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

Distributed Computation of Stochastic GNE with Partial Information: An Augmented Best-Response Approach

Yuanhanqing Huang and Jianghai Hu This work was supported by the National Science Foundation under Grant No. 2014816. A conference version has been submitted to ACC 2022 [1].The authors are with the Elmore Family School of Electrical and Computer Engineering, Purdue University, West Lafayette, IN, 47907, USA (e-mail: huan1282@purdue.edu; jianghai@purdue.edu).
Abstract

In this paper, we focus on the stochastic generalized Nash equilibrium problem (SGNEP) which is an important and widely-used model in many different fields. In this model, subject to certain global resource constraints, a set of self-interested players aim to optimize their local objectives that depend on their own decisions and the decisions of others and are influenced by some random factors. We propose a distributed stochastic generalized Nash equilibrium seeking algorithm in a partial-decision information setting based on the Douglas-Rachford operator splitting scheme, which relaxes assumptions in the existing literature. The proposed algorithm updates players’ local decisions through augmented best-response schemes and subsequent projections onto the local feasible sets, which occupy most of the computational workload. The projected stochastic subgradient method is applied to provide approximate solutions to the augmented best-response subproblems for each player. The Robbins-Siegmund theorem is leveraged to establish the main convergence results to a true Nash equilibrium using the proposed inexact solver. Finally, we illustrate the validity of the proposed algorithm via two numerical examples, i.e., a stochastic Nash-Cournot distribution game and a multi-product assembly problem with the two-stage model.

Index Terms:
Generalized Nash equilibrium, stochastic optimization, variational inequality, distributed algorithm, operator splitting

I Introduction

In the Nash equilibrium problem (NEP), a set of self-interested players aim to optimize their individual payoffs which depend not only on their own decisions but also on the decisions of others [2]. The generalized Nash equilibrium problem (GNEP) extends the NEP by considering additional global resource constraints that these players should collectively satisfy [3, 4]. In numerous practical applications, such as strategic behaviors in power markets [5, 6], engagement of multiple humanitarian organizations in disaster relief [7], and the traffic assignment of strategic risk-averse users [8], in addition to couplings in objectives and global constraints, there is also uncertainty involved in the objective functions. For example, in the target-rate power management problem for wireless networks, the power of battery-driven devices should be regulated in a real-time manner in the presence of inherent stochastic fluctuations of the underlying network [9]. These applications motivate us to consider an extension to the GNEP, called the stochastic generalized Nash equilibrium problem (SGNEP). In the SGNEP, instead of deterministic objective functions, players optimize the expected values of uncertain objective functions which are dependent on some random variables. Although the SGNEP can capture a wider range of applications, computing its Nash equilibria becomes a much more challenging problem compared to the GNEP, due to the absence of closed-form expressions of the objective functions. Fortunately, as has been shown in [10, Sec. 1.4], many SGNEPs can be formulated as stochastic variational inequalities (SVIs) and solved by leveraging existing results from this field, e.g. [11, 12, 13, 14, 15].

Our aim in this paper is to develop a distributed algorithm under the partial-decision information setting for solving SGNEPs over a network of players. In the context of non-cooperative games on networks, participants are self-interested and make rational decisions that maximize their own payoffs. It is unrealistic that these participants will adopt any centralized methods that require cooperation among them. Because of this, there is an enduring research interest in distributing the computation of Nash equilibria [16, 17], especially through the avenue of operator splitting technique [18, 19]. In addition to the distributed computation, under most circumstances, participants can only have access to local information and decisions of their neighbors, which constitutes a partial-decision information setting [20, 21, 22]. One reason is that these players are reluctant to send their local information and decisions to the general public out of privacy concerns while being willing to share these with their trusted neighbors on the whole network. Although there may exist some central node that collects and distributes the necessary information to each player, this central node is subject to performance limitations, such as single point of failure, and limited flexibility and scalability [23]. The limited capacity of communication channels also constraints information sharing over the network.

Significant efforts have been devoted to designing algorithms to solve SGNEPs distributedly under the full-decision information setting where each player has access to all other players’ decisions. The authors of [11] consider a Cartesian stochastic variational inequality problem with a monotone map. They propose a stochastic iterative Tikhonov regularization method and a stochastic iterative proximal-point method to solve it, which let players update their regularization parameters and centering term properly after each iteration. The authors of [24] propose a solution based on the preconditioned forward-backward (FB) operator splitting with the expected-value pseudogradient assumed to be restricted co-coercive and approximated via the stochastic approximation (SA) scheme. To accelerate game dynamics and relax the co-coercivity assumption, [25] adopts a forward-backward-forward framework. These works are mostly gradient-based which enjoy low complexity in implementation and updating per player step. Nevertheless, rational players would take best-response actions given others’ strategies and deviate from gradient-based schemes unless forced by some external authority. The work in [26] provides an inexact generalization of the proximal best-response (BR) schemes to the SNEP whose corresponding proximal BR map admits a contractive property. The authors of [27] further consider SNEPs with composite objectives and design a variable sample-size proximal BR scheme, under a contractive property on the proximal BR map. Yet, far less has been studied when it comes to the distributed solution to SGNEPs with merely partial information. The only existing work to our best knowledge is [28], which also relies on the FB framework along with the SA method. The convergence of the proposed algorithms has been analyzed under the assumption that the preconditioned forward operator is restricted co-coercive, which only allows comparatively small step sizes.

Our contributions can be summarized in the following aspects. First, we propose a distributed algorithm to solve the SGNEP under the partial-decision information setting based on the Douglas-Rachford splitting and the proximal mapping. In the proposed algorithm, the involved players are asked to update their decision vectors in two separate steps: solving the augmented best-response subproblems, and projecting onto the local feasible sets after some linear transformations. The updates of their local estimates and dual variables only require some trivial linear transformations. This algorithm can deal with cases where the scenario-based objectives of players are nonsmooth, and relaxes some commonly-made assumptions such as the α\alpha-cocoercivity with α>1/2\alpha>1/2 in the FB splitting. Second, we establish the convergence of the proposed algorithm under assumptions concerning the properties of the pseudogradient, the extended pseudogradient, and the stochastic subgradients of the objectives. Without relying on the contractive property, the proof in this paper is based on the Robbins-Siegmund theorem and extends the convergence results discussed in [26]. Drawing tools and techniques from stochastic approximation and convex analysis, we then construct a feasible inexact solver for the augmented best-response subproblems based on the projected stochastic subgradient method and discuss the prescribed accuracy within which the inexact solver should achieve such that the algorithm convergence is ensured. The proposed inexact solver based on the projected stochastic subgradient method requires the projection onto some bounded box sets rather than the (potentially complicated) local feasible sets of the original problem, which considerably improves the computational efficiency.

The remainder of this paper is organized as follows. In Sec. II, we formally formulate the SGNEP on networks and provide some basic definitions as well as assumptions. In this section, we recast the SGNEP as the zero-finding problem of a certain operator and prove that the solution of the latter is a "variational" solution of the former. In Sec. III, a distributed algorithm in a partial-decision information setting is proposed. Sec. IV focuses on the convergence analysis of the proposed algorithm. In this section, we also construct an inexact solver based on the projected stochastic subgradient method. In Sec. V, to demonstrate the theoretical findings and the proposed algorithm in practical applications, we include two numerical examples: a stochastic Nash-Cournot distribution game and a multi-product assembly problem with the two-stage model. Sec. VI concludes the paper and highlights potential extensions and applications.

Basic Notations: For a set of matrices {Vi}iS\{V_{i}\}_{i\in S}, we let blkd(V1,,V|S|)\text{blkd}(V_{1},\ldots,V_{|S|}) or blkd(Vi)iS\text{blkd}(V_{i})_{i\in S} denote the diagonal concatenation of these matrices, [V1,,V|S|][V_{1},\ldots,V_{|S|}] their horizontal stack, and [V1;;V|S|][V_{1};\cdots;V_{|S|}] their vertical stack. For a set of vectors {vi}iS\{v_{i}\}_{i\in S}, [vi]iS[v_{i}]_{i\in S} or [v1;;v|S|][v_{1};\cdots;v_{|S|}] denotes their vertical stack. For a matrix VV and a pair of positive integers (i,j)(i,j), [V](i,j)[V]_{(i,j)} denotes the entry on the ii-th row and the jj-th column of VV. For a vector vv and a positive integer ii, [v]i[v]_{i} denotes the ii-th entry of vv. Denote ¯{+}\overline{\mathbb{R}}\coloneqq\mathbb{R}\cup\{+\infty\}, +[0,+)\mathbb{R}_{+}\coloneqq[0,+\infty), and ++(0,+)\mathbb{R}_{++}\coloneqq(0,+\infty). 𝕊+n\mathbb{S}^{n}_{+} (resp. S++nS^{n}_{++}) represents the set of all n×nn\times n symmetric positive semi-definite (resp. definite) matrices. ι𝒮(x)\iota_{\mathcal{S}}(x) is defined to be the indicator function of a set 𝒮\mathcal{S}, i.e., if x𝒮x\in\mathcal{S}, then ι𝒮(x)=0\iota_{\mathcal{S}}(x)=0; otherwise, ι𝒮(x)=+\iota_{\mathcal{S}}(x)=+\infty. NS(x)N_{S}(x) denotes the normal cone to the set SnS\subseteq\mathbb{R}^{n} at the point xx: if xSx\in S, then NS(x){unsupzSu,zx0}N_{S}(x)\coloneqq\{u\in\mathbb{R}^{n}\mid\sup_{z\in S}\langle u,z-x\rangle\leq 0\}; otherwise, NS(x)N_{S}(x)\coloneqq\varnothing. If SnS\in\mathbb{R}^{n} is a closed and convex set, the map PjS:nS\operatorname{Pj}_{S}:\mathbb{R}^{n}\to S denotes the projection onto SS, i.e., PjS(x)argminvSvx2\operatorname{Pj}_{S}(x)\coloneqq\operatorname{argmin}_{v\in S}\lVert v-x\rVert_{2}. We use \rightrightarrows to indicate a point-to-set map. For an operator T:nnT:\mathbb{R}^{n}\rightrightarrows\mathbb{R}^{n}, Zer(T){xnTx𝟎}\text{Zer}(T)\coloneqq\{x\in\mathbb{R}^{n}\mid Tx\ni\boldsymbol{0}\} and Fix(T){xnTxx}\text{Fix}(T)\coloneqq\{x\in\mathbb{R}^{n}\mid Tx\ni x\} denote its zero set and fixed point set, respectively. We denote dom(T)\operatorname{dom}(T) the domain of the operator TT and gra(T)\text{gra}(T) the graph of it. The resolvent and reflected resolvent of TT are defined as JT(I+T)1J_{T}\coloneqq(I+T)^{-1} and RT2JTIR_{T}\coloneqq 2J_{T}-I, respectively.

II Problem Formulation

II-A Stochastic Game Formulation and SGNE

In this section, we formulate the stochastic generalized Nash equilibrium problem (SGNEP) on networks. There are NN players participating in the game, indexed by 𝒩{1,,N}\mathcal{N}\coloneqq\{1,\ldots,N\}. Each player i𝒩i\in\mathcal{N} needs to determine its local decision vector xi𝒳ix_{i}\in\mathcal{X}_{i} to optimize its objective, where 𝒳ini\mathcal{X}_{i}\subseteq\mathbb{R}^{n_{i}} is the local feasible set/action space of player ii. This Nash equilibrium seeking problem is generalized because, besides the local constraints {𝒳i}i𝒩\{\mathcal{X}_{i}\}_{i\in\mathcal{N}}, the decision vectors of all the players should satisfy some global resource constraints, i.e., i𝒩Aixic\sum_{i\in\mathcal{N}}A_{i}x_{i}\leq c. Here, we have the matrix Aim×niA_{i}\in\mathbb{R}^{m\times n_{i}} with mm denoting the number of the (global) affine coupling constraints, and the constant vector cmc\in\mathbb{R}^{m} representing the quantities of available resources. Altogether, for each player ii, the feasible set of the decision vector xix_{i} is given by

𝒳~i(xi)𝒳i{xiniAixi+j𝒩iAjxjc},\tilde{\mathcal{X}}_{i}(x_{-i})\coloneqq\mathcal{X}_{i}\cap\{x_{i}\in\mathbb{R}^{n_{i}}\mid A_{i}x_{i}+{\textstyle\sum}_{j\in\mathcal{N}_{-i}}A_{j}x_{j}\leq c\}, (1)

where 𝒩i𝒩\{i}\mathcal{N}_{-i}\coloneqq\mathcal{N}\backslash\{i\}, and xix_{-i} denotes the stack of decision vectors except that of player ii. Accordingly, the feasible set of the collective decision vector x[x1;;xN]x\coloneqq[x_{1};\cdots;x_{N}] is given by

𝒳~𝒳{xnAxc𝟎},\tilde{\mathcal{X}}\coloneqq\mathcal{X}\cap\{x\in\mathbb{R}^{n}\mid Ax-c\leq\boldsymbol{0}\}, (2)

where 𝒳i𝒩𝒳i\mathcal{X}\coloneqq\prod_{i\in\mathcal{N}}\mathcal{X}_{i}, ni𝒩nin\coloneqq\sum_{i\in\mathcal{N}}n_{i}, and A[A1,A2,,AN]A\coloneqq[A_{1},A_{2},\ldots,A_{N}].

To capture uncertainty in practical settings, we consider stochastic games where the objective function 𝕁i(xi;xi)\mathbb{J}_{i}(x_{i};x_{-i}) of each player ii is the expected value of certain function JiJ_{i}. Formally, given the decisions xix_{-i} of the other players, each player ii aims to solve the following local problem:

{minimizexi𝒳i𝕁i(xi;xi)=𝔼ξi[Ji(xi;xi,ξi)]subjecttoAixicj𝒩iAjxj,\begin{cases}\operatorname{minimize}_{x_{i}\in\mathcal{X}_{i}}\mathbb{J}_{i}(x_{i};x_{-i})=\mathbb{E}_{\xi_{i}}[J_{i}(x_{i};x_{-i},\xi_{i})]\\ \operatorname{subject\>to}\qquad A_{i}x_{i}\leq c-\sum_{j\in\mathcal{N}_{-i}}A_{j}x_{j}\end{cases}, (3)

where ξi:Ωinξi\xi_{i}:\Omega_{i}\to\mathbb{R}^{n_{\xi_{i}}} is a random variable in a well-defined probability space.

Given the above formulation of the SGNEP, we have the following standing assumptions that hold throughout the paper.

Assumption 1.

(Scenario-Based Objectives) For each player i𝒩i\in\mathcal{N}, given any fixed sample ωiΩi\omega_{i}\in\Omega_{i}, the scenario-based objective Ji(;,ξi(ωi))J_{i}(\cdot;\cdot,\xi_{i}(\omega_{i})) is proper and continuous. In addition, Ji(xi;xi,ξi(ωi))J_{i}(x_{i};x_{-i},\xi_{i}(\omega_{i})) is a convex function w.r.t. xix_{i} given any fixed xix_{-i} and ωiΩi\omega_{i}\in\Omega_{i}.

Assumption 2.

(Feasible Sets) Each local feasible set 𝒳i\mathcal{X}_{i} is nonempty, compact, and convex. The collective feasible set 𝒳~\tilde{\mathcal{X}} is nonempty, and the Mangasarian-Fromovitz constraint qualification (MFCQ) holds [10, Ch 3.2][29, Sec. 16.2.3].

Collectively solving the problems in (3) gives rise to the stochastic generalized Nash equilibrium (SGNE), the formal definition of which is given as follows [24, 30]:

Definition 1.

The collective decision x𝒳~x^{*}\in\tilde{\mathcal{X}} is a stochastic generalized Nash equilibrium (SGNE) if no player can benefit by unilaterally deviating from xx^{*}. Specifically, for all i𝒩i\in\mathcal{N}, 𝕁i(xi;xi)𝕁i(xi;xi)\mathbb{J}_{i}(x^{*}_{i};x^{*}_{-i})\leq\mathbb{J}_{i}(x_{i};x^{*}_{-i}) for any xiX~i(xi)x_{i}\in\tilde{X}_{i}(x^{*}_{-i}).

Under Assumption 1, the SGNE seeking problems can be transformed to the corresponding generalized quasi-variational inequality (GQVI) [29, Sec. 12.2]. As shown in [29, Prop. 12.3], Definition 1 of SGNE coincides with the following definition from the perspective of GQVI:

Definition 2.

The collective decision x𝒳~x^{*}\in\tilde{\mathcal{X}} is a stochastic generalized Nash equilibrium (SGNE) if xx^{*} along with a suitable gi𝒩xi𝕁i(xi;xi)g^{*}\in\prod_{i\in\mathcal{N}}\partial_{x_{i}}\mathbb{J}_{i}(x^{*}_{i};x^{*}_{-i}) is a solution of the problem GQVI(i𝒩𝒳~i,i𝒩xi𝕁i)\text{GQVI}(\prod_{i\in\mathcal{N}}\tilde{\mathcal{X}}_{i},\prod_{i\in\mathcal{N}}\partial_{x_{i}}\mathbb{J}_{i}), i.e.,

(xx)Tg0,xi𝒩𝒳~i(xi).\begin{split}(x-x^{*})^{T}g^{*}\geq 0,\forall x\in{\textstyle\prod}_{i\in\mathcal{N}}\tilde{\mathcal{X}}_{i}(x^{*}_{-i}).\end{split} (4)

As suggested in [29, Sec. 12.2.3], under Assumptions 1 and 2, we can equivalently recast the problem in (3) into a set of inclusions by considering the Karush-Kuhn-Tucker (KKT) conditions of the above GQVI such that i𝒩\forall i\in\mathcal{N}:

𝟎xi𝕁i(xi;xi)+AiTλi+N𝒳i(xi)𝟎(Axc)+N+m(λi),\begin{split}&\boldsymbol{0}\in\partial_{x_{i}}\mathbb{J}_{i}(x^{*}_{i};x^{*}_{-i})+A_{i}^{T}\lambda_{i}+N_{\mathcal{X}_{i}}(x^{*}_{i})\\ &\boldsymbol{0}\in-(Ax^{*}-c)+N_{\mathbb{R}^{m}_{+}}(\lambda_{i}),\end{split} (5)

where λi\lambda_{i} is the Lagrangian multiplier for the global resource constraints Aixicj𝒩iAjxjA_{i}x_{i}\leq c-\sum_{j\in\mathcal{N}_{-i}}A_{j}x_{j} for each player ii.

In this paper, we restrict our attention to a subset of these SGNEs where the players share the same coupled constraints, and hence all the Lagrangian multipliers are in consensus, i.e., λ1==λN\lambda_{1}=\ldots=\lambda_{N}. This gives rise to a generalized variational inequality (GVI) problem. This subclass of the SGNEs, known as the variational stochastic generalized Nash equilibria (v-SGNEs) [4, 3], enforces the idea of economic fairness and enjoys better social stability/sensitivity [31]. We will focus on this subclass since we can leverage a variety of tools that have been developed for solving (G)VIs [10, Ch. 10-12] and design the modified best-response dynamics based on it.

Definition 3.

