking the derivative with respect to Δ𝒕\Delta\boldsymbol{t}, we have

Δt0\displaystyle\frac{\partial\mathcal{L}}{\partial\Delta t_{0}^{*}} =n=1KρnhπnηP0+ξ=0,\displaystyle=-\sum_{n=1}^{K}\rho^{*}_{n}h_{\pi_{n}}\eta P_{0}+\xi^{*}=0, (9)
Δt1\displaystyle\frac{\partial\mathcal{L}}{\partial\Delta t^{*}_{1}} =ρ12λ(Aπ1)3hπ1(Δt1)3n=2KρnhπnηP0+ξ=0,\displaystyle=-\rho^{*}_{1}\frac{2\lambda(A_{\pi_{1}})^{3}}{h_{\pi_{1}}(\Delta t^{*}_{1})^{3}}-\sum_{n=2}^{K}\rho^{*}_{n}h_{\pi_{n}}\eta P_{0}+\xi^{*}=0, (10)
Δti\displaystyle\frac{\partial\mathcal{L}}{\partial\Delta t^{*}_{i}} =n=1i12κ(xn,i)3(Δti)3ρi2λ(Aπi)3hπi(Δti)3n=i+1KρnhπnηP0\displaystyle=-\sum_{n=1}^{i-1}2\kappa\frac{(x_{n,i})^{3}}{(\Delta t^{*}_{i})^{3}}-\rho^{*}_{i}\frac{2\lambda(A_{\pi_{i}})^{3}}{h_{\pi_{i}}(\Delta t^{*}_{i})^{3}}-\sum_{n=i+1}^{K}\rho^{*}_{n}h_{\pi_{n}}\eta P_{0}
+ξωiFmax=0,(2iK1),\displaystyle\quad\ +\xi^{*}-\omega^{*}_{i}F_{\max}=0,\quad(2\leq i\leq K-1), (11)
ΔtK\displaystyle\frac{\partial\mathcal{L}}{\partial\Delta t^{*}_{K}} =n=1K12κ(xn,K)3(ΔtK)3ρK2λ(AπK)3hπK(ΔtK)3+ξωKFmax=0,\displaystyle\!=\!-\!\!\sum_{n=1}^{K-1}\!2\kappa\frac{(x_{n,K})^{3}}{(\Delta t^{*}_{K})^{3}}\!-\!\rho^{*}_{K}\frac{2\lambda(A_{\pi_{K}})^{3}}{h_{\pi_{K}}(\Delta t^{*}_{K})^{3}}\!+\!\xi^{*}\!-\!\omega^{*}_{K}F_{\max}\!=\!0, (12)
ΔtK+1\displaystyle\frac{\partial\mathcal{L}}{\partial\Delta t^{*}_{K+1}} =n=1K2κ(xn,K+1)3(ΔtK+1)3+ξωK+1Fmax=0.\displaystyle=-\sum_{n=1}^{K}2\kappa\frac{(x_{n,K+1})^{3}}{(\Delta t^{*}_{K+1})^{3}}+\xi^{*}-\omega^{*}_{K+1}F_{\max}=0. (13)

Through solving above equations, the optimal Δti\Delta t^{*}_{i} (i)\left(\forall i\right) is obtained in the following proposition.

Proposition 4.

The optimal Δ𝒕\Delta\boldsymbol{t}^{*} is given by

Δt0=Ti=1K+1Δti,\displaystyle\Delta t_{0}^{*}=T-\sum_{i=1}^{K+1}\Delta t_{i}^{*}, (14)
Δti=2κhπin=1i1(xn,i)3+2ρiλ(Aπi)3ξhπin=i+1Kρn(hπn)2ηP0ωihπiFmax3,\displaystyle\Delta t^{*}_{i}=\sqrt[3]{\frac{2\kappa h_{\pi_{i}}\sum_{n=1}^{i-1}(x_{n,i})^{3}+2\rho^{*}_{i}\lambda(A_{\pi_{i}})^{3}}{\xi^{*}h_{\pi_{i}}-\sum_{n=i+1}^{K}\rho^{*}_{n}(h_{\pi_{n}})^{2}\eta P_{0}-\omega^{*}_{i}h_{\pi_{i}}F_{\max}}},
(i{2,,K1}),\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad(\forall i\in\{2,\cdots,K-1\}), (15)
ΔtK=2κhπKn=1K1(xn,K)3+2ρKλ(AπK)3ξhπKωKhπKFmax3,\displaystyle\Delta t^{*}_{K}=\sqrt[3]{\frac{2\kappa h_{\pi_{K}}\sum_{n=1}^{K-1}(x_{n,K})^{3}+2\rho_{K}^{*}\lambda(A_{\pi_{K}})^{3}}{\xi^{*}h_{\pi_{K}}-\omega^{*}_{K}h_{\pi_{K}}F_{\max}}}, (16)
ΔtK+1=2κn=1K(xn,K+1)3ξωK+1Fmax3,\displaystyle\Delta t_{K+1}^{*}=\sqrt[3]{\frac{2\kappa\sum_{n=1}^{K}(x_{n,K+1})^{3}}{\xi^{*}-\omega^{*}_{K+1}F_{\max}}}, (17)

and Δt1\Delta t_{1}^{*} is the null point of Ψ(x)\Psi(x), where Ψ(x)=(Ti=1K+1Δti)x2(hπ1)2ηP0λ(Aπ1)3\Psi(x)=\left(T-\sum_{i=1}^{K+1}\Delta t_{i}\right)x^{2}(h_{\pi_{1}})^{2}\eta P_{0}-\lambda(A_{\pi_{1}})^{3}, x(0,23(Ti=2K+1Δti)]x\in\Big{(}0,\frac{2}{3}\left(T-\sum_{i=2}^{K+1}\Delta t_{i}\right)\Big{]}.

Proof.

The proof of Proposition 4 is provided in Appendix H. ∎

Through iteratively optimizing time allocation and computation resource allocation, we can obtain the optimal solution of primal problem (7). However, if problem (7) is infeasible, we formulate the corresponding 1\ell_{1}-minimization problem as follows:

minΔ𝒕,𝒙,𝜻>0,𝜾>0\displaystyle\min_{\Delta\boldsymbol{t},\boldsymbol{x},\boldsymbol{\zeta}>0,\boldsymbol{\iota}>0} k=1K(ζk+ιk),\displaystyle\sum_{k=1}^{K}\left(\zeta_{k}+\iota_{k}\right), (18)
s.t.\displaystyle s.t.\quad\ n=1Kak,nλ(Ak)3hk(Δtn)2ζk+n=1Ki=0n1ak,nΔtihkηP0,\displaystyle\sum_{n=1}^{K}a_{k,n}\frac{\lambda(A_{k})^{3}}{h_{k}(\Delta t_{n})^{2}}\!\leq\!\zeta_{k}\!+\!\sum_{n=1}^{K}\sum_{i=0}^{n-1}a_{k,n}\Delta t_{i}h_{k}\eta P_{0},
k𝒦,\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\forall k\in\mathcal{K}, (18a)
ιk+n=1Km=n+1K+1ak,nxn,mAkIk,k𝒦,\displaystyle\iota_{k}\!+\!\sum_{n=1}^{K}\sum_{m=n+1}^{K+1}\!a_{k,n}x_{n,m}\!\geq\!A_{k}I_{k},\forall k\in\mathcal{K}, (18b)
(2d),(7a).\displaystyle\text{(\ref{P1}d)},\text{(\ref{P2.1}a)}. (18c)

Since problem (18) is convex and always feasible, we can use the interior point method to obtain the optimal solution and corresponding dual variables.

Furthermore, we can observe that the solution of primal problem always provides a performance upper bound for problem (2) since user scheduling is fixed. Then the upper bound is updated as UB(j)min{UB(j1),f(j)}UB^{(j)}\leftarrow\min\{UB^{(j-1)},f^{(j)}\}, where f(j)f^{(j)} denotes the objective value of primal problem (6). As can be seen, the upper bound is always non-increasing as iteration proceeds. Subsequently, we construct master problem using the solutions and dual variables of primal problem (7) and feasibility problem (18).

IV-B Master Problem

At each iteration, optimality cut or feasibility cut are added to master problem depending on whether the primal problem is feasible. Denote 𝒥1\mathcal{J}_{1} and 𝒥2\mathcal{J}_{2} as the set of iteration indexes indicating the primal problem is feasible and infeasible, respectively. Specifically, the optimality cut for each j𝒥1j\in\mathcal{J}_{1} of feasible iterations is defined as

θ(𝑨,𝝆(j),𝜷(j))=n=1Km=n+1K+1κ(xn,m(j))3(Δtm(j))2\displaystyle\theta(\boldsymbol{A},\boldsymbol{\rho}^{(j)},\boldsymbol{\beta}^{(j)})=\sum_{n=1}^{K}\sum_{m=n+1}^{K+1}\kappa\frac{\left(x^{(j)}_{n,m}\right)^{3}}{\left(\Delta t^{(j)}_{m}\right)^{2}}
+k=1Kρk(j)(n=1Kak,nλ(Ak)3hk(Δtn(j))2n=1Ki=0n1ak,nΔti(j)hkηP0)\displaystyle+\sum_{k=1}^{K}\rho^{(j)}_{k}\left(\sum_{n=1}^{K}a_{k,n}\frac{\lambda(A_{k})^{3}}{h_{k}\left(\Delta t^{(j)}_{n}\right)^{2}}-\sum_{n=1}^{K}\sum_{i=0}^{n-1}a_{k,n}\Delta t^{(j)}_{i}h_{k}\eta P_{0}\right)
+k=1Kβk(j)(AkIkn=1Km=n+1K+1ak,nxn,m(j)),\displaystyle+\sum_{k=1}^{K}\beta_{k}^{(j)}\left(A_{k}I_{k}-\sum_{n=1}^{K}\sum_{m=n+1}^{K+1}a_{k,n}x^{(j)}_{n,m}\right), (19)

where ρk(j)\rho_{k}^{(j)} and βk(j)\beta_{k}^{(j)} represent the dual variables related to primal problem at the jj-th iteration, xn,m(j)x_{n,m}^{(j)} and Δtm(j)\Delta t^{(j)}_{m} denote the solution of primal problem at the jj-th iteration. The terms irrelavant to 𝑨\boldsymbol{A} are omitted based on complementary slackness theorem [31]. Similarly, the feasibility cut for each j𝒥2j\in\mathcal{J}_{2} of infeasible iterations is defined as

