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

Differentially Private Distributed Mismatch Tracking Algorithm for Constraint-Coupled Resource Allocation Problems

Wenwen Wu1, Shanying Zhu1, Shuai Liu2 and Xinping Guan1 *This work was supported in part by the NSF of China under the Grants 61922058, 62173225 and the Key Program of the National Natural Science Foundation of China under Grant 62133008.1Wenwen Wu, Shanying Zhu and Xinping Guan are with the Department of Automation, Shanghai Jiao Tong University, Shanghai 200240, China; Key Laboratory of System Control and Information Processing, Ministry of Education of China,Shanghai 200240, China, and also Shanghai Engineering Research Center of Intelligent Control and Management, Shanghai 200240, China.2Shuai Liu is with the School of Control Science and Engineering, Shandong University, Jinan,China.
Abstract

This paper considers privacy-concerned distributed constraint-coupled resource allocation problems over an undirected network, where each agent holds a private cost function and obtains the solution via only local communication. With privacy concerns, we mask the exchanged information with independent Laplace noise against a potential attacker with potential access to all network communications. We propose a differentially private distributed mismatch tracking algorithm (diff-DMAC) to achieve cost-optimal distribution of resources while preserving privacy. Adopting constant stepsizes, the linear convergence property of diff-DMAC in mean square is established under the standard assumptions of Lipschitz gradients and strong convexity. Moreover, it is theoretically proven that the proposed algorithm is ϵ\epsilon-differentially private. And we also show the trade-off between convergence accuracy and privacy level. Finally, a numerical example is provided for verification.

I INTRODUCTION

Distributed resource allocation problem (DRAP) has received much attention due to its wide applications in various domains, including smart grids [1], wireless networks[2] and robot networks [3]. The goal of DRAP is to achieve the cost-optimal distribution of limited resources among users to meet their demands, local constraints, and possibly certain coupled global constraints, see [4] for example. Conventional centralized approaches are subject to performance limitations of the central entity, such as a single point of failure, limited scalability. Moreover, it may raise privacy concerns when the central agent is not reliable enough. Alternatively, distributed approaches which operate with only local information have better robustness and scalability, especially for large-scale systems [5, 6, 7].

To implement algorithms in a distributed manner, the information exchange between agents in the network is unavoidable which can raise concerns about privacy disclosure. For example, in the supply–demand optimization of the power grid, attackers can infer users’ life patterns from certain published information that is computed based on the demand from users[8]. While the aforementioned works consider an ambitious suite of topics under various constraints imposed by real-world applications, privacy is an aspect generally absent in their treatments.

Recently, privacy preservation becomes an increasingly critical issue for distributed systems. Encryption-based algorithms are proposed for distributed optimization problems [9], [10] and proved to be effective. However, the encryption technique leads to a high computational complexity which seems undesirable in large-scale networks. To preserve the privacy of the agents’ costs, an asynchronous heterogeneous-stepsize algorithm is proposed in [11], but the method assumes that the infomation of communication topology is unavailable to attackers.

Another thread to address distributed optimization problems is to adopt differential privacy technique which is robust to arbitrary auxiliary information exposed to attackers, including communication topology information, thus well suited for multi-agents scenarios. Motivated by the privacy concerns in EV charging networks, the work in [8] presents a differentially private distributed algorithm to solve distributed optimization problem, but an extra central agent is needed. In [12], the authors design a completely distributed algorithm that guarantees differential privacy by perturbing the states with Laplace noise. It requires the stepsize to be decaying to guarantee the convergence, resulting in a low convergence rate. Adopting the gradient-tracking scheme, a differentially private distributed algorithm is proposed in [13] which has a linear convergence rate with constant stepsizes. However, the proposed algorithm in [13] is only applicable to unconstrained problems. For solving the economic dispatch problem where both global and local constraints are considered, a privacy-preserving distributed optimization algorithm is proposed in [14] for quadratic cost functions and its effctiveness in preserving privacy is guaranteed through qualitative analysis. The extension to general DRAPs is performed in [15] for strongly convex and Lipschitz smooth cost functions. Additionally, quantitative analysis of privacy is provided. However, the work only considers scalar cost functions and the coupling within constraints is ignored.

In this paper, a differentially private distributed mismatch tracking algorithm (diff-DMAC) is proposed to solve constraint-coupled DRAPs with individual constraints and a global linear coupling constraint while preserving the privacy of the local cost functions. To meet the global coupled-constraint, a mismatch tracking step is introduced to obtain the supply-demand mismatch in the network. And the tracked global mismatch is then implemented for error compensation. With a constant stepsize, diff-DMAC has a provable linear convergence rate for strongly convex and Lipschitz smooth cost functions. We further present a rigorous analysis of the algorithm’s differential privacy and thus provide a stronger privacy guarantee. Finally, its convergence properties and effectiveness in privacy-preserving are numerically validated.

The rest of the paper is organized as follows. Section II formulates the constraint-coupled DRAP. The proposed algorithm is developed in Section III. And the theoretical analyses about its convergence and privacy are given in Section IV and V, respectively. Then the algorithm is numerically tested in Section VI. Finally, Section VII concludes the paper.

Notation: Vectors default to columns if not otherwise specified. Bold letter 𝐱n×p\mathbf{x}\in\mathbb{R}^{n\times p} is defined as 𝐱=[x1,,xn]\mathbf{x}=[x_{1}^{\intercal},\cdots,x_{n}^{\intercal}]^{\intercal}. The Kronecker product is denoted by \otimes. Let 𝟏n\mathbf{1}_{n} be the nn-dimension vector with all one entries. For vectors, we use \|\cdot\| to denote the 2-norm. And for matrices, \|\cdot\| denotes the spectral norm. λ¯(X)\underline{\lambda}(X) denotes the minimum eigenvalue of matrix XX. We use blkdiag(X1,,Xn)blkdiag(X_{1},\cdots,X_{n}) to refer to the block-diagonal matrix with X1,,XnX_{1},\cdots,X_{n} as blocks. For a random variable xx\in\mathbb{R}, Lap(θ)Lap(\theta) denotes the Laplace distribution with pdf f=(x|θ)=12θe|x|θf=(x|\theta)=\frac{1}{2\theta}e^{-\frac{|x|}{\theta}}, where θ>0\theta>0. If xLap(θ)x\sim Lap(\theta), we have 𝔼[|x|]=θ\mathbb{E}[|x|]=\theta, 𝔼[x2]=2θ2\mathbb{E}[x^{2}]=2\theta^{2}. ()\mathbb{P}(\cdot) is used to denote the probability. 𝔹(S)\mathbb{B}(S) denotes the set of Borel subsets of topological space SS.

II Problem Formulation

We consider the following general constraint-coupled distributed resource allocation problem

min𝒙npf(𝒙)\displaystyle\min_{\boldsymbol{x}\in\mathbb{R}^{np}}f(\boldsymbol{x}) =i=1nfi(xi)\displaystyle=\sum_{i=1}^{n}f_{i}(x_{i})
s.t.i=1nAixi\displaystyle s.t.\sum_{i=1}^{n}A_{i}x_{i} =i=1ndi,xi𝒳i,i,\displaystyle=\sum_{i=1}^{n}d_{i},\ \,x_{i}\in\mathcal{X}_{i},\forall i, (1)

where fi:pf_{i}:\mathbb{R}^{p}\rightarrow\mathbb{R} is the agent ii’s private cost function, xipx_{i}\in\mathbb{R}^{p} is the decision vector of agent ii and did_{i} denotes the local resource demand. 𝒳i\mathcal{X}_{i} is a local convex and closed set which encodes local constraints of agent ii. Aim×p(pm)A_{i}\in\mathbb{R}^{m\times p}\ (p\geq m) is the coupling matrix with full row rank and AiAiA_{i}^{\intercal}A_{i} is invertible, i.e., λ\lambda(AiAi)>0(A_{i}^{\intercal}A_{i})>0.

Many practical problems take the form of problem (1), e.g.,[1, 16]. One typical example is the economic dispatch of multi-microgrid (multi-MG) systems as shown in Fig. 1, where both microturbines and non-dispatchable distributed generators are involved to provide energy to meet the loads, and the outputs of non-dispatchable distributed generators can fluctuate significantly [16]. When the supply-demand balance in the MG cannot be maintained, the coordination among MGs is needed and both within-MG, between-MG optimization problems should be considered to improve systems’ reliability [17].

Refer to caption
Figure 1: Multi-microgrid systems with distributed generators.

Assumption 1: Each cost function fi:pf_{i}:\mathbb{R}^{p}\rightarrow\mathbb{R} is strongly convex and Lipschitz smooth, i.e.,

fi(x)fi(x)Lixx\|\nabla f_{i}(x)-\nabla f_{i}(x^{\prime})\|\leq L_{i}\|x-x^{\prime}\| (2)

and

(xx)(fi(x)fi(x))φixx2(x-x^{\prime})^{\intercal}\left(\nabla f_{i}(x)-\nabla f_{i}(x^{\prime})\right)\geq\varphi_{i}\|x-x^{\prime}\|^{2} (3)

x,xp\forall x,x^{\prime}\in\mathbb{R}^{p}, where φi>0\varphi_{i}>0 and Li<L_{i}<\infty are the strong convexity and Lipschitz constants, respectively. Define φ¯=mini{φi}\underline{\varphi}=\min_{i}\{\varphi_{i}\} and L¯=maxi{Li}\overline{L}=\max_{i}\{L_{i}\}.

The communication network over which agents exchange information can be represented by an undirected graph 𝒢=(𝒩,)\mathcal{G}=\left(\mathcal{N},\mathcal{E}\right) where 𝒩={1,,n}\mathcal{N}=\{1,\cdots,n\} is the set of agents and 𝒩×𝒩\mathcal{E}\subseteq\mathcal{N}\times\mathcal{N} denotes the set of edges, accompanied with a nonnegative weighted matrix 𝒲\mathcal{W}. For any i,j𝒩i,j\in\mathcal{N} in the network, wij>0w_{ij}>0 denotes agent jj can exchange information with agent ii. The collection of all individual agents that agent ii can communicate with is defined as its neighbors set 𝒩i\mathcal{N}_{i}.