The collective decision x𝒳~x^{*}\in\tilde{\mathcal{X}} is a variational stochastic generalized Nash equilibrium (v-SGNE) if xx^{*} along with a suitable gi𝒩xi𝕁i(xi;xi)g^{*}\in\prod_{i\in\mathcal{N}}\partial_{x_{i}}\mathbb{J}_{i}(x^{*}_{i};x^{*}_{-i}) is a solution of the GVI(𝒳~,i𝒩xi𝕁i)\text{GVI}(\tilde{\mathcal{X}},\prod_{i\in\mathcal{N}}\partial_{x_{i}}\mathbb{J}_{i}), i.e.,

(xx)Tg0,x𝒳~.\begin{split}(x-x^{*})^{T}g^{*}\geq 0,\forall x\in\tilde{\mathcal{X}}.\end{split} (6)

Similarly, the KKT system of the above GVI is given by:

𝟎xi𝕁i(xi;xi)+AiTλ+N𝒳i(xi)𝟎(Axc)+N+m(λ),\begin{split}&\boldsymbol{0}\in\partial_{x_{i}}\mathbb{J}_{i}(x^{*}_{i};x^{*}_{-i})+A_{i}^{T}\lambda+N_{\mathcal{X}_{i}}(x^{*}_{i})\\ &\boldsymbol{0}\in-(Ax^{*}-c)+N_{\mathbb{R}^{m}_{+}}(\lambda),\end{split} (7)

where λ\lambda is the Lagrangian multiplier for the global constraints in (3). Notice that the GVI in (6) is not completely equivalent to the initial SGNEP in (3) as the game may admit an SGNE while the GVI has no solution. We make the following assumption concerning the existence of v-SGNEs.

Assumption 3.

(Existence of v-SGNE) The SGNEP considered admits a nonempty set of v-SGNEs.

Remark 1.

The existence and multiplicity of solutions of GNEPs with continuously differentiable objectives have been extensively studied, and the related theories can be found in [10, Ch. 2&3]. For the GNEPs with nonsmooth objectives, we can check the existence of v-GNEs of these GNEPs by [29, Prop. 12.11]. If the closed-form expressions of the objectives 𝕁i(xi;xi)\mathbb{J}_{i}(x_{i};x_{-i}) for any i𝒩i\in\mathcal{N} are unavailable and we cannot apply the above results, [30, Sec. 4] provides sufficient conditions to guarantee the existence of v-SGNEs based on the properties of scenario-based objectives.

II-B Network Game Formulation

In network games, there exists an underlying communication graph 𝒢=(𝒩g,g)\mathcal{G}=(\mathcal{N}_{g},\mathcal{E}_{g}), where players can communicate with their neighbors through arbitrators on the edges. The node set 𝒩g\mathcal{N}_{g} denotes the set of all players, and g𝒩g×𝒩g\mathcal{E}_{g}\subseteq\mathcal{N}_{g}\times\mathcal{N}_{g} is the set of directed edges. The cardinalities |𝒩g||\mathcal{N}_{g}| and |g||\mathcal{E}_{g}| are denoted by NgN_{g} and EgE_{g}. In this case, 𝒩g=𝒩\mathcal{N}_{g}=\mathcal{N} and Ng=NN_{g}=N. We use (i,j)(i,j) to denote a directed edge having node/player ii as its tail and node/player jj as its head. For notational brevity, let 𝒩i\mathcal{N}_{i} denote the set of immediate neighbors of player ii who can directly communicate with it, 𝒩i+{j𝒩(j,i)g}\mathcal{N}^{+}_{i}\coloneqq\{j\in\mathcal{N}\mid(j,i)\in\mathcal{E}_{g}\} the set of in-neighbors of player ii, and 𝒩i{j𝒩(i,j)g}\mathcal{N}^{-}_{i}\coloneqq\{j\in\mathcal{N}\mid(i,j)\in\mathcal{E}_{g}\} the set of out-neighbors of player ii. Note that although the multipliers we are going to introduce are defined in a directed fashion, we assume each node can send messages to both its in- and out-neighbors, and 𝒢\mathcal{G} should satisfy the following assumption.

Assumption 4.

(Communicability) The underlying communication graph 𝒢=(𝒩g,g)\mathcal{G}=(\mathcal{N}_{g},\mathcal{E}_{g}) is undirected and connected. Besides, it has no self-loops.

We next recast the SGNEP in (3) as the zero-finding problem of a certain operator that can be carried out distributedly over the communication graph 𝒢\mathcal{G} via the network Lagrangian of this game and refer the interested reader to [32] for more details. Now for each player i𝒩i\in\mathcal{N}, besides its local decision vector yii𝒳iy^{i}_{i}\in\mathcal{X}_{i}, it keeps a local estimate yijnjy^{j}_{i}\in\mathbb{R}^{n_{j}} of the player jj’s decision for all j𝒩ij\in\mathcal{N}_{-i}, which together constitutes its augmented decision vector yiy_{i}. Here, we use yiiy^{i}_{i} to denote the local decision of each player ii to distinguish from the case where only local decision xix_{i} are maintained and considered. We denote yii[yij]j𝒩iy^{-i}_{i}\coloneqq[y^{j}_{i}]_{j\in\mathcal{N}_{-i}} the vertical stack of {yij}j𝒩i\{y^{j}_{i}\}_{j\in\mathcal{N}_{-i}} and yi[yij]j𝒩y_{i}\coloneqq[y^{j}_{i}]_{j\in\mathcal{N}} the vertical stack of {yij}j𝒩\{y^{j}_{i}\}_{j\in\mathcal{N}}, both in prespecified orders. Denote n<i=j𝒩,j<injn_{<i}=\sum_{j\in\mathcal{N},j<i}n_{j} and n>i=j𝒩,j>injn_{>i}=\sum_{j\in\mathcal{N},j>i}n_{j}. The extended feasible set of 𝒚[yi]i𝒩\boldsymbol{y}\coloneqq[y_{i}]_{i\in\mathcal{N}} is defined as 𝒳^𝒳^1×𝒳^2××𝒳^N\hat{\mathcal{X}}\coloneqq\hat{\mathcal{X}}_{1}\times\hat{\mathcal{X}}_{2}\times\cdots\times\hat{\mathcal{X}}_{N} with each one defined as 𝒳^in<i×𝒳i×n>i\hat{\mathcal{X}}_{i}\coloneqq\mathbb{R}^{n_{<i}}\times\mathcal{X}_{i}\times\mathbb{R}^{n_{>i}}. For brevity, we shall write {yi}\{y_{i}\} in replacement of the more cumbersome notation {yi}i𝒩\{y_{i}\}_{i\in\mathcal{N}} and similarly for other variables on nodes and edges (e.g. the dual variables {μji}(j,i)g\{\mu_{ji}\}_{(j,i)\in\mathcal{E}_{g}} to be introduced below will be denoted simply by {μji}\{\mu_{ji}\}), unless otherwise specified. In the reformulated zero-finding problem, we introduced a set of dual variables {λi}\{\lambda_{i}\} to enforce the global resource constraints. Moreover, another two sets of dual variables {μji}\{\mu_{ji}\} and {zji}\{z_{ji}\} are introduced to guarantee the consensus of {yi}\{y_{i}\} and {λi}\{\lambda_{i}\}. It is worth mentioning that {yi}\{y_{i}\} and {λi}\{\lambda_{i}\} are maintained by players while {μji}\{\mu_{ji}\} and {zji}\{z_{ji}\} are maintained by arbitrators on the edges.

We next give a brief introduction to two commonly used operators in the distributed solution of GNEP: the pseudogradient 𝔽:nn\operatorname{\mathbb{F}}:\mathbb{R}^{n}\rightrightarrows\mathbb{R}^{n} and the extended pseudogradient 𝔽~:nNn\operatorname{\tilde{\mathbb{F}}}:\mathbb{R}^{nN}\rightrightarrows\mathbb{R}^{n}. The pseudogradient 𝔽\operatorname{\mathbb{F}} is the vertical stack of the partial subgradients of the objective functions of all players, which is given as follows:

𝔽:x[xi𝕁i(xi;xi)]i𝒩.\operatorname{\mathbb{F}}:x\mapsto[\partial_{x_{i}}\mathbb{J}_{i}(x_{i};x_{-i})]_{i\in\mathcal{N}}. (8)

In contrast, the extended pseudogradient 𝔽~\operatorname{\tilde{\mathbb{F}}} defined in (9) is a commonly used operator under the partial-decision information setting, where each player keeps the local estimates of others’ decisions and then uses these estimates as the parametric inputs:

𝔽~:[yi]i𝒩[yii𝕁i(yii;yii)]i𝒩.\operatorname{\tilde{\mathbb{F}}}:[y_{i}]_{i\in\mathcal{N}}\mapsto[\partial_{y^{i}_{i}}\mathbb{J}_{i}(y^{i}_{i};y^{-i}_{i})]_{i\in\mathcal{N}}. (9)

To incorporate the extended pseudogradient 𝔽~\operatorname{\tilde{\mathbb{F}}} into a fixed-point iteration, we then introduce the individual selection matrices {i}i𝒩\{\mathcal{R}_{i}\}_{i\in\mathcal{N}} and their diagonal concatenation n×nN\mathcal{R}\in\mathbb{R}^{n\times nN}:

i=[𝟎ni×n<i,𝐈ni,𝟎ni×n>i],=blkd(1,,N).\begin{split}&\mathcal{R}_{i}=[\boldsymbol{0}_{n_{i}\times n_{<i}},\mathbf{I}_{n_{i}},\boldsymbol{0}_{n_{i}\times n_{>i}}],\;\mathcal{R}=\text{blkd}(\mathcal{R}_{1},\ldots,\mathcal{R}_{N}).\end{split} (10)

Notice that yii=iyiy^{i}_{i}=\mathcal{R}_{i}y_{i} and iiT=Ini\mathcal{R}_{i}\mathcal{R}_{i}^{T}=I_{n_{i}}. Finally, the set-valued operator 𝕋\operatorname{\mathbb{T}} we are going to study is given below:

𝕋:[𝒚𝝀𝝁𝒛][T(𝔽~(𝒚)+ΛT𝝀)+Bn𝝁+ρμLn𝒚+N𝒳^(𝒚)N+mN(𝝀)Λ𝒚+𝒄+Bm𝒛+ρzLm𝝀BnT𝒚BmT𝝀],\small\operatorname{\mathbb{T}}:\begin{bmatrix}\boldsymbol{y}\\ \boldsymbol{\lambda}\\ \boldsymbol{\mu}\\ \boldsymbol{z}\end{bmatrix}\mapsto\begin{bmatrix}\mathcal{R}^{T}(\operatorname{\tilde{\mathbb{F}}}(\boldsymbol{y})+\Lambda^{T}\boldsymbol{\lambda})+B_{n}\boldsymbol{\mu}+\rho_{\mu}L_{n}\boldsymbol{y}+N_{\hat{\mathcal{X}}}(\boldsymbol{y})\\ N_{\mathbb{R}^{mN}_{+}}(\boldsymbol{\lambda})-\Lambda\mathcal{R}\boldsymbol{y}+\boldsymbol{c}+B_{m}\boldsymbol{z}+\rho_{z}L_{m}\boldsymbol{\lambda}\\ -B_{n}^{T}\cdot\boldsymbol{y}\\ -B_{m}^{T}\cdot\boldsymbol{\lambda}\end{bmatrix},\normalsize (11)

where Λ\Lambda is the diagonal concatenation of {Ai}i𝒩\{A_{i}\}_{i\in\mathcal{N}}, i.e., Λblkd(A1,,AN)\Lambda\coloneqq\text{blkd}(A_{1},\ldots,A_{N}); 𝒄\boldsymbol{c} is the vertical stack of {ci}i𝒩\{c_{i}\}_{i\in\mathcal{N}} with i𝒩ci=c\sum_{i\in\mathcal{N}}c_{i}=c; Bn(BIn)B_{n}\coloneqq(B\otimes I_{n}), Ln(LIn)L_{n}\coloneqq(L\otimes I_{n}), Bm(BIm)B_{m}\coloneqq(B\otimes I_{m}), Lm(LIm)L_{m}\coloneqq(L\otimes I_{m}), BB and LL are the incidence matrix and Laplacian matrix of the underlying communication graph, respectively, with L=BBTL=B\cdot B^{T}; and 𝒚\boldsymbol{y}, 𝝀\boldsymbol{\lambda}, 𝝁\boldsymbol{\mu}, and 𝒛\boldsymbol{z} are the stack vectors of {yi}\{y_{i}\}, {λi}\{\lambda_{i}\}, {μji}\{\mu_{ji}\}, and {zji}\{z_{ji}\}, respectively; ψ\psi denotes the stack of the primal and dual variables, i.e., ψ[𝒚;𝝀;𝝁;𝒛]\psi\coloneqq[\boldsymbol{y};\boldsymbol{\lambda};\boldsymbol{\mu};\boldsymbol{z}].

Theorem 1.

Suppose Assumptions 1 to 4 hold, and there exists ψ[𝐲;𝛌;𝛍;𝐳]Zer(𝕋)\psi^{*}\coloneqq[\boldsymbol{y}^{*};\boldsymbol{\lambda}^{*};\boldsymbol{\mu}^{*};\boldsymbol{z}^{*}]\in\text{Zer}(\operatorname{\mathbb{T}}). Then 𝐲=𝟏Ny\boldsymbol{y}^{*}=\boldsymbol{1}_{N}\otimes y^{*}, 𝛌=𝟏Nλ\boldsymbol{\lambda}^{*}=\boldsymbol{1}_{N}\otimes\lambda^{*}, and (y,λ)(y^{*},\lambda^{*}) satisfies the KKT conditions (7) for v-GNE with xx^{*} replaced with yy^{*}. Conversely, for a solution (y,λ)(y^{\dagger},\lambda^{\dagger}) of the KKT problem in (7), there exist 𝛍\boldsymbol{\mu}^{\dagger} and 𝐳\boldsymbol{z}^{\dagger} such that ψ[𝟏Ny;𝟏Nλ;𝛍;𝐳]Zer(𝕋)\psi^{\dagger}\coloneqq[\boldsymbol{1}_{N}\otimes y^{\dagger};\boldsymbol{1}_{N}\otimes\lambda^{\dagger};\boldsymbol{\mu}^{\dagger};\boldsymbol{z}^{\dagger}]\in\text{Zer}(\operatorname{\mathbb{T}}).

Proof.

See the proof of [32, Thm. 1]. ∎

Thus, finding a v-SGNE of the game in (3) is equivalent to solving for a zero point of the operator 𝕋\mathbb{T}. To facilitate the convergence analysis of the algorithm to be proposed for the latter task, we make two parallel assumptions, either of which is instrumental for the convergence proof in Sect. IV.

Assumption 5.

(Convergence Condition) At least one of the following statements holds: {outline}[enumerate] \1 the operator T𝔽~+ρμ2Ln\mathcal{R}^{T}\operatorname{\tilde{\mathbb{F}}}+\frac{\rho_{\mu}}{2}L_{n} is maximally monotone; \1 the pseudogradient 𝔽\operatorname{\mathbb{F}} is strongly monotone and Lipschitz continuous, i.e., there exist η>0\eta>0 and θ1>0\theta_{1}>0, such that x,xn\forall x,x^{\prime}\in\mathbb{R}^{n}, xx,𝔽(x)𝔽(x)ηxx2\langle x-x^{\prime},\operatorname{\mathbb{F}}(x)-\operatorname{\mathbb{F}}(x^{\prime})\rangle\geq\eta\lVert x-x^{\prime}\rVert^{2} and 𝔽(x)𝔽(x)θ1xx\lVert\operatorname{\mathbb{F}}(x)-\operatorname{\mathbb{F}}(x^{\prime})\rVert\leq\theta_{1}\lVert x-x^{\prime}\rVert. The operator T𝔽~\mathcal{R}^{T}\operatorname{\tilde{\mathbb{F}}} is Lipschitz continuous, i.e., there exists θ2>0\theta_{2}>0, such that 𝐲,𝐲nN\forall\boldsymbol{y},\boldsymbol{y}^{\prime}\in\mathbb{R}^{nN}, 𝔽~(𝐲)𝔽~(𝐲)θ2𝐲𝐲\lVert\operatorname{\tilde{\mathbb{F}}}(\boldsymbol{y})-\operatorname{\tilde{\mathbb{F}}}(\boldsymbol{y}^{\prime})\rVert\leq\theta_{2}\lVert\boldsymbol{y}-\boldsymbol{y}^{\prime}\rVert. Moreover, ρμ2σ1((θ1+θ2)24η+θ2)\rho_{\mu}\geq\frac{2}{\sigma_{1}}(\frac{(\theta_{1}+\theta_{2})^{2}}{4\eta}+\theta_{2}), where σ1\sigma_{1} is the smallest positive eigenvalue of LL.

III An Augmented Best-Response Scheme

To compute the zeros of the operator 𝕋\operatorname{\mathbb{T}} given in the previous section, we leverage the Douglas-Rachford (DR) splitting method which combines operator splitting and the Krasnoselskii-Mann (K-M) schemes. Given a nonexpansive operator QQ with a nonempty fixed point set Fix(Q)\text{Fix}(Q), the K-M scheme [33, Sec. 5.2] suggests the following iteration:

ψ(k+1)ψ(k)+γ(k)(Qψ(k)ψ(k)),\psi^{(k+1)}\coloneqq\psi^{(k)}+\gamma^{(k)}(Q\psi^{(k)}-\psi^{(k)}), (12)

where (γ(k))k(\gamma^{(k)})_{k\in\mathbb{N}} is a sequence such that γ(k)[0,1]\gamma^{(k)}\in[0,1] for all kk\in\mathbb{N} and kγ(k)(1γ(k))=\sum_{k\in\mathbb{N}}\gamma^{(k)}(1-\gamma^{(k)})=\infty. Here, we introduce a set of local bounded box constraints {𝒳iB}\{\mathcal{X}^{B}_{i}\} which can be chosen manually as long as it satisfies 𝒳i𝒳iB\mathcal{X}_{i}\subseteq\mathcal{X}^{B}_{i} for all i𝒩i\in\mathcal{N}. We similarly define the extended box set 𝒳^B𝒳^1B×𝒳^2B×𝒳^NB\hat{\mathcal{X}}^{B}\coloneqq\hat{\mathcal{X}}^{B}_{1}\times\hat{\mathcal{X}}^{B}_{2}\times\cdots\hat{\mathcal{X}}^{B}_{N} where the extended box set of each player ii is defined as 𝒳^iBn<i×𝒳iB×n>i\hat{\mathcal{X}}^{B}_{i}\coloneqq\mathbb{R}^{n_{<i}}\times\mathcal{X}^{B}_{i}\times\mathbb{R}^{n_{>i}}. It is easy to see that the normal cones of 𝒳^B\hat{\mathcal{X}}^{B} and 𝒳^\hat{\mathcal{X}} satisfy N𝒳^B+N𝒳^=N𝒳^N_{\hat{\mathcal{X}}^{B}}+N_{\hat{\mathcal{X}}}=N_{\hat{\mathcal{X}}}. The motivation for introducing these box sets is to simplify the computation while maintaining boundedness for the convergence analysis as we will show later in this paper. We split the operator 𝕋\operatorname{\mathbb{T}} into the following operators 𝔸\mathbb{A} and \mathcal{B}:

𝔸:ψ(D+𝔸y)ψand:ψ(D+y)ψ\operatorname{\mathbb{A}}:\psi\mapsto(D+\mathbb{A}_{y})\psi\;\text{and}\;\operatorname{\mathcal{B}}:\psi\mapsto(D+\mathcal{B}_{y})\psi (13)

with DD, 𝒜y\mathcal{A}_{y}, and y\mathcal{B}_{y} defined by

D=[ρμ2Ln12(Λ)T12Bn012Λρz2Lm012Bm12BnT000012BmT00],D=\begin{bmatrix}\frac{\rho_{\mu}}{2}L_{n}&\frac{1}{2}(\Lambda\mathcal{R})^{T}&\frac{1}{2}B_{n}&0\\ -\frac{1}{2}\Lambda\mathcal{R}&\frac{\rho_{z}}{2}L_{m}&0&\frac{1}{2}B_{m}\\ -\frac{1}{2}B_{n}^{T}&0&0&0\\ 0&-\frac{1}{2}B_{m}^{T}&0&0\end{bmatrix}, (14)
𝔸y:ψ[T𝔽~(𝒚)+N𝒳^(𝒚)𝒄00],y:ψ[N𝒳^(𝒚)N+mN(𝝀)00].\small\mathbb{A}_{y}:\psi\mapsto\begin{bmatrix}\mathcal{R}^{T}\operatorname{\tilde{\mathbb{F}}}(\boldsymbol{y})+N_{\mathcal{\hat{X}^{B}}}(\boldsymbol{y})\\ \boldsymbol{c}\\ 0\\ 0\end{bmatrix},\;\mathcal{B}_{y}:\psi\mapsto\begin{bmatrix}N_{\mathcal{\hat{X}}}(\boldsymbol{y})\\ N_{\mathbb{R}^{mN}_{+}}(\boldsymbol{\lambda})\\ 0\\ 0\end{bmatrix}.\normalsize (15)

Furthermore, we introduce the following design matrix Φ\Phi for distributedly computing the resolvents JΦ1𝔸J_{\Phi^{-1}\operatorname{\mathbb{A}}} and JΦ1J_{\Phi^{-1}\operatorname{\mathcal{B}}}:

Φ=[𝝉11ρμ2Ln12(Λ)T12Bn012Λ𝝉21ρz2Lm012Bm12BnT0𝝉310012BmT0𝝉41],\small\Phi=\begin{bmatrix}\boldsymbol{\tau}_{1}^{-1}-\frac{\rho_{\mu}}{2}L_{n}&-\frac{1}{2}(\Lambda\mathcal{R})^{T}&-\frac{1}{2}B_{n}&0\\ -\frac{1}{2}\Lambda\mathcal{R}&\boldsymbol{\tau}_{2}^{-1}-\frac{\rho_{z}}{2}L_{m}&0&-\frac{1}{2}B_{m}\\ -\frac{1}{2}B_{n}^{T}&0&\boldsymbol{\tau}_{3}^{-1}&0\\ 0&-\frac{1}{2}B_{m}^{T}&0&\boldsymbol{\tau}_{4}^{-1}\end{bmatrix},\normalsize (16)

where 𝝉1blkd(τ11In,,τ1NIn)\boldsymbol{\tau}_{1}\coloneqq\text{blkd}(\tau_{11}I_{n},\ldots,\tau_{1N}I_{n}) with τ11++,,τ1N++\tau_{11}\in\mathbb{R}_{++},\ldots,\tau_{1N}\in\mathbb{R}_{++}; similarly for 𝝉2\boldsymbol{\tau}_{2}, 𝝉3\boldsymbol{\tau}_{3} and 𝝉4\boldsymbol{\tau}_{4}. Notice that these step sizes 𝝉1,,𝝉4\boldsymbol{\tau}_{1},\ldots,\boldsymbol{\tau}_{4} should be small enough to guarantee that Φ\Phi is positive definite. Conservative upper bounds for these step sizes can be derived using the Gershgorin circle theorem [34].