θ^(𝑨,𝝆^(j),𝜷^(j))=k=1Kβ^k(j)(AkIkn=1Km=n+1K+1ak,nx^n,m(j))\displaystyle\hat{\theta}(\boldsymbol{A},\hat{\boldsymbol{\rho}}^{(j)},\hat{\boldsymbol{\beta}}^{(j)})=\sum_{k=1}^{K}\hat{\beta}_{k}^{(j)}\left(A_{k}I_{k}-\sum_{n=1}^{K}\sum_{m=n+1}^{K+1}a_{k,n}\hat{x}^{(j)}_{n,m}\right)
+k=1Kρ^k(j)(n=1Kak,nλ(Ak)3hk(Δt^n(j))2n=1Ki=0n1ak,nΔt^i(j)hkηP0),\displaystyle+\sum_{k=1}^{K}\hat{\rho}^{(j)}_{k}\left(\sum_{n=1}^{K}a_{k,n}\frac{\lambda(A_{k})^{3}}{h_{k}\left(\Delta\hat{t}^{(j)}_{n}\right)^{2}}-\sum_{n=1}^{K}\sum_{i=0}^{n-1}a_{k,n}\Delta\hat{t}^{(j)}_{i}h_{k}\eta P_{0}\right), (20)

where ρ^k(j)\hat{\rho}_{k}^{(j)} and β^k(j)\hat{\beta}_{k}^{(j)} represent the dual variables related to feasibility problem at the jj-th iteration, x^n,m(j)\hat{x}_{n,m}^{(j)} and Δt^m(j)\Delta\hat{t}^{(j)}_{m} denote the solution of feasibility problem at the jj-th iteration. Therefore, master problem is formulated as

min𝑨,ψ\displaystyle\min_{\boldsymbol{A},\psi}\quad ψ,\displaystyle\psi, (21)
s.t.\displaystyle s.t.\quad θ(𝑨,𝝆(j),𝜷(j))ψ,j𝒥1,\displaystyle\theta(\boldsymbol{A},\boldsymbol{\rho}^{(j)},\boldsymbol{\beta}^{(j)})\leq\psi,\quad\forall j\in\mathcal{J}_{1}, (21a)
θ^(𝑨,𝝆^(j),𝜷^(j))0,j𝒥2,\displaystyle\hat{\theta}(\boldsymbol{A},\hat{\boldsymbol{\rho}}^{(j)},\hat{\boldsymbol{\beta}}^{(j)})\leq 0,\quad\forall j\in\mathcal{J}_{2}, (21b)
(2e)(6g).\displaystyle\text{(\ref{P1}e)}-\text{(\ref{P2}g)}. (21c)

In particular, (21a) and (21b) denote the set of hyperplanes spanned by the optimality cut and feasibility cut from the first to the jj-th iteration, respectively. The two different types of cuts are exploited to reduce the search region for the global optimal solution [36]. Master problem (21) is a standard mixed-integer linear programming (MILP) problem, which can be solved by numerical solvers such as Gurobi[37] and Mosek[38]. Since master problem is the relaxing problem of MINLP problem (2), solving master problem provides a performance lower bound for problem (2). The lower bound is given by LB(j)ψLB^{(j)}\leftarrow\psi. Since at each iteration, an additional cut (optimality cut or feasibility cut) is added to master problem which narrows the feasible zone, the lower bound is always non-decreasing. As a consequence, the performance upper bound obtained by primal problem and the performance lower bound obtained by the master problem are non-increasing and non-decreasing w.r.t. the iteration index, respectively. As a result, the performance upper bound and the performance lower bound go to converge [29]. Therefore, through iteratively solving primal problem and master problem, we can obtain the optimal solution when the upper bound and lower bound are sufficiently close [33, 36]. The detailed algorithm is summarized in Algorithm 2.

IV-C Complexity Analysis

The complexity of solving problem (2) by Algorithm 2 lies in solving the primal problem, feasibility problem, and master problem at each iteration. For primal problem, where we iteratively update time allocation variables and frequency variables. The frequency optimization method is given in Algorithm 1, whose complexity is 𝒪(K+2ϰϵ1log2(1/ϵ0))\mathcal{O}\left(\frac{K+2-\varkappa}{\sqrt{\epsilon_{1}}}\log_{2}(1/\epsilon_{0})\right) as analyzed in Section III. The time allocation optimization is according to Proposition 4, whose complexity is estimated as 𝒪(Klog2(T))\mathcal{O}\left(K\log_{2}(T)\right). Therefore, the total complexity of solving primal problem is 𝒪(K+2ϰϵ1log2(1/ϵ0)Klog2(T)L1)\mathcal{O}\left(\frac{K+2-\varkappa}{\sqrt{\epsilon_{1}}}\log_{2}(1/\epsilon_{0})K\log_{2}(T)L_{1}\right), where L1L_{1} denotes the iteration number in the primal problems. For the feasibility problem, the complexity is given by 𝒪(((K+1)K2+3K+1)3.5)\mathcal{O}\left(\left(\frac{(K+1)K}{2}+3K+1\right)^{3.5}\right) by the interior point method. For the master problem, the computational complexity is 𝒪(2K)\mathcal{O}\left(2^{K}\right) by the Branch and Bound (BnB) method [39].

1
2Initialize arbitrary feasible user scheduling 𝑨(j)\boldsymbol{A}^{(j)}, and set j=1j=1, UB=+UB=+\infty, LB=LB=-\infty, 𝒥1=𝒥2=\mathcal{J}_{1}=\mathcal{J}_{2}=\emptyset;
3repeat
4      
5      if problem (7) is feasible then
6            
7            repeat
8                  
9                  Obtain the optimal computation resource 𝒇(j)\boldsymbol{f}^{(j)} and dual variable βk(j)\beta_{k}^{(j)} according to Algorithm 1;
10                  Obtain the optimal time allcation Δ𝒕(j)\Delta\boldsymbol{t}^{(j)} and dual variable ρk(j)\rho_{k}^{(j)};
11            until the objective of (7) converges;
12            Update UBUB and 𝒥1\mathcal{J}_{1};
13      else
14             Solve feasibility problem (18) and update 𝒥2\mathcal{J}_{2};
15            Obtain the corresponding optimal solution 𝒙^(j)\hat{\boldsymbol{x}}^{(j)} and Δ𝒕^(j)\Delta\hat{\boldsymbol{t}}^{(j)} as well as dual variables ρ^k(j)\hat{\rho}_{k}^{(j)} and β^K(j)\hat{\beta}_{K}^{(j)};
16       end if
17      
18      Solve master problem (21) by adding optimality cuts (IV-B) and feasibility cuts (IV-B);
19      Set jj+1j\leftarrow j+1;
20      Update 𝑨(j)\boldsymbol{A}^{(j)} and LBLB;
21until UBUB and LBLB are sufficiently close;
Output the optimal 𝒇\boldsymbol{f}^{*}, Δ𝒕\Delta\boldsymbol{t}^{*} and 𝑨\boldsymbol{A}^{*}.
Algorithm 2 Joint User Scheduling and Resource Allocation Algorithm

V Simulations

In this section, we perform simulations to validate the proposed scheme and algorithm. There are K=10K=10 devices around the server. The task size AkA_{k} and computation intensity obey uniform distribution on [10,50][10,50] Kbits and [500,1500][500,1500] cycles/bit, respectively. The transmit power of BS is P0=3P_{0}=3 W. The energy coefficient of the MEC server and energy conversion factor of devices are set as κ=1026\kappa=10^{-26} and η=0.51\eta=0.51. We set the energy constant of transmission λ=1025\lambda=10^{-25}. Furthermore, the maximum computation resource is Fmax=1F_{\max}=1 GHz and the allowable delay is T=1T=1 second. In channel model, we set antenna gain A=3A=3, carrier frequency fc=915f_{c}=915 MHz, path-loss factor =3\ell=3, speed of light c=3×108c=3\times 10^{8} m/s, and the Rician factor is γ=0.3\gamma=0.3. The following benchmarking schemes are provided:

  • JSORA[26]: The joint sensing-and-offloading resource allocation algorithm, where the allocated frequency for each device keeps unchanged during its computation duration, i.e., fn,n+1==fn,K+1f_{n,n+1}=\cdots=f_{n,K+1} (n𝒦)(\forall n\in\mathcal{K}).

  • JCCRM-Sync[14, 15, 16, 17, 11, 12, 13]: The joint communication and computation resource management algorithm, where the computing of server will not begin until all tasks are received, which is adopted by most of the literature.

  • Random scheduling scheme[25, 22]: We randomly set the user scheduling for offloading.

  • Exhaustive search: We randomly choose multiple initial points for Algorithm 2 and select the smallest result as output. The results of exhaustive search method can be regarded as global optimal solutions.

Besides, all accuracies used in the simulations are set as 10510^{-5} for fairness.

Refer to caption
Figure 4: Energy consumption versus number of devices.
Refer to caption
Figure 5: Energy consumption versus maximum delay.

Fig. 5 demonstrates the energy consumption performance comparisons between different schemes under different numbers of devices. We can observe that the energy consumption of the proposed scheme as well as benchmark schemes increases with the number of devices getting large. This is because that devices have to compete for fixed communication and computation resource. As the number of devices increases, the average transmission time and computation time of each device get small, thus average computation resource becomes large. Therefore, larger energy consumption of server is required in order to finish devices’ tasks within the required delay. Moreover, as can be seen in Fig. 5, the gap between the proposed algorithm and exhaustive search scheme is small. This indicates that the proposed algorithm achieves close-to-optimal solutions. Compared with JSORA scheme, random scheduling scheme and JCCRM-Sync scheme, the proposed scheme achieves 30.88%30.88\%, 19.51%19.51\%, 87.87%87.87\% energy reductions, respectively. This can be explained by that the proposed algorithm can take full advantage of the flexibility of asynchronous computing and user scheduling. Particularly, compared with the proposed scheme, JCCRM-Sync scheme wastes the idle computation resource from time slots t2t_{2} to tKt_{K}. Similarly, JSORA scheme can not make full use of computation resource from time slots t2t_{2} to tKt_{K}. Hence, its performance is better than JCCRM-Sync but worse than the proposed scheme. Additionally, random scheduling, as most of the existing literature does, can not utilize the heterogeneity of tasks size and computation intensity well in MEC networks.

In Fig. 5, we depict the energy consumption curves of different schemes versus the maximum allowable delay. As can be seen, the energy consumption of all schemes decreases as the maximum delay becomes large. This is because as delay gets large, the server has more time to finish tasks. Thus, the fewer computation resource is allowable. Hence, energy consumption can be reduced. From Fig. 5, it can be verified that the proposed algorithm outperforms benchmarking schemes in terms of energy consumption in the considered region of delay, especially in resource-scarce scenarios. This phenomenon can be observed in Fig. 5 and Fig. 5 that the difference in energy consumption between the proposed algorithm and benchmark schemes gets small when resource is abundant. This is because the flaws of benchmark schemes compared with the proposed algorithm can be appropriately compensated by utilizing additional sources. Furthermore, it should be noticed that JSORA scheme is equivalent to the proposed scheme when computation resource is abundant according to Theorem 3.

Refer to caption
(a) Transition point: t3t_{3}
Refer to caption
(b) Transition point: t5t_{5}
Refer to caption
(c) Transition point: t6t_{6}
Refer to caption
(d) No transition point
Figure 6: Illustrations of different transition points under different maximum frequencies.

Fig. 6 illustrates a specific case of the allocated frequency of each device at each computation slot under different maximum computation frequencies when K=5K=5. It can be seen that as the maximum frequency FmaxF_{\max} becomes large, the transition point is gradually postponed, and finally no transition point exists when computation resource is sufficiently large which is in accordance with Theorem 3. Specifically, for each subfigure, we can find that before the transition point, the allocated frequency for each device being computed remains unchanged and the maximum frequency constraints do not work. From the transition point to the end, the allocated frequency for each device becomes small and the maximum frequency of the server is used. This verifies Theorem 3.

Refer to caption
Figure 7: Comparisons of computational complexity with K=10K=10.

Fig. 7 depicts the computational complexity comparisons between the proposed Algorithm 1 and the interior point method under different transition points. As can be seen, the computational complexity of Algorithm 1 is significantly reduced compared with the interior point method, by more than 100100 times on average. As the transition point becomes larger, the complexity further decreases. For example, when the transition point ϰ=11\varkappa=11, the complexity of Algorithm 1 is reduced by 10001000 times. This is because the proposed computation resource allocation algorithm fully utilizes the properties in Theorem 3 to reduce algorithm complexity, especially when the computation resource of the server is abundant.