Assumption 2: The graph 𝒢\mathcal{G} is undirected and connected and the weight matrix 𝒲n×n\mathcal{W}\in\mathbb{R}^{n\times n} is doubly stochastic, i.e., 𝒲𝟏n=𝟏n\mathcal{W}\mathbf{1}_{n}=\mathbf{1}_{n} and 𝟏n𝒲=𝟏n\mathbf{1}_{n}^{\intercal}\mathcal{W}=\mathbf{1}_{n}^{\intercal}.

Assumptions 1-2 are standard when solving related problems.

In this problem, agents want to achieve the optimum while preserving the privacy of local cost functions which can be commercially sensitive in practice. To preserve privacy of functions, we add noise 𝜻\boldsymbol{\zeta} to mask the exchanged information 𝐱(k)\mathbf{x}(k) at each time instant as 𝐳(k)=𝐱(k)+𝜻(k)\mathbf{z}(k)=\mathbf{x}(k)+\boldsymbol{\zeta}(k). Let \mathcal{F} be the function set {fi}i=1n\{f_{i}\}_{i=1}^{n}, the update of 𝐱\mathbf{x} can be expressed in impact form

𝐱(k+1)=g(𝐱(k),𝜻(k),𝒲,),\displaystyle\mathbf{x}(k+1)=g(\mathbf{x}(k),\boldsymbol{\zeta}(k),\mathcal{W},\mathcal{F}),

where g()g(\cdot) denotes the update rule. Therefore, exchanged information 𝐳={𝐳(k)}k=0\mathbf{z}=\{\mathbf{z}(k)\}_{k=0}^{\infty} can be determined if the initialization 𝐱(0)\mathbf{x}(0), the added noises 𝜻={𝜻(k)}k=0\boldsymbol{\zeta}=\{\boldsymbol{\zeta}(k)\}_{k=0}^{\infty}, the communication topology 𝒲\mathcal{W} and the function set \mathcal{F} are known. To provide theoretical guarantees on the privacy for their cost functions, the concept of differential privacy is introduced, see [15].

Definition 1: Given δ>0\delta>0 and two function sets (1)={fi(1)}i=1n\mathcal{F}^{(1)}=\{f_{i}^{(1)}\}_{i=1}^{n} and (2)={fi(2)}i=1n\mathcal{F}^{(2)}=\{f_{i}^{(2)}\}_{i=1}^{n}, the two subsets are δ\delta-adjacent if there exists δ(δ,δ)\delta^{\prime}\in(-\delta,\delta) and some i0i_{0} such that

fi(1)=fi(2),ii0;fi0(1)(x)=fi0(2)(x+δ),x𝒳i0(1),f_{i}^{(1)}=f_{i}^{(2)},\forall i\neq i_{0};\nabla f_{i_{0}}^{(1)}(x)=\nabla f_{i_{0}}^{(2)}(x+\delta^{\prime}),\forall x\in\mathcal{X}_{i_{0}}^{(1)},

where 𝒳i0(l)\mathcal{X}_{i_{0}}^{(l)} is the domain of fi0(l),l=1,2f^{(l)}_{i_{0}},l=1,2 and they satisfy that 𝒳i0(2)={x+δ|x𝒳i0(1)}\mathcal{X}_{i_{0}}^{(2)}=\{x+\delta^{\prime}|x\in\mathcal{X}_{i_{0}}^{(1)}\}.

Definition 2: Given δ,ϵ>0\delta,\epsilon>0, for any two δ\delta-adjacent function sets (1),(2)\mathcal{F}^{(1)},\mathcal{F}^{(2)} and any observation 𝒪𝔹((nm))\mathcal{O}\subset\mathbb{B}((\mathbb{R}^{nm})^{\mathbb{N}}), the process keeps ϵ\epsilon-differentially private if

{𝜻Ω|ZT(1)(𝜻)𝒪}eϵ{𝜻Ω|ZT(2)(𝜻)𝒪},\displaystyle\mathbb{P}\{\boldsymbol{\zeta}\in\Omega|Z_{T^{(1)}}(\boldsymbol{\zeta})\in\mathcal{O}\}\leq e^{\epsilon}\mathbb{P}\{\boldsymbol{\zeta}\in\Omega|Z_{T^{(2)}}(\boldsymbol{\zeta})\in\mathcal{O}\},

where we define T(l)={x(0),𝒲,(l)}T^{(l)}=\{x(0),\mathcal{W},\mathcal{F}^{(l)}\}, ZT(1)(𝜻)=𝐳Z_{T^{(1)}}(\boldsymbol{\zeta})=\mathbf{z} and the observation 𝒪\mathcal{O} encodes all the information collected by the potential attacker.

The attacker is assumed to know all auxiliary information, including the exchanged information between agents, communication topology, etc., except the parameter δ\delta^{\prime}. Intuitively, differential privacy guarantees that the two function sets can not be distinguished through the released information and thus avoids the leakage of the optimal operating point.

III Algorithm Development

In this section, a distributed algorithm is designed to solve constraint-coupled DRAPs with privacy concerns. The following Proposition is presented to give the optimal conditions for constraint-coupled DRAP.

Proposition 1: x=[x1,,xn]n×px^{\star}=[{x_{1}^{\star}}^{\intercal},\cdots,{x_{n}^{\star}}^{\intercal}]\in\mathbb{R}^{n\times p} is the optimal solution of (1) if it satisfies
i) i=1nAixi=i=1ndi\ \sum_{i=1}^{n}A_{i}x_{i}^{\star}=\sum_{i=1}^{n}d_{i},
ii) there exists μm\mu^{\star}\in\mathbb{R}^{m} such that

x=argmin𝐳𝒳{f(𝐳)(𝟏nμ)𝐀𝐳},\displaystyle x=\arg\min_{\mathbf{z}\in\mathcal{X}}\{f(\mathbf{z})-(\mathbf{1}_{n}\otimes\mu^{\star})^{\intercal}\mathbf{A}\mathbf{z}\}, (4)

where 𝒳\mathcal{X} is defined as the Cartesian product 𝒳=𝒳1××𝒳n\mathcal{X}=\mathcal{X}_{1}\times\cdots\times\mathcal{X}_{n} and the matrix 𝐀=blkdiag(A1,,An)\mathbf{A}=blkdiag(A_{1},\cdots,A_{n}).

Proof: i) is one of the feasible conditions of (1) which must be satisfied. Problem (4) can be decomposed into nn subproblems xi=argminz𝒳i{fi(z)μAiz},i=1,,nx_{i}^{\star}=\arg\min_{z\in\mathcal{X}_{i}}\{f_{i}(z)-{\mu^{\star}}^{\intercal}A_{i}z\},\forall i=1,\cdots,n. It follows that

fi(xi)μAixifi(xi)μAixi,xi𝒳i.\displaystyle f_{i}(x_{i}^{\star})-{\mu^{\star}}^{\intercal}A_{i}x_{i}^{\star}\leq f_{i}(x_{i})-{\mu^{\star}}^{\intercal}A_{i}x_{i},\ \forall x_{i}\in\mathcal{X}_{i}.

Summing the above inequality from i=1i=1 to nn yields

i=1nfi(xi)i=1nfi(xi)μ(i=1nAixii=1nAixi).\displaystyle\sum\limits_{i=1}^{n}f_{i}(x_{i}^{\star})\leq\sum\limits_{i=1}^{n}f_{i}(x_{i})-{\mu^{\star}}^{\intercal}(\sum\limits_{i=1}^{n}A_{i}x_{i}-\sum\limits_{i=1}^{n}A_{i}x_{i}^{\star}).

If 𝐱\mathbf{x}^{\star} satisfies condition i), i=1nAixii=1nAixi=0\sum_{i=1}^{n}A_{i}x_{i}-\sum_{i=1}^{n}A_{i}x_{i}^{\star}=0 holds for any feasible solution 𝐱\mathbf{x} of (1). Thus, we have f(𝐱)f(𝐱)f(\mathbf{x}^{\star})\leq f(\mathbf{x}) which completes the proof. \hfill\blacksquare

In the following, we will use Proposition 1 to develop a distributed algorithm for solving the DRAP (1). Based on the condition ii) of Proposition 1, the 𝐱\mathbf{x}-update is designed and the consensus protocol [18] is adopted to drive μi\mu_{i} in each agent to same value

𝝁(k+1)=𝐖𝝁(k),\displaystyle\boldsymbol{\mu}(k+1)=\mathbf{W}\boldsymbol{\mu}(k), (5)
𝐱(k+1)=argmin𝐳𝒳{f(𝐳)𝝁(k+1)𝐀𝐳},\displaystyle\mathbf{x}(k+1)=\arg\min_{\mathbf{z}\in\mathcal{X}}\{f(\mathbf{z})-\boldsymbol{\mu}^{\intercal}(k+1)\mathbf{A}\mathbf{z}\}, (6)

where we define 𝐖=𝒲Im\mathbf{W}=\mathcal{W}\otimes I_{m}.