Assumption 6.

The step sizes 𝛕1,,𝛕4\boldsymbol{\tau}_{1},\ldots,\boldsymbol{\tau}_{4} are chosen properly such that the design matrix Φ\Phi in (16) is positive definite. Specifically, it suffices to choose τ1i1>12Ai1+(12+ρμ)di\tau_{1i}^{-1}>\frac{1}{2}\lVert A_{i}\rVert_{1}+(\frac{1}{2}+\rho_{\mu})d_{i}, τ2i1>12Ai+(12+ρz)di,i𝒩\tau_{2i}^{-1}>\frac{1}{2}\lVert A_{i}\rVert_{\infty}+(\frac{1}{2}+\rho_{z})d_{i},\forall i\in\mathcal{N}, and τ3j1>1\tau_{3j}^{-1}>1, τ4j1>1\tau_{4j}^{-1}>1, jg\forall j\in\mathcal{E}_{g}.

Here, did_{i} denotes the degree of node/player ii. In general, determining the above step sizes requires some global information acquired through coordination among players such as a proper ρμ\rho_{\mu}. After the incorporation of the design matrix Φ\Phi, we now work in the inner product space 𝒦\mathcal{K} which is a real vector space endowed with the inner product ψ1,ψ2𝒦=ψ1TΦψ2\langle\psi_{1},\psi_{2}\rangle_{\mathcal{K}}=\psi_{1}^{T}\Phi\psi_{2}. For brevity, let 𝔸¯Φ1𝔸\operatorname{\bar{\mathbb{A}}}\coloneqq\Phi^{-1}\operatorname{\mathbb{A}} and ¯Φ1\operatorname{\bar{\mathcal{B}}}\coloneqq\Phi^{-1}\operatorname{\mathcal{B}}. In the DR splitting scheme, the general operator QQ in (12) is given by R¯R𝔸¯\mathscr{R}_{*}\coloneqq R_{\operatorname{\bar{\mathcal{B}}}}\circ R_{\operatorname{\bar{\mathbb{A}}}} and it suggests the following exact iteration:

ψ~(k+1)𝒫(ψ~(k)), with 𝒫=Id+γ(k)(Id).\tilde{\psi}^{(k+1)}\coloneqq\mathscr{P}_{*}(\tilde{\psi}^{(k)}),\text{ with }\mathscr{P}_{*}=\text{Id}+\gamma^{(k)}(\mathscr{R}_{*}-\text{Id}). (17)

Given a generic single-valued operator QQ, we say that QQ is restricted nonexpansive w.r.t. a set SS if, for all ψdomQ\psi\in\operatorname{dom}{Q} and ψS\psi^{*}\in S, QψQψψψ\lVert Q\psi-Q\psi^{*}\rVert\leq\lVert\psi-\psi^{*}\rVert [20]; if, in addition, S=Fix(Q)S=\text{Fix}(Q), then QQ is quasinonexpansive [33, Def. 4.1(v)]. From the main convergence results in [32, Thm. 2&3], if Assumptions 1 to 6 hold, even though \mathscr{R}_{*} is not nonexpansive in a general sense, it possesses quasinonexpansiveness in the inner-product space 𝒦\mathcal{K}, and hence the sequence (yi(k))k(y_{i}^{(k)})_{k\in\mathbb{N}} generated by the exact iteration above (see [32, Algorithm 1] for detailed implementations) will converge to a v-SGNE of the original problem defined in (3).

However, unlike the problem setting in [32] where each player has a closed-form objective function, here the objective function is expected-value, and all too often its closed-form expression may be unavailable. Consequently, the argmin\operatorname{argmin} operation in the first player loop of [32, Algorithm 1] can not be carried out exactly. In this case, we need a desirable inexact solver such that, although at each iteration step, it can only get an approximate solution, the computed sequence can still eventually converge to a v-SGNE. We let R𝒜¯R_{\operatorname{\bar{\mathcal{A}}}} denote the (scenario-based) approximate operator to the exact reflected resolvent R𝔸¯R_{\operatorname{\bar{\mathbb{A}}}}, and \mathscr{R} denote the corresponding composite R¯R𝒜¯R_{\operatorname{\bar{\mathcal{B}}}}\circ R_{\operatorname{\bar{\mathcal{A}}}}. Substituting the operator \mathscr{R}_{*} with \mathscr{R} in [32, Algorithm 1] gives rise to the following approximate iteration:

ψ~(k+1)𝒫(ψ~(k)), with 𝒫=Id+γ(k)(Id).\tilde{\psi}^{(k+1)}\coloneqq\mathscr{P}(\tilde{\psi}^{(k)}),\text{ with }\mathscr{P}=\text{Id}+\gamma^{(k)}(\mathscr{R}-\text{Id}). (18)

The updating steps of (18) are presented in Algorithm 1. For brevity, let y~iLi(k)j𝒩i(y~ii(k)y~ji(k)){\tilde{y}}^{-i(k)}_{iL}\coloneqq\sum_{\scriptscriptstyle j\in\mathcal{N}_{i}}({\tilde{y}}^{-i(k)}_{i}-{\tilde{y}}^{-i(k)}_{j}), and similarly for y~iLi(k){\tilde{y}}^{i(k)}_{iL}, λ~iL(k){\tilde{\lambda}}^{(k)}_{iL}, y^iLi(k+1){\hat{y}}^{i(k+1)}_{iL}, and λ^iL(k+1){\hat{\lambda}}^{(k+1)}_{iL}; let μ~iBi(k)j𝒩i+μ~jii(k)j𝒩iμ~iji(k){\tilde{\mu}}^{-i(k)}_{iB}\coloneqq\sum_{\scriptscriptstyle j\in\mathcal{N}^{+}_{i}}{\tilde{\mu}}^{-i(k)}_{ji}-\sum_{\scriptscriptstyle j\in\mathcal{N}^{-}_{i}}{\tilde{\mu}}^{-i(k)}_{ij}, and similarly for μ~iBi(k){\tilde{\mu}}^{i(k)}_{iB}, z~iB(k){\tilde{z}}^{(k)}_{iB}, μ^iB(k+1){\hat{\mu}}^{(k+1)}_{iB}, and z^iB(k+1){\hat{z}}^{(k+1)}_{iB}; let y^ji(k+1)y^i(k+1)y^j(k+1){\hat{y}}^{(k+1)}_{ji}\coloneqq{\hat{y}}^{(k+1)}_{i}-{\hat{y}}^{(k+1)}_{j}, and similarly for λ^ji(k+1){\hat{\lambda}}^{(k+1)}_{ji}, y¯ji(k+1){\bar{y}}^{(k+1)}_{ji}, and λ¯ji(k+1){\bar{\lambda}}^{(k+1)}_{ji}.

1 Initialize: {y~i(0)},{λ~i(0)},{μ~ji(0)},{z~ji(0)}\{{\tilde{y}}^{(0)}_{i}\},\{{\tilde{\lambda}}^{(0)}_{i}\},\{{\tilde{\mu}}^{(0)}_{ji}\},\{{\tilde{z}}^{(0)}_{ji}\};
2 Iterate until convergence:
3 for player i𝒩i\in\mathcal{N} do
4      
5      Communicate with neighboring players and edges to obtain y~iL(k){\tilde{y}}^{(k)}_{iL}, μ~iB(k){\tilde{\mu}}^{(k)}_{iB}, λ~iL(k){\tilde{\lambda}}^{(k)}_{iL}, z~iB(k){\tilde{z}}^{(k)}_{iB};
6      
7      yii(k+1)=y~ii(k)τ1i2(ρμy~iLi(k)+μ~iBi(k)){y}^{-i(k+1)}_{i}={\tilde{y}}^{-i(k)}_{i}-\frac{{\tau}_{1i}}{2}({\rho}_{\mu}{\tilde{y}}^{-i(k)}_{iL}+{\tilde{\mu}}^{-i(k)}_{iB}) ;
8      
9      Obtain yii(k+1)y^{i(k+1)}_{i} via Subroutine 2 that approximately solves: argminvi𝒳iB[𝕁i(vi;yii(k+1))+12(λ~i(k))TAivi{\operatorname{argmin}}_{v_{i}\in\mathcal{X}^{B}_{i}}[{\mathbb{J}}_{i}({v}_{i};{y}^{-i(k+1)}_{i})+\frac{1}{2}{(\tilde{\lambda}}^{(k)}_{i})^{T}A_{i}{v}_{i}
10       +12(ρμy~iLi(k)+μiBi(k))Tvi+12τ1iviy~ii(k)2]\qquad\qquad+\frac{1}{2}{({\rho}_{\mu}{\tilde{y}}^{i(k)}_{iL}+{\mu}^{i(k)}_{iB})}^{T}{v}_{i}+\frac{1}{2{\tau}_{1i}}\lVert{v}_{i}-{\tilde{y}}^{i(k)}_{i}\rVert^{2}];
11       λi(k+1)=λ~i(k)+τ2i(Ai(yii(k+1)12y~ii(k))ρz2λ~iL(k)12z~iB(k)ci){\lambda}^{(k+1)}_{i}={\tilde{\lambda}}^{(k)}_{i}+{\tau}_{2i}({A}_{i}({y}^{i(k+1)}_{i}-\frac{1}{2}{\tilde{y}}^{i(k)}_{i})-\frac{{\rho}_{z}}{2}{\tilde{\lambda}}^{(k)}_{iL}-\frac{1}{2}{\tilde{z}}^{(k)}_{iB}-{c}_{i});
12       y^i(k+1)=2yi(k+1)y~i(k),λ^i(k+1)=2λi(k+1)λ~i(k){\hat{y}}^{(k+1)}_{i}=2{y}^{(k+1)}_{i}-{\tilde{y}}^{(k)}_{i},\;{\hat{\lambda}}^{(k+1)}_{i}=2{\lambda}^{(k+1)}_{i}-{\tilde{\lambda}}^{(k)}_{i};
13      
14 end for
15for edge (j,i)g(j,i)\in\mathcal{E}_{g} do
16       Comm. with player ii & jj to obtain y^ji(k+1){\hat{y}}^{(k+1)}_{ji}, λ^ji(k+1){\hat{\lambda}}^{(k+1)}_{ji};
17       μji(k+1)=μ~ji(k)+τ3i2y^ji(k+1),μ^ji(k+1)=2μji(k+1)μ~ji(k){\mu}^{(k+1)}_{ji}={\tilde{\mu}}^{(k)}_{ji}+\frac{{\tau}_{3i}}{2}{\hat{y}}^{(k+1)}_{ji},\;{\hat{\mu}}^{(k+1)}_{ji}=2{\mu}^{(k+1)}_{ji}-{\tilde{\mu}}^{(k)}_{ji};
18       zji(k+1)=z~ji(k)+τ4i2λ^ji(k+1),z^ji(k+1)=2zji(k+1)z~ji(k){z}^{(k+1)}_{ji}={\tilde{z}}^{(k)}_{ji}+\frac{{\tau}_{4i}}{2}{\hat{\lambda}}^{(k+1)}_{ji},\;{\hat{z}}^{(k+1)}_{ji}=2{z}^{(k+1)}_{ji}-{\tilde{z}}^{(k)}_{ji};
19      
20 end for
21
22for player i𝒩i\in\mathcal{N} do
23       Comm. & obtain y^iL(k+1){\hat{y}}^{(k+1)}_{iL}, μ^iB(k+1){\hat{\mu}}^{(k+1)}_{iB}, λ^iL(k+1){\hat{\lambda}}^{(k+1)}_{iL}, z^iB(k+1){\hat{z}}^{(k+1)}_{iB};
24       y¯i(k+1)=Pj𝒳^i[y^i(k+1)τ1i2(iTAiTλ^i(k+1)+ρμy^iL(k+1)+μ^iB(k+1))]{\bar{y}}^{(k+1)}_{i}=\operatorname{Pj}_{\scriptscriptstyle\hat{\mathcal{X}}_{i}}[{\hat{y}}^{(k+1)}_{i}-\frac{{\tau}_{1i}}{2}({\mathcal{R}}^{T}_{i}{A}^{T}_{i}{\hat{\lambda}}^{(k+1)}_{i}+{\rho}_{\mu}{\hat{y}}^{(k+1)}_{iL}+{\hat{\mu}}^{(k+1)}_{iB})];
25      
26      λ¯i(k+1)=Pj+m[λ^i(k+1)+τ2i(Ai(y¯ii(k+1)12y^ii(k)){\bar{\lambda}}^{(k+1)}_{i}=\operatorname{Pj}_{\scriptscriptstyle\mathbb{R}^{\scriptscriptstyle m}_{\scriptscriptstyle+}}[{\hat{\lambda}}^{(k+1)}_{i}+{\tau}_{2i}({A}_{i}({\bar{y}}^{i(k+1)}_{i}-\frac{1}{2}{\hat{y}}^{i(k)}_{i})
27       ρz2λ^iL(k+1)12z^iB(k+1))]\quad-\frac{{\rho}_{z}}{2}{\hat{\lambda}}^{(k+1)}_{iL}-\frac{1}{2}{\hat{z}}^{(k+1)}_{iB})];
28      
29 end for
30
31for edge (j,i)g(j,i)\in\mathcal{E}_{g} do
32       Comm. & obtain y¯ji(k+1){\bar{y}}^{(k+1)}_{ji}, λ¯ji(k+1){\bar{\lambda}}^{(k+1)}_{ji};
33       μ¯ji(k+1)=μ^ji(k+1)+τ3i(y¯ji(k+1)12y^ji(k+1)){\bar{\mu}}^{(k+1)}_{ji}={\hat{\mu}}^{(k+1)}_{ji}+{\tau}_{3i}({\bar{y}}^{(k+1)}_{ji}-\frac{1}{2}{\hat{y}}^{(k+1)}_{ji});
34       z¯ji(k+1)=z^ji(k+1)+τ4i(λ¯ji(k+1)12λ^ji(k+1)){\bar{z}}^{(k+1)}_{ji}={\hat{z}}^{(k+1)}_{ji}+{\tau}_{4i}({\bar{\lambda}}^{(k+1)}_{ji}-\frac{1}{2}{\hat{\lambda}}^{(k+1)}_{ji});
35      
36 end for
37K-M updates: ψ~(k+1)=ψ~(k)+2γ(k)(ψ¯(k+1)ψ(k+1));{\tilde{\psi}}^{(k+1)}={\tilde{\psi}}^{(k)}+2{\gamma}^{(k)}({\bar{\psi}}^{(k+1)}-{\psi}^{(k+1)});
Return: {y~i(k)}\{{\tilde{y}}^{(k)}_{i}\}.
Algorithm 1 Distributed v-SGNE Seeking under the Partial-Decision Information Setting

Depending on the inexact solver adopted, R𝒜¯R_{\operatorname{\bar{\mathcal{A}}}} usually admits no explicit formulas. Yet, as will be shown later in the next section, we can still establish the convergence of Algorithm 1 based on some specific properties of R𝒜¯R_{\operatorname{\bar{\mathcal{A}}}}.

IV Convergence Analysis and Construction of Inexact Solver

IV-A General Convergence Results Using Approximate Solution

We start by stating the Robbins-Siegmund theorem [35, Thm. 1], which plays a significant role in analyzing the convergence of algorithms in the field of stochastic optimization. In this subsection, we study the sufficient conditions from a generic perspective to guarantee the convergence of Algorithm 1 to a v-SGNE of the problem (3). We first define the approximate error and its norm for each iteration as

ϵ(k)(ψ~(k))(ψ~(k))andε(k)ϵ(k)𝒦,\displaystyle{\epsilon}^{(k)}\coloneqq\mathscr{R}({\tilde{\psi}}^{(k)})-\mathscr{R}_{*}({\tilde{\psi}}^{(k)})\;\text{and}\;{\varepsilon}^{(k)}\coloneqq\lVert{\epsilon}^{(k)}\rVert_{\mathcal{K}}, (19)

where ψ~(k)[𝒚~(k);𝝀~(k);𝝁~(k);𝒛~(k)]{\tilde{\psi}}^{(k)}\coloneqq[{\tilde{\boldsymbol{y}}}^{(k)};{\tilde{\boldsymbol{\lambda}}}^{(k)};{\tilde{\boldsymbol{\mu}}}^{(k)};{\tilde{\boldsymbol{z}}}^{(k)}]. We next introduce the residual function res(ψ~)ψ~(ψ~)𝒦\text{res}(\tilde{\psi})\coloneqq\lVert\tilde{\psi}-\mathscr{R}_{*}(\tilde{\psi})\rVert_{\mathcal{K}} such that res(ψ~)=0\text{res}(\tilde{\psi}^{*})=0 is a necessary condition for ψ~\tilde{\psi}^{*} to belong to the fixed-point set of \mathscr{R}_{*}. This relation can be easily checked by using [33, Prop. 26.1(iii)]. Let k\mathcal{F}_{k} denote the σ\sigma-field comprised of {ψ~(0),{ξi(0)}i𝒩,,{ξi(k1)}i𝒩}\{{\tilde{\psi}}^{(0)},\{{\xi}^{(0)}_{i}\}_{i\in\mathcal{N}},\ldots,\{{\xi}^{(k-1)}_{i}\}_{i\in\mathcal{N}}\}, where for each major iteration kk\in\mathbb{N}, ξi(k)={ξi,0(k),,ξi,Ti(k)1(k)}{\xi}^{(k)}_{i}=\{{\xi}^{(k)}_{i,0},\ldots,{\xi}^{(k)}_{i,{T}^{(k)}_{i}-1}\} and Ti(k){T}^{(k)}_{i} denotes the number of noise realizations that player ii observes at the kk-th iteration.