Refer to caption
Figure 8: Energy consumption versus minimum task size.

To test the compatibility of the proposed algorithm under different task scale differences, the energy consumption versus minimum task size AminA_{\min} is shown in Fig. 8, where task size obeys uniform distribution on [Amin,Amax][A_{\min},A_{\max}] with fixed mean value Amin+Amax2=30\frac{A_{\min}+A_{\max}}{2}=30 Kbits. With a large AminA_{\min}, the task scale difference is small. In Fig. 8, the proposed scheme and exhaustive search scheme achieve nearly the same performance, and outperform other schemes. One can observe that the energy consumption increases as task scale difference gets large. This can be explained by that the resources have to be tilted towards the devices with large task sizes, thus resulting in more energy consumption.

VI Conclusion

In this paper, we have investigated a joint user scheduling and resource optimization framework for MEC networks with asynchronous computing. An optimization problem of joint user scheduling, communication and computation resource management has been solved aiming to minimize the energy consumption of server under the delay constraint. Simulations verified that the proposed algorithm yields significant performance gains compared with benchmark schemes. This work establishes a new principle of asynchronous computing and verifies the superiority over its synchronous counterpart. For future works, we will generalize the proposed asynchronous computing framework to heterogeneous task deadlines scenarios so as to further activate its potential. As another direction, the extension to online algorithm design and accommodate new coming devices deserve further investigation.

Appendix A Proof of Proposition 1

The feasibility problem of (3) can be constructed as

minfn,m\displaystyle\min_{f_{n,m}}\quad maxm{2,,K+1}n=1m1fn,m,\displaystyle\max_{m\in\{2,\cdots,K+1\}}\sum_{n=1}^{m-1}f_{n,m}, (A.1)
s.t.\displaystyle s.t.\quad m=n+1K+1fn,mΔtmFn,n𝒦,\displaystyle\sum_{m=n+1}^{K+1}f_{n,m}\Delta t_{m}\geq F_{n},\quad\forall n\in\mathcal{K}, (A.1a)
fn,m0,n𝒦,m=n+1,,K+1.\displaystyle f_{n,m}\geq 0,\quad\forall n\in\mathcal{K},\forall m=n+1,\cdots,K+1. (A.1b)

If the optimal objective of problem (A.1) is less than or equal to FmaxF_{\max}, problem (3) is feasible; otherwise, it is infeasible. Subsequently, we analyze the optimal solution of problem (A.1). First, when K=1K=1, i.e., there exists only one task, the optimal objective of problem (A.1) is FKΔtK+1\frac{F_{K}}{\Delta t_{K+1}}. When K=2K=2, we consider two cases: 1) If F1Δt2F2Δt3\frac{F_{1}}{\Delta t_{2}}\leq\frac{F_{2}}{\Delta t_{3}}, this indicates that the optimal scheme is computing task 22 after task 11 is finished. Therefore, the optimal solution is f1,2=F1Δt2f_{1,2}=\frac{F_{1}}{\Delta t_{2}} and f2,3=F2Δt3f_{2,3}=\frac{F_{2}}{\Delta t_{3}}. Hence, the optimal objective is F2Δt3\frac{F_{2}}{\Delta t_{3}}. 2) If F1Δt2>F2Δt3\frac{F_{1}}{\Delta t_{2}}>\frac{F_{2}}{\Delta t_{3}}, this implies that part of task 11 can be processed in parallel with task 22. Hence, the optimal solution is given by f1,2=f1,3+f2,3=F1+F2Δt2+Δt3f_{1,2}=f_{1,3}+f_{2,3}=\frac{F_{1}+F_{2}}{\Delta t_{2}+\Delta t_{3}}. Since F1+F2Δt2+Δt3>F2Δt3\frac{F_{1}+F_{2}}{\Delta t_{2}+\Delta t_{3}}>\frac{F_{2}}{\Delta t_{3}}, the optimal objective is F1+F2Δt2+Δt3\frac{F_{1}+F_{2}}{\Delta t_{2}+\Delta t_{3}}. In conclusion, when K=2K=2, the optimal objective is max{F1+F2Δt2+Δt3,F2Δt3}\max\{\frac{F_{1}+F_{2}}{\Delta t_{2}+\Delta t_{3}},\frac{F_{2}}{\Delta t_{3}}\}. Similarly, by recursion, we can deduce that when there exist KK TD, the optimal solution is max{F1++FKΔt2++ΔtK+1,F2++FKΔt3++ΔtK+1,,FKΔtK+1}\max\{\frac{F_{1}+\cdots+F_{K}}{\Delta t_{2}+\cdots+\Delta t_{K+1}},\frac{F_{2}+\cdots+F_{K}}{\Delta t_{3}+\cdots+\Delta t_{K+1}},\cdots,\frac{F_{K}}{\Delta t_{K+1}}\}. That completes the proof. \blacksquare

Appendix B Proof of Property 1) in Theorem 3

Before that, we give the following two corollaries to facilitate the proof.

Corollary 5.

[Row property] The optimal computation resource of each task is non-increasing during its computation period, i.e., fn,n+1fn,n+2fn,K+1f_{n,n+1}^{*}\geq f_{n,n+2}^{*}\geq\cdots\geq f_{n,K+1}^{*} (n𝒦)(\forall n\in\mathcal{K}).

Proof.

Please refer to Appendix E. ∎

Corollary 6.

[Column property] Denote the sum computation cycles of the nn-th offloading device in time slots tmt_{m} and tm+1t_{m+1} by FnF_{n} for m=3,,Km=3,\cdots,K and n=1,,m1n=1,\cdots,m-1. If Fn>0F_{n}>0 holds for all n=1,,m1n=1,\cdots,m-1, the optimal frequency shifts δnfn,mfn,m+1\delta^{*}_{n}\triangleq f^{*}_{n,m}-f^{*}_{n,m+1} (n=1,,m1)(n=1,\cdots,m-1) are either all zeros or all positive, i.e., have the coincident zero or positive characteristics.

Proof.

Please refer to Appendix F. ∎

First, applying the KKT conditions gives

3κ(fn,m)2Δtm+αmβnΔtmγn,m=0,\displaystyle 3\kappa(f^{*}_{n,m})^{2}\Delta t_{m}+\alpha_{m}-\beta_{n}\Delta t_{m}-\gamma_{n,m}=0, (B.1)
αm(n=1m1fn,mFmax)=0,\displaystyle\alpha_{m}\left(\sum_{n=1}^{m-1}f^{*}_{n,m}-F_{\max}\right)=0, (B.2)
βn(Fnm=n+1K+1fn,mΔtm)=0,\displaystyle\beta_{n}\left(F_{n}-\sum_{m=n+1}^{K+1}f^{*}_{n,m}\Delta t_{m}\right)=0, (B.3)
γn,mfn,m=0,n𝒦,\displaystyle\gamma_{n,m}f^{*}_{n,m}=0,\quad\forall n\in\mathcal{K}, (B.4)
αm0,βn0,γn,m0.\displaystyle\alpha_{m}\!\geq\!0,\beta_{n}\!\geq\!0,\gamma_{n,m}\!\geq\!0. (B.5)

Based on (B.1), we obtain that

fn,m=βn3κ+γn,mαm3κΔtm.\displaystyle f^{*}_{n,m}=\sqrt{\frac{\beta_{n}}{3\kappa}+\frac{\gamma_{n,m}-\alpha_{m}}{3\kappa\Delta t_{m}}}. (B.6)

In case of fn,m>0f_{n,m}>0 (n𝒦,m=n+1,,K+1)(\forall n\in\mathcal{K},\forall m=n+1,\cdots,K+1), we have γn,m=0\gamma_{n,m}=0 according to (B.4). Furthermore, βn>0\beta_{n}>0 is derived from (B.6). According to Corollary 5, the optimal solution satisfies fn,mfn,m+1f^{*}_{n,m}\geq f^{*}_{n,m+1}. Thus, we have αmΔtmαm+1Δtm+1\frac{\alpha_{m}}{\Delta t_{m}}\leq\frac{\alpha_{m+1}}{\Delta t_{m+1}}. Assume that there exists a certain i{3,,K}i\in\{3,\cdots,K\} such that fn,i=fn,i+1f^{*}_{n,i}=f^{*}_{n,i+1}. We have αiΔti=αi+1Δti+1\frac{\alpha_{i}}{\Delta t_{i}}=\frac{\alpha_{i+1}}{\Delta t_{i+1}}. If αiΔti=αi+1Δti+1>0\frac{\alpha_{i}}{\Delta t_{i}}=\frac{\alpha_{i+1}}{\Delta t_{i+1}}>0, i.e., αi>0\alpha_{i}>0 and αi+1>0\alpha_{i+1}>0, we should have n=1i1fn,i=n=1ifn,i+1=Fmax\sum_{n=1}^{i-1}f_{n,i}=\sum_{n=1}^{i}f_{n,i+1}=F_{\max} according to (B.2). Furthermore, due to fn,m>0f_{n,m}>0 (n𝒦,m=n+1,,K+1)(\forall n\in\mathcal{K},\forall m\in=n+1,\cdots,K+1), the computation cycles fn,iΔti+fn,i+1Δti+1>0f_{n,i}\Delta t_{i}+f_{n,i+1}\Delta t_{i+1}>0 (n1,,i1)(\forall n\in 1,\cdots,i-1). According to Corollary 6, we have fn,i=fn,i+1f_{n,i}=f_{n,i+1} (n=1,,i1)(n=1,\cdots,i-1). Hence, it can be derived that fi,i+1=0f_{i,i+1}=0 which contradicts that fi,i+1f_{i,i+1} is positive. Therefore, we have αi=αi+1=0\alpha_{i}=\alpha_{i+1}=0, i.e., αiΔti=αi+1Δti+1=0\frac{\alpha_{i}}{\Delta t_{i}}=\frac{\alpha_{i+1}}{\Delta t_{i+1}}=0. Since α2Δt2αiΔti\frac{\alpha_{2}}{\Delta t_{2}}\leq\cdots\leq\frac{\alpha_{i}}{\Delta t_{i}} and α2,,αi10\alpha_{2},\cdots,\alpha_{i-1}\geq 0, we can further obtain that α2==αi+1=0\alpha_{2}=\cdots=\alpha_{i+1}=0. This indicates that if there exists a certain i{3,,K}i\in\{3,\cdots,K\} such that fn,i=fn,i+1>0f^{*}_{n,i}=f^{*}_{n,i+1}>0, we have fn,n+1=fn,n+2==fn,i+1f^{*}_{n,n+1}=f^{*}_{n,n+2}=\cdots=f^{*}_{n,i+1}.

Additionally, if there exists a certain i<jK+1i<j\leq K+1 such that fn,j=0f^{*}_{n,j}=0, we can deduce that fn,j=fn,j+1==fn,K+1=0f^{*}_{n,j}=f^{*}_{n,j+1}=\cdots=f_{n,K+1}^{*}=0 since fn,jfn,j+1fn,K+1f^{*}_{n,j}\geq f^{*}_{n,j+1}\geq\cdots\geq f^{*}_{n,K+1}.