Inspired by the gradient-tracking scheme [19] , a tracking step is introduced here to track the global supply-demand mismatch as follows

𝐲(k+1)=𝐖𝐲(k)+𝐀𝐱(k+1)𝐀𝐱(k).\displaystyle\mathbf{y}(k+1)=\mathbf{W}\mathbf{y}(k)+\mathbf{A}\mathbf{x}(k+1)-\mathbf{A}\mathbf{x}(k). (7)

Under Assumption 2, if the initialization condition (𝟏nIm)(𝐀𝐱(0)𝐝)=(𝟏nIm)𝐲(0)(\mathbf{1}_{n}\otimes I_{m})^{\intercal}(\mathbf{A}\mathbf{x}(0)-\mathbf{d})=(\mathbf{1}_{n}\otimes I_{m})^{\intercal}\mathbf{y}(0) holds, iteration (7) can infer that

i=1nyi(k)=i=1nAixi(k)i=1ndi,k>0.\displaystyle\sum_{i=1}^{n}y_{i}(k)=\sum_{i=1}^{n}A_{i}x_{i}(k)-\sum_{i=1}^{n}d_{i},\ \forall k>0. (8)

To guarantee the condition i) of Proposition 1 holds, the tracked mismatch is used as a compensation term for 𝝁\boldsymbol{\mu}-update (5). Thus, the proposed algorithm is obtained

𝝁(k+1)=𝐖𝝁(k)α𝐲(k),\displaystyle\boldsymbol{\mu}(k+1)=\mathbf{W}\boldsymbol{\mu}(k)-\alpha\mathbf{y}(k), (9)
𝐱(k+1)=argmin𝐳𝒳{f(𝐳)𝝁(k+1)𝐀𝐳},\displaystyle\mathbf{x}(k+1)=\arg\min_{\mathbf{z}\in\mathcal{X}}\{f(\mathbf{z})-\boldsymbol{\mu}^{\intercal}(k+1)\mathbf{A}\mathbf{z}\}, (10)
𝐲(k+1)=𝐖𝐲(k)+𝐀𝐱(k+1)𝐀𝐱(k).\displaystyle\mathbf{y}(k+1)=\mathbf{W}\mathbf{y}(k)+\mathbf{A}\mathbf{x}(k+1)-\mathbf{A}\mathbf{x}(k). (11)

The following Lemma shows the fixed point of the proposed algorithm is consistent with the optimal solution of the problem when Assumption 2 holds.

Lemma 1: Under Assumption 2, any fixed point of the proposed algorithm satisfies the optimal conditions of constraint-coupled DRAP.

Proof: It is not difficult to check that any fixed point (𝝁F,𝐱F,𝐲F)(\boldsymbol{\mu}_{F},\mathbf{x}_{F},\mathbf{y}_{F}) of the proposed algorithm satisfies

(Inm𝐖)𝝁F=α𝐲F,\displaystyle(I_{nm}-\mathbf{W})\boldsymbol{\mu}_{F}=-\alpha\mathbf{y}_{F}, (12)
𝐱F=argmin𝐳𝒳{f(𝐳)𝝁F𝐀𝐳},\displaystyle\mathbf{x}_{F}=\arg\min_{\mathbf{z}\in\mathcal{X}}\{f(\mathbf{z})-\boldsymbol{\mu}_{F}^{\intercal}\mathbf{A}\mathbf{z}\}, (13)
(Inm𝐖)𝐲F=0.\displaystyle(I_{nm}-\mathbf{W})\mathbf{y}_{F}=0. (14)

Under Assumption 2, from (12) and (14), we have

(𝟏nIm)𝐲F=0,𝐲F=𝟏nyF.\displaystyle(\mathbf{1}_{n}\otimes I_{m})^{\intercal}\mathbf{y}_{F}=0,\ \ \mathbf{y}_{F}=\mathbf{1}_{n}\otimes y_{F}. (15)

Thus, 𝐲F=0\mathbf{y}_{F}=0 holds and combining it with (12) yields

𝝁F=𝟏nμF.\displaystyle\boldsymbol{\mu}_{F}=\mathbf{1}_{n}\otimes\mu_{F}. (16)

Moreover, with proper initialization, from (8) we have

(𝟏nIm)(𝐀𝐱F𝐝)=(𝟏nIm)𝐲F=0.\displaystyle(\mathbf{1}_{n}\otimes I_{m})^{\intercal}(\mathbf{A}\mathbf{x}_{F}-\mathbf{d})=(\mathbf{1}_{n}\otimes I_{m})^{\intercal}\mathbf{y}_{F}=0. (17)

where (13), (16) and (17) correspond to the optimal conditions in Proposition 1 which completes the proof. \hfill\blacksquare

With privacy concerns, agents exchange noise-masked information against potential attackers to prevent information disclosure. In [20], Laplace noise is shown to satisfy the necessary and sufficient condition of differential privacy. Furthermore, due to the noise accumulation in the tracking process as i=1nyi=i=1nAixii=1ndi+i=1nt=0n1ζi(t)\sum_{i=1}^{n}y_{i}=\sum_{i=1}^{n}A_{i}x_{i}-\sum_{i=1}^{n}d_{i}+\sum_{i=1}^{n}\sum_{t=0}^{n-1}\zeta_{i}(t), the added noise ζi\zeta_{i} needs to decay to ensure the convergence of yiy_{i}. Likewise, ηi\eta_{i} also needs to decay for the convergence of μi\mu_{i} and yiy_{i}. Thus, the added noises are set to ηi(k)Lap(θikη)\eta_{i}(k)\sim Lap(\theta_{ik}^{\eta}), ζi(k)Lap(θikζ)\zeta_{i}(k)\sim Lap(\theta_{ik}^{\zeta}) where θikη=dηiqik,θikζ=dζiqik,qi(0,1)\theta_{ik}^{\eta}=d_{\eta_{i}}{q_{i}}^{k},\theta_{ik}^{\zeta}=d_{\zeta_{i}}{q_{i}}^{k},q_{i}\in(0,1). That is, two diminishing Laplace noises are added to the communication process of μi\mu_{i}, yiy_{i}, respectively. Both of them are shown to be necessary in Section V.

Based on the mismatch tracking scheme and the information-masked protocol, a differentially private distributed mismatch tracking algorithm (diff-DMAC) is proposed which is shown in Algorithm 1 in details.

Algorithm 1 : diff-DMAC Algorithm
1:  Initialization: xi(0)x_{i}(0) and μi(0)\mu_{i}(0) are arbitrarily assigned and yi(0)=Aixi(0)diy_{i}(0)=A_{i}x_{i}(0)-d_{i}.
2:  for k=0,1,k=0,1,... do
3:     Adds noise ηi(k),ζi(k)\eta_{i}(k),\zeta_{i}(k) to get zμi(k)z_{\mu_{i}}(k) and zyi(k)z_{y_{i}}(k) as zμj(k)=μj(k)+ηj(k)z_{\mu_{j}}(k)=\mu_{j}(k)+\eta_{j}(k), zyj(k)=yj(k)+ζj(k)z_{y_{j}}(k)=y_{j}(k)+\zeta_{j}(k).
4:     Broadcasts zμi(k)z_{\mu_{i}}(k) and zyi(k)z_{y_{i}}(k) to j𝒩ij\in\mathcal{N}_{i}.
5:     Receives zμj(k)z_{\mu_{j}}(k) and zyj(k)z_{y_{j}}(k) from j𝒩ij\in\mathcal{N}_{i}
6:     Updates μi(k+1)\mu_{i}(k+1) through μi(k+1)=j=1nwijzμj(k)αyi(k)\mu_{i}(k+1)=\sum_{j=1}^{n}w_{ij}z_{\mu_{j}}(k)-\alpha y_{i}(k)
7:     Updates xi(k+1)x_{i}(k+1) through xi(k+1)=argminz𝒳i{fi(z)μi(k+1)Aiz}x_{i}(k+1)=\arg\min_{z\in\mathcal{X}_{i}}\{f_{i}(z)-\mu_{i}^{\intercal}(k+1)A_{i}z\}
8:     Updates yi(k+1)y_{i}(k+1) through yi(k+1)=j=1nwijzyj(k)+Aixi(k+1)Aixi(k).y_{i}(k+1)=\sum_{j=1}^{n}w_{ij}z_{y_{j}}(k)+A_{i}x_{i}(k+1)-A_{i}x_{i}(k).
9:  end for

IV Convergence Analysis for diff-DMAC

In this section, we will establish the linear convergence rate of the proposed algorithm in mean square. And the upper and lower bounds of the convergence accuracy is also provided. We set di=0d_{i}=0, i\forall i in the following analysis for simplicity while nonzero scenarios can be analyzed similarly.

Let f(x)=[f1(x1),,fn(xn)]\nabla f(x)=[{\nabla f_{1}(x_{1})}^{\intercal},\cdots,{\nabla f_{n}(x_{n})}^{\intercal}]^{\intercal} and 𝐖¯=𝟏n𝟏nnIm\mathbf{\bar{W}}=\frac{\mathbf{1}_{n}^{\intercal}\mathbf{1}_{n}}{n}\otimes I_{m}, 𝐖ˇ=Inm𝐖¯\mathbf{\check{W}}=I_{nm}-\mathbf{\bar{W}}, 𝐖~=𝐖𝐖ˇ\mathbf{\tilde{W}}=\mathbf{W}\mathbf{\check{W}}. From (7), (9), we have that

𝝁¯(k+1)=𝝁¯(k)+𝜼¯(k)α𝒚¯(k),\displaystyle\bar{\boldsymbol{\mu}}(k+1)=\bar{\boldsymbol{\mu}}(k)+\bar{\boldsymbol{\eta}}(k)-\alpha\bar{\boldsymbol{y}}(k), (18)
𝒚¯(k+1)=𝒚¯(k)+𝜻¯(k)+𝐖¯𝐀(𝐱(k+1)𝐱(k)),\displaystyle\bar{\boldsymbol{y}}(k+1)=\bar{\boldsymbol{y}}(k)+\bar{\boldsymbol{\zeta}}(k)+\mathbf{\bar{W}}\mathbf{A}({\mathbf{x}}(k+1)-{\mathbf{x}}(k)), (19)
𝝁ˇ(k+1)=𝐖~(𝝁ˇ(k)+𝜼ˇ(k))α𝒚ˇ(k),\displaystyle\check{\boldsymbol{\mu}}(k+1)=\mathbf{\tilde{W}}(\check{\boldsymbol{\mu}}(k)+\check{\boldsymbol{\eta}}(k))-\alpha\check{\boldsymbol{y}}(k), (20)
𝒚ˇ(k+1)=𝐖~(𝒚ˇ(k)+𝜻ˇ(k))+𝐖ˇ𝐀(𝐱(𝐤+𝟏)𝐱(𝐤)),\displaystyle\check{\boldsymbol{y}}(k+1)=\mathbf{\tilde{W}}(\check{\boldsymbol{y}}(k)+\check{\boldsymbol{\zeta}}(k))+\mathbf{\check{W}}\mathbf{A}(\mathbf{x(k+1)-x(k)}), (21)

where 𝝁¯=𝐖¯𝝁\bar{\boldsymbol{\mu}}=\mathbf{\bar{W}}{\boldsymbol{\mu}}, 𝝁ˇ=𝐖ˇ𝝁\check{\boldsymbol{\mu}}=\mathbf{\check{W}}{\boldsymbol{\mu}} and the doubly stochasticity of the weight matrix is used here. Under Assumption 2, as a consequence of the Perron–Frobenius theorem, we have that 𝒲𝟏𝟏n<1\|\mathcal{W}-\frac{\mathbf{1}^{\intercal}\mathbf{1}}{n}\|<1, and thus