Theorem 2.

Consider the SGNEP given in (3), and suppose Assumptions 1 to 6 hold. Moreover, (γ(k))k(\gamma^{(k)})_{k\in\mathbb{N}} is a sequence such that γ(k)[0,1]\gamma^{(k)}\in[0,1] and kγ(k)(1γ(k))=+\sum_{k\in\mathbb{N}}\gamma^{(k)}(1-\gamma^{(k)})=+\infty. If the sequence (ψ~(k))(\tilde{\psi}^{(k)}) generated by the inexact solver satisfies {outline}[enumerate] \1 (ψ~(k)𝒦)k(\lVert\tilde{\psi}^{(k)}\rVert_{\mathcal{K}})_{k\in\mathbb{N}} is bounded a.s.; \1 kγ(k)𝔼[ε(k)(k)]<, a.s. \sum_{k\in\mathbb{N}}\gamma^{(k)}\mathbb{E}[\varepsilon^{(k)}\mid\mathcal{F}^{(k)}]<\infty,\text{ a.s. }, then (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} converges to a fixed point of \mathscr{R}_{*} a.s., and limkJ𝔸¯(ψ~(k))=ψ\lim_{k\to\infty}J_{\operatorname{\bar{\mathbb{A}}}}(\tilde{\psi}^{(k)})=\psi^{*} a.s. Also, the corresponding entries of ψ\psi^{*} satisfy 𝐲=(𝟏Ny)\boldsymbol{y}^{*}=(\boldsymbol{1}_{N}\otimes y^{*}) and 𝛌(k)=(𝟏λ)\boldsymbol{\lambda}^{(k)}=(\boldsymbol{1}\otimes\lambda^{*}). Here, yy^{*} is a v-SGNE of the original SGNEP (3) and (y,λ)(y^{*},\lambda^{*}) together is a solution to the KKT conditions (6) of the SGNEP.

Proof.

See Appendix A. ∎

Before proceeding, it is worth highlighting why we need to keep both the condition (i) and (ii) to hold in Theorem 2. Although the condition (i), i.e., (ψ~(k)𝒦)k(\lVert\tilde{\psi}^{(k)}\rVert_{\mathcal{K}})_{k\in\mathbb{N}} is bounded a.s., is a necessary condition for the summability statement in (ii), as has been showed in [33, Prop. 5.34] for deterministic cases, under the partial-information setting, a natural strategy is to prove the condition (i) first using a more primitive condition, and then establish the condition (ii) based on (i). The specific conditions regarding the algorithm parameters to ensure the convergence will be later discussed in Theorem 3.

Remark 2.

When proving Theorem 2, the inequalities invoked follow from the quasinonexpansiveness of the exact operator \mathscr{R}_{*} and the Cauchy-Schwarz inequality. The proof and conclusion in Theorem 2 thus can be applied to the analysis of a general continuous operator QQ in (12) and its approximation other than the operators \mathscr{R}_{*} and \mathscr{R} in this paper, as long as the operator QQ is quasinonexpansive and the conditions regarding (γ(k))k(\gamma^{(k)})_{k\in\mathbb{N}}, (ε(k))k(\varepsilon^{(k)})_{k\in\mathbb{N}}, and (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} in Theorem 2 are satisfied.

IV-B Construction of a Desirable Inexact Solver

As we discussed at the end of Section III, it is challenging to solve the augmented best-response subproblems that involve the exact expected-value objectives (the argmin\operatorname{argmin} problems in the first player for-loop of Algorithm 1). Theorem 2 suggests that we can still obtain a v-SGNE by solving these augmented best-response subproblems not precisely but up to some prescribed accuracy. In this subsection, we consider a specific scenario-based solver using the projected stochastic subgradient method [36, Ch. 2]. As has been shown in the existing literature[37], the weighted average of the projected stochastic subgradient method possesses an O(1/t)O(1/t) convergence rate if the subgradient is unbiased and the variance of the subgradient is finite. Here, we study the explicit conditions that the projected stochastic subgradient solver should satisfy to serve as a feasible inexact solver in the context of distributed SGNEP with only partial-decision information, as suggested in Theorem 2.

We first assume the unbiasedness and finite-variance properties of a general projected stochastic subgradient method. Throughout this subsection, we use kk to index the major iterations (the iteration of the v-SGNE seeking Algorithm 1) and tt to index the minor iterations (the iteration of the inexact solver in the first player for-loop of Algorithm 1). Furthermore, at each major iteration kk, for each player ii, let the augmented scenario-based objective function be denoted by J^i(k)(vi;ξi,t(k))Ji(vi;yii(k+1),ξi,t(k))+(φ~i(k))Tyii+12τ1iviy~ii(k)22\hat{J}^{(k)}_{i}(v_{i};\xi^{(k)}_{i,t})\coloneqq J_{i}(v_{i};y^{-i(k+1)}_{i},\xi^{(k)}_{i,t})+(\tilde{\varphi}^{(k)}_{i})^{T}y^{i}_{i}+\frac{1}{2\tau_{1i}}\lVert v_{i}-\tilde{y}^{i(k)}_{i}\rVert^{2}_{2}, and the augmented expected-value objective function be denoted by 𝕁^i(k)(vi)𝕁i(vi;yii(k+1))+(φ~i(k))Tvi+12τ1iviy~ii(k)22\hat{\mathbb{J}}^{(k)}_{i}(v_{i})\coloneqq\mathbb{J}_{i}(v_{i};y^{-i(k+1)}_{i})+(\tilde{\varphi}^{(k)}_{i})^{T}v_{i}+\frac{1}{2\tau_{1i}}\lVert v_{i}-\tilde{y}^{i(k)}_{i}\rVert^{2}_{2}, where φ~i(k)12(AiTλ~i(k)+μ~iBi(k)+ρμy~iLi(k))\tilde{\varphi}^{(k)}_{i}\coloneqq\frac{1}{2}(A_{i}^{T}\tilde{\lambda}^{(k)}_{i}+\tilde{\mu}^{i(k)}_{iB}+\rho_{\mu}\tilde{y}^{i(k)}_{iL}). Note that 𝕁^i(k)()\hat{\mathbb{J}}^{(k)}_{i}(\cdot) is the objective in the first player-loop of Algorithm 1 that needs to be inexactly solved. Here, the vector φ~i(k)\tilde{\varphi}^{(k)}_{i} represents some augmented terms that enforce the consensus constraints and the global resource constraints. For brevity, the local estimates of the other players’ decisions yii(k+1)y^{-i(k+1)}_{i} are omitted from the arguments of the augmented functions defined above. Let Ti(k)T^{(k)}_{i} denote the total number of the projected stochastic subgradient steps taken in the kk-th major iteration by player ii. The subgradient of the scenario-based objective function at the kk-th major iteration and the tt-th minor iteration is denoted by gi,t(k)yiiJ^i(k)(yi,ti(k+1);ξi,t(k))g^{(k)}_{i,t}\in\partial_{y^{i}_{i}}\hat{J}^{(k)}_{i}(y^{i(k+1)}_{i,t};\xi^{(k)}_{i,t}), where t=0,1,,Ti(k)1t=0,1,\ldots,T^{(k)}_{i}-1.

Assumption 7.

For each player i𝒩i\in\mathcal{N}, at each major iteration kk of Algorithm 1, the following statements hold:
(i) (Unbiasedness) At each minor iteration tt, there exists a gi,t(k)yiiJ^i(k)(yi,ti(k+1);ξi,t(k))g^{(k)}_{i,t}\in\partial_{y^{i}_{i}}\hat{J}^{(k)}_{i}(y^{i(k+1)}_{i,t};\xi^{(k)}_{i,t}) such that 𝔼[gi,t(k)σ{k,ξi,[t](k)}]\mathbb{E}[g^{(k)}_{i,t}\mid\sigma\{\mathcal{F}_{k},\xi^{(k)}_{i,[t]}\}] is a.s. a subgradient of the expected-value augmented objective 𝕁^i(k)()\hat{\mathbb{J}}_{i}^{(k)}(\cdot) at yi,ti(k+1)y^{i(k+1)}_{i,t}, where ξi,[t](k){ξi,0(k),,ξi,t1(k)}\xi^{(k)}_{i,[t]}\coloneqq\{\xi^{(k)}_{i,0},\ldots,\xi^{(k)}_{i,t-1}\} with ξi,[0](k)\xi^{(k)}_{i,[0]}\coloneqq\varnothing;
(ii) (Upper-bounded variance) For any yii𝒳iBy^{i}_{i}\in\mathcal{X}^{B}_{i}, there exists a gi(k)yiiJ^i(k)(yii;ξi)g^{(k)}_{i}\in\partial_{y^{i}_{i}}\hat{J}^{(k)}_{i}(y^{i}_{i};\xi_{i}) such that 𝔼[gi(k)22k]αg,i2ψ~(k)22+βg,i2\mathbb{E}[\lVert g^{(k)}_{i}\rVert^{2}_{2}\mid\mathcal{F}_{k}]\leq\alpha_{g,i}^{2}\lVert\tilde{\psi}^{(k)}\rVert^{2}_{2}+\beta_{g,i}^{2} a.s. for some positive constants αg,i\alpha_{g,i} and βg,i\beta_{g,i}.

We refer the reader to the paragraph before Theorem 2 for the definitions of the stack vector ψ~(k)\tilde{\psi}^{(k)} and the filtration (k)k(\mathcal{F}_{k})_{k\in\mathbb{N}} as a reminder. The proposed inexact solver for the first player for-loop of Algorithm 1 is given in Subroutine 2. Note that the DR splitting scheme ensures local feasibility with single projection onto local feasible sets, and requires multiple projections onto relaxed bounded box sets, which considerable reduces the computational complexity compared with other methods such as proximal-point scheme [21].

1 For each i𝒩i\in\mathcal{N}, at the kk-th major iteration of Alg. 1:
2 Initialize: yi,0i(k+1)y~ii(k);y^{i(k+1)}_{i,0}\coloneqq\tilde{y}^{i(k)}_{i};
3 for t=0 to Ti(k)1t=0\text{ to }T^{(k)}_{i}-1 do
4       yi,t+1i(k+1)Pj𝒳iB[yi,ti(k+1)κi,tgi,t(k)]y^{i(k+1)}_{i,t+1}\coloneqq\operatorname{Pj}_{\mathcal{X}^{B}_{i}}[y^{i(k+1)}_{i,t}-\kappa_{i,t}\cdot g^{(k)}_{i,t}], with κi,t2τ1it+2\kappa_{i,t}\coloneqq\frac{2\tau_{1i}}{t+2};
5      
6 end for
Return: yii(k+1)yi,Ti(k)i(k+1)y^{i(k+1)}_{i}\coloneqq y^{i(k+1)}_{i,T^{(k)}_{i}}.
Subroutine 2 Projected Stochastic Subgradient Solver

The following lemma discusses the convergence rate of Subroutine 2 as a minor updating routine inside Algorithm 1. We use yi,i(k+1)y^{i(k+1)}_{i,*} to denote the accurate minimizer of the expected-value augmented function 𝕁^i(k)()\hat{\mathbb{J}}^{(k)}_{i}(\cdot).

Lemma 1.

Suppose Assumptions 1 to 7 hold. Then, for any T=1,,Ti(k)T=1,\ldots,T^{(k)}_{i}, the distance between the approximate solution by Subroutine 2 and the accurate solution satisfies 𝔼[yi,Ti(k+1)yi,i(k+1)22k]4τ1i2T1(αg,i2ψ~(k)22+βg,i2)\mathbb{E}[\lVert y^{i(k+1)}_{i,T}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}]\leq 4\tau_{1i}^{2}T^{-1}(\alpha^{2}_{g,i}\lVert\tilde{\psi}^{(k)}\rVert^{2}_{2}+\beta^{2}_{g,i}) a.s.

Proof.

See Appendix B. ∎

From Lemma 1, we can conclude that for each player i𝒩i\in\mathcal{N}, after the kk-th major iteration of Algorithm 1 where player ii implements Ti(k)T^{(k)}_{i} projected stochastic subgradient steps in Subroutine 2, 𝔼[yii(k+1)yi,i(k+1)22k](2τ1i)2Ti(k)(αg,i2ψ~(k)22+βg,i2)\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}\leq\frac{(2\tau_{1i})^{2}}{T^{(k)}_{i}}(\alpha^{2}_{g,i}\lVert\tilde{\psi}^{(k)}\rVert^{2}_{2}+\beta^{2}_{g,i}). Based on this result, it is straightforward to derive an upper bound for the approximate error ε(k)(ψ~(k))(ψ~(k))𝒦\varepsilon^{(k)}\coloneqq\lVert\mathscr{R}(\tilde{\psi}^{(k)})-\mathscr{R}_{*}(\tilde{\psi}^{(k)})\rVert_{\mathcal{K}}. As will be shown later, this upper bound can be treated as a function of \stackunder[1.2pt]T (k)min{Ti(k):i𝒩}\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)}\coloneqq\min\{T^{(k)}_{i}:i\in\mathcal{N}\} which we can tune to provide a desirable sequence of approximation accuracies.

Lemma 2.

Consider (ε(k))k(\varepsilon^{(k)})_{k\in\mathbb{N}} generated by Algorithm 1 using Subroutine 2 as the inexact solver. Suppose Assumptions 1 to 7 hold. Then there exist some positive constants αψ\alpha_{\psi} and βψ\beta_{\psi} such that the following relation holds a.s.:

𝔼[ε(k)k](\stackunder[1.2pt]T (k))1/2(αψψ~(k)𝒦+βψ).\mathbb{E}\big{[}\varepsilon^{(k)}\mid\mathcal{F}_{k}\big{]}\leq(\stackunder[1.2pt]{T}{\rule{3.44444pt}{0.32289pt}}^{(k)})^{-1/2}(\alpha_{\psi}\lVert\tilde{\psi}^{(k)}\rVert_{\mathcal{K}}+\beta_{\psi}). (20)
Proof.

See Appendix C. ∎

Lemma 2 establishes the relationship between the approximate error ε(k)\varepsilon^{(k)} and the stack vector ψ~(k)\tilde{\psi}^{(k)} at each major iteration kk. We define γT(k)γ(k)(\stackunder[1.2pt]T (k))1/2\gamma^{(k)}_{T}\coloneqq\gamma^{(k)}(\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)})^{-1/2}. From Theorem 2, it suffices to have the sequence (γT(k))k(\gamma^{(k)}_{T})_{k\in\mathbb{N}} summable and (ψ~(k))k(\lVert\tilde{\psi}^{(k)}\rVert)_{k\in\mathbb{N}} bounded. To this end, we next focus on proving the conditions needed to guarantee the boundedness of (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} and finally derive the sufficient conditions to ensure the convergence of Algorithm 1.

Theorem 3.

Consider the sequence (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} generated by Algorithm 1 using Subroutine 2 as an inexact solver. Suppose Assumptions 1 to 7 hold. In addition, the sequence (γ(k))k({\gamma}^{(k)})_{k\in\mathbb{N}} satisfies 0γ(k)10\leq\gamma^{(k)}\leq 1 and kγ(k)(1γ(k))=+\sum_{k\in\mathbb{N}}\gamma^{(k)}(1-\gamma^{(k)})=+\infty, and the sequence (γT(k))k(\gamma^{(k)}_{T})_{k\in\mathbb{N}} is absolutely summable. Then (ψ~(k)𝒦)k(\lVert\tilde{\psi}^{(k)}\rVert_{\mathcal{K}})_{k\in\mathbb{N}} is bounded a.s., and kγ(k)𝔼[ε(k)k]<\sum_{k\in\mathbb{N}}\gamma^{(k)}\mathbb{E}[\varepsilon^{(k)}\mid\mathcal{F}_{k}]<\infty a.s. As a result, the sequence (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} will converge to a fixed point of \mathscr{R}_{*} and the associated sequence (y(k))k(y^{(k)})_{k\in\mathbb{N}} generated by J𝔸¯(ψ~(k))J_{\operatorname{\bar{\mathbb{A}}}}(\tilde{\psi}^{(k)}) will converge to a v-SGNE of the problem (3).

Proof.

See Appendix D. ∎

Based on Theorem 3, to get a solution arbitrarily closed to a v-SGNE, we will run Algorithm 1 for a sufficiently large number of major iterations. Then, we use the obtained last-iterate ψ~(k)\tilde{\psi}^{(k)} to run Subroutine 2 for another sufficiently large number of minor iterations \stackunder[1.2pt]T (k)\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)} [38, Lemma 3]. To ensure the convergence of Algorithm 1, it suffices to properly choose (γ(k))k(\gamma^{(k)})_{k\in\mathbb{N}} and (\stackunder[1.2pt]T (k))k(\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)})_{k\in\mathbb{N}} such that (γT(k))k(\gamma^{(k)}_{T})_{k\in\mathbb{N}} is a summable sequence. As an example for admissible parameters, we can choose γ(k)=1/ka\gamma^{(k)}=1/k^{a} and \stackunder[1.2pt]T (k)=kb\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)}=k^{b}, with 0<a10<a\leq 1 and a+b/2>1a+b/2>1. This can be manipulated to make the proposed algorithm work under different practical settings. For instance, if these players are working in a feedback-parsimonious setting, i.e., the available realizations of noisy first-order/gradient information per iteration are scarce, one can choose a faster decaying rate for (γ(k))k(\gamma^{(k)})_{k\in\mathbb{N}} as long as kγ(k)=+\sum_{k\in\mathbb{N}}\gamma^{(k)}=+\infty and let (\stackunder[1.2pt]T (k))k(\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)})_{k\in\mathbb{N}} grow linearly or even sublinearly. In contrast, if the available realizations are abundant, one can let (\stackunder[1.2pt]T (k))k(\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)})_{k\in\mathbb{N}} grow superlinearly while fixing γ(k)\gamma^{(k)} to be some constant such that the proposed algorithm can enjoy a faster convergence rate.

V Case Study and Numerical Simulations

V-A Stochastic Nash-Cournot Distribution Game

We evaluate the performance of the proposed algorithm with a Nash Cournot distribution problem [10, Sec. 1.4.3][39] over a transport network. Several firms (indexed by 𝒩{1,,N}\mathcal{N}\coloneqq\{1,\ldots,N\}), who produce a common homogeneous commodity, participate in this game. These firms try to optimize their own payoffs by deciding the quantity of the commodity to produce at each factory and the quantities to distribute to different markets. A transport network is provided, with markets as the nodes and roads as the edges. Let 𝒩T\mathcal{N}_{T} denote the node set of this network and T\mathcal{E}_{T} the edge set, distinguished from 𝒩g\mathcal{N}_{g} and g\mathcal{E}_{g} of the underlying communication network 𝒢\mathcal{G}. Denote cardinalities of 𝒩T\mathcal{N}_{T} and T\mathcal{E}_{T} by NTN_{T} and ETE_{T}, and the incident matrix of this transport network by BTNT×ETB_{T}\in\mathbb{R}^{N_{T}\times E_{T}}.

