Learning Provably Stable Local Volt/Var Controllers for Efficient Network Operation

Zhenyi Yuan, , Guido Cavraro, ,
Manish K. Singh, , and Jorge Cortés
This work was authored by the National Renewable Energy Laboratory, operated by Alliance for Sustainable Energy, LLC, for the U.S. Department of Energy (DOE) under Contract No. DE-AC36-08GO28308. Funding provided by the NREL Laboratory Directed Research and Development Program. The views expressed in the article do not necessarily represent the views of the DOE or the U.S. Government. The U.S. Government retains and the publisher, by accepting the article for publication, acknowledges that the U.S. Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for U.S. Government purposes. This work was also partially supported by NSF Award ECCS-1947050.Z. Yuan and J. Cortés are with the Department of Mechanical and Aerospace Engineering, UC San Diego. {z7yuan,cortes}@ucsd.edu. G. Cavraro is with the Power Systems Engineering Center, National Renewable Energy Laboratory. [email protected]. M. K. Singh is with the Department of Electrical and Computer Engineering, University of Minnesota. [email protected].
Abstract

This paper develops a data-driven framework to synthesize local Volt/Var control strategies for distributed energy resources (DERs) in power distribution grids (DGs). Aiming to improve DG operational efficiency, as quantified by a generic optimal reactive power flow (ORPF) problem, we propose a two-stage approach. The first stage involves learning the manifold of optimal operating points determined by an ORPF instance. To synthesize local Volt/Var controllers, the learning task is partitioned into learning local surrogates (one per DER) of the optimal manifold with voltage input and reactive power output. Since these surrogates characterize efficient DG operating points, in the second stage, we develop local control schemes that steer the DG to these operating points. We identify the conditions on the surrogates and control parameters to ensure that the locally acting controllers collectively converge, in a global asymptotic sense, to a DG operating point agreeing with the local surrogates. We use neural networks to model the surrogates and enforce the identified conditions in the training phase. AC power flow simulations on the IEEE 37-bus network empirically bolster the theoretical stability guarantees obtained under linearized power flow assumptions. The tests further highlight the optimality improvement compared to prevalent benchmark methods.

Index Terms:
Distributed energy resources, global stability, local control, Volt/Var control.

I Introduction

The deployment of a massive number of distributed energy resources (DERs) in power distribution grids (DGs) is dramatically changing the electric power grid. Primarily driven by sustainability and economic incentives, DERs present additional opportunities, including reductions in the power generation costs and of greenhouse gas emissions. Nevertheless, DERs’ uncoordinated power injections or sudden generation changes could pose challenges to system operations and stability, e.g., induce undesirable voltage deviations in DGs. To facilitate their integration in power grids, DERs are being provided with sensing and computational capabilities, hence becoming smart agents. DERs can exploit the flexibility of their power electronic interface to perform, among other ancillary services, reactive power control. They can also take advantage of the widespread availability of data from DGs and the increased capabilities for storing and processing the data to learn effective control policies. This paper aims to leverage learning in the synthesis of local Volt/Var controllers for voltage regulation incorporating optimality considerations and rigorous performance guarantees.

Literature Review

The main goal of Volt/Var control strategies is to keep voltages within safe preassigned limits by commanding DERs’ reactive power injections. Classically, DERs’ reactive power outputs are computed, in an open-loop fashion, by the system operator solving optimal power flow (OPF) problems. Efficient and advanced solvers for OPF problems are available, see, e.g. [1, 2, 3]. However, high penetration of renewable generation and increased variability of DGs require solving numerous instances of OPF problems within a limited time frame. Aiming to tackle this challenge, several learning-based approaches have been proposed to predict OPF solutions, see, e.g., [4, 5, 6, 7, 8], to mention a few. Once trained, the inference time for these approaches when presented with a new input is minimal. Nevertheless, the solution of OPF problems requires information from all the buses. Specifically, power demands from loads and generation limits from generators must be precisely known. Such requirements are prohibitive for practical DGs because, in general, not all the buses are monitored in real time, individual loads are unlikely to announce their demand profiles in advance, and the availability of small size generators is hard to predict.

This has motivated the development of closed-loop strategies, which compensate for the lack of information with measurements retrieved from the field. Given the massive number of controllable devices envisioned to be hosted in future DGs, decentralized approaches are often advocated for practical applications. There are two notable classes of decentralized algorithms. The first consists of distributed algorithms in which agents cooperate and share information with peers. Distributed algorithms can achieve optimal performance in the sense that they can be designed to exactly solve a given OPF instance, see, e.g., [9]. Optimization-based feedback controllers that steer the network toward solutions to OPF problems based on the cyclical alternation of sensing, communication, and actuation have recently become popular [10, 11, 12]. Nevertheless, distributed strategies are suitable for systems endowed with a reliable real-time communication network meeting precise and strict requirements, which are rarely satisfied in practice for DGs. For instance, in many works, each generator is required to share information with all its neighbors in the power network before every power output update [10]. The second class of decentralized algorithms consists of local approaches, in which each agent makes decisions based only on information available locally. In local schemes, reactive power compensations are adjusted based merely on measurements taken locally. Even though pertinent standards allow DERs to provide reactive power compensations following static Volt/Var control rules, see IEEE standard 1547 [13], the literature has provided a variety of options for local voltage regulation [10, 14, 15, 16]. However, local schemes have intrinsic performance limitations, e.g., they might fail to regulate voltages even if the overall generation resources are enough [17].