Combing the above two cases, we complete the proof. \blacksquare

Appendix C Proof of Property 2) in Theorem 3

According to Appendix B, we can obtain that α2==αi=0\alpha^{*}_{2}=\cdots=\alpha^{*}_{i}=0. Moreover, since fn1,n>fn1,n+10f^{*}_{n-1,n}>f^{*}_{n-1,n+1}\geq 0 (inK)(i\leq n\leq K), we have βn13κ+γn1,nαn3κΔtn>βn13κ+γn1,n+1αn+13κΔtn+1\sqrt{\frac{\beta_{n-1}}{3\kappa}+\frac{\gamma_{n-1,n}-\alpha_{n}}{3\kappa\Delta t_{n}}}>\sqrt{\frac{\beta_{n}-1}{3\kappa}+\frac{\gamma_{n-1,n+1}-\alpha_{n+1}}{3\kappa\Delta t_{n+1}}} according to (B.6). Due to that γn1,n=0\gamma_{n-1,n}=0 and γn1,n+1>0\gamma_{n-1,n+1}>0, we can deduce that αnΔtn>γn1,n+1αn+1Δtn+1>αn+1Δtn+1-\frac{\alpha_{n}}{\Delta t_{n}}>\frac{\gamma_{n-1,n+1}-\alpha_{n+1}}{\Delta t_{n+1}}>-\frac{\alpha_{n+1}}{\Delta t_{n+1}}, i.e., αnΔtn<αn+1Δtn+1\frac{\alpha_{n}}{\Delta t_{n}}<\frac{\alpha_{n+1}}{\Delta t_{n+1}} (inK)(i\leq n\leq K). Overall, we can conclude that 0=α2Δt2==αiΔti<αi+1Δti+1<αK+1ΔtK+10=\frac{\alpha^{*}_{2}}{\Delta t_{2}}=\cdots=\frac{\alpha^{*}_{i}}{\Delta t_{i}}<\frac{\alpha^{*}_{i+1}}{\Delta t_{i+1}}\cdots<\frac{\alpha^{*}_{K+1}}{\Delta t_{K+1}}. \blacksquare

Appendix D Proof of Property 3) in Theorem 3

We first prove the “only if” part. According to property 1) in Theorem 3, if tit_{i} is the transition point, we have fn,n+1==fn,i1>fn,i>fn,j==fn,K+1=0f_{n,n+1}^{*}=\cdots=f_{n,i-1}^{*}>f_{n,i}^{*}\cdots>f^{*}_{n,j}=\cdots=f_{n,K+1}^{*}=0 (n+1i<jK+1)(n+1\leq i<j\leq K+1) and n=1m1fn,m=Fmax\sum_{n=1}^{m-1}f_{n,m}=F_{\max} (imK+1)(i\leq m\leq K+1). Since m=n+1K+1fn,mΔtm=Fn\sum_{m=n+1}^{K+1}f^{*}_{n,m}\Delta t_{m}=F_{n}, we can obtain that m=iK+1fn,mΔtm<Fn(m=iK+1Δtm)m=n+1K+1Δtm\sum_{m=i}^{K+1}f^{*}_{n,m}\Delta t_{m}<\frac{F_{n}(\sum_{m=i}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}} (1ni2)(1\leq n\leq i-2). Thus, we have (m=iK+1Δtm)Fmax=n=1i2m=iK+1fn,mΔtm+n=i1Km=n+1K+1fn,mΔtm<n=1i2Fn(m=iK+1Δtm)m=n+1K+1Δtm+n=i1KFn(\sum_{m=i}^{K+1}\Delta t_{m})F_{\max}=\sum_{n=1}^{i-2}\sum_{m=i}^{K+1}f^{*}_{n,m}\Delta t_{m}+\sum_{n=i-1}^{K}\sum_{m=n+1}^{K+1}f^{*}_{n,m}\Delta t_{m}<\sum_{n=1}^{i-2}\frac{F_{n}(\sum_{m=i}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}}+\sum_{n=i-1}^{K}F_{n}, i.e., Fmax<ϝ(i)F_{\max}<\digamma(i).

Similarly, if ti1t_{i-1} is the transition point, we have Fmax<ϝ(i1)F_{\max}<\digamma(i-1). Since ϝ(i)ϝ(i1)=Fi2m=i1K+1Δtm+n=i1KFnm=iK+1Δtmn=i2KFnm=i1K+1Δtm=n=i1KFnm=iK+1Δtmn=i1KFnm=i1K+1Δtm>0\digamma(i)-\digamma(i-1)=\frac{F_{i-2}}{\sum_{m=i-1}^{K+1}\Delta t_{m}}+\frac{\sum_{n=i-1}^{K}F_{n}}{\sum_{m=i}^{K+1}\Delta t_{m}}-\frac{\sum_{n=i-2}^{K}F_{n}}{\sum_{m=i-1}^{K+1}\Delta t_{m}}=\frac{\sum_{n=i-1}^{K}F_{n}}{\sum_{m=i}^{K+1}\Delta t_{m}}-\frac{\sum_{n=i-1}^{K}F_{n}}{\sum_{m=i-1}^{K+1}\Delta t_{m}}>0, we should have ϝ(i1)Fmax<ϝ(i)\digamma(i-1)\leq F_{\max}<\digamma(i).

For the “if” part, if Fmax<ϝ(i)F_{\max}<\digamma(i), we can deduce that (m=iK+1Δtm)Fmax<n=1i2Fn(m=iK+1Δtm)m=n+1K+1Δtm+n=i1KFn(\sum_{m=i}^{K+1}\Delta t_{m})F_{\max}<\sum_{n=1}^{i-2}\frac{F_{n}(\sum_{m=i}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}}+\sum_{n=i-1}^{K}F_{n}. Moreover, since n=1i2m=iK+1fn,mΔtm+n=i1KFn(m=iK+1Δtm)Fmax\sum_{n=1}^{i-2}\sum_{m=i}^{K+1}f^{*}_{n,m}\Delta t_{m}+\sum_{n=i-1}^{K}F_{n}\leq(\sum_{m=i}^{K+1}\Delta t_{m})F_{\max}, we have n=1i2m=iK+1fn,mΔtm+n=i1KFn<n=1i2Fn(m=iK+1Δtm)m=n+1K+1Δtm+n=i1KFn\sum_{n=1}^{i-2}\sum_{m=i}^{K+1}f^{*}_{n,m}\Delta t_{m}+\sum_{n=i-1}^{K}F_{n}<\sum_{n=1}^{i-2}\frac{F_{n}(\sum_{m=i}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}}+\sum_{n=i-1}^{K}F_{n}, i.e., n=1i2m=iK+1fn,mΔtm<n=1i2Fn(m=iK+1Δtm)m=n+1K+1Δtm\sum_{n=1}^{i-2}\sum_{m=i}^{K+1}f^{*}_{n,m}\Delta t_{m}<\sum_{n=1}^{i-2}\frac{F_{n}(\sum_{m=i}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}}. Since fn,n+1fn,i1fn,K+1f_{n,n+1}^{*}\geq\cdots\geq f_{n,i-1}^{*}\geq\cdots\geq f_{n,K+1}^{*} (n+1i<K+1)(n+1\leq i<K+1), we can obtain that m=iK+1fn,mΔtmFn(m=iK+1Δtm)m=n+1K+1Δtm\sum_{m=i}^{K+1}f^{*}_{n,m}\Delta t_{m}\leq\frac{F_{n}(\sum_{m=i}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}} (1ni2)(1\leq n\leq i-2). Therefore, to let n=1i2m=iK+1fn,mΔtm<n=1i2Fn(m=iK+1Δtm)m=n+1K+1Δtm\sum_{n=1}^{i-2}\sum_{m=i}^{K+1}f^{*}_{n,m}\Delta t_{m}<\sum_{n=1}^{i-2}\frac{F_{n}(\sum_{m=i}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}} hold, we should have m=iK+1fn,mΔtm<Fn(m=iK+1Δtm)m=n+1K+1Δtm\sum_{m=i}^{K+1}f^{*}_{n,m}\Delta t_{m}<\frac{F_{n}(\sum_{m=i}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}} (1ni2)(1\leq n\leq i-2). Further, it can be deduced that m=i1K+1fn,mΔtm<Fn(m=i1K+1Δtm)m=n+1K+1Δtm\sum_{m=i-1}^{K+1}f^{*}_{n,m}\Delta t_{m}<\frac{F_{n}(\sum_{m=i-1}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}} (1ni3)(1\leq n\leq i-3).

Additionally, since ϝ(i1)Fmax\digamma(i-1)\leq F_{\max}, we can deduce that n=1i3Fn(m=i1K+1Δtm)m=n+1K+1Δtm+n=i2KFn(m=i1K+1Δtm)Fmax\sum_{n=1}^{i-3}\frac{F_{n}(\sum_{m=i-1}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}}+\sum_{n=i-2}^{K}F_{n}\leq(\sum_{m=i-1}^{K+1}\Delta t_{m})F_{\max}. Thus, we have n=1i3m=i1K+1fn,mΔtm+n=i2KFn<n=1i3Fn(m=i1K+1Δtm)m=n+1K+1Δtm+n=i2KFn(m=i1K+1Δtm)Fmax\sum_{n=1}^{i-3}\sum_{m=i-1}^{K+1}f^{*}_{n,m}\Delta t_{m}+\sum_{n=i-2}^{K}F_{n}<\sum_{n=1}^{i-3}\frac{F_{n}(\sum_{m=i-1}^{K+1}\Delta t_{m})}{\sum_{m=n+1}^{K+1}\Delta t_{m}}+\sum_{n=i-2}^{K}F_{n}\leq(\sum_{m=i-1}^{K+1}\Delta t_{m})F_{\max}, which indicates that the computation resource is abundant from ti1t_{i-1} to tK+1t_{K+1}. Therefore, we can deduce that fn,n+1==fn,i1f^{*}_{n,n+1}=\cdots=f^{*}_{n,i-1} (ni2)(n\leq i-2). Assume that ti~t_{\tilde{i}} (i~>i)(\tilde{i}>i) is the transition point. We have ϝ(i~1)Fmax\digamma(\tilde{i}-1)\leq F_{\max}. Since ϝ(i~1)ϝ(i)0\digamma(\tilde{i}-1)-\digamma(i)\geq 0, FmaxF_{\max} is infeasible, which breaks the assumption. Therefore, we can conclude that tit_{i} is the transition point.

Combining the proofs of “if” and “only if” part, we complete the proof. \blacksquare

Appendix E Proof of Corollary 5

To find out the optimal computation resource allocation scheme, we first investigate the property of the most energy-efficient scheme without the maximum frequency restriction in Lemma 7, whose proof is provided in Appendix G.

Lemma 7.

Regardless of Δtm\Delta t_{m}, Δtm+1\Delta t_{m+1} and with given computation cycles FnF_{n} in time slots tmt_{m} and tm+1t_{m+1}, scheme fn,m=fn,m+1f^{*}_{n,m}=f^{*}_{n,m+1} consumes the least energy among all the solutions satisfying fn,mΔtm+fn,m+1Δtm+1=Fnf_{n,m}\Delta t_{m}+f_{n,m+1}\Delta t_{m+1}=F_{n}.

For Corollary 5, we first prove that f1,2f1,3f_{1,2}^{*}\geq f_{1,3}^{*} with given computation cycle F1F_{1} and f2,3f_{2,3}. Denote the sum computation cycles in t2t_{2} and t3t_{3} of the first offloading device as F1F_{1}. Through relaxing the maximum computation resource constraint, the energy consumption is the least when f1,2=f1,3=F1Δt2+Δt3f_{1,2}=f_{1,3}=\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}} according to Lemma 7. Since constraint (3a) should be satisfied, we have

f1,2Fmax,f1,3+f2,3Fmax.\displaystyle f_{1,2}\leq F_{\max},\quad f_{1,3}+f_{2,3}\leq F_{\max}. (E.1)

We consider the following two cases; otherwise, f1,2f_{1,2} and f1,3f_{1,3} have no feasible solution with the given F1F_{1}.