Each firm has NTiN_{T_{i}} factories at certain nodes on this transport network, given by the set 𝒩Ti\mathcal{N}_{T_{i}}. Its decision vector xiET+NTix_{i}\in\mathbb{R}^{E_{T}+N_{T_{i}}} is comprised of two parts (xi[ui;vi]x_{i}\coloneqq[u_{i};v_{i}]): each entry of ui+ETu_{i}\in\mathbb{R}^{E_{T}}_{+} represents the quantity of the commodity delivered through a road in T\mathcal{E}_{T}; each entry of vi+NTiv_{i}\in\mathbb{R}^{N_{T_{i}}}_{+} represents the quantity of the commodity produced by one of its factories in 𝒩Ti\mathcal{N}_{T_{i}}. The indicator matrix which maps from each entry of viv_{i} to the corresponding node on the transport network is denoted by EiNT×NTiE_{i}\in\mathbb{R}^{N_{T}\times N_{T_{i}}}, and we let Ai[BT,Ei]A_{i}\coloneqq[B_{T},E_{i}]. These two parts (uiu_{i} and viv_{i}) together uniquely determine the distribution of commodity AixiA_{i}x_{i} over the markets. If we assume that the factories owned by firm ii have maximum production capacities bi++NTib_{i}\in\mathbb{R}^{N_{T_{i}}}_{++}, then each entry of the vector uiETu_{i}\in\mathbb{R}^{E_{T}} is upper-bounded by bi1\lVert b_{i}\rVert_{1}, and the local feasible set 𝒳i\mathcal{X}_{i} is a polytope which can be written as: 𝒳i{xiET+NTi0vibi,0uibi1𝟏ET,Aixi0}\mathcal{X}_{i}\coloneqq\{x_{i}\in\mathbb{R}^{E_{T}+N_{T_{i}}}\mid 0\leq v_{i}\leq b_{i},0\leq u_{i}\leq\lVert b_{i}\rVert_{1}\otimes\boldsymbol{1}_{E_{T}},A_{i}x_{i}\geq 0\}. The objective function of each firm ii is given by: Ji(xi;xi,ξi)=xiTQixi+Ct(ui)+Cpi(vi)(P(Ax)+ξi)TAixiJ_{i}(x_{i};x_{-i},\xi_{i})=x_{i}^{T}Q_{i}x_{i}+C_{t}(u_{i})+C^{i}_{p}(v_{i})-(P(Ax)+\xi_{i})^{T}A_{i}x_{i}, where Qi𝕊++ET+NTiQ_{i}\in\mathbb{S}^{E_{T}+N_{T_{i}}}_{++}, A[A1,,AN]A\coloneqq[A_{1},\ldots,A_{N}], x[x1;;xN]nx\coloneqq[x_{1};\ldots;x_{N}]\in\mathbb{R}^{n} with nNET+i𝒩NTin\coloneqq NE_{T}+\sum_{i\in\mathcal{N}}N_{T_{i}}, and P(Ax)wΣAxP(Ax)\coloneqq w-\Sigma Ax maps from the total quantities AxAx of the commodity at markets to their unit prices with w+NTw\in\mathbb{R}^{N_{T}}_{+} and Σ𝕊++NT\Sigma\in\mathbb{S}^{N_{T}}_{++}. The transport cost CtC_{t} is defined as the sum of the costs at all roads, i.e., Ct(ui)kTCtk([ui]k)C_{t}(u_{i})\coloneqq\sum_{k\in\mathcal{E}_{T}}C^{k}_{t}([u_{i}]_{k}), where each road kTk\in\mathcal{E}_{T} has Ctk([ui]k)ηk([ui]k(111+[ui]k))C^{k}_{t}([u_{i}]_{k})\coloneqq\eta_{k}([u_{i}]_{k}-(1-\frac{1}{1+[u_{i}]_{k}})). The production cost CpiC^{i}_{p} is also defined as the sum of the costs at all factories, i.e., Cpi(vi)k𝒩TiCpi,k(vi)C^{i}_{p}(v_{i})\coloneqq\sum_{k\in\mathcal{N}_{T_{i}}}C^{i,k}_{p}(v_{i}), where each factory k𝒩Tik\in\mathcal{N}_{T_{i}} has Cpi,k([vi]k)κi,k([vi]k(111+[vi]k))C^{i,k}_{p}([v_{i}]_{k})\coloneqq\kappa_{i,k}([v_{i}]_{k}-(1-\frac{1}{1+[v_{i}]_{k}})). The total income (P(Ax)+ξi)TAixi(P(Ax)+\xi_{i})^{T}A_{i}x_{i} captures uncertainty in the unit prices through the random vector ξi\xi_{i}, which has its entries independently identically distributed with mean zero.

Furthermore, we assume that each market has a maximum capacity for the commodity, and the decision vectors of the players should collectively satisfy the global resource constraints i𝒩Aixic\sum_{i\in\mathcal{N}}A_{i}x_{i}\leq c where c++NTc\in\mathbb{R}^{N_{T}}_{++}. Building on the discussed setups, each firm i𝒩i\in\mathcal{N}, given the production and distribution strategies of the other players (xix_{-i}), aims to solve the following stochastic optimization problem:

{minimizexi𝒳i𝔼ξi[Ji(xi;xi,ξi)]subjecttoAixicj𝒩iAjxj.\begin{cases}\operatorname{minimize}_{x_{i}\in\mathcal{X}_{i}}&\mathbb{E}_{\xi_{i}}[J_{i}(x_{i};x_{-i},\xi_{i})]\\ \operatorname{subject\>to}&A_{i}x_{i}\leq c-\sum_{j\in\mathcal{N}_{-i}}A_{j}x_{j}.\end{cases} (21)

V-A1 Assumptions Verification

We use the transport network of the city of Oldenburg [40] (Fig. 2 top): it consists of NT=29N_{T}=29 nodes (markets) and ET=2×34E_{T}=2\times 34 directed edges (roads). Five firms (N=5N=5) participates in this game, and each firm has a single factory at a given location/node {8,14,21,10,29}\{8,14,21,10,29\}. Each factory has its maximum production capacity uniformly sampled from the interval [10,14][10,14], and QiQ_{i} is a diagonal matrix with the diagonal entries uniformly sampled from [2,3][2,3]. In the transport costs, we have 18ηk(0,1]\frac{1}{8}\eta_{k}\in(0,1] being the ratio between the length of road kk and the maximum length of the roads in T\mathcal{E}_{T}. In the production costs, we fix the coefficients κi,k=2\kappa_{i,k}=2. In the price function P()P(\cdot), we draw each entry of the vector ww uniformly at random from the interval [26,30][26,30] and set the matrix Σ\Sigma to have [Σ]ii1[\Sigma]_{ii}\coloneqq 1 for all i𝒩Ti\in\mathcal{N}_{T} and [Σ]ji0.3(118η(j,i))[\Sigma]_{ji}\coloneqq 0.3\cdot(1-\frac{1}{8}\eta_{(j,i)}) for all (j,i)T(j,i)\in\mathcal{E}_{T}.

For each player i𝒩i\in\mathcal{N}, it is easy to check by definition that Ji(xi;xi,ξi)J_{i}(x_{i};x_{-i},\xi_{i}) and 𝕁i(xi;xi)\mathbb{J}_{i}(x_{i};x_{-i}) are smooth and proper, and they are convex in xix_{i}. Moreover, the pseudogradient 𝔽\mathbb{F} is strongly monotone on the local compact feasible sets i𝒩𝒳i\prod_{i\in\mathcal{N}}\mathcal{X}_{i} (detailed verifications are omitted due to space limit). Then by [10, Thm. 2.3.3], this problem admits a unique v-SGNE. We set the communication graph of the players to be composed of an undirected circle plus two randomly selected edges. Therefore, Assumptions 1 to 4 are fulfilled. We choose ρμ=8\rho_{\mu}=8 and then appropriately set the step sizes to be 𝝉1=0.0285INn\boldsymbol{\tau}_{1}=0.0285\otimes I_{Nn}, 𝝉2=0.09INm\boldsymbol{\tau}_{2}=0.09\otimes I_{Nm}, 𝝉3=0.5IEn\boldsymbol{\tau}_{3}=0.5\otimes I_{En}, and 𝝉4=0.5IEm\boldsymbol{\tau}_{4}=0.5\otimes I_{Em}. It can be checked numerically that the conditions in Assumptions 5 and 6 are satisfied. We further set ξiU[2,2]\xi_{i}\sim U[-2,2] and can easily verify the conditions in Assumption 7.

V-A2 Simulation Results

The sequence (γ(k))k(\gamma^{(k)})_{k\in\mathbb{N}} is fixed to be 1/21/2, and the subgradient steps taken is chosen as T(k)=104k2.1+20T^{(k)}=\lceil 10^{-4}k^{2.1}\rceil+20. We compare the performance of Algorithm 1 with that of [28], with c=4c=4 and the relaxed step sizes chosen as 0.040.04. Theses step sizes are empirically pushed to near the upper limit of convergence; otherwise, [20, Lemma 6] suggests a set of miniscule and conservative step sizes (3×105\approx 3\times 10^{-5}). The performances of the proposed algorithm are shown in Fig. 1. We use the thick and semi-transparent lines to illustrate the real fluctuations of the metrics throughout the iterations, while using the thin lines to exhibit the simple moving averages of the metrics with a window size of 3030. The averages of the normalized distances to the v-SGNE are presented in Fig. 1(a), where the unique v-SGNE is calculated using the centralized method from [41]. Note that yj(k)y^{(k)}_{j} denotes the stack of player jj’s local decision and local estimates at the kk-th iteration, and yy^{*} the v-SGNE of the game. Fig. 1(b) shows the relative lengths of the updating step at each iteration. Let y¯(k)1Nj𝒩yj(k)\bar{y}^{(k)}\coloneqq\frac{1}{N}\sum_{j\in\mathcal{N}}y^{(k)}_{j}. Fig. 1(c) exhibits how the sums of the standard deviations of the local estimates {yj}\{y_{j}\}, i.e., =1n(1Nj𝒩([yj(k)][y¯(k)])2)12\sum_{\ell=1}^{n}(\frac{1}{N}\sum_{j\in\mathcal{N}}([y^{(k)}_{j}]_{\ell}-[\bar{y}^{(k)}]_{\ell})^{2})^{\frac{1}{2}}, evolve over the iterations. It measures the level of consensus among different local estimates yjy_{j}. Fig. 1(d) is almost the same as Fig. 1(c) except that we are now investigating the consensus of local dual variables {λj}\{\lambda_{j}\}. The computed v-SGNE of this problem is illustrated in Fig. 2, where we use five different colors to represent the different players/firms. The top panel includes a geographic illustration, with the locations of the factories denoted by the colored letters and the total quantities transported on the roads illustrated by the brightness of the edges. The bottom panel shows the commodity contributions from the players at each market on this transport network.

Refer to caption
Figure 1: Performances of Alg. 1 in a Nash-Cournot Game
Refer to caption
Figure 2: The v-SGNE Obtained by Alg. 1

V-B Multi-Product Assembly Game with the Two-Stage Model

The two-stage stochastic programming problem originated from the work of [42] and found its applications in fields such as financial planning and control [43, Sec. 1.2], investment in power plants [43, Sec. 1.3], transportation planning during emergency response [44], etc. In this paper, we consider a multi-product assembly problem using the two-stage model [45, Sec. 1.3.1]. In a game network with NN manufacturers/players indexed by 𝒩{1,,N}\mathcal{N}\coloneqq\{1,\ldots,N\}, each player ii produces i\ell_{i} types of commodities. There are in total mm different subassemblies which have to be ordered from a third-party vendor. For each player ii, it needs nin_{i} different types of subassemblies in total, and a unit of commodity jj requires hi,(j,v)h_{i,(j,v)} units of subassembly vv, where j=1,,ij=1,\ldots,\ell_{i} and v=1,,niv=1,\ldots,n_{i}. The demands for player ii’s commodities are modeled as a random vector Di[Di,1;;Di,i]D_{i}\coloneqq[D_{i,1};\cdots;D_{i,\ell_{i}}], which has its range 𝒟i\mathcal{D}_{i} inside a bounded set in the positive orthant.

We start by formulating the second-stage problem. Let the numbers of subassemblies ordered by player ii be denoted by xi+nix_{i}\in\mathbb{R}^{n_{i}}_{+}, which is treated as a parameter in the second-stage problem. In this stage, player ii makes a production plan about the quantity of each commodity to produce based on the realized demand vector di+id_{i}\in\mathbb{R}^{\ell_{i}}_{+}. This production plan should maximize the profit and at the same time not exceed the quantities of available subassemblies. The income of player ii is comprised of the unit selling prices of the commodities piip_{i}\in\mathbb{R}^{\ell_{i}} and the unit salvage values of subassemblies that are not used sms\in\mathbb{R}^{m}. Denote the numbers of produced units by zi+iz_{i}\in\mathbb{R}^{\ell_{i}}_{+}, and the numbers of subassemblies left in inventory by yi+niy_{i}\in\mathbb{R}^{n_{i}}_{+}. We introduce the matrix Hii×niH_{i}\in\mathbb{R}^{\ell_{i}\times n_{i}} with each entry [Hi](j,v)=hi,(j,v)[H_{i}]_{(j,v)}=h_{i,(j,v)} and a binary matrix Aim×niA_{i}\in\mathbb{R}^{m\times n_{i}} mapping each entry of yiy_{i} to one among the mm subassemblies. In addition, assume the full-row-rank matrix HiH_{i} has ini\ell_{i}\leq n_{i} and no column sums to zero. Then we can define the nonsmooth function 𝒬i(xi;di)=min{piTzisTAiyiyi=xiHiTzi,𝟎zidi,yi𝟎}\mathcal{Q}_{i}(x_{i};d_{i})=\min\{-p_{i}^{T}z_{i}-s^{T}A_{i}y_{i}\mid y_{i}=x_{i}-H_{i}^{T}z_{i},\boldsymbol{0}\leq z_{i}\leq d_{i},y_{i}\geq\boldsymbol{0}\}, the minimizer of which is the best production plan.

With 𝒬i(xi;di)\mathcal{Q}_{i}(x_{i};d_{i}) defined, we can then formulate the first-stage problem. The price of subassembly vv per unit consists of the base cost CvC_{v} which is a random variable and the additional cost with the increasing ratio [Σ](ν,ν)[\Sigma]_{(\nu,\nu)} per ordered unit. At this stage, when making decisions about the pre-order quantities xix_{i} to maximize the profit, each player ii is uncertain about the base prices of subassemblies and the demands for its commodities. Each player ii has an expected-value objective w.r.t. the random vectors C[Cν]ν=1,,mC\coloneqq[C_{\nu}]_{\nu=1,\ldots,m} and DiD_{i}. Moreover, their decisions should collectively satisfy the global constraints concerning the available subassemblies. Altogether, the first-stage problem for each player ii can be expressed as:

{minimizexi𝒳i𝔼[12xiTQixi+(C+ΣAx)TAixi+𝒬i(xi;Di)]subjecttoAixicj𝒩iAjxj,\begin{cases}\underset{x_{i}\in\mathcal{X}_{i}}{\operatorname{minimize}}&\mathbb{E}[\frac{1}{2}x_{i}^{T}Q_{i}x_{i}+(C+\Sigma Ax)^{T}A_{i}x_{i}+\mathcal{Q}_{i}(x_{i};D_{i})]\\ \operatorname{subject\>to}&A_{i}x_{i}\leq c-\sum_{j\in\mathcal{N}_{-i}}A_{j}x_{j},\end{cases} (22)

where A[A1,,AN]A\coloneqq[A_{1},\ldots,A_{N}], x[x1;;xN]x\coloneqq[x_{1};\ldots;x_{N}], 𝒳i\mathcal{X}_{i} is the local feasible set of the decision vector xix_{i} which is compact and convex, QiQ_{i} and Σ\Sigma are diagonal matrices with each diagonal entry positive, and the constant vector cmc\in\mathbb{R}^{m} denotes the quantities of available subassemblies.

Suppose N=5N=5 players participate in this game to compete for m=10m=10 types of subassemblies. The decision vector of each player ii has dimension nin_{i} chosen uniformly at random from {7,8,9,10}\{7,8,9,10\}. The local feasible set 𝒳i\mathcal{X}_{i} is the direct product of nin_{i} connected compact intervals. The communication graph consists of a directed circle and two randomly selected edges.

V-B1 Assumptions Verification

We claim that the function 𝒬i(xi;di)\mathcal{Q}_{i}(x_{i};d_{i}) is a piecewise linear function in xi𝒳ix_{i}\in\mathcal{X}_{i} given any fixed di𝒟id_{i}\in\mathcal{D}_{i}, where 𝒟i\mathcal{D}_{i} and 𝒳i\mathcal{X}_{i} are both bounded. We first introduce the residual variable ri=dizir_{i}=d_{i}-z_{i} and convert the inequality constraints in 𝒬i(xi;di)\mathcal{Q}_{i}(x_{i};d_{i}) to equality ones as follows:

{minimizezi𝟎,hi𝟎,ri𝟎piTzisTAihisubjecttohi=xiHiTzi,zi+ri=di.\begin{cases}\operatorname{minimize}_{z_{i}\geq\boldsymbol{0},h_{i}\geq\boldsymbol{0},r_{i}\geq\boldsymbol{0}}\;-p_{i}^{T}z_{i}-s^{T}A_{i}h_{i}\\ \operatorname{subject\>to}\;h_{i}=x_{i}-H_{i}^{T}z_{i},z_{i}+r_{i}=d_{i}.\end{cases} (23)

By letting Bi=[HiTIni𝟎ni×iIi𝟎i×niIi]B_{i}=\Big{[}\begin{smallmatrix}H_{i}^{T}&I_{n_{i}}&\boldsymbol{0}_{n_{i}\times\ell_{i}}\\ I_{\ell_{i}}&\boldsymbol{0}_{\ell_{i}\times n_{i}}&I_{\ell_{i}}\end{smallmatrix}\Big{]}, ui[zi;hi;ri]u_{i}\coloneqq[z_{i};h_{i};r_{i}], qi[pi;AiTs;𝟎i]q_{i}\coloneqq[-p_{i};-A_{i}^{T}s;\boldsymbol{0}_{\ell_{i}}], I~i=[Ini;𝟎i×ni]\tilde{I}_{i}=[I_{n_{i}};\boldsymbol{0}_{\ell_{i}\times n_{i}}], and d~i=[𝟎ni;di]\tilde{d}_{i}=[\boldsymbol{0}_{n_{i}};d_{i}], the above constrained linear programming can be presented as: minimizeuiqiTui\operatorname{minimize}_{u_{i}}q_{i}^{T}u_{i}, while subjecttoBiui=I~ixi+d~i\operatorname{subject\>to}B_{i}u_{i}=\tilde{I}_{i}x_{i}+\tilde{d}_{i} and ui𝟎u_{i}\geq\boldsymbol{0}. Its dual problem can then be derived as:

maximizevi(I~ixi+d~i)Tvi,subjecttoBiTviqi.\operatorname{maximize}_{v_{i}}(\tilde{I}_{i}x_{i}+\tilde{d}_{i})^{T}v_{i},\;\operatorname{subject\>to}B_{i}^{T}v_{i}\leq q_{i}. (24)

We progress with the dual problem which only has xix_{i} as the coefficients of the objective function. Since the feasible set 𝒳i\mathcal{X}_{i} is compact inside the non-negative orthant, the simplex method will identify a vertex solution to the problem (24), even though the problem may admit unbounded solutions. Note that the polyhedral 𝒫i{vini+iBiTviqi}\mathcal{P}_{i}\coloneqq\{v_{i}\in\mathbb{R}^{n_{i}+\ell_{i}}\mid B_{i}^{T}v_{i}\leq q_{i}\} only admits a finite number of vertices 𝒱i{V1,V2,,VM}\mathcal{V}_{i}\coloneqq\{V_{1},V_{2},\ldots,V_{M}\} (-\infty excluded). Thus, 𝒬i(xi;di)maxVj𝒱iVjT[xi;di]\mathcal{Q}_{i}(x_{i};d_{i})\coloneqq\max_{V_{j}\in\mathcal{V}_{i}}V_{j}^{T}\cdot[x_{i};d_{i}], which completes the proof that 𝒬i(xi;di)\mathcal{Q}_{i}(x_{i};d_{i}) is a piecewise linear function in xix_{i}. It follows that the expected value function 𝔼Di[𝒬i(xi;Di)]\mathbb{E}_{D_{i}}[\mathcal{Q}_{i}(x_{i};D_{i})] is a convex function in xix_{i} [46, Sec. 3.2.1]. Applying the arguments in [32, Sec. V] to the remaining parts of 𝕁i(xi;xi)\mathbb{J}_{i}(x_{i};x_{-i}), we can show that the pseudogradient 𝔽\mathbb{F} is strongly monotone. By [29, Prop. 12.11], this multi-product assembly problem admits a unique Nash equilibrium. It can also be checked numerically that there exists a ρμ>0\rho_{\mu}>0 such that the operator T𝔽~+ρμ2Ln\mathcal{R}^{T}\operatorname{\tilde{\mathbb{F}}}+\frac{\rho_{\mu}}{2}L_{n} is maximally monotone. These arguments guarantee that Assumptions 1, 3 and 5 hold for this SGNEP.