λ¯=𝐖~=(𝒲𝟏𝟏n)Im<1.\displaystyle\bar{\lambda}=\|\mathbf{\tilde{W}}\|=\|(\mathcal{W}-\frac{\mathbf{1}^{\intercal}\mathbf{1}}{n})\otimes I_{m}\|<1. (22)

Lemma 2: For positive a1a_{1}, a2a_{2}, a3>0a_{3}>0, if a1a2a3>1a_{1}a_{2}a_{3}>1 holds, there exist γ1\gamma_{1}, γ2\gamma_{2}, γ3>0\gamma_{3}>0 such that

γ1<a1γ2,γ2<a2γ3,γ3<a3γ1.\displaystyle\gamma_{1}<a_{1}\gamma_{2},\ \gamma_{2}<a_{2}\gamma_{3},\ \gamma_{3}<a_{3}\gamma_{1}. (23)

The proof is omitted. Note that for any β>0\beta>0, γi=βγi\gamma_{i}^{\prime}=\beta\gamma_{i}, i=1,2,3i=1,2,3, also satisfy the condition (23). Define C={1+(𝐀2α2φ¯22αL¯)λ¯(AA)}12C=\{1+(\frac{\|\mathbf{A}\|^{2}\alpha^{2}}{\underline{\varphi}^{2}}-\frac{2\alpha}{\overline{L}})\cdot\underline{\lambda}(A^{\intercal}A)\}^{\frac{1}{2}}. Based on the above results, the following Theorem is proved by induction argument.

Theorem 1: Given a constant r(rLB,)r\in(r_{LB},\infty), where rLBr_{LB} is defined as rLB=max{qζi,qηi,C,λ¯}r_{LB}=\max\{q_{\zeta_{i}},q_{\eta_{i}},C,\bar{\lambda}\}. Suppose Assumptions 1,2 hold, α\alpha satisfies

α<φ¯22𝐀2L¯ and (rC)φ¯α𝐀((rλ¯)2φ¯2α𝐀1)>1.\displaystyle\alpha<\frac{\underline{\varphi}^{2}}{2\|\mathbf{A}\|^{2}\overline{L}}\text{ and }\frac{(r-C)\underline{\varphi}}{\alpha\|\mathbf{A}\|}(\frac{(r-\bar{\lambda})^{2}\underline{\varphi}}{2\alpha\|\mathbf{A}\|}-1)>1. (24)

then the sequences generated by diff-DMAC satisfy

𝔼[𝝁ˇ(k)]γ1rk,\displaystyle\mathbb{E}[\|\check{\boldsymbol{\mu}}(k)\|]\leq\gamma_{1}r^{k}, (25)
𝔼[𝒚ˇ(k)]γ2rk,\displaystyle\mathbb{E}[\|\check{\boldsymbol{y}}(k)\|]\leq\gamma_{2}r^{k}, (26)
𝔼[𝒙(k)𝒙]γ3rk,\displaystyle\mathbb{E}[\|\boldsymbol{x}(k)-\boldsymbol{x}^{\infty}\|]\leq\gamma_{3}r^{k}, (27)
𝔼[𝝁¯(k)𝝁]γ4rk,\displaystyle\mathbb{E}[\|\bar{\boldsymbol{\mu}}(k)-\boldsymbol{\mu}^{\infty}\|]\leq\gamma_{4}r^{k}, (28)

for some constant γ1,γ2,γ3,γ4>0\gamma_{1},\gamma_{2},\gamma_{3},\gamma_{4}>0, where 𝐱\mathbf{x}^{\infty} and 𝝁\boldsymbol{\mu}^{\infty} satisfy

𝐖¯𝐀𝐱=k=1𝜻¯(k),𝐖¯𝝁=𝝁,\displaystyle\bar{\mathbf{W}}\mathbf{A}{\mathbf{x}}^{\infty}=-\sum\limits_{k=1}^{\infty}\bar{\boldsymbol{\zeta}}(k),\ \bar{\mathbf{W}}{\boldsymbol{\mu}}^{\infty}={\boldsymbol{\mu}}^{\infty}, (29)
xi=argminz𝒳i{fi(z)μiAiz},i.\displaystyle x_{i}^{\infty}=\arg\min_{z\in\mathcal{X}_{i}}\{\nabla f_{i}(z)-{\mu_{i}^{\infty}}^{\intercal}A_{i}z\},\forall i. (30)

Proof: Combining (18) and (29), we have that

𝝁¯(k+1)𝝁¯=\displaystyle\bar{\boldsymbol{\mu}}(k+1)-\bar{\boldsymbol{\mu}}^{\infty}= 𝝁¯(k)𝝁¯α𝐖¯𝐀(𝐱(k)𝐱)\displaystyle\bar{\boldsymbol{\mu}}(k)-\bar{\boldsymbol{\mu}}^{\infty}-\alpha\bar{\mathbf{W}}\mathbf{A}({\mathbf{x}}(k)-{\mathbf{x}}^{\infty})
+αt=k𝜻¯(t)+𝜼¯(k).\displaystyle+\alpha\sum\limits_{t=k}^{\infty}\bar{\boldsymbol{\zeta}}(t)+\bar{\boldsymbol{\eta}}(k). (31)

Under Assumption 1, the function

fi(Aiμi)=argmaxz𝒳i{μiAizfi(z)}\displaystyle f_{i}^{*}(A_{i}^{\intercal}\mu_{i})=\arg\max_{z\in\mathcal{X}_{i}}\{\mu_{i}^{\intercal}A_{i}z-f_{i}(z)\}

is differentiable and fif_{i}^{*} satisfies 1Li\frac{1}{L_{i}}-strongly convex and 1φi\frac{1}{\varphi_{i}}-Lipschitz smooth [21]. Furthermore, it follows (10) and the Proposition B.25 in [22] that

xi(k)=fi(Aiμi(k)),i.\displaystyle x_{i}(k)=\nabla f_{i}^{*}(A_{i}^{\intercal}\mu_{i}(k)),\forall i. (32)

From (30) and (32), we can obtain that

𝝁¯(k)𝝁α𝐖¯𝐀(𝐱(k)𝐱)\displaystyle\|\bar{\boldsymbol{\mu}}(k)-\boldsymbol{\mu}^{\infty}-\alpha\bar{\mathbf{W}}\mathbf{A}({\mathbf{x}}(k)-{\mathbf{x}}^{\infty})\|
=𝝁¯(k)𝝁α𝐖¯𝐀(f(𝐀𝝁(k))f(𝐀𝝁))\displaystyle=\|\bar{\boldsymbol{\mu}}(k)-\boldsymbol{\mu}^{\infty}-\alpha\bar{\mathbf{W}}\mathbf{A}(\nabla f^{*}(\mathbf{A}^{\intercal}\boldsymbol{\mu}(k))-\nabla f^{*}(\mathbf{A}^{\intercal}\boldsymbol{\mu}^{\infty}))\|
{𝝁¯(k)𝝁2+α2f(𝐀𝝁¯(k))f(𝐀𝝁)\displaystyle\leq\left\{\|\bar{\boldsymbol{\mu}}(k)-\boldsymbol{\mu}^{\infty}\|^{2}+\alpha^{2}\|\nabla f^{*}(\mathbf{A}^{\intercal}\bar{\boldsymbol{\mu}}(k))-\nabla f^{*}(\mathbf{A}^{\intercal}\boldsymbol{\mu}^{\infty})\|\right.
2α(𝝁¯(k)𝝁)𝐀(f(𝐀𝝁¯(k))f(𝐀𝝁))}12\displaystyle\ \ \ \left.-2\alpha(\bar{\boldsymbol{\mu}}(k)-\boldsymbol{\mu}^{\infty})^{\intercal}\mathbf{A}(\nabla f^{*}(\mathbf{A}^{\intercal}\bar{\boldsymbol{\mu}}(k))-\nabla f^{*}(\mathbf{A}^{\intercal}\boldsymbol{\mu}^{\infty}))\right\}^{\frac{1}{2}}
+αf(𝐀𝝁(k))f(𝐀𝝁¯(k))\displaystyle\ \ \ +\alpha\|\nabla f^{*}({\mathbf{A}^{\intercal}\boldsymbol{\mu}}(k))-\nabla f^{*}(\mathbf{A}^{\intercal}\bar{\boldsymbol{\mu}}(k))\|
C𝝁¯(k)𝝁+𝐀2αφ¯𝝁ˇ(k),\displaystyle\leq C\|\bar{\boldsymbol{\mu}}(k)-\boldsymbol{\mu}^{\infty}\|+\frac{\|\mathbf{A}\|^{2}\alpha}{\underline{\varphi}}\|\check{\boldsymbol{\mu}}(k)\|, (33)

where C<1C<1 under (24) and the last inequality follows from the strongly convexity and Lipschitz smoothness of ff^{*}.