Case 1: F1Δt2+Δt3+f2,3Fmax\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}}+f_{2,3}\leq F_{\max}. In this case, we can deduce that f1,2=f1,3=F1Δt2+Δt3f_{1,2}=f_{1,3}=\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}} satisfies (E.1). Since f1,2=f1,3=F1Δt2+Δt3f_{1,2}=f_{1,3}=\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}} is the most energy efficient solution, the optimal solution in this case is f1,2=f1,3=F1Δt2+Δt3f^{*}_{1,2}=f^{*}_{1,3}=\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}}.

Case 2: F1Δt2+Δt3+f2,3>Fmax\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}}+f_{2,3}>F_{\max}. Obviously, f1,2=f1,3=F1Δt2+Δt3f_{1,2}=f_{1,3}=\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}} is infeasible in this case. We then prove that f1,2<f1,3f_{1,2}<f_{1,3} is also impossible. Since f1,2Δt2+f1,3Δt3=F1f_{1,2}\Delta t_{2}+f_{1,3}\Delta t_{3}=F_{1}, f1,2f_{1,2} increases as f1,3f_{1,3} decreases. If f1,2<f1,3f_{1,2}<f_{1,3}, we can deduce that f1,3>F1Δt2+Δt3f_{1,3}>\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}}. Therefore, we have f1,3+f2,3>F1Δt2+Δt3+f2,3>Fmaxf_{1,3}+f_{2,3}>\frac{F_{1}}{\Delta t_{2}+\Delta t_{3}}+f_{2,3}>F_{\max} which violates the maximum frequency constraint. As a consequence, the optimal solution is f1,2>f1,3f_{1,2}^{*}>f_{1,3}^{*}. According to Lemma 7, the energy consumption EE increases with δ1=f1,2f1,3\delta_{1}=f_{1,2}-f_{1,3} in the considered region 0<δ1F1/Δt20<\delta_{1}\leq F_{1}/\Delta t_{2}. Therefore, in order to achieve the fewest energy consumption, we should let δ1\delta_{1} as small as possible. Hence, we can obtain that f1,3=Fmaxf2,3f^{*}_{1,3}=F_{\max}-f_{2,3} and f1,2=F1f1,3Δt3Δt2f^{*}_{1,2}=\frac{F_{1}-f^{*}_{1,3}\Delta t_{3}}{\Delta t_{2}}. The corresponding optimal δ1=F1(Fmaxf2,3)(Δt2+Δt3)Δt2\delta_{1}^{*}=\frac{F_{1}-(F_{\max}-f_{2,3})(\Delta t_{2}+\Delta t_{3})}{\Delta t_{2}}.

Summarizing the above two cases, we can obtain that f1,2f1,3f^{*}_{1,2}\geq f^{*}_{1,3}. Subsequently, we prove that in tmt_{m} and tm+1t_{m+1} (m=3,,K)(m=3,\cdots,K), we always have fn,mfn,m+1f^{*}_{n,m}\geq f^{*}_{n,m+1} for all n=1,,m1n=1,\cdots,m-1.

Denote the sum computation cycles of the nn-th offloading device in time slots tmt_{m} and tm+1t_{m+1} by FnF_{n}, i.e.,

{f1,mΔtm+f1,m+1Δtm+1=F1,fn,mΔtm+fn,m+1Δtm+1=Fn,fm1,mΔtm+fm1,m+1Δtm+1=Fm1.\displaystyle\left\{\begin{aligned} &f_{1,m}\Delta t_{m}+f_{1,m+1}\Delta t_{m+1}=F_{1},\\ &\quad\quad\quad\quad\quad\cdots\\ &f_{n,m}\Delta t_{m}+f_{n,m+1}\Delta t_{m+1}=F_{n},\\ &\quad\quad\quad\quad\quad\cdots\\ &f_{m-1,m}\Delta t_{m}+f_{m-1,m+1}\Delta t_{m+1}=F_{m-1}.\end{aligned}\right. (E.2)

According to Lemma 7, when fn,m=fn,m+1=FnΔtm+Δtm+1f_{n,m}=f_{n,m+1}=\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}} for all n=1,,m1n=1,\cdots,m-1, the minimum energy consumption of the nn-th offloading device can be achieved, thus the total energy consumption is minimum. Moreover, the following constraints should be satisfied:

n=1m1fn,mFmax,n=1mfn,m+1Fmax.\displaystyle\sum_{n=1}^{m-1}f_{n,m}\leq F_{\max},\quad\sum_{n=1}^{m}f_{n,m+1}\leq F_{\max}. (E.3)

We consider two cases.

Case 1: n=1m1FnΔtm+Δtm+1+fm,m+1Fmax\sum_{n=1}^{m-1}\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}}+f_{m,m+1}\leq F_{\max}. In this case, we can deduce that fn,m=fn,m+1=FnΔtm+Δtm+1f_{n,m}=f_{n,m+1}=\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}} for all n=1,,m1n=1,\cdots,m-1 satisfies (E.3). Thus, the optimal solution in this case is fn,m=fn,m+1=FnΔtm+Δtm+1f^{*}_{n,m}=f^{*}_{n,m+1}=\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}} for all n=1,,m1n=1,\cdots,m-1.

Case 2: n=1m1FnΔtm+Δtm+1+fm,m+1>Fmax\sum_{n=1}^{m-1}\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}}+f_{m,m+1}>F_{\max}. Obviously, fn,m=fn,m+1=FnΔtm+Δtm+1f_{n,m}=f_{n,m+1}=\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}} is infeasible in this case. We then prove that n=1m1fn,m>n=1m1fn,m+1\sum_{n=1}^{m-1}f^{*}_{n,m}>\sum_{n=1}^{m-1}f^{*}_{n,m+1} by contradiction. By summing all the equalities in (E.2), we have Δtmn=1m1fn,m+Δtm+1n=1m1fn,m+1=n=1m1Fn\Delta t_{m}\sum_{n=1}^{m-1}f_{n,m}+\Delta t_{m+1}\sum_{n=1}^{m-1}f_{n,m+1}=\sum_{n=1}^{m-1}F_{n}. Thus, n=1m1fn,m\sum_{n=1}^{m-1}f_{n,m} is negatively correlated with n=1m1fn,m+1\sum_{n=1}^{m-1}f_{n,m+1}. If n=1m1fn,m<n=1m1fn,m+1\sum_{n=1}^{m-1}f_{n,m}<\sum_{n=1}^{m-1}f_{n,m+1}, it can be inferred that n=1m1fn,m+1>n=1m1FnΔtm+Δtm+1\sum_{n=1}^{m-1}f_{n,m+1}>\sum_{n=1}^{m-1}\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}}. We can further have n=1m1fn,m+1+fm,m+1>n=1m1FnΔtm+Δtm+1+fm,m+1>Fmax\sum_{n=1}^{m-1}f_{n,m+1}+f_{m,m+1}>\sum_{n=1}^{m-1}\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}}+f_{m,m+1}>F_{max} which violates constraint (E.3). Similarly, if n=1m1fn,m=n=1m1fn,m+1\sum_{n=1}^{m-1}f_{n,m}=\sum_{n=1}^{m-1}f_{n,m+1}, we can obtain that n=1m1fn,m=n=1m1fn,m+1=n=1m1FnΔtm+Δtm+1\sum_{n=1}^{m-1}f_{n,m}=\sum_{n=1}^{m-1}f_{n,m+1}=\sum_{n=1}^{m-1}\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}}. Hence, we have n=1m1fn,m+1+fm,m+1=n=1m1FnΔtm+Δtm+1+fm,m+1>Fmax\sum_{n=1}^{m-1}f_{n,m+1}+f_{m,m+1}=\sum_{n=1}^{m-1}\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}}+f_{m,m+1}>F_{max} which breaks constraint (E.3). As a consequence, the optimal solution in this case satisfies n=1m1fn,m>n=1m1fn,m+1\sum_{n=1}^{m-1}f^{*}_{n,m}>\sum_{n=1}^{m-1}f^{*}_{n,m+1}.

Next, we prove that fn,mfn,m+1f^{*}_{n,m}\geq f^{*}_{n,m+1} for all n=1,,m1n=1,\cdots,m-1. With given Δtm\Delta t_{m} and Δtm+1\Delta t_{m+1}, we denote the energy consumption of the nn-th offloading device by En(Fn,δn)E_{n}(F_{n},\delta_{n}), where δn=fn,mfn,m+1\delta_{n}=f_{n,m}-f_{n,m+1}. According to (G), En(Fn,δn)E_{n}(F_{n},\delta_{n}) is expressed by

En(Fn,δn)=κ(Fn+δnΔtm+1)3(Δtm+Δtm+1)3Δtm\displaystyle E_{n}(F_{n},\delta_{n})=\kappa\frac{(F_{n}+\delta_{n}\Delta t_{m+1})^{3}}{(\Delta t_{m}+\Delta t_{m+1})^{3}}\Delta t_{m}
+κ(FnδnΔtm)3(Δtm+Δtm+1)3Δtm+1,(n=1,,m1),\displaystyle+\kappa\frac{(F_{n}-\delta_{n}\Delta t_{m})^{3}}{(\Delta t_{m}+\Delta t_{m+1})^{3}}\Delta t_{m+1},(n=1,\cdots,m-1), (E.4)

which decreases when Fn/Δtm+1δn0-F_{n}/\Delta t_{m+1}\leq\delta_{n}\leq 0 while increases when 0<δnFn/Δtm0<\delta_{n}\leq F_{n}/\Delta t_{m}.

Furthermore, since n=1m1fn,m+1Fmaxfm,m+1\sum_{n=1}^{m-1}f_{n,m+1}\leq F_{\max}-f_{m,m+1}, we have

n=1m1fn,mn=1m1fn,m+1\displaystyle\sum_{n=1}^{m-1}f_{n,m}-\sum_{n=1}^{m-1}f_{n,m+1}
n=1m1FnΔtm+1n=1m1fn,m+1Δtmn=1m1fn,m+1\displaystyle\geq\frac{\sum_{n=1}^{m-1}F_{n}-\Delta t_{m+1}\sum_{n=1}^{m-1}f_{n,m+1}}{\Delta t_{m}}-\sum_{n=1}^{m-1}f_{n,m+1}
n=1m1Fn(Δtm+Δtm+1)(Fmaxfm,m+1)Δtm.\displaystyle\geq\frac{\sum_{n=1}^{m-1}F_{n}-(\Delta t_{m}+\Delta t_{m+1})\left(F_{\max}-f_{m,m+1}\right)}{\Delta t_{m}}. (E.5)

According to (E), we have n=1m1δnn=1m1Fn(Δtm+Δtm+1)(Fmaxfm,m+1)ΔtmΩ\sum_{n=1}^{m-1}\delta_{n}\geq\frac{\sum_{n=1}^{m-1}F_{n}-(\Delta t_{m}+\Delta t_{m+1})\left(F_{\max}-f_{m,m+1}\right)}{\Delta t_{m}}\triangleq\Omega. Next, we utilize contradiction to prove that δn0\delta_{n}^{*}\geq 0 for all n=1,,m1n=1,\cdots,m-1. Assume in the optimal solution δn\delta^{*}_{n} there exists a certain δn<0\delta_{n}<0. We can suitably decrease other positive δn\delta^{*}_{n} and increase the negative δn<0\delta_{n}<0 to zero while keeping n=1m1δn\sum_{n=1}^{m-1}\delta_{n} unchanged. In this case, the total energy consumption is effectively reduced, which contradicts the optimality. That completes the proof of δn0\delta^{*}_{n}\geq 0, i.e., fn,mfn,m+1f^{*}_{n,m}\geq f^{*}_{n,m+1} for all n=1,,m1n=1,\cdots,m-1.