To guarantee that Assumption 7 holds, it suffices to verify that the nonsmooth parts of the objectives fulfill these conditions. We can establish the interchangeability of subdifferential and integral using [45, Thm. 7.52]. We then consider the function ϕi(χ)maxvi𝒫i(viTχ)\phi_{i}(\chi)\coloneqq\max_{v_{i}\in\mathcal{P}_{i}}(v_{i}^{T}\cdot\chi), where 𝒫i{vini+iBiTviqi}\mathcal{P}_{i}\coloneqq\{v_{i}\in\mathbb{R}^{n_{i}+\ell_{i}}\mid B_{i}^{T}v_{i}\leq q_{i}\}. Since the set 𝒫i\mathcal{P}_{i} is nonempty, ϕi(χ)\phi_{i}(\chi) is the support function of 𝒫i\mathcal{P}_{i}. By definition, the support function ϕi(χ)\phi_{i}(\chi) is the conjugate function of the indicator function ι𝒫i(χ)\iota_{\mathcal{P}_{i}}(\chi), i.e., ϕi(χ)=maxvi𝒫i(viTχ)=maxvi(viTχι𝒫i(vi))\phi_{i}(\chi)=\max_{v_{i}\in\mathcal{P}_{i}}(v_{i}^{T}\cdot\chi)=\max_{v_{i}}(v_{i}^{T}\cdot\chi-\iota_{\mathcal{P}_{i}}(v_{i})). Since the set 𝒫i\mathcal{P}_{i} is convex and closed, the function ι𝒫i(χ)\iota_{\mathcal{P}_{i}}(\chi) is convex, lower semicontinuous and proper. By [45, Thm. 7.5 and (7.24)], we obtain ϕi(χ)=argmaxvi{viTχι𝒫i(vi)}=argmaxvi𝒫i{viTχ}\partial\phi_{i}(\chi)=\operatorname{argmax}_{v_{i}}\{v_{i}^{T}\cdot\chi-\iota_{\mathcal{P}_{i}}(v_{i})\}=\operatorname{argmax}_{v_{i}\in\mathcal{P}_{i}}\{v_{i}^{T}\cdot\chi\}. Moreover, by the chain rule, the subdifferential should be 𝒬i(xi;di)=I~iTargmaxvi𝒫i{(I~ixi+d~i)Tvi}\partial\mathcal{Q}_{i}(x_{i};d_{i})=\tilde{I}^{T}_{i}\cdot\operatorname{argmax}_{v_{i}\in\mathcal{P}_{i}}\{(\tilde{I}_{i}x_{i}+\tilde{d}_{i})^{T}\cdot v_{i}\}. As we discussed in the verification of Assumption 1, the solution set of argmaxvi𝒫i{(I~ixi+d~i)Tvi}\operatorname{argmax}_{v_{i}\in\mathcal{P}_{i}}\{(\tilde{I}_{i}x_{i}+\tilde{d}_{i})^{T}\cdot v_{i}\} must contain at least one of 𝒫i\mathcal{P}_{i}’s vertices. Hence, we can always find a bounded subgradient of 𝒬i\mathcal{Q}_{i} such that Assumption 7 (ii) holds.

V-B2 Simulation Results

We restrict each random variable DiD_{i} to having a finite range {d1,,dL}\{d_{1},\ldots,d_{L}\} with the probability distribution {P1,,PL}\{P_{1},\ldots,P_{L}\}. Under this restriction, the objective function of each player ii can be explicitly written as: 𝕁i(xi;xi)=12xiTQixi+(𝔼[C]+ΣAx)TAixi+l=1LPl𝒬i(xi;dl)\mathbb{J}_{i}(x_{i};x_{-i})=\frac{1}{2}x_{i}^{T}Q_{i}x_{i}+(\mathbb{E}[C]+\Sigma Ax)^{T}A_{i}x_{i}+\sum_{l=1}^{L}P_{l}\mathcal{Q}_{i}(x_{i};d_{l}). The method proposed in [32] can then be applied to compute the unique v-SGNE for reference. The performance of Algorithm 1 when solving this multi-product assembly problem is illustrated in Fig. 3. The thin lines reflect the simple moving averages of these metrics with a window size of 2020. The curves of T(k)k2.1T^{(k)}\propto k^{2.1} illustrate a steady convergence towards the v-GNE as suggested in Theorem 3, while the trajectories of T(k)=20T^{(k)}=20 stop decreasing after some iterations. The curves of T(k)kT^{(k)}\propto k also keep descending yet with a gentler trend compared with those of T(k)k2.1T^{(k)}\propto k^{2.1}, which suggests the possibility of some relaxations to the current conditions in Theorems 2 and 3. For the detailed figure descriptions, please refer to Sec. V-A2.

Refer to caption
Figure 3: Performances of Alg. 1 for the Two-Stage Model

VI Conclusion and Future Directions

In this paper, we study the stochastic generalized Nash equilibrium problem and propose a distributed stochastic algorithm under the partial-decision information setting based on solving augmented best-response subproblems induced by the Douglas-Rachford scheme. The proposed algorithm is proved to converge to a true variational stochastic generalized Nash equilibrium if the sequence of inertial step sizes and the inverse of the number of realizations per major iteration decrease altogether at a proper rate. This raises the question if there exists a less conservative bound for this decreasing rate such that the proposed algorithm can still converge yet with a faster convergence rate and fewer observations needed per major iteration. Another interesting work remains concerning the convergence rate analysis of the proposed algorithm. As we have previously mentioned, the fixed point iteration discussed in this paper engages two reflected resolvent operators, which merely admit quasinonexpansiveness rather than contractiveness. The convergence rate analysis under this setting remains an under-explored yet increasingly active direction [47, 48, 49]. Finally, although we only analyze the projected stochastic subgradient method, the main convergence result in Theorem 2 actually allows a lot of possibilities. It would be interesting to develop inexact solvers based on different stochastic optimization schemes, e.g. proximal point methods [50], that are more sample-efficient or further relax the assumptions made.

Appendix

A Proof of Theorem 2

Proof.

The following proof is largely inspired from that of [33, Prop. 5.34] for deterministic sequences and nonexpansive operators case with suitable modifications. Given an arbitrary initial point ψ~(0)=ψ~(0)\tilde{\psi}^{(0)}=\tilde{\psi}^{(0)}_{*}, we let (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} denote the sequence generated by the approximate iteration (18). Another auxiliary sequence (ψ~(k))k(\tilde{\psi}^{(k)}_{*})_{k\in\mathbb{N}} is constructed by letting ψ~(k+1)𝒫(ψ~(k))\tilde{\psi}^{(k+1)}_{*}\coloneqq\mathscr{P}_{*}(\tilde{\psi}^{(k)}). We next try to extract a recursive relationship w.r.t. ψ~(k+1)ψ~𝒦2\lVert\tilde{\psi}^{(k+1)}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}} to establish that kres(ψ~(k))<\sum_{k\in\mathbb{N}}\text{res}(\tilde{\psi}^{(k)})<\infty a.s., where ψ~\tilde{\psi}^{*} is a fixed point of \mathscr{R}_{*}. Writing the explicit updating formula of ψ~(k+1)ψ~𝒦2\lVert\tilde{\psi}^{(k+1)}_{*}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}} yields

ψ~(k+1)ψ~𝒦2=(1γ(k))ψ~(k)+γ(k)(ψ~(k))ψ~𝒦2\displaystyle\lVert{\tilde{\psi}}^{(k+1)}_{*}-{\tilde{\psi}}^{*}\rVert^{2}_{\mathcal{K}}=\lVert(1-{\gamma}^{(k)}){\tilde{\psi}}^{(k)}+{\gamma}^{(k)}\mathscr{R}_{*}({\tilde{\psi}}^{(k)})-{\tilde{\psi}}^{*}\rVert^{2}_{\mathcal{K}}
=(1γ(k))ψ~(k)ψ~𝒦2+γ(k)(ψ~(k))(ψ~)𝒦2\displaystyle=(1-{\gamma}^{(k)})\lVert{\tilde{\psi}}^{(k)}-{\tilde{\psi}}^{*}\rVert^{2}_{\mathcal{K}}+{\gamma}^{(k)}\lVert\mathscr{R}_{*}({\tilde{\psi}}^{(k)})-\mathscr{R}_{*}({\tilde{\psi}}^{*})\rVert^{2}_{\mathcal{K}}
γ(k)(1γ(k))(ψ~(k))ψ~(k)𝒦2\displaystyle\qquad-{\gamma}^{(k)}(1-{\gamma}^{(k)})\lVert\mathscr{R}_{*}({\tilde{\psi}}^{(k)})-{\tilde{\psi}}^{(k)}\rVert^{2}_{\mathcal{K}}
ψ~(k)ψ~𝒦2γ(k)(1γ(k))(res(ψ~(k)))2.\displaystyle\leq\lVert{\tilde{\psi}}^{(k)}-{\tilde{\psi}}^{*}\rVert^{2}_{\mathcal{K}}-{\gamma}^{(k)}(1-{\gamma}^{(k)})(\text{res}({\tilde{\psi}}^{(k)}))^{2}.

where the inequality follows from the fact that \mathscr{R}_{*} is quasinonexpansive. Next, we derive a recursive relationship for ψ~(k+1)ψ~𝒦2\lVert\tilde{\psi}^{(k+1)}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}} as follows:

ψ~(k+1)ψ~𝒦2=γ(k)ϵ(k)+ψ~(k+1)ψ~𝒦2\displaystyle\lVert\tilde{\psi}^{(k+1)}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}}=\lVert\gamma^{(k)}\epsilon^{(k)}+\tilde{\psi}^{(k+1)}_{*}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}}
=ψ~(k+1)ψ~𝒦2+(γ(k)ε(k))2+2γ(k)ϵ(k),ψ~(k+1)ψ~𝒦\displaystyle=\lVert\tilde{\psi}^{(k+1)}_{*}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}}+(\gamma^{(k)}\varepsilon^{(k)})^{2}+2\langle\gamma^{(k)}\epsilon^{(k)},\tilde{\psi}^{(k+1)}_{*}-\tilde{\psi}^{*}\rangle_{\mathcal{K}}
ψ~(k)ψ~𝒦2γ(k)(1γ(k))(res(ψ~(k)))2\displaystyle\leq\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}}-\gamma^{(k)}(1-\gamma^{(k)})(\text{res}(\tilde{\psi}^{(k)}))^{2}
+(γ(k)ε(k))2+2γ(k)ε(k)ψ~(k)ψ~𝒦,\displaystyle\qquad+(\gamma^{(k)}\varepsilon^{(k)})^{2}+2\gamma^{(k)}\varepsilon^{(k)}\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}},

where the last inequality follows from the relation derived above and the Cauchy-Schwarz inequality. Taking conditional expectation 𝔼[(k)]\mathbb{E}[\cdot\mid\mathcal{F}^{(k)}] on both sides yields:

𝔼[ψ~(k+1)ψ~𝒦2(k)]ψ~(k)ψ~𝒦2γ(k)(1γ(k))(res(ψ~(k)))2+𝔼[2γ(k)ε(k)ψ~(k)ψ~𝒦+(γ(k)ε(k))2(k)].\begin{split}&\mathbb{E}[\lVert\tilde{\psi}^{(k+1)}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}}\mid\mathcal{F}^{(k)}]\\ &\leq\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert^{2}_{\mathcal{K}}-\gamma^{(k)}(1-\gamma^{(k)})(\text{res}(\tilde{\psi}^{(k)}))^{2}\\ &\quad+\mathbb{E}[2\gamma^{(k)}\varepsilon^{(k)}\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}}+(\gamma^{(k)}\varepsilon^{(k)})^{2}\mid\mathcal{F}^{(k)}].\end{split} (25)

Based on the fact that kγ(k)𝔼[ε(k)(k)]<+\sum_{k\in\mathbb{N}}\gamma^{(k)}\mathbb{E}[\varepsilon^{(k)}\mid\mathcal{F}^{(k)}]<+\infty a.s. and (ψ~(k)ψ~𝒦)k(\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}})_{k\in\mathbb{N}} is bounded a.s., we can obtain that k𝔼[2γ(k)ε(k)ψ~(k)ψ~𝒦+(γ(k)ε(k))2(k)]<\sum_{k\in\mathbb{N}}\mathbb{E}[2\gamma^{(k)}\varepsilon^{(k)}\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}}+(\gamma^{(k)}\varepsilon^{(k)})^{2}\mid\mathcal{F}^{(k)}]<\infty a.s. By applying the Robbins-Siegmund theorem, we can then conclude that on a set Ω^\hat{\Omega} which has probability one, kγ(k)(1γ(k))res(ψ~(k))2<\sum_{k\in\mathbb{N}}\gamma^{(k)}(1-\gamma^{(k)})\text{res}(\tilde{\psi}^{(k)})^{2}<\infty with γ(k)[0,1]\gamma^{(k)}\in[0,1] and kγ(k)(1γ(k))=+\sum_{k\in\mathbb{N}}\gamma^{(k)}(1-\gamma^{(k)})=+\infty. Now we fix an arbitrary sample path ω^Ω^\hat{\omega}\in\hat{\Omega} for subsequent analysis, while omitting ω^\hat{\omega} for brevity. In the following we will prove by contradiction that lim infkres(ψ~(k))2=0\liminf_{k\to\infty}\text{res}(\tilde{\psi}^{(k)})^{2}=0 Suppose otherwise that lim infkres(ψ~(k))2=δ\liminf_{k\to\infty}\text{res}(\tilde{\psi}^{(k)})^{2}=\delta, where δ>0\delta>0 is some positive constant. Then there exists a sufficiently large KδK_{\delta} such that for any k>Kδk>K_{\delta}, res(ψ~(k))2>δ/2\text{res}(\tilde{\psi}^{(k)})^{2}>\delta/2. By this lower bound, we have k>Kδγ(k)(1γ(k))res(ψ~(k))2>δ/2k>Kδγ(k)(1γ(k))=+\sum_{k>K_{\delta}}\gamma^{(k)}(1-\gamma^{(k)})\text{res}(\tilde{\psi}^{(k)})^{2}>\delta/2\sum_{k>K_{\delta}}\gamma^{(k)}(1-\gamma^{(k)})=+\infty, which contradicts the previous statement that kγ(k)(1γ(k))res(ψ~(k))2<\sum_{k\in\mathbb{N}}\gamma^{(k)}(1-\gamma^{(k)})\text{res}(\tilde{\psi}^{(k)})^{2}<\infty. This shows that lim infkres(ψ~(k))2=0\liminf_{k\to\infty}\text{res}(\tilde{\psi}^{(k)})^{2}=0. As a result, there exists a subsequence, denoted by (ψ~(ki))i(\tilde{\psi}^{(k_{i})})_{i\in\mathbb{N}}, such that limires(ψ~(ki))2=0\lim_{i\to\infty}\text{res}(\tilde{\psi}^{(k_{i})})^{2}=0.

Moreover, the above subsequence (ψ~(ki))ki(\tilde{\psi}^{(k_{i})})_{k_{i}\in\mathbb{N}} is bounded and thus has a convergent subsequence (ψ~(li))i(\tilde{\psi}^{(l_{i})})_{i\in\mathbb{N}} where (li)i(ki)i(l_{i})_{i\in\mathbb{N}}\subseteq(k_{i})_{i\in\mathbb{N}} such that limiψ~(li)=ψ~\lim_{i\to\infty}\tilde{\psi}^{(l_{i})}=\tilde{\psi}^{\dagger}. If Assumption 5(i) holds, by definition, \mathscr{R}_{*} is a nonexpansive mapping. It then follows from [33, Cor. 4.28] that ψ~Fix()\tilde{\psi}^{\dagger}\in\text{Fix}(\mathscr{R}_{*}). If Assumption 5(ii) holds instead, from [32, Lemma 6], \mathscr{R}_{*} is a continuous mapping, i.e., limires(ψ~(li))=0\lim_{i\to\infty}\text{res}(\tilde{\psi}^{(l_{i})})=0 implies (ψ~)=ψ~\mathscr{R}_{*}(\tilde{\psi}^{\dagger})=\tilde{\psi}^{\dagger} and hence ψ~Fix()\tilde{\psi}^{\dagger}\in\text{Fix}(\mathscr{R}_{*}). Therefore we can substitute ψ~\tilde{\psi}^{*} in (25) with ψ~\tilde{\psi}^{\dagger}. By [35, Thm. 1], limkψ~(k)ψ~𝒦2\lim_{k\to\infty}\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{\dagger}\rVert^{2}_{\mathcal{K}} exists. Since (ψ~(li))i(\tilde{\psi}^{(l_{i})})_{i\in\mathbb{N}} is a subsubsequence of (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} converging to the fixed point ψ~\tilde{\psi}^{\dagger}, we can conclude that limkψ~(k)ψ~𝒦2=0\lim_{k\to\infty}\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{\dagger}\rVert^{2}_{\mathcal{K}}=0, and hence limkψ~(k)=ψ~\lim_{k\to\infty}\tilde{\psi}^{(k)}=\tilde{\psi}^{\dagger}. Altogether, ψJ𝔸¯(ψ~)\psi^{\dagger}\coloneqq J_{\operatorname{\bar{\mathbb{A}}}}(\tilde{\psi}^{\dagger}) belongs to the zero set of 𝕋\operatorname{\mathbb{T}} in (11). Combining this with the conclusions of Theorem 1, the proof is complete. ∎

B Proof of Lemma 1

Proof.