We will prove (25)-(28) by induction. When κ=0\kappa=0, it is not difficult to find γ1,γ2,γ3,γ4>0\gamma_{1},\gamma_{2},\gamma_{3},\gamma_{4}>0 such that (25)-(28) hold. We assume that for all κk\kappa\leq k, (25)-(28) hold. Considering κ=k+1\kappa=k+1, we obtain from (31) and (33) that

𝔼[𝝁¯(k+1)𝝁]\displaystyle\mathbb{E}[\|\bar{\boldsymbol{\mu}}(k+1)-\boldsymbol{\mu}^{\infty}\|]
C𝔼[𝝁¯(k)𝝁]+𝐀αφ¯𝔼[𝝁ˇ(k)]\displaystyle\leq C\mathbb{E}[\|\bar{\boldsymbol{\mu}}(k)-\boldsymbol{\mu}^{\infty}\|]+\frac{\|\mathbf{A}\|\alpha}{\underline{\varphi}}\mathbb{E}[\|\check{\boldsymbol{\mu}}(k)\|]
+𝔼[αt=k𝜻¯(t)]𝔼[𝜼¯(k)]\displaystyle\ \ \ +\mathbb{E}[\|\alpha\sum\limits_{t=k}^{\infty}\bar{\boldsymbol{\zeta}}(t)\|]-\mathbb{E}[\|\bar{\boldsymbol{\eta}}(k)\|]
Cγ4rk+𝐀αφ¯γ1rk+nmq¯kd¯η+αnmq¯kd¯ζ1q¯.\displaystyle\leq C\gamma_{4}r^{k}+\frac{\|\mathbf{A}\|\alpha}{\bar{\varphi}}\gamma_{1}r^{k}+nm\bar{q}^{k}\bar{d}_{\eta}+\frac{\alpha nm\bar{q}^{k}\bar{d}_{\zeta}}{1-\bar{q}}. (34)

Combining (20)-(22) gives

𝔼[𝝁ˇ(k+1)]\displaystyle\mathbb{E}[\|\check{\boldsymbol{\mu}}(k+1)\|] λ¯𝔼[𝝁ˇ(k)]+α𝔼[𝒚ˇ(k)]+λ¯𝔼[𝜼ˇ(k)]\displaystyle\leq\bar{\lambda}\mathbb{E}[\|\check{\boldsymbol{\mu}}(k)\|]+\alpha\mathbb{E}[\|\check{\boldsymbol{y}}(k)\|]+\bar{\lambda}\mathbb{E}[\|\check{\boldsymbol{\eta}}(k)\|]
λ¯γ1rk+αγ2rk+λ¯nmq¯kd¯η,\displaystyle\leq\bar{\lambda}\gamma_{1}r^{k}+\alpha\gamma_{2}r^{k}+\bar{\lambda}nm\bar{q}^{k}\bar{d}_{\eta}, (35)
𝔼[𝒚ˇ(k+1)]\displaystyle\mathbb{E}[\|\check{\boldsymbol{y}}(k+1)\|] λ¯𝔼[𝒚ˇ(k)]+𝐀𝔼[𝒙(k)𝒙]\displaystyle\leq\bar{\lambda}\mathbb{E}[\|\check{\boldsymbol{y}}(k)\|]+\|\mathbf{A}\|\mathbb{E}[\|\boldsymbol{x}(k)-\boldsymbol{x}^{\infty}\|]
+𝐀𝔼[𝒙(k+1)𝒙]+λ¯𝔼[𝜻ˇ(k)]\displaystyle\ \ \ +\|\mathbf{A}\|\mathbb{E}[\|\boldsymbol{x}(k+1)-\boldsymbol{x}^{\infty}\|]+\bar{\lambda}\mathbb{E}[\|\check{\boldsymbol{\zeta}}(k)\|]
λ¯γ2rk+𝐀γ3rk+λ¯nmq¯kd¯ζ\displaystyle\leq\bar{\lambda}\gamma_{2}r^{k}+\|\mathbf{A}\|\gamma_{3}r^{k}+\bar{\lambda}nm\bar{q}^{k}\bar{d}_{\zeta}
+𝐀𝔼[𝒙(k+1)𝒙].\displaystyle\ \ \ +\|\mathbf{A}\|\mathbb{E}[\|\boldsymbol{x}(k+1)-\boldsymbol{x}^{\infty}\|]. (36)

Under Assumption 1, (32) implies that

𝔼[𝒙(k+1)𝒙]=𝔼[f(𝐀𝝁(k+1))f(𝐀𝝁)]\displaystyle\mathbb{E}[\|\boldsymbol{x}(k+1)-\boldsymbol{x}^{\infty}\|]=\mathbb{E}[\|\nabla f^{*}(\mathbf{A}\boldsymbol{\mu}(k+1))-\nabla f^{*}(\mathbf{A}\boldsymbol{\mu}^{\infty})\|]
𝐀φ¯𝔼[𝝁ˇ(k+1)]+𝐀φ¯𝔼[𝝁¯(k+1)𝝁],\displaystyle\leq\frac{\|\mathbf{A}\|}{\underline{\varphi}}\mathbb{E}[\|\check{\boldsymbol{\mu}}(k+1)\|]+\frac{\|\mathbf{A}\|}{\underline{\varphi}}\mathbb{E}[\|\bar{\boldsymbol{\mu}}(k+1)-\boldsymbol{\mu}^{\infty}\|], (37)

where the last inequality is due to the Lipschitz smoothness of ff^{*}.

Note that constant λ¯,C(0,1)\bar{\lambda},C\in(0,1), due to Lemma 2, there exist γ1,γ2,γ3,γ4>0\gamma_{1},\gamma_{2},\gamma_{3},\gamma_{4}>0 such that

Cγ4+𝐀αφ¯γ1<γ4r,\displaystyle C\gamma_{4}+\frac{\|\mathbf{A}\|\alpha}{\underline{\varphi}}\gamma_{1}<\gamma_{4}r, (38)
λ¯γ1+αγ2<γ1r,\displaystyle\bar{\lambda}\gamma_{1}+\alpha\gamma_{2}<\gamma_{1}r, (39)
λ¯γ2+2𝐀γ3<γ2r,\displaystyle\bar{\lambda}\gamma_{2}+2\|\mathbf{A}\|\gamma_{3}<\gamma_{2}r, (40)
𝐀φ¯(γ1+γ4)<γ3.\displaystyle\frac{\|\mathbf{A}\|}{\underline{\varphi}}(\gamma_{1}+\gamma_{4})<\gamma_{3}. (41)

Basd on equations (38)-(41), taking suffciently large γ1,γ2,γ3,γ4>0\gamma_{1},\gamma_{2},\gamma_{3},\gamma_{4}>0 yields

Cγ4rk+𝐀αφ¯γ1rk+nmq¯kd¯η+αnmq¯kd¯ζ1q¯\displaystyle C\gamma_{4}r^{k}+\frac{\|\mathbf{A}\|\alpha}{\bar{\varphi}}\gamma_{1}r^{k}+nm\bar{q}^{k}\bar{d}_{\eta}+\frac{\alpha nm\bar{q}^{k}\bar{d}_{\zeta}}{1-\bar{q}} γ4rk+1,\displaystyle\leq\gamma_{4}r^{k+1},
λ¯γ1rk+αγ2rk+λ¯nmq¯kd¯η\displaystyle\bar{\lambda}\gamma_{1}r^{k}+\alpha\gamma_{2}r^{k}+\bar{\lambda}nm\bar{q}^{k}\bar{d}_{\eta} γ1rk+1,\displaystyle\leq\gamma_{1}r^{k+1},
λ¯γ2rk+2𝐀γ3rk\displaystyle\bar{\lambda}\gamma_{2}r^{k}+2\|\mathbf{A}\|\gamma_{3}r^{k} γ2rk+1,\displaystyle\leq\gamma_{2}r^{k+1},
𝐀φ¯(γ1+γ4)rk+1\displaystyle\frac{\|\mathbf{A}\|}{\underline{\varphi}}(\gamma_{1}+\gamma_{4})r^{k+1} γ3rk+1,\displaystyle\leq\gamma_{3}r^{k+1},

where we use the definition of rLBr_{LB} and substituting these equations into (34)-(37) completes the proof. \hfill\blacksquare

Based on Theorem 1, we prove the linear convergence property of diff-DMAC and quantify its convergence accuracy.

Theorem 2: Suppose the conditions in Theorem 1 are satisfied. If α\alpha further satisfies

α<Φ¯[(1C)+(1C)2+2(1C)(1λ¯)2]2𝐀,\displaystyle\alpha<\frac{\underline{\varPhi}[-(1-C)+\sqrt{(1-C)^{2}+2(1-C)(1-\bar{\lambda})^{2}}]}{2\|\mathbf{A}\|}, (42)

the sequence {𝐱(𝐤)}k0\{\mathbf{x(k)}\}_{k\geq 0} generated by diff-DMAC will converge linearly to the neighbor of the optimum 𝐱\mathbf{x}^{\star} in mean square and we have

1n2𝐀2Nζ𝔼{𝒙𝒙2}L¯2nφ¯2[λ¯(𝐀𝐀)]2Nζ,\displaystyle\frac{1}{n^{2}\|\mathbf{A}\|^{2}}N_{\zeta}\leq\mathbb{E}\{\|\boldsymbol{x}^{\infty}-\boldsymbol{x}^{\star}\|^{2}\}\leq\frac{\overline{L}^{2}}{n\underline{\varphi}^{2}[\underline{\lambda}(\mathbf{A}^{\intercal}\mathbf{A})]^{2}}N_{\zeta}, (43)

where NζN_{\zeta} is defined as Nζ=i=1n2mdζi21qζi2N_{\zeta}=\sum_{i=1}^{n}\frac{2md_{\zeta i}^{2}}{1-q_{\zeta i}^{2}} .

Proof: Note that λ¯,C(0,1)\bar{\lambda},C\in(0,1), it is not difficult to see that with α\alpha chosen to satisfy (42), we can always find a constant r(rLB,1)r\in(r_{LB},1) such that equations (25)-(28) hold. Thus, the convergence of diff-DMAC is proved by Theorem 1.