In summary, since we have proven f1,2f1,3f_{1,2}^{*}\geq f_{1,3}^{*} and fn,mfn,m+1f_{n,m}^{*}\geq f_{n,m+1}^{*} for m=3,,Km=3,\cdots,K and n=1,,m1n=1,\cdots,m-1, we can deduce Corollary 5. \blacksquare

Appendix F Proof of Corollary 6

In case of n=1m1FnΔtm+Δtm+1+fm,m+1Fmax\sum_{n=1}^{m-1}\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}}+f_{m,m+1}\leq F_{\max}, the optimal δn\delta^{*}_{n} (n=1,,m1)(n=1,\cdots,m-1) are all zeros according to Corollary 5. Therefore, we only need to justify the case of n=1m1FnΔtm+Δtm+1+fm,m+1>Fmax\sum_{n=1}^{m-1}\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}}+f_{m,m+1}>F_{\max}. In this case, we first prove that the optimal n=1m1δn=Ω\sum_{n=1}^{m-1}\delta^{*}_{n}=\Omega. Assume that the optimal n=1m1δn>Ω\sum_{n=1}^{m-1}\delta^{*}_{n}>\Omega. We can suitably reduce the positive δn\delta^{*}_{n} such that the energy consumption is further reduced, which contradicts the optimality. Therefore, we can construct the following energy consumption minimization problem:

min𝜹\displaystyle\min_{\boldsymbol{\delta}}\quad n=1m1En(Fn,δn),\displaystyle\sum_{n=1}^{m-1}E_{n}(F_{n},\delta_{n}), (F.1)
s.t.\displaystyle s.t.\quad n=1m1δn=Ω,\displaystyle\sum_{n=1}^{m-1}\delta_{n}=\Omega, (F.1a)
0δnFn/Δtm,n=1,,m1,\displaystyle 0\leq\delta_{n}\leq F_{n}/\Delta t_{m},\quad\forall n=1,\cdots,m-1, (F.1b)

where 𝜹=[δ1,,δm1]T\boldsymbol{\delta}=[\delta_{1},\cdots,\delta_{m-1}]^{T}.

Based on (G.3), the second derivative of En(Fn,δn)E_{n}(F_{n},\delta_{n}) with respect to δn\delta_{n} is given by

d2En(Fn,δn)d(δn)2\displaystyle\frac{\mathrm{d}^{2}E_{n}(F_{n},\delta_{n})}{\mathrm{d}(\delta_{n})^{2}} =6κΔtmΔtm+1(Δtm+Δtm+1)2[δn(Δtm+1Δtm)+Fn].\displaystyle\!=\!\frac{6\kappa\Delta t_{m}\Delta t_{m+1}}{(\Delta t_{m}+\Delta t_{m+1})^{2}}\left[\delta_{n}(\Delta t_{m+1}\!-\!\Delta t_{m})\!+\!F_{n}\right]. (F.2)

We can infer that the second derivative of En(Fn,δn)E_{n}(F_{n},\delta_{n}) is always positive in the considered region Fn/Δtm+1δnFn/Δtm-F_{n}/\Delta t_{m+1}\leq\delta_{n}\leq F_{n}/\Delta t_{m}, no matter Δtm\Delta t_{m} is larger than or smaller than, or equal to Δtm+1\Delta t_{m+1}. Hence, En(Fn,δn)E_{n}(F_{n},\delta_{n}) is convex with respect to δn\delta_{n}. Thus, problem (F.1) is convex. The partial Lagrangian function of this problem is expressed as

min𝜹\displaystyle\min_{\boldsymbol{\delta}}\quad n=1m1En(Fn,δn)+Υ(n=1m1δnΩ),\displaystyle\sum_{n=1}^{m-1}E_{n}(F_{n},\delta_{n})+\Upsilon\left(\sum_{n=1}^{m-1}\delta_{n}-\Omega\right), (F.3)
s.t.\displaystyle s.t.\quad 0δnFn/Δtm,n=1,,m1,\displaystyle 0\leq\delta_{n}\leq F_{n}/\Delta t_{m},\quad\forall n=1,\cdots,m-1, (F.3a)

where Υ\Upsilon is the dual variable with respect to constraint (F.1a). Problem (F.3) can be decomposed into a series of (m1)(m-1) parallel problems:

minδn\displaystyle\min_{\delta_{n}}\quad En(Fn,δn)+Υδn,\displaystyle E_{n}(F_{n},\delta_{n})+\Upsilon\delta_{n}, (F.4)
s.t.\displaystyle s.t.\quad 0δnFn/Δtm,\displaystyle 0\leq\delta_{n}\leq F_{n}/\Delta t_{m}, (F.4a)

Denote the objective of (F.4) by JnJ_{n}. Taking the derivative of JnJ_{n} with respect to δn\delta_{n}, we have

dJndδn=dEn(Fn,δn)dδn+Υ,\displaystyle\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}}=\frac{\mathrm{d}E_{n}(F_{n},\delta_{n})}{\mathrm{d}\delta_{n}}+\Upsilon,
=3κΔtmΔtm+1(Δtm+Δtm+1)2δn(δn(Δtm+1Δtm)+2Fn)+Υ.\displaystyle=\frac{3\kappa\Delta t_{m}\Delta t_{m+1}}{(\Delta t_{m}+\Delta t_{m+1})^{2}}\delta_{n}\left(\delta_{n}(\Delta t_{m+1}\!-\!\Delta t_{m})\!+\!2F_{n}\right)\!+\!\Upsilon. (F.5)

It can be deduced that dEn(Fn,δn)dδn\frac{\mathrm{d}E_{n}(F_{n},\delta_{n})}{\mathrm{d}\delta_{n}} is non-negative when 0δnFn/Δtm0\leq\delta_{n}\leq F_{n}/\Delta t_{m}. If Υ0\Upsilon\geq 0, we have dJndδn>=0\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}}>=0. Thus, the optimal solution is achieved when δn=0\delta_{n}=0 for all n=1,,m1n=1,\cdots,m-1, which contradicts (F.1a). Hence, we should have Υ<0\Upsilon<0. Due to d2Jnd(δn)2=d2End(δn)2>0\frac{\mathrm{d}^{2}J_{n}}{\mathrm{d}(\delta_{n})^{2}}=\frac{d^{2}E_{n}}{d(\delta_{n})^{2}}>0 in the region of [0,Fn/Δtm][0,F_{n}/\Delta t_{m}], we can obtain that dJndδn\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}} monotonously increases. Moreover, we have dJndδn|δn=0=Υ<0\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}}|_{\delta_{n}=0}=\Upsilon<0. Therefore, we consider the following two cases.

Case 1: dJndδn|δn=Fn/Δtm0\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}}|_{\delta_{n}=F_{n}/\Delta t_{m}}\leq 0, i.e., Υ3κΔtm+1Fn2(Δtm+Δtm+1)Δtm\Upsilon\leq-\frac{3\kappa\Delta t_{m+1}F_{n}^{2}}{(\Delta t_{m}+\Delta t_{m+1})\Delta t_{m}}. In this case, dJndδn<=0\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}}<=0 in the region of 0δnFn/Δtm0\leq\delta_{n}\leq F_{n}/\Delta t_{m}. Therefore, the optimal solution is δn=Fn/Δtm\delta_{n}^{*}=F_{n}/\Delta t_{m}.

Case 2: dJndδn|δn=Fn/Δtm>0\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}}|_{\delta_{n}=F_{n}/\Delta t_{m}}>0, i.e., Υ>3κΔtm+1Fn2(Δtm+Δtm+1)Δtm\Upsilon>-\frac{3\kappa\Delta t_{m+1}F_{n}^{2}}{(\Delta t_{m}+\Delta t_{m+1})\Delta t_{m}}. In this case, dJndδn\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}} has a null point in the region of 0δnFn/Δtm0\leq\delta_{n}\leq F_{n}/\Delta t_{m}. Thus, JnJ_{n} decreases first and then increases. Through solving dJndδn=0\frac{\mathrm{d}J_{n}}{\mathrm{d}\delta_{n}}=0, we obtain that

δn={Fn+Fn2ΥΞ(Δtm+1Δtm)Δtm+1Δtm,if Δtm<Δtm+1 or Δtm>Δtm+1,ΥΞ2Fn,if Δtm=Δtm+1,\displaystyle\delta_{n}^{*}=\left\{\begin{aligned} &\frac{-F_{n}+\sqrt{F_{n}^{2}-\Upsilon\Xi(\Delta t_{m+1}-\Delta t_{m})}}{\Delta t_{m+1}-\Delta t_{m}},\text{if }\Delta t_{m}<\Delta t_{m+1}\\ &\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\ \ \text{ or }\Delta t_{m}>\Delta t_{m+1},\\ &-\frac{\Upsilon\Xi}{2F_{n}},\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\ \text{if }\Delta t_{m}=\Delta t_{m+1},\end{aligned}\right. (F.6)

where Ξ=(Δtm+Δtm+1)23κΔtmΔtm+1\Xi=\frac{(\Delta t_{m}+\Delta t_{m+1})^{2}}{3\kappa\Delta t_{m}\Delta t_{m+1}}. Meanwhile, the optimal Υ\Upsilon^{*} should satisfy constraint (F.1a). Obviously, both the above two cases satisfy δn>0\delta^{*}_{n}>0 (n=1,,m1)(n=1,\cdots,m-1), completing the proof. \blacksquare

Appendix G Proof of Lemma 7

Refer to caption
(a) Δtm>Δtm+1\Delta t_{m}>\Delta t_{m+1}
Refer to caption
(b) Δtm<Δtm+1\Delta t_{m}<\Delta t_{m+1}
Refer to caption
(c) Δtm=Δtm+1\Delta t_{m}=\Delta t_{m+1}
Figure 9: Illustrations of three cases for the analysis of energy consumption in Appendix G.

Denote δn=fn,mfn,m+1\delta_{n}=f_{n,m}-f_{n,m+1}. Since fn,mf_{n,m} and fn,m+1f_{n,m+1} should be larger than or equal to zero, we can deduce that Fn/Δtm+1δnFn/Δtm-F_{n}/\Delta t_{m+1}\leq\delta_{n}\leq F_{n}/\Delta t_{m}. Hence, we can obtain that

fn,m=Fn+δnΔtm+1Δtm+Δtm+1,fn,m+1=FnδnΔtmΔtm+Δtm+1.\displaystyle f_{n,m}=\frac{F_{n}+\delta_{n}\Delta t_{m+1}}{\Delta t_{m}+\Delta t_{m+1}},\ f_{n,m+1}=\frac{F_{n}-\delta_{n}\Delta t_{m}}{\Delta t_{m}+\Delta t_{m+1}}. (G.1)

Therefore, the energy consumption in time slot tmt_{m} and tm+1t_{m+1} can be given by

En\displaystyle E_{n} =κfn,m3Δtm+κfn,m+13Δtm+1,\displaystyle=\kappa f_{n,m}^{3}\Delta t_{m}+\kappa f_{n,m+1}^{3}\Delta t_{m+1},
=κ(Fn+δnΔtm+1)3(Δtm+Δtm+1)3Δtm+κ(FnδnΔtm)3(Δtm+Δtm+1)3Δtm+1.\displaystyle=\kappa\frac{(F_{n}+\delta_{n}\Delta t_{m+1})^{3}}{(\Delta t_{m}+\Delta t_{m+1})^{3}}\Delta t_{m}+\kappa\frac{(F_{n}-\delta_{n}\Delta t_{m})^{3}}{(\Delta t_{m}+\Delta t_{m+1})^{3}}\Delta t_{m+1}. (G.2)

Taking the first derivative of EnE_{n} with respect to δn\delta_{n}, we have

dEndδn\displaystyle\frac{\mathrm{d}E_{n}}{\mathrm{d}\delta_{n}} =3κΔtmΔtm+1(Δtm+Δtm+1)2δn(δn(Δtm+1Δtm)+2Fn).\displaystyle=\frac{3\kappa\Delta t_{m}\Delta t_{m+1}}{(\Delta t_{m}+\Delta t_{m+1})^{2}}\delta_{n}\left(\delta_{n}(\Delta t_{m+1}-\Delta t_{m})+2F_{n}\right). (G.3)

Equation (G.3) has two null points: 0 and 2FnΔtmΔtm+1\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}}. We consider the following three cases.