For each player i𝒩i\in\mathcal{N}, at an arbitrary major iteration kk and its minor iteration t=0,,Ti(k)1t=0,\ldots,T^{(k)}_{i}-1, by applying the update inside the for-loop of Algorithm 2 and using the nonexpansiveness of the projection operator onto a convex set, we can obtain the following inequality of the distance between the approximate minimizer after the ttth minor iteration yi,t+1i(k+1)y^{i(k+1)}_{i,t+1} and the accurate minimizer yi,i(k+1)y^{i(k+1)}_{i,*}:

yi,t+1i(k+1)yi,i(k+1)22yi,ti(k+1)κi,tgi,t(k)yi,i(k+1)22.\lVert y^{i(k+1)}_{i,t+1}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\leq\lVert y^{i(k+1)}_{i,t}-\kappa_{i,t}\cdot g^{(k)}_{i,t}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}. (26)

Expanding the 2\ell^{2} norm and taking conditional expectation 𝔼[σ{k,ξi,[t](k)}]\mathbb{E}\big{[}\cdot\mid\sigma\{\mathcal{F}_{k},\xi^{(k)}_{i,[t]}\}\big{]} on both sides of (26) yields:

𝔼[yi,t+1i(k+1)yi,i(k+1)22σ{k,ξi,[t](k)}]κi,t2𝔼[gi,t(k)22σ{k,ξi,[t](k)}]+yi,ti(k+1)yi,i(k+1)222κi,tyi,ti(k+1)yi,i(k+1),𝕘i,t(k),\displaystyle\begin{split}&\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i,t+1}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\sigma\{\mathcal{F}_{k},\xi^{(k)}_{i,[t]}\}\big{]}\\ &\leq\kappa_{i,t}^{2}\mathbb{E}\big{[}\lVert g^{(k)}_{i,t}\rVert^{2}_{2}\mid\sigma\{\mathcal{F}_{k},\xi^{(k)}_{i,[t]}\}\big{]}+\lVert y^{i(k+1)}_{i,t}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\\ &\qquad-2\kappa_{i,t}\langle y^{i(k+1)}_{i,t}-y^{i(k+1)}_{i,*},\mathbbm{g}^{(k)}_{i,t}\rangle,\\ \end{split} (27)

where 𝕘i,t(k)𝔼[gi,t(k)σ{k,ξi,[t](k+1)}]yii𝕁^i(k)(yi,ti(k+1))\mathbbm{g}^{(k)}_{i,t}\coloneqq\mathbb{E}[g^{(k)}_{i,t}\mid\sigma\{\mathcal{F}_{k},\xi^{(k+1)}_{i,[t]}\}]\in\partial_{y^{i}_{i}}\hat{\mathbb{J}}^{(k)}_{i}(y^{i(k+1)}_{i,t}) by Assumption 7. Using the 1τ1i\frac{1}{\tau_{1i}}-strong convexity of 𝕁^i(k)\hat{\mathbb{J}}^{(k)}_{i}, the inner product inside the inequality (27) satisfies yi,ti(k+1)yi,i(k+1),𝕘i,t(k)𝕁^i(k)(yi,ti(k+1))𝕁^i(k)(yi,i(k+1))+12τ1iyi,ti(k+1)yi,i(k+1)22\langle y^{i(k+1)}_{i,t}-y^{i(k+1)}_{i,*},\mathbbm{g}^{(k)}_{i,t}\rangle\geq\hat{\mathbb{J}}^{(k)}_{i}(y^{i(k+1)}_{i,t})-\hat{\mathbb{J}}^{(k)}_{i}(y^{i(k+1)}_{i,*})+\frac{1}{2\tau_{1i}}\lVert y^{i(k+1)}_{i,t}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}. We then take conditional expectations 𝔼[k]\mathbb{E}[\cdot\mid\mathcal{F}_{k}] on both sides of the above inequality. By the rule of successive conditioning and the fact that yi,i(k+1)y^{i(k+1)}_{i,*} minimizes 𝕁^i(k)()\hat{\mathbb{J}}^{(k)}_{i}(\cdot), the following inequality holds a.s.:

𝔼[yi,t+1i(k+1)yi,i(k+1)22k]κi,t2𝔼[gi,t(k)22k]+(1κi,tτ1i)𝔼[yi,ti(k+1)yi,i(k+1)22k].\begin{split}&\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i,t+1}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}\leq\kappa_{i,t}^{2}\mathbb{E}\big{[}\lVert g^{(k)}_{i,t}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}\\ &\quad+(1-\frac{\kappa_{i,t}}{\tau_{1i}})\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i,t}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}.\end{split} (28)

Re-arranging (28) and applying Assumption 7(ii), we have that the following holds a.s.:

1κi,t𝔼[yi,t+1i(k+1)yi,i(k+1)22k](1κi,t1τ1i)𝔼[yi,ti(k+1)yi,i(k+1)22k]κi,t(αg,i2ψ~(k)22+βg,i2).\begin{split}&\frac{1}{\kappa_{i,t}}\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i,t+1}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}-(\frac{1}{\kappa_{i,t}}-\frac{1}{\tau_{1i}})\cdot\\ &\quad\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i,t}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}\leq\kappa_{i,t}(\alpha_{g,i}^{2}\lVert\tilde{\psi}^{(k)}\rVert^{2}_{2}+\beta^{2}_{g,i}).\end{split} (29)

By setting κi,t2τ1it+2\kappa_{i,t}\coloneqq\frac{2\tau_{1i}}{t+2}, multiplying both sides of (29) by (t+1)/2(t+1)/2, and summing (29) for t=0,,T1t=0,\ldots,T-1, for an arbitrary T{1,,Ti(k)}T\in\{1,\ldots,T^{(k)}_{i}\}, we obtain a telescoping sum and have that the following holds a.s.:

(T+1)T4τ1i𝔼[yi,Ti(k+1)yi,i(k+1)22k]t=0T1t+122τ1it+2(αg,i2ψ~(k)22+βg,i2).\begin{split}&\frac{(T+1)T}{4\tau_{1i}}\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i,T}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}\\ &\qquad\qquad\leq\sum_{t=0}^{T-1}\frac{t+1}{2}\cdot\frac{2\tau_{1i}}{t+2}(\alpha_{g,i}^{2}\lVert\tilde{\psi}^{(k)}\rVert^{2}_{2}+\beta^{2}_{g,i}).\end{split} (30)

Simplifying (30), we deduce that 𝔼[yi,Ti(k+1)yi,i(k+1)22k]4τ1i2T1(αg,i2ψ~(k)22+βg,i2)\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i,T}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}\leq 4\tau_{1i}^{2}T^{-1}(\alpha^{2}_{g,i}\lVert\tilde{\psi}^{(k)}\rVert^{2}_{2}+\beta^{2}_{g,i}) a.s. ∎

C Proof of Lemma 2

Proof.

By the nonexpansiveness of the reflected resolvent R¯R_{\operatorname{\bar{\mathcal{B}}}}, the approximate error ε(k)\varepsilon^{(k)} should satisfy:

𝔼[ε(k)k]𝔼[R𝒜¯(ψ~(k))R𝔸¯(ψ~(k))𝒦k]=2𝔼[ψ(k+1)ψ(k+1)𝒦k],\begin{split}\mathbb{E}[\varepsilon^{(k)}\mid\mathcal{F}_{k}]&\leq\mathbb{E}\big{[}\lVert R_{\operatorname{\bar{\mathcal{A}}}}(\tilde{\psi}^{(k)})-R_{\operatorname{\bar{\mathbb{A}}}}(\tilde{\psi}^{(k)})\rVert_{\mathcal{K}}\mid\mathcal{F}_{k}\big{]}\\ &=2\mathbb{E}\big{[}\lVert\psi^{(k+1)}-\psi^{(k+1)}_{*}\rVert_{\mathcal{K}}\mid\mathcal{F}_{k}\big{]},\end{split} (31)

where ψ(k+1)=[𝒚(k+1);𝝀(k+1);𝝁(k+1);𝒛(k+1)]J𝒜¯(ψ~(k))\psi^{(k+1)}=[\boldsymbol{y}^{(k+1)};\boldsymbol{\lambda}^{(k+1)};\boldsymbol{\mu}^{(k+1)};\boldsymbol{z}^{(k+1)}]\coloneqq J_{\operatorname{\bar{\mathcal{A}}}}(\tilde{\psi}^{(k)}) is the stack vector obtained by using the inexact solver suggested in Algorithm 2 and ψ(k+1)=[𝒚(k+1);𝝀(k+1);𝝁(k+1);𝒛(k+1)]J𝔸¯(ψ~(k))\psi^{(k+1)}_{*}=[\boldsymbol{y}^{(k+1)}_{*};\boldsymbol{\lambda}^{(k+1)}_{*};\boldsymbol{\mu}^{(k+1)}_{*};\boldsymbol{z}^{(k+1)}_{*}]\coloneqq J_{\operatorname{\bar{\mathbb{A}}}}(\tilde{\psi}^{(k)}) is the one using the accurate solver. Given the conclusion of Lemma 1 and the first two for-loops in Algorithm 1, the approximate error of the dual variables 𝝀\boldsymbol{\lambda} has the following upper bound:

𝔼[𝝀(k+1)𝝀(k+1)2k]𝔼[τ2Λ(𝒚(k+1)𝒚(k+1))2k]\displaystyle\mathbb{E}\big{[}\lVert\boldsymbol{\lambda}^{(k+1)}-\boldsymbol{\lambda}^{(k+1)}_{*}\rVert_{2}\mid\mathcal{F}_{k}\big{]}\leq\mathbb{E}\big{[}\lVert\tau_{2}\Lambda\mathcal{R}(\boldsymbol{y}^{(k+1)}-\boldsymbol{y}^{(k+1)}_{*})\rVert_{2}\mid\mathcal{F}_{k}\big{]}
τ2Λ2𝔼[𝒚(k+1)𝒚(k+1)2k].\displaystyle\qquad\leq\lVert\tau_{2}\Lambda\mathcal{R}\rVert_{2}\cdot\mathbb{E}\big{[}\lVert\boldsymbol{y}^{(k+1)}-\boldsymbol{y}^{(k+1)}_{*}\rVert_{2}\mid\mathcal{F}_{k}\big{]}.

Similar results can be trivially derived for 𝝁\boldsymbol{\mu} and 𝒛\boldsymbol{z}, the details of which are omitted for brevity. Altogether, we have that the following relation 𝔼[ψ(k+1)ψ(k+1)2k]C1𝔼[𝒚(k+1)𝒚(k+1)2k]\mathbb{E}\big{[}\lVert\psi^{(k+1)}-\psi^{(k+1)}_{*}\rVert_{2}\mid\mathcal{F}_{k}\big{]}\leq C_{1}\cdot\mathbb{E}\big{[}\lVert\boldsymbol{y}^{(k+1)}-\boldsymbol{y}^{(k+1)}_{*}\rVert_{2}\mid\mathcal{F}_{k}\big{]} holds for some constant C1C_{1}. For each player i𝒩i\in\mathcal{N}, the local estimates of others’ decisions are the same in yi(k+1)y^{i(k+1)} and yi(k+1)y^{i(k+1)}_{*}, while the local decisions, by Lemma 1, satisfy 𝔼[yii(k+1)yi,i(k+1)22k]4τ1i2(Ti(k))1(αg,i2ψ~(k)22+βg,i2)\mathbb{E}\big{[}\lVert y^{i(k+1)}_{i}-y^{i(k+1)}_{i,*}\rVert^{2}_{2}\mid\mathcal{F}_{k}\big{]}\leq 4\tau_{1i}^{2}(T^{(k)}_{i})^{-1}(\alpha^{2}_{g,i}\lVert\tilde{\psi}^{(k)}\rVert^{2}_{2}+\beta^{2}_{g,i}) a.s. for each i𝒩i\in\mathcal{N}. Picking the maximum coefficients α¯gmax{αg,i:i𝒩}\bar{\alpha}_{g}\coloneqq\max\{\alpha_{g,i}:i\in\mathcal{N}\}, β¯gmax{βg,i:i𝒩}\bar{\beta}_{g}\coloneqq\max\{\beta_{g,i}:i\in\mathcal{N}\}, τ¯1max{τ1,i:i𝒩}\bar{\tau}_{1}\coloneqq\max\{\tau_{1,i}:i\in\mathcal{N}\} and the minimum minor steps taken \stackunder[1.2pt]T (k)min{Ti(k):i𝒩}\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)}\coloneqq\min\{T^{(k)}_{i}:i\in\mathcal{N}\} over all players. By Jensen’s inequality and the non-negativity of αg,i\alpha_{g,i}, βg,i\beta_{g,i}, and ψ~(k)\lVert\tilde{\psi}^{(k)}\rVert, an upper bound for the stacked local decisions and estimates is given by:

𝔼[𝒚(k+1)𝒚(k+1)2k](𝔼[i𝒩yii(k+1)yii(k+1)22k])1/22Nτ¯1(\stackunder[1.2pt]T (k))1/2(α¯gψ~(k)2+β¯g), a.s.\begin{split}&\mathbb{E}[\lVert\boldsymbol{y}^{(k+1)}-\boldsymbol{y}^{(k+1)}_{*}\rVert_{2}\mid\mathcal{F}_{k}]\leq(\mathbb{E}[\sum_{i\in\mathcal{N}}\lVert y^{i(k+1)}_{i}-y^{i(k+1)}_{i*}\rVert^{2}_{2}\mid\mathcal{F}_{k}])^{1/2}\\ &\qquad\leq 2\sqrt{N}\bar{\tau}_{1}(\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)})^{-1/2}(\bar{\alpha}_{g}\lVert\tilde{\psi}^{(k)}\rVert_{2}+\bar{\beta}_{g}),\text{ a.s.}\end{split}

Combining the above inequalities, we derive the following a.s. upper bound in the Euclidean space:

𝔼[ψ(k+1)ψ(k+1)2k]C2(\stackunder[1.2pt]T (k))1/2(α¯gψ~(k)2+β¯g),\mathbb{E}\big{[}\lVert\psi^{(k+1)}-\psi^{(k+1)}_{*}\rVert_{2}\mid\mathcal{F}_{k}\big{]}\leq C_{2}(\stackunder[1.2pt]{T}{\rule{3.44444pt}{0.32289pt}}^{(k)})^{-1/2}(\bar{\alpha}_{g}\lVert\tilde{\psi}^{(k)}\rVert_{2}+\bar{\beta}_{g}), (32)

where C22C1τ¯1NC_{2}\coloneqq 2C_{1}\bar{\tau}_{1}\sqrt{N}. We convert the above conclusion from the Euclidean space to the inner product space 𝒦\mathcal{K} defined by the positive definite design matrix Φ\Phi. The maximum (resp. minimum) eigenvalue of Φ\Phi is denoted by σ¯Φ\bar{\sigma}_{\Phi} (resp. \stackunder[1.2pt]σ Φ\stackunder[1.2pt]{$\sigma$}{\rule{3.44444pt}{0.32289pt}}_{\Phi}). Then (32) implies the following relation holds a.s. in 𝒦\mathcal{K}:

𝔼[ψ(k+1)ψ(k+1)𝒦k]C2σ¯Φ(\stackunder[1.2pt]T (k))1/2(α¯g\stackunder[1.2pt]σ Φψ~(k)𝒦+β¯g).\mathbb{E}\big{[}\lVert\psi^{(k+1)}-\psi^{(k+1)}_{*}\rVert_{\mathcal{K}}\mid\mathcal{F}_{k}\big{]}\leq\frac{C_{2}\sqrt{\bar{\sigma}_{\Phi}}}{(\stackunder[1.2pt]{T}{\rule{3.44444pt}{0.32289pt}}^{(k)})^{1/2}}(\frac{\bar{\alpha}_{g}}{\sqrt{\stackunder[1.2pt]{\sigma}{\rule{3.44444pt}{0.32289pt}}_{\Phi}}}\lVert\tilde{\psi}^{(k)}\rVert_{\mathcal{K}}+\bar{\beta}_{g}). (33)

Hence, there exist positive constants αψ\alpha_{\psi} and βψ\beta_{\psi} independent of kk such that 𝔼[ε(k)k](\stackunder[1.2pt]T (k))1/2(αψψ~(k)𝒦+βψ)\mathbb{E}[\varepsilon^{(k)}\mid\mathcal{F}_{k}]\leq(\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)})^{-1/2}(\alpha_{\psi}\lVert\tilde{\psi}^{(k)}\rVert_{\mathcal{K}}+\beta_{\psi}) a.s. ∎

D Proof of Theorem 3

Proof.

Consider a sequence of augmented vectors (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} generated by the approximate iteration 𝒫=Id+γ(k)(Id)\mathscr{P}=\text{Id}+\gamma^{(k)}(\mathscr{R}-\text{Id}) and a sequence (ψ~(k))k(\tilde{\psi}^{(k)}_{*})_{k\in\mathbb{N}} generated by ψ~(k+1)𝒫(ψ~(k))\tilde{\psi}^{(k+1)}_{*}\coloneqq\mathscr{P}_{*}(\tilde{\psi}^{(k)}). Let ψ~\tilde{\psi}^{*} denote one of the fixed points of \mathscr{R}_{*}. To prove that (ψ~(k))k(\tilde{\psi}^{(k)})_{k\in\mathbb{N}} is bounded a.s., note that

𝔼[ψ~(k+1)ψ~𝒦k]=𝔼[ψ~(k+1)ψ~(k+1)+ψ~(k+1)ψ~𝒦k]\displaystyle\mathbb{E}[\lVert\tilde{\psi}^{(k+1)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}}\mid\mathcal{F}_{k}]=\mathbb{E}[\lVert\tilde{\psi}^{(k+1)}-\tilde{\psi}^{(k+1)}_{*}+\tilde{\psi}^{(k+1)}_{*}-\tilde{\psi}^{*}\rVert_{\mathcal{K}}\mid\mathcal{F}_{k}]
γ(k)𝔼[ε(k)k]+𝔼[𝒫(ψ~(k))𝒫(ψ~)𝒦k].\displaystyle\leq\gamma^{(k)}\mathbb{E}[\varepsilon^{(k)}\mid\mathcal{F}_{k}]+\mathbb{E}[\lVert\mathscr{P}_{*}(\tilde{\psi}^{(k)})-\mathscr{P}_{*}(\tilde{\psi}^{*})\rVert_{\mathcal{K}}\mid\mathcal{F}_{k}].

Let γT(k)γ(k)(\stackunder[1.2pt]T (k))1/2\gamma^{(k)}_{T}\coloneqq\gamma^{(k)}(\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)})^{-1/2}. By applying Lemma 2 and using the fact that 𝒫\mathscr{P}_{*} is (quasi)nonexpansive, we have:

𝔼[ψ~(k+1)ψ~𝒦k]\displaystyle\mathbb{E}[\lVert\tilde{\psi}^{(k+1)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}}\mid\mathcal{F}_{k}]
γT(k)(αψψ~(k)𝒦+βψ)+𝔼[ψ~(k)ψ~𝒦k]\displaystyle\leq\gamma^{(k)}_{T}(\alpha_{\psi}\lVert\tilde{\psi}^{(k)}\rVert_{\mathcal{K}}+\beta_{\psi})+\mathbb{E}[\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}}\mid\mathcal{F}_{k}]
=γT(k)(αψψ~(k)ψ~+ψ~𝒦+βψ)+ψ~(k)ψ~𝒦\displaystyle=\gamma^{(k)}_{T}(\alpha_{\psi}\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}+\tilde{\psi}^{*}\rVert_{\mathcal{K}}+\beta_{\psi})+\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}}
(1+αψγT(k))ψ~(k)ψ~𝒦+γT(k)(αψψ~𝒦+βψ),a.s.\displaystyle\leq(1+\alpha_{\psi}\gamma^{(k)}_{T})\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}}+\gamma^{(k)}_{T}(\alpha_{\psi}\lVert\tilde{\psi}^{*}\rVert_{\mathcal{K}}+\beta_{\psi}),\text{a.s.}

Since ψ~𝒦<\lVert\tilde{\psi}^{*}\rVert_{\mathcal{K}}<\infty and we assume that (γT(k))k(\gamma^{(k)}_{T})_{k\in\mathbb{N}} is a summable sequence, the Robbins-Siegmund Theorem ([35, Thm. 1]) can be applied to show limkψ~(k)ψ~𝒦\lim_{k\to\infty}\lVert\tilde{\psi}^{(k)}-\tilde{\psi}^{*}\rVert_{\mathcal{K}} exists and is finite a.s. Consequently, there exists a set Ω^\hat{\Omega} which has probability one, such that for any ω^Ω^\hat{\omega}\in\hat{\Omega}, the sequence (ψ~(k)(ω^)ψ~𝒦)k(\lVert\tilde{\psi}^{(k)}(\hat{\omega})-\tilde{\psi}^{*}\rVert_{\mathcal{K}})_{k\in\mathbb{N}} is bounded. Therefore, we can find some constant B(ω^)B(\hat{\omega}) which satisfies, for all kk\in\mathbb{N}, ψ~(k)(ω^)𝒦=ψ~(k)(ω^)ψ~+ψ~𝒦ψ~(k)(ω^)ψ~𝒦+ψ~𝒦B(ω^)\lVert\tilde{\psi}^{(k)}(\hat{\omega})\rVert_{\mathcal{K}}=\lVert\tilde{\psi}^{(k)}(\hat{\omega})-\tilde{\psi}^{*}+\tilde{\psi}^{*}\rVert_{\mathcal{K}}\leq\lVert\tilde{\psi}^{(k)}(\hat{\omega})-\tilde{\psi}^{*}\rVert_{\mathcal{K}}+\lVert\tilde{\psi}^{*}\rVert_{\mathcal{K}}\leq B(\hat{\omega}).