Due to (29) and 𝐖¯𝐀𝐱=0\bar{\mathbf{W}}\mathbf{A}{\mathbf{x}}^{\star}=0, we have

𝔼[𝐖¯𝐀(𝐱𝐱)2]=𝔼[t=0𝜻¯(t)2]=1n2i=1n2mdζi21qζi2.\displaystyle\mathbb{E}[\|\bar{\mathbf{W}}\mathbf{A}({\mathbf{x}}^{\infty}-{\mathbf{x}}^{\star})\|^{2}]=\mathbb{E}[\|\sum\limits_{t=0}^{\infty}\bar{\boldsymbol{\zeta}}(t)\|^{2}]=\frac{1}{n^{2}}\sum\limits_{i=1}^{n}\frac{2md_{\zeta_{i}}^{2}}{1-q_{\zeta_{i}}^{2}}. (44)

Since 𝐖¯=𝟏𝟏nIm\mathbf{\bar{W}}=\frac{\mathbf{1}^{\intercal}\mathbf{1}}{n}\otimes I_{m}, it is easy to obtain that

𝔼[𝐖¯𝐀(𝐱𝐱)2]𝐀2𝔼[𝐱𝐱2].\displaystyle\mathbb{E}[\|\bar{\mathbf{W}}\mathbf{A}({\mathbf{x}}^{\infty}-{\mathbf{x}}^{\star})\|^{2}]\leq\|\mathbf{A}\|^{2}\mathbb{E}[\|\mathbf{x}^{\infty}-\mathbf{x}^{\star}\|^{2}]. (45)

The optimality condition of the 𝐱\mathbf{x}-update (10) implies that

(fi(xi)Aiμi)(xixi)\displaystyle(\nabla f_{i}(x_{i}^{\infty})-A_{i}^{\intercal}\mu_{i}^{\infty})^{\intercal}(x_{i}^{\star}-x_{i}^{\infty}) 0,\displaystyle\geq 0, (46)
(fi(xi)Aiμi)(xixi)\displaystyle(\nabla f_{i}(x_{i}^{\star})-A_{i}^{\intercal}\mu_{i}^{\star})^{\intercal}(x_{i}^{\infty}-x_{i}^{\star}) 0,i.\displaystyle\geq 0,\forall i. (47)

Based on (46) and (47), we obtain

(μiμi)Ai(xixi)\displaystyle(\mu_{i}^{\infty}-\mu_{i}^{\star})^{\intercal}A_{i}(x_{i}^{\infty}-x_{i}^{\star})
(fi(xi)fi(xi))(xixi)\displaystyle\geq(\nabla f_{i}(^{\intercal}x_{i}^{\infty})-\nabla f_{i}(^{\intercal}x_{i}^{\star}))^{\intercal}(x_{i}^{\infty}-x_{i}^{\star})
Φ¯AiL¯μiμixixi,i.\displaystyle\geq\frac{\underline{\varPhi}\|A_{i}\|}{\bar{L}}\|\mu_{i}^{\infty}-\mu_{i}^{\star}\|\|x_{i}^{\infty}-x_{i}^{\star}\|,\forall i. (48)

The facts that 𝐖¯𝝁=𝝁\bar{\mathbf{W}}{\boldsymbol{\mu}}^{\infty}={\boldsymbol{\mu}}^{\infty} and 𝐖¯𝝁=𝝁\bar{\mathbf{W}}{\boldsymbol{\mu}}^{\star}={\boldsymbol{\mu}}^{\star} imply

μiμi\displaystyle\mu_{i}^{\infty}-\mu_{i}^{\star} =1ni=1nμi1ni=1nμi,i𝒱,\displaystyle=\frac{1}{n}\sum\limits_{i=1}^{n}\mu_{i}^{\infty}-\frac{1}{n}\sum\limits_{i=1}^{n}\mu_{i}^{\star},\ \forall i\in\mathcal{V}, (49)
(𝝁𝝁)𝐀(𝐱𝐱)\displaystyle(\boldsymbol{\mu}^{\infty}-\boldsymbol{\mu}^{\star})^{\intercal}\mathbf{A}(\mathbf{x}^{\infty}-\mathbf{x}^{\star}) =(𝝁𝝁)𝐖¯𝐀(𝐱𝐱)\displaystyle=(\boldsymbol{\mu}^{\infty}-\boldsymbol{\mu}^{\star})^{\intercal}\bar{\mathbf{W}}\mathbf{A}({\mathbf{x}}^{\infty}-{\mathbf{x}}^{\star})
𝝁𝝁𝐖¯𝐀(𝐱𝐱),\displaystyle\leq\|\boldsymbol{\mu}^{\infty}-\boldsymbol{\mu}^{\star}\|\|\bar{\mathbf{W}}\mathbf{A}({\mathbf{x}}^{\infty}-{\mathbf{x}}^{\star})\|, (50)

Then combining (48)-(50) yields

𝔼[𝐱𝐱2]L¯2nφ¯2[λ¯(𝐀𝐀)]2𝔼[𝐖¯𝐀(𝐱𝐱)2].\displaystyle\mathbb{E}[\|\mathbf{x}^{\infty}-\mathbf{x}^{\star}\|^{2}]\leq\frac{\bar{L}^{2}}{n\underline{\varphi}^{2}[\underline{\lambda}(\mathbf{A}^{\intercal}\mathbf{A})]^{2}}\mathbb{E}[\|\bar{\mathbf{W}}\mathbf{A}({\mathbf{x}}^{\infty}-{\mathbf{x}}^{\star})\|^{2}]. (51)

Substituting (44) into (45), (51) completes the proof. \hfill\blacksquare

V Differential Privacy

In this section, we present the privacy preserving performance of the proposed algorithm by theoretical analysis.

Theorem 3: Suppose the conditions in Theorem 2 hold, if qi0q_{i_{0}} satisfies

qi0(αAi02+Ai0α2Ai02+4αφi02φi0,1),\displaystyle q_{i_{0}}\in(\frac{\alpha\|A_{i_{0}}\|^{2}+\|A_{i_{0}}\|\sqrt{\alpha^{2}\|A_{i_{0}}\|^{2}+4\alpha\varphi_{i_{0}}}}{2\varphi_{i_{0}}},1), (52)

the proposed diff-DMAC preservses ϵi0\epsilon_{i_{0}}-differential privacy for the cost function of agent i0i_{0} , where

ϵi0=(1αdζi0+1dηi0)αφi0δAi0φi0qi02αqi0α.\displaystyle\epsilon_{i_{0}}=(\frac{1}{\alpha d_{\zeta_{i_{0}}}}+\frac{1}{d_{\eta_{i_{0}}}})\frac{\alpha\varphi_{i_{0}}\delta\|A_{i_{0}}\|}{\varphi_{i_{0}}q_{i_{0}}^{2}-\alpha q_{i_{0}}-\alpha}. (53)

Proof: Two function sets have the same initialization and all auxiliary variables are known to attackers as mentioned in Section II. Therefore, exchanged variables must be same, i.e., 𝒛𝝁(1)(k)=𝒛𝝁(2)(k)\boldsymbol{z}_{\boldsymbol{\mu}}^{(1)}(k)=\boldsymbol{z}_{\boldsymbol{\mu}}^{(2)}(k) and 𝒛𝒚(1)(k)=𝒛𝒚(2)(k)\boldsymbol{z}_{\boldsymbol{y}}^{(1)}(k)=\boldsymbol{z}_{\boldsymbol{y}}^{(2)}(k), otherwise, the two function sets will be easily distinguished. Therefore, for any ii0i\neq i_{0}, due to the realation fi(1)=fi(2)f_{i}^{(1)}=f_{i}^{(2)}, the added noise should satisfy

ζi(1)(k)=ζi(2)(k),ηi(1)(k)=ηi(2)(k),k,ii0.\zeta_{i}^{(1)}(k)=\zeta_{i}^{(2)}(k),\ \eta_{i}^{(1)}(k)=\eta_{i}^{(2)}(k),\forall k,\forall i\neq i_{0}.

And for i=i0i=i_{0}, since fi(1)fi(2)f_{i}^{(1)}\neq f_{i}^{(2)} the noise should satisfy

Δηi0(k)=Δμi0(k),Δζi0(k)=Δyi0(k),\displaystyle\Delta\eta_{i_{0}}(k)=-\Delta\mu_{i_{0}}(k),\ \Delta\zeta_{i_{0}}(k)=-\Delta y_{i_{0}}(k), (54)

where Δζi0(k)=ζi0(1)(k)ζi0(2)(k)\Delta\zeta_{i_{0}}(k)=\zeta_{i_{0}}^{(1)}(k)-\zeta_{i_{0}}^{(2)}(k), Δηi0(k)=ηi0(1)(k)ηi0(2)(k)\Delta\eta_{i_{0}}(k)=\eta_{i_{0}}^{(1)}(k)-\eta_{i_{0}}^{(2)}(k), Δμi0(k)=μi0(1)(k)μi0(2)(k)\Delta\mu_{i_{0}}(k)=\mu_{i_{0}}^{(1)}(k)-\mu_{i_{0}}^{(2)}(k) and Δyi0(k)=yi0(1)(k)yi0(2)(k)\Delta y_{i_{0}}(k)=y_{i_{0}}^{(1)}(k)-y_{i_{0}}^{(2)}(k). Furthermore, it follows that

Δμi0(k+1)=αΔyi0(k),Δyi0(k+1)=ΔAi0gi0(k),\displaystyle\Delta\mu_{i_{0}}(k+1)=-\alpha\Delta y_{i_{0}}(k),\ \Delta y_{i_{0}}(k+1)=\Delta A_{i_{0}}g_{i_{0}}(k), (55)

where Δgi0(k)=Δxi0(k+1)Δxi0(k)\Delta g_{i_{0}}(k)=\Delta x_{i_{0}}(k+1)-\Delta x_{i_{0}}(k) and Δxi0(k)=xi0(1)(k)xi0(2)(k)\Delta x_{i_{0}}(k)=x_{i_{0}}^{(1)}(k)-x_{i_{0}}^{(2)}(k).