Case 1: Δtm>Δtm+1\Delta t_{m}>\Delta t_{m+1}. In this case, we have 0<2FnΔtmΔtm+10<\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}}. The energy consumption decreases when δn<0\delta_{n}<0 and δn>2FnΔtmΔtm+1\delta_{n}>\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}} while increases when 0δn2FnΔtmΔtm+10\leq\delta_{n}\leq\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}}, as shown in Fig. 9(a). Since we can easily prove that 2FnΔtmΔtm+1>Fn/Δtm\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}}>F_{n}/\Delta t_{m}, the minimum energy consumption is obtained when δn=0\delta_{n}=0, i.e., fn,m=fn,m+1f^{*}_{n,m}=f^{*}_{n,m+1}.

Case 2: Δtm<Δtm+1\Delta t_{m}<\Delta t_{m+1}. In this case, we have 2FnΔtmΔtm+1<0\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}}<0. The energy consumption increases when δn<2FnΔtmΔtm+1\delta_{n}<\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}} and δn>0\delta_{n}>0 while decreases when 2FnΔtmΔtm+1δn0\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}}\leq\delta_{n}\leq 0, as shown in Fig. 9(b). Similarly, since we can prove that 2FnΔtmΔtm+1<Fn/Δtm+1\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}}<-F_{n}/\Delta t_{m+1}, the minimum energy consumption is obtained when δn=0\delta_{n}=0.

Case 3: Δtm=Δtm+1\Delta t_{m}=\Delta t_{m+1}. In this case, two null points coincide, i.e., 2FnΔtmΔtm+1=0\frac{2F_{n}}{\Delta t_{m}-\Delta t_{m+1}}=0. Therefore, the energy consumption decreases when δn<0\delta_{n}<0 while increases when δn0\delta_{n}\geq 0, as shown in Fig. 9(c). δn=0\delta_{n}=0 is the solution that minimizes energy consumption.

In summary, the energy consumption when fn,m=fn,m+1=FnΔtm+Δtm+1f^{*}_{n,m}=f^{*}_{n,m+1}=\frac{F_{n}}{\Delta t_{m}+\Delta t_{m+1}} is the most energy efficient solution. \blacksquare

Appendix H Proof of Proposition 4

According to Theorem 3, we have xi1,i>0x_{i-1,i}>0 (i{2,,K+1})(\forall i\in\{2,\cdots,K+1\}). Hence, according to (11), we have ξn=i+1KρnhπnηP0ωiFmax=n=1i12κ(xn,i)3(Δti)3+ρi2λ(Aπi)3hπi(Δti)3>0\xi^{*}-\sum_{n=i+1}^{K}\rho^{*}_{n}h_{\pi_{n}}\eta P_{0}-\omega^{*}_{i}F_{\max}=\sum_{n=1}^{i-1}2\kappa\frac{(x_{n,i})^{3}}{(\Delta t^{*}_{i})^{3}}+\rho^{*}_{i}\frac{2\lambda(A_{\pi_{i}})^{3}}{h_{\pi_{i}}(\Delta t^{*}_{i})^{3}}>0. Therefore, it can be derived that Δti=2κhπin=1i1(xn,i)3+2ρiλ(Aπi)3ξhπin=i+1Kρn(hπn)2ηP0ωihπiFmax3\Delta t^{*}_{i}=\sqrt[3]{\frac{2\kappa h_{\pi_{i}}\sum_{n=1}^{i-1}(x_{n,i})^{3}+2\rho^{*}_{i}\lambda(A_{\pi_{i}})^{3}}{\xi^{*}h_{\pi_{i}}-\sum_{n=i+1}^{K}\rho^{*}_{n}(h_{\pi_{n}})^{2}\eta P_{0}-\omega^{*}_{i}h_{\pi_{i}}F_{\max}}} (i{2,,K1})(\forall i\in\{2,\cdots,K-1\}). Similarly, based on (12) and (13), we have ΔtK=2κhπKn=1K1(xn,K)3+2ρKλ(AπK)3ξhπKωKhπKFmax3\Delta t^{*}_{K}=\sqrt[3]{\frac{2\kappa h_{\pi_{K}}\sum_{n=1}^{K-1}(x_{n,K})^{3}+2\rho_{K}^{*}\lambda(A_{\pi_{K}})^{3}}{\xi^{*}h_{\pi_{K}}-\omega^{*}_{K}h_{\pi_{K}}F_{\max}}} and ΔtK+1=2κn=1K(xn,K+1)3ξωK+1Fmax3\Delta t_{K+1}^{*}=\sqrt[3]{\frac{2\kappa\sum_{n=1}^{K}(x_{n,K+1})^{3}}{\xi^{*}-\omega^{*}_{K+1}F_{\max}}}.

Besides, according to (13), we have ξ=n=1K2κ(xn,K+1)3(ΔtK+1)3+ωK+1Fmax>0\xi^{*}=\sum_{n=1}^{K}2\kappa\frac{(x_{n,K+1})^{3}}{(\Delta t^{*}_{K+1})^{3}}+\omega^{*}_{K+1}F_{\max}>0 since n=1K(xn,K+1)3>(xK,K+1)3=(AπKIπK)3>0\sum_{n=1}^{K}(x_{n,K+1})^{3}>(x_{K,K+1})^{3}=(A_{\pi_{K}}I_{\pi_{K}})^{3}>0. Therefore, we can obtain that i=0K+1Δti=T\sum_{i=0}^{K+1}\Delta t^{*}_{i}=T. Furthermore, based on (10), we have ξ=n=1KρnhπnηP0\xi^{*}=\sum_{n=1}^{K}\rho^{*}_{n}h_{\pi_{n}}\eta P_{0}. Due to that ξ\xi^{*} is positive, there exists at least an n𝒦n\in\mathcal{K} such that ρn>0\rho_{n}^{*}>0. This indicates that for energy causality constraints (6a), at least a device is run out of energy after offloading, i.e., this device uses all the harvested energy for transmission. Substituting (9) into (10), we have ρ12λ(Aπ1)3hπ1(Δt1)3=ρ1hπ1ηP0\rho^{*}_{1}\frac{2\lambda(A_{\pi_{1}})^{3}}{h_{\pi_{1}}(\Delta t^{*}_{1})^{3}}=\rho^{*}_{1}h_{\pi_{1}}\eta P_{0}. If ρ1>0\rho_{1}^{*}>0, we can deduce that Δt1=2λ(hπ1)2ηP03Aπ1\Delta t^{*}_{1}=\sqrt[3]{\frac{2\lambda}{(h_{\pi_{1}})^{2}\eta P_{0}}}A_{\pi_{1}}. Moreover, since λ(Aπ1)3hπ1(Δt1)2=Δt0hπ1ηP0\frac{\lambda(A_{\pi_{1}})^{3}}{h_{\pi_{1}}(\Delta t_{1})^{2}}=\Delta t_{0}h_{\pi_{1}}\eta P_{0}, we have Δt0=λ4(hπ1)2ηP03Aπ1\Delta t_{0}^{*}=\sqrt[3]{\frac{\lambda}{4(h_{\pi_{1}})^{2}\eta P_{0}}}A_{\pi_{1}}. If ρ1=0\rho^{*}_{1}=0, ddΔt1=0\frac{\mathrm{d}\mathcal{L}}{\mathrm{d}\Delta t^{*}_{1}}=0 can be guaranteed for arbitrary Δt1\Delta t_{1} satisfying λ(Aπ1)3hπ1(Δt1)2Δt0hπ1ηP0\frac{\lambda(A_{\pi_{1}})^{3}}{h_{\pi_{1}}(\Delta t_{1})^{2}}\leq\Delta t_{0}h_{\pi_{1}}\eta P_{0}. That means any pairs of Δt0\Delta t_{0} and Δt1\Delta t_{1} satisfying λ(Aπ1)3hπ1(Δt1)2Δt0hπ1ηP0\frac{\lambda(A_{\pi_{1}})^{3}}{h_{\pi_{1}}(\Delta t_{1})^{2}}\leq\Delta t_{0}h_{\pi_{1}}\eta P_{0} and Δt0+Δt1=Ti=2K+1Δti\Delta t_{0}+\Delta t_{1}=T-\sum_{i=2}^{K+1}\Delta t^{*}_{i} are the optimal solutions. Hence, according to Δt0+Δt1=Ti=2K+1Δti\Delta t_{0}+\Delta t_{1}=T-\sum_{i=2}^{K+1}\Delta t_{i}, we should have λ(Aπ1)3hπ1(Δt1)2(Ti=2K+1ΔtiΔt1)hπ1ηP0\frac{\lambda(A_{\pi_{1}})^{3}}{h_{\pi_{1}}(\Delta t_{1})^{2}}\leq\left(T-\sum_{i=2}^{K+1}\Delta t_{i}-\Delta t_{1}\right)h_{\pi_{1}}\eta P_{0}, i.e., Ψ(Δt1)(Ti=2K+1ΔtiΔt1)(Δt1)2(hπ1)2ηP0λ(Aπ1)30\Psi(\Delta t_{1})\triangleq\left(T-\sum_{i=2}^{K+1}\Delta t_{i}-\Delta t_{1}\right)(\Delta t_{1})^{2}(h_{\pi_{1}})^{2}\eta P_{0}-\lambda(A_{\pi_{1}})^{3}\geq 0. Taking the derivative of Ψ(Δt1)\Psi(\Delta t_{1}), we have Ψ(Δt1)=Δt1(2T2i=2K+1Δti3Δt1)(hπ1)2ηP0\Psi^{\prime}(\Delta t_{1})=\Delta t_{1}\left(2T-2\sum_{i=2}^{K+1}\Delta t_{i}-3\Delta t_{1}\right)(h_{\pi_{1}})^{2}\eta P_{0}, which has two null points Δt1=0\Delta t_{1}=0 and Δt1=23(Ti=2K+1Δti)\Delta t_{1}=\frac{2}{3}\left(T-\sum_{i=2}^{K+1}\Delta t_{i}\right). Thus, Ψ(Δt1)\Psi(\Delta t_{1}) increases in the region of (0,23(Ti=2K+1Δti)]\Big{(}0,\frac{2}{3}\left(T-\sum_{i=2}^{K+1}\Delta t_{i}\right)\Big{]} and decreases in (23(Ti=2K+1Δti),(Ti=2K+1Δti))\left(\frac{2}{3}\left(T-\sum_{i=2}^{K+1}\Delta t_{i}\right),\left(T-\sum_{i=2}^{K+1}\Delta t_{i}\right)\right). Additionally, we can obtain that Ψ(0)=Ψ(Ti=2K+1Δti)\Psi(0)=\Psi\left(T-\sum_{i=2}^{K+1}\Delta t_{i}\right). Therefore, we can choose Δt1\Delta t^{*}_{1} as the unique null point of Ψ(Δt1)\Psi(\Delta t_{1}) in the range of Δt1(0,23(Ti=2K+1Δti)]\Delta t_{1}\in\Big{(}0,\frac{2}{3}\left(T-\sum_{i=2}^{K+1}\Delta t_{i}\right)\Big{]} without loss of generality. \blacksquare