Since the deterministic sequence (ψ~(k)(ω^)𝒦)k(\lVert\tilde{\psi}^{(k)}(\hat{\omega})\rVert_{\mathcal{K}})_{k\in\mathbb{N}} is upper bounded by a constant B(ω^)B(\hat{\omega}) for any ω^Ω^\hat{\omega}\in\hat{\Omega}, combining Lemma 2 and the summability of (γT(k))k(\gamma^{(k)}_{T})_{k\in\mathbb{N}}, we finally can conclude that kγ(k)𝔼[ε(k)k](ω^)kγT(k)(αψψ~(k)(ω^)𝒦+βψ)kγT(k)(αψB(ω^)+βψ)<\sum_{k\in\mathbb{N}}\gamma^{(k)}\mathbb{E}[\varepsilon^{(k)}\mid\mathcal{F}_{k}](\hat{\omega})\leq\sum_{k\in\mathbb{N}}\gamma^{(k)}_{T}(\alpha_{\psi}\lVert\tilde{\psi}^{(k)}(\hat{\omega})\rVert_{\mathcal{K}}+\beta_{\psi})\leq\sum_{k\in\mathbb{N}}\gamma^{(k)}_{T}(\alpha_{\psi}B(\hat{\omega})+\beta_{\psi})<\infty. ∎

E Almost-Sure Convergence of Subroutine 2

We now let J𝒜¯(k)J^{(k)}_{\operatorname{\bar{\mathcal{A}}}} denote the (scenario-based) approximate operator for the exact resolvent J𝔸¯J_{\operatorname{\bar{\mathbb{A}}}} at the kk-th iteration. As a reminder, note that the explicit steps of J𝒜¯(k)J^{(k)}_{\operatorname{\bar{\mathcal{A}}}} are presented in the first player and edge loops in Algorithm 1, before we implement the reflected steps. In the following lemma, we are going to establish that the result of J𝒜¯(k)J^{(k)}_{\operatorname{\bar{\mathcal{A}}}} can approximate that of J𝔸¯J_{\operatorname{\bar{\mathbb{A}}}} with arbitrary accuracy almost surely when a sufficiently large number of stochastic subgradient steps have been taken.

Lemma 3.

Suppose Assumptions 1, 2, 6 , and 7 hold, and ψ~\tilde{\psi} is an arbitrary bounded stack vector. In addition, let the number of subgradient steps taken per iteration satisfy limk\stackunder[1.2pt]T (k)=\lim_{k\to\infty}\stackunder[1.2pt]{$T$}{\rule{3.44444pt}{0.32289pt}}^{(k)}=\infty. Then limkJ𝒜¯(k)(ψ~)=J𝔸¯(ψ~)\lim_{k\to\infty}J^{(k)}_{\operatorname{\bar{\mathcal{A}}}}(\tilde{\psi})=J_{\operatorname{\bar{\mathbb{A}}}}(\tilde{\psi}) a.s.

Proof.

From the explicit updating steps presented in Algorithm 1, it is straightforward that, in the resulting vectors, the entries associated with the local estimates {yii}\{y^{-i}_{i}\} keep the same for J𝒜¯(k)J^{(k)}_{\operatorname{\bar{\mathcal{A}}}} and J𝔸¯J_{\operatorname{\bar{\mathbb{A}}}}, and the entries associated with the dual variables 𝝀\boldsymbol{\lambda}, 𝝁\boldsymbol{\mu}, and 𝒛\boldsymbol{z} are some linear transformations of ψ~\tilde{\psi} and those associated with {yii}\{y^{i}_{i}\}. Hence, it suffices to prove limtyi,tiyi,i22=0\lim_{t\to\infty}\lVert y^{i}_{i,t}-y^{i}_{i,*}\rVert^{2}_{2}=0 a.s. for all i𝒩i\in\mathcal{N}. We let gi,tg_{i,t} denote the scenario-based gradient evaluated at the point yi,tiy^{i}_{i,t}, and gi,tg^{*}_{i,t} the gradient corresponding to the expected-value augmented objective. Start by noting that the distance between yi,tiy^{i}_{i,t} the point obtained after tt minor iteration steps and yi,iy^{i}_{i,*} the minimizer of the expected-valued augmented objective 𝕁^i(k)\hat{\mathbb{J}}^{(k)}_{i} satisfies:

yi,t+1iyi,i22=Pj𝒳iB(yi,tiκi,tgi,t)Pj𝒳iB(yi,i)22\displaystyle\lVert y^{i}_{i,t+1}-y^{i}_{i,*}\rVert^{2}_{2}=\lVert\operatorname{Pj}_{\mathcal{X}^{B}_{i}}(y^{i}_{i,t}-\kappa_{i,t}g_{i,t})-\operatorname{Pj}_{\mathcal{X}^{B}_{i}}(y^{i}_{i,*})\rVert^{2}_{2}
yi,tiκi,tgi,tyi,i22\displaystyle\qquad\leq\lVert y^{i}_{i,t}-\kappa_{i,t}g_{i,t}-y^{i}_{i,*}\rVert^{2}_{2}
=yi,tiyi,i222κi,tyi,tiyi,i,gi,t+(κi,t)2gi,t22\displaystyle\qquad=\lVert y^{i}_{i,t}-y^{i}_{i,*}\rVert^{2}_{2}-2\kappa_{i,t}\langle y^{i}_{i,t}-y^{i}_{i,*},g_{i,t}\rangle+(\kappa_{i,t})^{2}\lVert g_{i,t}\rVert^{2}_{2}
=yi,tiyi,i222κi,tyi,tiyi,i,gi,t\displaystyle\qquad=\lVert y^{i}_{i,t}-y^{i}_{i,*}\rVert^{2}_{2}-2\kappa_{i,t}\langle y^{i}_{i,t}-y^{i}_{i,*},g^{*}_{i,t}\rangle
+2yi,tiyi,i,gi,tgi,t+(κi,t)2gi,t22.\displaystyle\qquad\qquad+2\langle y^{i}_{i,t}-y^{i}_{i,*},g^{*}_{i,t}-g_{i,t}\rangle+(\kappa_{i,t})^{2}\lVert g_{i,t}\rVert^{2}_{2}.

We then construct the following σ\sigma-field:

¯tσ{ξi,0,,ξi,t1}.\displaystyle\bar{\mathcal{F}}_{t}\coloneqq\sigma\{\xi_{i,0},\ldots,\xi_{i,t-1}\}.

Taking the conditional expectation 𝔼[¯t]\mathbb{E}[\cdot\mid\bar{\mathcal{F}}_{t}] on both side of the above inequality yields:

𝔼[yi,t+1iyi,i22¯t]\displaystyle\mathbb{E}[\lVert y^{i}_{i,t+1}-y^{i}_{i,*}\rVert^{2}_{2}\mid\bar{\mathcal{F}}_{t}]
(1κi,tτ1i)yi,tiyi,i22+(κi,t)2𝔼[gi,t22¯t]\displaystyle\qquad\leq(1-\frac{\kappa_{i,t}}{\tau_{1i}})\lVert y^{i}_{i,t}-y^{i}_{i,*}\rVert^{2}_{2}+(\kappa_{i,t})^{2}\mathbb{E}[\lVert g_{i,t}\rVert^{2}_{2}\mid\bar{\mathcal{F}}_{t}]
(1κi,tτ1i)yi,tiyi,i22+κi,t2(αg,i2ψ~22+βg,i2)\displaystyle\qquad\leq(1-\frac{\kappa_{i,t}}{\tau_{1i}})\lVert y^{i}_{i,t}-y^{i}_{i,*}\rVert^{2}_{2}+\kappa_{i,t}^{2}(\alpha^{2}_{g,i}\lVert\tilde{\psi}\rVert^{2}_{2}+\beta^{2}_{g,i})

where the first inequality follows from the strong convexity of 𝕁^i(k)\hat{\mathbb{J}}^{(k)}_{i}, and the second is based on Assumption 7(ii). By leveraging [51, Lemma 2.2.10], we can conclude that limtyi,tiyi,i22=0\lim_{t\to\infty}\lVert y^{i}_{i,t}-y^{i}_{i,*}\rVert^{2}_{2}=0 a.s., which completes the proof. ∎

References

  • [1] Y. Huang and J. Hu, “A distributed Douglas-Rachford based algorithm for stochastic GNE seeking with partial information,” in The 2022 American Control Conference (ACC), 2022, submitted.
  • [2] J. F. Nash et al., “Equilibrium points in N-person games,” Proceedings of the national academy of sciences, vol. 36, no. 1, pp. 48–49, 1950.
  • [3] F. Facchinei, A. Fischer, and V. Piccialli, “On generalized Nash games and variational inequalities,” Operations Research Letters, vol. 35, no. 2, pp. 159–164, 2007.
  • [4] F. Facchinei and C. Kanzow, “Generalized Nash equilibrium problems,” Annals of Operations Research, vol. 175, no. 1, pp. 177–211, 2010.
  • [5] A. Kannan, U. V. Shanbhag, and H. M. Kim, “Strategic behavior in power markets under uncertainty,” Energy Systems, vol. 2, no. 2, pp. 115–141, 2011.
  • [6] ——, “Addressing supply-side risk in uncertain power markets: stochastic Nash models, scalable algorithms and error analysis,” Optimization Methods and Software, vol. 28, no. 5, pp. 1095–1138, 2013.
  • [7] A. Nagurney, M. Salarpour, J. Dong, and L. S. Nagurney, “A stochastic disaster relief game theory network model,” in SN Operations Research Forum, vol. 1, no. 2.   Springer, 2020, pp. 1–33.
  • [8] E. Nikolova and N. E. Stier-Moses, “A mean-risk model for the traffic assignment problem with stochastic travel times,” Operations Research, vol. 62, no. 2, pp. 366–382, 2014.
  • [9] Z. Zhou, P. Mertikopoulos, A. L. Moustakas, N. Bambos, and P. Glynn, “Robust power management via learning and game design,” Operations Research, vol. 69, no. 1, pp. 331–345, 2021.
  • [10] F. Facchinei and J.-S. Pang, Finite-dimensional variational inequalities and complementarity problems.   Springer Science & Business Media, 2007.
  • [11] J. Koshal, A. Nedic, and U. V. Shanbhag, “Regularized iterative stochastic approximation methods for stochastic variational inequality problems,” IEEE Transactions on Automatic Control, vol. 58, no. 3, pp. 594–609, 2012.
  • [12] R. Bot, P. Mertikopoulos, M. Staudigl, and P. Vuong, “Mini-batch forward-backward-forward methods for solving stochastic variational inequalities,” Stochastic Systems, 2020.
  • [13] F. Yousefian, A. Nedić, and U. V. Shanbhag, “On smoothing, regularization, and averaging in stochastic approximation methods for stochastic variational inequality problems,” Mathematical Programming, vol. 165, no. 1, pp. 391–431, 2017.
  • [14] S. Cui and U. V. Shanbhag, “On the analysis of reflected gradient and splitting methods for monotone stochastic variational inequality problems,” in 2016 IEEE 55th Conference on Decision and Control (CDC).   IEEE, 2016, pp. 4510–4515.
  • [15] A. Kannan and U. V. Shanbhag, “Optimal stochastic extragradient schemes for pseudomonotone stochastic variational inequality problems and their variants,” Computational Optimization and Applications, vol. 74, no. 3, pp. 779–820, 2019.
  • [16] F. Salehisadaghiani and L. Pavel, “Distributed Nash equilibrium seeking: A gossip-based algorithm,” Automatica, vol. 72, pp. 209–216, 2016.
  • [17] F. Parise, S. Grammatico, B. Gentile, and J. Lygeros, “Distributed convergence to Nash equilibria in network and average aggregative games,” Automatica, vol. 117, p. 108959, 2020.
  • [18] P. Yi and L. Pavel, “An operator splitting approach for distributed generalized Nash equilibria computation,” Automatica, vol. 102, pp. 111–121, 2019.
  • [19] ——, “Distributed generalized Nash equilibria computation of monotone games via double-layer preconditioned proximal-point algorithms,” IEEE Transactions on Control of Network Systems, vol. 6, no. 1, pp. 299–311, 2018.
  • [20] L. Pavel, “Distributed GNE seeking under partial-decision information over networks via a doubly-augmented operator splitting approach,” IEEE Transactions on Automatic Control, vol. 65, no. 4, pp. 1584–1597, 2019.
  • [21] M. Bianchi, G. Belgioioso, and S. Grammatico, “Fast generalized nash equilibrium seeking under partial-decision information,” Automatica, vol. 136, p. 110080, 2022.
  • [22] G. Belgioioso, A. Nedich, and S. Grammatico, “Distributed generalized nash equilibrium seeking in aggregative games on time-varying networks,” IEEE Transactions on Automatic Control, 2020.
  • [23] T. Yang, X. Yi, J. Wu, Y. Yuan, D. Wu, Z. Meng, Y. Hong, H. Wang, Z. Lin, and K. H. Johansson, “A survey of distributed optimization,” Annual Reviews in Control, vol. 47, pp. 278–305, 2019.
  • [24] B. Franci and S. Grammatico, “A distributed forward-backward algorithm for stochastic generalized Nash equilibrium seeking,” IEEE Transactions on Automatic Control, 2020.
  • [25] S. Cui, B. Franci, S. Grammatico, U. V. Shanbhag, and M. Staudigl, “A relaxed-inertial forward-backward-forward algorithm for stochastic generalized Nash equilibrium seeking,” arXiv preprint arXiv:2103.13115, 2021.
  • [26] J. Lei, U. V. Shanbhag, J.-S. Pang, and S. Sen, “On synchronous, asynchronous, and randomized best-response schemes for stochastic Nash games,” Mathematics of Operations Research, vol. 45, no. 1, pp. 157–190, 2020.
  • [27] J. Lei and U. V. Shanbhag, “Distributed variable sample-size gradient-response and best-response schemes for stochastic nash equilibrium problems over graphs,” arXiv preprint arXiv:1811.11246, 2018.
  • [28] B. Franci and S. Grammatico, “Stochastic generalized Nash equilibrium seeking under partial-decision information,” arXiv preprint arXiv:2011.05357, 2020.
  • [29] D. P. Palomar and Y. C. Eldar, Convex optimization in signal processing and communications.   Cambridge university press, 2010.
  • [30] U. Ravat and U. V. Shanbhag, “On the characterization of solution sets of smooth and nonsmooth convex stochastic Nash games,” SIAM Journal on Optimization, vol. 21, no. 3, pp. 1168–1199, 2011.
  • [31] A. A. Kulkarni and U. V. Shanbhag, “On the variational equilibrium as a refinement of the generalized Nash equilibrium,” Automatica, vol. 48, no. 1, pp. 45–55, 2012.
  • [32] Y. Huang and J. Hu, “Distributed solution of GNEP over networks via the Douglas-Rachford splitting method,” in 2021 IEEE 60th Conference on Decision and Control (CDC).   IEEE, 2021, to appear, a full version is available at https://arxiv.org/abs/2103.09393.
  • [33] H. H. Bauschke, Convex Analysis and Monotone Operator Theory in Hilbert Spaces, 2nd ed., ser. CMS Books in Mathematics, Ouvrages de mathématiques de la SMC, 2017.
  • [34] H. E. Bell, “Gershgorin’s theorem and the zeros of polynomials,” The American Mathematical Monthly, vol. 72, no. 3, pp. 292–295, 1965.
  • [35] H. Robbins and D. Siegmund, “A convergence theorem for non-negative almost supermartingales and some applications,” in Optimizing methods in statistics.   Elsevier, 1971, pp. 233–257.
  • [36] N. Z. Shor, Minimization methods for non-differentiable functions.   Springer Science & Business Media, 2012, vol. 3.
  • [37] S. Lacoste-Julien, M. Schmidt, and F. Bach, “A simpler approach to obtaining an O(1/t) convergence rate for the projected stochastic subgradient method,” arXiv preprint arXiv:1212.2002, 2012.
  • [38] Y. Huang and J. Hu, “Distributed computation of stochastic GNE with partial information: An augmented best-response approach,” arXiv preprint arXiv:2109.12290, 2021.
  • [39] F. Parise, B. Gentile, and J. Lygeros, “A distributed algorithm for almost-nash equilibria of average aggregative games with coupling constraints,” IEEE Transactions on Control of Network Systems, vol. 7, no. 2, pp. 770–782, 2019.
  • [40] T. Brinkhoff, “A framework for generating network-based moving objects,” GeoInformatica, vol. 6, no. 2, pp. 153–180, 2002.
  • [41] G. Belgioioso and S. Grammatico, “Projected-gradient algorithms for generalized equilibrium seeking in aggregative games are preconditioned forward-backward methods,” in 2018 European Control Conference (ECC).   IEEE, 2018, pp. 2188–2193.
  • [42] G. B. Dantzig, “Linear programming under uncertainty,” Management science, vol. 1, no. 3-4, pp. 197–206, 1955.
  • [43] J. R. Birge and F. Louveaux, Introduction to stochastic programming.   Springer Science & Business Media, 2011.
  • [44] G. Barbarosoǧlu and Y. Arda, “A two-stage stochastic programming framework for transportation planning in disaster response,” Journal of the operational research society, vol. 55, no. 1, pp. 43–53, 2004.
  • [45] A. Shapiro, D. Dentcheva, and A. Ruszczyński, Lectures on stochastic programming: modeling and theory.   SIAM, 2014.
  • [46] S. Boyd, S. P. Boyd, and L. Vandenberghe, Convex optimization.   Cambridge university press, 2004.
  • [47] D. Boley, “Local linear convergence of the alternating direction method of multipliers on quadratic or linear programs,” SIAM Journal on Optimization, vol. 23, no. 4, pp. 2183–2207, 2013.
  • [48] J. Liang, J. Fadili, and G. Peyré, “Local linear convergence analysis of primal-dual splitting methods,” Optimization, vol. 67, no. 6, pp. 821–853, 2018.
  • [49] S. Cui and U. V. Shanbhag, “Variance-reduced proximal and splitting schemes for monotone stochastic generalized equations,” arXiv preprint arXiv:2008.11348, 2020.
  • [50] H. Asi and J. C. Duchi, “The importance of better models in stochastic optimization,” Proceedings of the National Academy of Sciences, vol. 116, no. 46, pp. 22 924–22 930, 2019.
  • [51] B. T. Polyak, “Introduction to optimization. optimization software,” Inc., Publications Division, New York, vol. 1, 1987.