Following the analysis in [13], we need to obtain the upper bound of fζη(𝜻(1),𝜼(1))fζη((𝜻(1),𝜼(1)))\frac{f_{\zeta\eta}(\boldsymbol{\zeta}^{(1)},\boldsymbol{\eta}^{(1)})}{f_{\zeta\eta}(\mathcal{B}(\boldsymbol{\zeta}^{(1)},\boldsymbol{\eta}^{(1)}))} to quantify the privacy level ϵi0\epsilon_{i_{0}},

{𝜻,𝜼Ω|ZT(1)(𝜻,𝜼)𝒪}{𝜻,𝜼Ω|ZT(2)(𝜻,𝜼)𝒪}sup(𝜻,𝜼)fζη(𝜻(1),𝜼(1))fζη((𝜻(1),𝜼(1))),\displaystyle\frac{\mathbb{P}\{\boldsymbol{\zeta},\boldsymbol{\eta}\in\Omega|Z_{T^{(1)}}(\boldsymbol{\zeta},\boldsymbol{\eta})\in\mathcal{O}\}}{\mathbb{P}\{\boldsymbol{\zeta},\boldsymbol{\eta}\in\Omega|Z_{T^{(2)}}(\boldsymbol{\zeta},\boldsymbol{\eta})\in\mathcal{O}\}}\leq\sup_{(\boldsymbol{\zeta},\boldsymbol{\eta})}\frac{f_{\zeta\eta}(\boldsymbol{\zeta}^{(1)},\boldsymbol{\eta}^{(1)})}{f_{\zeta\eta}(\mathcal{B}(\boldsymbol{\zeta}^{(1)},\boldsymbol{\eta}^{(1)}))}, (56)

where ZT(l)(𝜻(l)𝜼(l))={𝒛𝝁(l)(k),𝒛𝒚(l)(k)}k0,l=1,2Z_{T^{(l)}}(\boldsymbol{\zeta}^{(l)}\boldsymbol{\eta}^{(l)})=\{\boldsymbol{z}_{\boldsymbol{\mu}}^{(l)}(k),\boldsymbol{z}_{\boldsymbol{y}}^{(l)}(k)\}_{k\geq 0},l=1,2, belongs to attacker’s observation.

For l{1,2}l\in\{1,2\}, the 𝐱\mathbf{x}-update can be reritten as

xi0(l)(k+1)=argminz𝒳i0(l){fi0(l)(z)μi0(l)(k+1)Ai0z}.\displaystyle x_{i_{0}}^{(l)}(k+1)=\arg\min_{z\in\mathcal{X}_{i_{0}}^{(l)}}\{f_{i_{0}}^{(l)}(z)-{\mu_{i_{0}}^{(l)}}^{\intercal}(k+1)A_{i_{0}}z\}.

From Definition 1, we have xi0(1),xi0(2)δ𝒳i0(1)x_{i_{0}}^{(1)},x_{i_{0}}^{(2)}-\delta^{\prime}\in\mathcal{X}_{i_{0}}^{(1)} and xi0(2),xi0(1)+δ𝒳i0(1)x_{i_{0}}^{(2)},x_{i_{0}}^{(1)}+\delta^{\prime}\in\mathcal{X}_{i_{0}}^{(1)}. Similar to the analysis of Theorem 2, the optimal condition of the written 𝐱\mathbf{x}-update yields

(μi0(1)(k)μi0(2)(k))Ai0(xi0(1)(k)+δxi0(2)(k))\displaystyle(\mu_{i_{0}}^{(1)}(k)-\mu_{i_{0}}^{(2)}(k))^{\intercal}A_{i_{0}}(x_{i_{0}}^{(1)}(k)+\delta^{\prime}-x_{i_{0}}^{(2)}(k))
(fi0(1)(xi0(1)(k))fi0(2)(xi0(2)(k)))(xi0(1)(k)+δxi0(2)(k))\displaystyle\geq(\nabla f_{i_{0}}^{(1)}(x_{i_{0}}^{(1)}(k))-\nabla f_{i_{0}}^{(2)}(x_{i_{0}}^{(2)}(k)))^{\intercal}(x_{i_{0}}^{(1)}(k)+\delta^{\prime}-x_{i_{0}}^{(2)}(k))
φi0Ai02Ai0(xi0(1)(k)+δxi0(2)(k))2.\displaystyle\geq\frac{\varphi_{i_{0}}}{\|A_{i_{0}}\|^{2}}\|A_{i_{0}}(x_{i_{0}}^{(1)}(k)+\delta^{\prime}-x_{i_{0}}^{(2)}(k))\|^{2}. (57)

The last inequality uses the relation fi0(1)(xi0(1)(k))=fi0(2)(xi0(1)(k)+δ)\nabla f_{i_{0}}^{(1)}(x_{i_{0}}^{(1)}(k))=\nabla f_{i_{0}}^{(2)}(x_{i_{0}}^{(1)}(k)+\delta^{\prime}). Then applying Cauchy–Schwarz inequality to the LHS of (57) yields

Ai0(xi0(1)(k)+δxi0(2)(k))Ai02φi0Δμi0(k).\displaystyle\|A_{i_{0}}(x_{i_{0}}^{(1)}(k)+\delta^{\prime}-x_{i_{0}}^{(2)}(k))\|\leq\frac{\|A_{i_{0}}\|^{2}}{\varphi_{i_{0}}}\|\Delta\mu_{i_{0}}(k)\|. (58)

From (54), (55) and the relation (58), we have

Δηi0(k+1)=\displaystyle\left\|\Delta\eta_{i_{0}}(k+1)\right\|= αAi0Δgi0(k1)2\displaystyle\alpha\left\|A_{i_{0}}\Delta g_{i_{0}}(k-1)\right\|_{2}
=\displaystyle= αAi0(xi0(1)(k)xi0(2)(k)δ)\displaystyle\alpha\Big{\|}A_{i_{0}}(x_{i_{0}}^{(1)}(k)-x_{i_{0}}^{(2)}(k)-\delta)
Ai0(xi0(1)(k1)xi0(2)(k1)δ)\displaystyle-A_{i_{0}}\left(x_{i_{0}}^{(1)}(k-1)-x_{i_{0}}^{(2)}(k-1)-\delta\right)\Big{\|}
\displaystyle\leq αAi02φi0(Δηi0(k)2+Δηi0(k1)).\displaystyle\frac{\alpha\|A_{i_{0}}\|^{2}}{\varphi_{i_{0}}}\left(\left\|\Delta\eta_{i_{0}}(k)\right\|_{2}+\left\|\Delta\eta_{i_{0}}(k-1)\right\|\right). (59)

The two roots of φi0x2αAi02xαAi02=0\varphi_{i_{0}}x^{2}-\alpha\|A_{i_{0}}\|^{2}x-\alpha\|A_{i_{0}}\|^{2}=0, denoted as τ1,τ2\tau_{1},\tau_{2} are different and all lie in (1,1)(-1,1). Therefore, from (59), we have

Δηi0(k)2\displaystyle\|\Delta\eta_{i_{0}}(k)\|_{2} αδAi0τ1τ2(τ1k1τ2k1),\displaystyle\leq\frac{\alpha\delta\|A_{i_{0}}\|}{\tau_{1}-\tau_{2}}(\tau_{1}^{k-1}-\tau_{2}^{k-1}), (60)

since Δηi0(1)=0\|\Delta\eta_{i_{0}}(1)\|=0 and Δηi0(2)<αδAi0\|\Delta\eta_{i_{0}}(2)\|<\alpha\delta\|A_{i_{0}}\| which can be derived from (54) and (55). These two equations also give that α|Δζi0(k)|=|Δηi0(k+1)|,k\alpha|\Delta\zeta_{i_{0}}(k)|=|\Delta\eta_{i_{0}}(k+1)|,\forall k. Thus,

fζη(𝜻(1),𝜼(1))fζη((𝜻(1),𝜼(1)))\displaystyle\frac{f_{\zeta\eta}(\boldsymbol{\zeta}^{(1)},\boldsymbol{\eta}^{(1)})}{f_{\zeta\eta}\left(\mathcal{B}(\boldsymbol{\zeta}^{(1)},\boldsymbol{\eta}^{(1)})\right)} =k=1fL(ζi0(1)(k),θikζ)fL(ηi0(1)(k),θikη)fL(ζi0(2)(k),θikζ)fL(ηi0(2)(k),θikη)\displaystyle=\prod_{k=1}^{\infty}\frac{f_{L}(\zeta_{i_{0}}^{(1)}(k),\theta_{ik}^{\zeta})f_{L}(\eta_{i_{0}}^{(1)}(k),\theta_{ik}^{\eta})}{f_{L}(\zeta_{i_{0}}^{(2)}(k),\theta_{ik}^{\zeta})f_{L}(\eta_{i_{0}}^{(2)}(k),\theta_{ik}^{\eta})}
k=1e|Δζi0(k)|dζi0qi0kk=2e|Δηi0(k)|dηi0qi0k.\displaystyle\leq\prod_{k=1}^{\infty}e^{\frac{|\Delta\zeta_{i_{0}}(k)|}{d_{\zeta i_{0}}q_{i_{0}}^{k}}}\prod_{k=2}^{\infty}e^{\frac{|\Delta\eta_{i_{0}}(k)|}{d_{\eta i_{0}}q_{i_{0}}^{k}}}.
e(1αdζi0+1dηi0)αφi0δAi0φi0qi02αqi0α=eϵi0,\displaystyle\leq e^{\left(\frac{1}{\alpha d_{\zeta i_{0}}}+\frac{1}{d_{\eta i_{0}}}\right)\frac{\alpha\varphi_{i_{0}}\delta\|A_{i_{0}}\|}{\varphi_{i_{0}}q_{i_{0}}^{2}-\alpha q_{i_{0}}-\alpha}}=e^{\epsilon_{i_{0}}}, (61)

where the last inequality follows the the sum formula of geometric series. And the inequality φi0qi02αAi02qi0αAi02>0\varphi_{i_{0}}q_{i_{0}}^{2}-\alpha\|A_{i_{0}}\|^{2}q_{i_{0}}-\alpha\|A_{i_{0}}\|^{2}>0 holds under (52).