References

  • [1] Q.-V. Pham, F. Fang, V. N. Ha, M. J. Piran, M. Le, L. B. Le, W.-J. Hwang, and Z. Ding, “A Survey of Multi-Access Edge Computing in 5G and Beyond: Fundamentals, Technology Integration, and State-of-the-Art,” IEEE Access, vol. 8, pp. 116 974–117 017, June 2020.
  • [2] Y. Mao, C. You, J. Zhang, K. Huang, and K. B. Letaief, “A Survey on Mobile Edge Computing: The Communication Perspective,” IEEE Commun. Surv. Tutor., vol. 19, no. 4, pp. 2322–2358, Fourthquarter 2017.
  • [3] H. Zhou, Z. Zhang, D. Li, and Z. Su, “Joint Optimization of Computing Offloading and Service Caching in Edge Computing-Based Smart Grid,” IEEE Trans. Cloud Comp., vol. 11, no. 2, pp. 1122–1132, Apr. 2023.
  • [4] X. Li, H. Zhang, H. Zhou, N. Wang, K. Long, S. Al-Rubaye, and G. K. Karagiannidis, “Multi-Agent DRL for Resource Allocation and Cache Design in Terrestrial-Satellite Networks,” IEEE Trans. Wirel. Commun., vol. 22, no. 8, pp. 5031–5042, Aug. 2023.
  • [5] H. Zhou, T. Wu, X. Chen, S. He, D. Guo, and J. Wu, “Reverse Auction-Based Computation Offloading and Resource Allocation in Mobile Cloud-Edge Computing,” IEEE Trans. Mobile Comp., vol. 22, no. 10, pp. 6144–6159, Oct. 2023.
  • [6] X. Cao, F. Wang, J. Xu, R. Zhang, and S. Cui, “Joint Computation and Communication Cooperation for Energy-Efficient Mobile Edge Computing,” IEEE Int. Things J., vol. 6, no. 3, pp. 4188–4200, June 2019.
  • [7] Y. Pan, M. Chen, Z. Yang, N. Huang, and M. Shikh-Bahaei, “Energy-Efficient NOMA-Based Mobile Edge Computing Offloading,” IEEE Commun. Letters, vol. 23, no. 2, pp. 310–313, Feb. 2019.
  • [8] Z. Yu, Y. Tang, L. Zhang, and H. Zeng, “Deep Reinforcement Learning Based Computing Offloading Decision and Task Scheduling in Internet of Vehicles,” in Proc. IEEE/CIC Int. Conf. Commun. China (ICCC), Xiamen, China, July 2021, pp. 1166–1171.
  • [9] X. Lu, P. Wang, D. Niyato, D. I. Kim, and Z. Han, “Wireless networks with RF energy harvesting: A contemporary survey,” IEEE Commun. Surv. Tutor., vol. 17, no. 2, pp. 757–789, 2014.
  • [10] Z. Zhang, H. Pang, A. Georgiadis, and C. Cecati, “Wireless Power Transfer—An Overview,” IEEE Trans. Industrial Electron., vol. 66, no. 2, pp. 1044–1058, Feb. 2019.
  • [11] F. Wang, J. Xu, X. Wang, and S. Cui, “Joint Offloading and Computing Optimization in Wireless Powered Mobile-Edge Computing Systems,” IEEE Trans. Wirel. Commun., vol. 17, no. 3, pp. 1784–1797, Mar. 2018.
  • [12] S. Bi and Y. J. Zhang, “Computation Rate Maximization for Wireless Powered Mobile-Edge Computing With Binary Computation Offloading,” IEEE Trans. Wirel. Commun., vol. 17, no. 6, pp. 4177–4190, June 2018.
  • [13] P. Chen, B. Lyu, Y. Liu, H. Guo, and Z. Yang, “Multi-IRS Assisted Wireless-Powered Mobile Edge Computing for Internet of Things,” IEEE Trans. Green Commun. Netw., pp. 1–1, Sep. 2022.
  • [14] X. Hu, K.-K. Wong, and K. Yang, “Wireless Powered Cooperation-Assisted Mobile Edge Computing,” IEEE Trans. Wirel. Commun., vol. 17, no. 4, pp. 2375–2388, Apr. 2018.
  • [15] K. Zhang, Y. Mao, S. Leng, S. Maharjan, and Y. Zhang, “Optimal delay constrained offloading for vehicular edge computing networks,” in Proc. IEEE Int. Conf. Commun. (ICC), Paris, France, July 2017, pp. 1–6.
  • [16] Z. Zhu, J. Peng, X. Gu, H. Li, K. Liu, Z. Zhou, and W. Liu, “Fair Resource Allocation for System Throughput Maximization in Mobile Edge Computing,” IEEE Access, vol. 6, pp. 5332–5340, Jan. 2018.
  • [17] F. Zhou and R. Q. Hu, “Computation Efficiency Maximization in Wireless-Powered Mobile Edge Computing Networks,” IEEE Trans. Wirel. Commun., vol. 19, no. 5, pp. 3170–3184, Feb. 2020.
  • [18] Y. Xu, P. Cheng, Z. Chen, M. Ding, Y. Li, and B. Vucetic, “Task offloading for large-scale asynchronous mobile edge computing: An index policy approach,” IEEE Trans. Signal Proc., vol. 69, pp. 401–416, Dec. 2020.
  • [19] Y. Dai, M. Sheng, J. Liu, N. Cheng, and X. Shen, “Delay-efficient offloading for NOMA-MEC with asynchronous uploading completion awareness,” in Proc. IEEE Global Commun. Conf. (GLOBECOM).   Waikoloa, HI, USA: IEEE, Feb. 2019, pp. 1–6.
  • [20] C. You, Y. Zeng, R. Zhang, and K. Huang, “Asynchronous Mobile-Edge Computation Offloading: Energy-Efficient Resource Management,” IEEE Trans. Wirel. Commun., vol. 17, no. 11, pp. 7590–7605, Nov. 2018.
  • [21] Y. Hu, M. Chen, Y. Wang, Z. Li, M. Pei, and Y. Cang, “Discrete-Time Joint Scheduling of Uploading and Computation for Deterministic MEC Systems Allowing for Task Interruptions and Insertions,” IEEE Wirel. Commun. Letters, vol. 12, no. 1, pp. 21–25, Jan. 2023.
  • [22] S. Eom, H. Lee, J. Park, and I. Lee, “Asynchronous Protocol Designs for Energy Efficient Mobile Edge Computing Systems,” IEEE Trans. Veh. Techn., vol. 70, no. 1, pp. 1013–1018, Jan. 2021.
  • [23] K. Guo and T. Q. S. Quek, “On the Asynchrony of Computation Offloading in Multi-User MEC Systems,” IEEE Trans. Commun., vol. 68, no. 12, pp. 7746–7761, Dec. 2020.
  • [24] Z. Kuang, L. Li, J. Gao, L. Zhao, and A. Liu, “Partial Offloading Scheduling and Power Allocation for Mobile Edge Computing Systems,” IEEE Internet Things J., vol. 6, no. 4, pp. 6774–6785, Aug. 2019.
  • [25] P. Cai, F. Yang, J. Wang, X. Wu, Y. Yang, and X. Luo, “JOTE: Joint Offloading of Tasks and Energy in Fog-Enabled IoT Networks,” IEEE Internet of Things Journal, vol. 7, no. 4, pp. 3067–3082, April 2020.
  • [26] Z. Liang, H. Chen, Y. Liu, and F. Chen, “Data Sensing and Offloading in Edge Computing Networks: TDMA or NOMA?” IEEE Trans. Wirel. Commun., vol. 21, no. 6, pp. 4497–4508, June 2022.
  • [27] S. Bi, L. Huang, H. Wang, and Y.-J. A. Zhang, “Lyapunov-Guided Deep Reinforcement Learning for Stable Online Computation Offloading in Mobile-Edge Computing Networks,” IEEE Trans. Wirel. Commun., vol. 20, no. 11, pp. 7519–7537, Nov. 2021.
  • [28] F. Wang, J. Xu, and S. Cui, “Optimal Energy Allocation and Task Offloading Policy for Wireless Powered Mobile Edge Computing Systems,” IEEE Trans. Wireless Commun., vol. 19, no. 4, pp. 2443–2459, Apr. 2020.
  • [29] J. Liu, K. Xiong, D. W. K. Ng, P. Fan, Z. Zhong, and K. B. Letaief, “Max-Min Energy Balance in Wireless-Powered Hierarchical Fog-Cloud Computing Networks,” IEEE Trans. Wirel. Commun., vol. 19, no. 11, pp. 7064–7080, Nov. 2020.
  • [30] M. Li, N. Cheng, J. Gao, Y. Wang, L. Zhao, and X. Shen, “Energy-Efficient UAV-Assisted Mobile Edge Computing: Resource Allocation and Trajectory Optimization,” IEEE Trans. Veh. Techn., vol. 69, no. 3, pp. 3424–3438, Mar. 2020.
  • [31] D. P. Bertsekas, Convex optimization Theory.   Athena Scientific Belmont, 2009.
  • [32] Y. Yu, X. Bu, K. Yang, H. Yang, X. Gao, and Z. Han, “UAV-Aided Low Latency Multi-Access Edge Computing,” IEEE Trans. Veh. Techn., vol. 70, no. 5, pp. 4955–4967, May 2021.
  • [33] A. Ibrahim, O. A. Dobre, T. M. N. Ngatched, and A. G. Armada, “Bender’s Decomposition for Optimization Design Problems in Communication Networks,” IEEE Netw., vol. 34, no. 3, pp. 232–239, May 2020.
  • [34] D. W. K. Ng, Y. Wu, and R. Schober, “Power Efficient Resource Allocation for Full-Duplex Radio Distributed Antenna Networks,” IEEE Trans. Wirel. Commun., vol. 15, no. 4, pp. 2896–2911, Apr. 2016.
  • [35] S. Boyd, S. P. Boyd, and L. Vandenberghe, Convex optimization.   Cambridge university press, 2004.
  • [36] D. W. K. Ng and R. Schober, “Secure and Green SWIPT in Distributed Antenna Networks With Limited Backhaul Capacity,” IEEE Trans. Wirel. Commun., vol. 14, no. 9, pp. 5082–5097, Sep. 2015.
  • [37] “GUROBI Optimization, State-of-the-Art Mathematical Programming Solver, v5.6”. Apr. 2014. [Online]. Available: http://www.gurobi.com/
  • [38] “MOSEK ApS: Software for Large-Scale Mathematical Optimization Problems, Version 7.0.0.111”. Apr. 2014. [Online]. Available: http://www.mosek.com/
  • [39] Y. Mezentsev, “Binary cut-and-branch method for solving mixed integer programming problems,” in Proc. Construct. Nonsmooth Analy. Related Topics (CNSA), St. Petersburg, Russia, May 2017, pp. 1–3.