Combining (56) and (61) yields the result corresponding to the definition of differential privacy of the cost function in Definition 2. Thus, the proof completes. \hfill\blacksquare

Note that the privacy level ϵ\epsilon in Theorem 3 is related to both noise ζ\zeta and η\eta. Both of them are necessary, since if at least one of them are set to zero, there does not exist a finite number ϵ>0\epsilon>0, i.e., the differential privacy cannot be preserved.

Since the convergence accuracy is only related to the noise 𝜻\boldsymbol{\zeta}, the optimal value of ϵi\epsilon_{i} can be obtained by setting dηid_{\eta_{i}}\rightarrow\infty and it gives

ϵi=φiδAidζi(φiqi2αqiα),i{1,,n}.\displaystyle\epsilon_{i}^{\star}=\frac{\varphi_{i}\delta\|A_{i}\|}{d_{\zeta i}(\varphi_{i}q_{i}^{2}-\alpha q_{i}-\alpha)},\forall i\in\{1,{\ldots},n\}. (62)

Combining (43) with (62), we can derive that increasing added noise leads to a high level privacy but a low accuracy which shows the trade-off between accuracy and privacy.

VI Numerical Experiments

In this section, we conduct numerical experiments to verify our theoretical analysis. The proposed algorithm is tested on a multi-MG system with 14 microgrids [16]. Regarding the communication network, we generate an undirected graph by adding random links to a ring network. We consider the following problem with quadratic cost functions fi(xi)=uixi2+vixi+wi,if_{i}(x_{i})=u_{i}x_{i}^{2}+v_{i}x_{i}+w_{i},\forall i

min𝒙n×1f(𝒙)\displaystyle\min_{\boldsymbol{x}\in\mathbb{R}^{n\times 1}}f(\boldsymbol{x}) =i=1nfi(xi)\displaystyle=\sum_{i=1}^{n}f_{i}(x_{i})
s.t.i=1naixi=i=1ndi,\displaystyle s.t.\sum_{i=1}^{n}a_{i}x_{i}=\sum_{i=1}^{n}d_{i}, ximinxiximax,i.\displaystyle\ x_{i}^{\min}\leq x_{i}\leq x_{i}^{\max},\forall i. (63)

where coefficients aia_{i} are randomly chosen and the parameters of the generators are adopted from [1].

Set total load demand dtotal=231MWd_{total}=231MW, the exchanged information is masked by independent Laplace noise with q=0.98,dη=dζ=1q=0.98,d_{\eta}=d_{\zeta}=1 and we apply diff-DMAC to problem (63). Fig. 2 (a) shows how the mean square error (MSE) 𝔼[𝐱(k)𝐱2]\mathbb{E}[\|\mathbf{x}(k)-\mathbf{x}^{\star}\|^{2}] changes with iteration time kk, where the expected errors are approximated by averaging over 100 simulation results. The results validate the linear convergence rate of the proposed algorithm. Fix q=0.98q=0.98, the relation between error and dζd_{\zeta} is shown in Fig. 2 (b). We find that as dζd_{\zeta} becomes larger, the MSE of 𝐱\mathbf{x} increases, which means larger noise brings lower accuracy. Moreover, the experimental result is strictly between the lower bound and upper bound given in Theorem 2.

Refer to caption
Refer to caption
Figure 2: (a) MSE versus Iteration time kk with different stepsizes; (b) Relation between MSE and dζ.d_{\zeta}.

As for the differential privacy of diff-DMAC, Theorem 3 is numerically tested with different dζd_{\zeta} as shown in Fig. 3(a). The experimental result is upper bounded by the theoretical result, i.e. ϵeϵ\epsilon_{e}\leq\epsilon, which verifies Theorem 3. Combining Fig. 3(a) with Fig. 3(b) shows that a bigger and slower decaying noise leads to a better privacy level. However, large noise also influences the convergence accuracy and Fig. 4 validates the aforementioned trade-off between privacy and accuracy.

Refer to caption
Refer to caption
Figure 3: (a) Privacy level versus dζd_{\zeta}; (b) ϵe\epsilon_{e} versus decay coefficient qq.
Refer to caption
Figure 4: Trade-off between convergence accuracy and privacy level.

VII Conclusion

In this paper, a differentially private distributed mismatch tracking algorithm has been proposed to solve constraint-coupled resource allocation problems with privacy concerns. Its linear convergence property has been established for strongly convex and Lipschitz-smooth cost functions. Then, the differential privacy of the proposed algorithm is theoretically proved and we also characterize the trade-off between accuracy and privacy. The theoretical results have been examined by numerical experiments.

References

  • [1] S. Kar and G. Hug, “Distributed robust economic dispatch in power systems: A consensus + innovations approach,” in Proc. IEEE Power and Energy Society General Meeting, San Diego, CA, USA, July 2012, pp. 1–8.
  • [2] L. Xiao, M. Johansson, and S. Boyd, “Simultaneous routing and resource allocation via dual decomposition,” IEEE Transactions on Communications, vol. 52, no. 7, pp. 1136–1144, 2004.
  • [3] L. Jin and S. Li, “Distributed task allocation of multiple robots: A control perspective,” IEEE Transactions on Systems, Man, and Cybernetics: Systems, vol. 48, no. 5, pp. 693–701, 2018.
  • [4] A. Falsone, I. Notarnicola, G. Notarstefano, and M. Prandini, “Tracking-ADMM for distributed constraint-coupled optimization,” Automatica, vol. 117, p. 108962, 2020.
  • [5] J. Xu, S. Zhu, Y. C. Soh, and L. Xie, “A dual splitting approach for distributed resource allocation with regularization,” IEEE Transactions on Control of Network Systems, vol. 6, no. 1, pp. 403–414, 2019.
  • [6] J. Zhang, K. You, and K. Cai, “Distributed dual gradient tracking for resource allocation in unbalanced networks,” IEEE Transactions on Signal Processing, vol. 68, pp. 2186–2198, 2020.
  • [7] N. S. Aybat and E. Y. Hamedani, “A distributed ADMM-like method for resource sharing over time-varying networks,” SIAM Journal on Optimization, vol. 29, no. 4, pp. 3036–3068, 2019.
  • [8] S. Han, U. Topcu, and G. J. Pappas, “Differentially private distributed constrained optimization,” IEEE Transactions on Automatic Control, vol. 62, no. 1, pp. 50–64, 2017.
  • [9] M. Ruan, H. Gao, and Y. Wang, “Secure and privacy-preserving consensus,” IEEE Transactions on Automatic Control, vol. 64, no. 10, pp. 4035–4049, 2019.
  • [10] X. Huo and M. Liu, “Privacy-preserving distributed multi-agent cooperative optimization—paradigm design and privacy analysis,” IEEE Control Systems Letters, vol. 6, pp. 824–829, 2022.
  • [11] Y. Lou, L. Yu, S. Wang, and P. Yi, “Privacy preservation in distributed subgradient optimization algorithms,” IEEE Transactions on Cybernetics, vol. 48, no. 7, pp. 2154–2165, 2018.
  • [12] J. Zhu, C. Xu, J. Guan, and D. O. Wu, “Differentially private distributed online algorithms over time-varying directed networks,” IEEE Transactions on Signal and Information Processing over Networks, vol. 4, no. 1, pp. 4–17, 2018.
  • [13] T. Ding, S. Zhu, J. He, C. Chen, and X. Guan, “Differentially private distributed optimization via state and direction perturbation in multiagent systems,” IEEE Transactions on Automatic Control, vol. 67, no. 2, pp. 722–737, 2022.
  • [14] S. Mao, Y. Tang, Z. Dong, K. Meng, Z. Y. Dong, and F. Qian, “A privacy preserving distributed optimization algorithm for economic dispatch over time-varying directed networks,” IEEE Transactions on Industrial Informatics, vol. 17, no. 3, pp. 1689–1701, 2021.
  • [15] T. Ding, S. Zhu, C. Chen, J. Xu, and X. Guan, “Differentially private distributed resource allocation via deviation tracking,” IEEE Transactions on Signal and Information Processing over Networks, vol. 7, pp. 222–235, 2021.
  • [16] K. Wu, Q. Li, Z. Chen, J. Lin, Y. Yi, and M. Chen, “Distributed optimization method with weighted gradients for economic dispatch problem of multi-microgrid systems,” Energy, vol. 222, p. 119898, 2021.
  • [17] D. O. Amoateng, M. Al Hosani, M. S. Elmoursi, K. Turitsyn, and J. L. Kirtley, “Adaptive voltage and frequency control of islanded multi-microgrids,” IEEE Transactions on Power Systems, vol. 33, no. 4, pp. 4454–4465, 2018.
  • [18] R. Olfati-Saber and R. Murray, “Consensus problems in networks of agents with switching topology and time-delays,” IEEE Transactions on Automatic Control, vol. 49, no. 9, pp. 1520–1533, 2004.
  • [19] J. Xu, S. Zhu, Y. C. Soh, and L. Xie, “Augmented distributed gradient methods for multi-agent optimization under uncoordinated constant stepsizes,” in Proc. 54th IEEE Conference on Decision and Control, Osaka, Japan, Dec. 2015, pp. 2055–2060.
  • [20] J. He and L. Cai, “Differential private noise adding mechanism: Basic conditions and its application,” in Proc. American Control Conference, Seattle, WA, USA, May 2017, pp. 1673–1678.
  • [21] J.-B. Hiriart-Urruty and C. Lemaréchal, Convex Analysis and Minimization Algorithms II: Advanced Theory and Bundle Methods.   Berlin, Heidelberg: Springer Berlin Heidelberg, 1993, vol. 306.
  • [22] D. Bertsekas, Nonlinear Programming.   Belmont, Massachusetts:Athena Scientific, 2016.