Recent advances in data-driven and learning-based control seek to leverage data from the plant to learn optimal controllers. Though impressive results have been demonstrated, it has not yet been widely used in engineering practice due to the lack of closed-loop stability guarantees. Various results have been developed to tackle this problem, see e.g., a comprehensive review [18]. Although most of these works achieve the stability of neural network control by using penalty functions to integrate the stability requirement as soft constraints, recent work in power systems, primarily on frequency control, seeks to explicitly engineer the neural network structure to integrate the stability requirement as hard constraints, see, e.g., [19, 20]. In the context of Volt/Var control, the goal is to leverage learning techniques to enhance the performance of local control schemes and reduce the gap with distributed and/or optimal controllers, while retaining closed-loop system stability. Related works include [21] and [22], where reinforcement learning is used to learn stability-guaranteed local Volt/Var control schemes. However, the former enforces stringent derivative constraints on the policies to be searched, whereas the latter only guarantees that the voltages converge to a region, instead of an equilibrium point, and both of them require the control policy to be continuously differentiable. Furthermore, neither of them takes accounts for the reactive power capacity limitations, which are critical when dealing with small-size generators. Recent works [23, 24] optimize the local Volt/Var control schemes though interval optimization with a day-ahead schedule, while preserving closed-loop system stability. However, the considered local Volt/Var curves are limited to a piecewise linear form, and might suffer non-negligible optimality gaps. Other works provide interesting insights on learning Volt/Var rules, but do not assess the stability of the overall system and hence are not straightforwardly suitable for practical applications: [25, 26] leverage segmented linear regression techniques to learn local surrogates that predict OPF solutions, [27] proposes to learn the local controller by taking as an input both voltages and active power setpoints, and [28] proposes a framework for tuning the parameters of standard piecewise linear local voltage regulators.

Statement of Contributions

We propose a framework for designing local Volt/Var scheme whose goal is to not only regulate voltages but also act as local surrogates of optimal reactive power flow (ORPF) problem solvers. ORPF problems are particular instances of OPF problem in which the goal is to optimize the generator’s reactive power injections. We base our work on the distinction between the control function and the equilibrium function. The first represents the reactive power update rule, and the latter describes the possible system equilibrium points. In many works, the equilibrium functions coincide with the control functions, but this is not the case for our framework. We advocate for a two-stage strategy. In the first stage, for each controllable node, the equilibrium function providing the ORPF solution surrogates is learned from historical data. Precisely, such a function receives as input the local voltage and gives as an output an approximation of the optimal reactive power setpoint. In the second stage, we devise a control function whose equilibrium points are exactly the ORPF approximated solutions provided by the equilibrium functions. The novelties of our paper with respect to the recent literature can be summarized as:

  • The equilibrium functions are not forced to be (piecewise) linear and are not subject to slope limitations [10, 14, 15]. This relaxes several restrictive constraints on equilibrium functions, which leads to an enlarged search space of potential candidates of desired OPRF surrogates, thus reducing the optimality gap.

  • The control rule is globally asymptotically stable, as opposed to algorithms that are locally asymptotically stable [16, 21] or whose stability is not analytically characterized. Our design provably steers the system to the desirable configuration described by the equilibrium functions irrespective of the initial reactive power injection.

  • We introduce the idea of pseudo data points to enhance the voltage regulation capability of learned controllers when voltages are not within the desired limits.

With respect to its preliminary version [29], this work differs as follows: it relaxes the differentiability requirement on the control rule; it establishes global, rather than local, stability guarantees for the control scheme; it provides a more general design of neural networks in the learning process; and it uses of pseudo data points.

Outline

The paper is organized as follows. In Section II, we model a power distribution network and define the problem of interest. Section III introduces the control scheme, states its stability properties, and identifies conditions on the equilibrium functions needed for the system stability. Sections IV describes the learning process: precisely, it reports how we construct the data set used for learning and how we parameterize the equilibrium function with a neural network to meet the required conditions by design. Numerical simulations in Section V validate the proposed approach and show significant improvements with respect to prevalent benchmark methods. Finally, Section VI concludes this work.

Notation

Throughout the paper, \mathbb{R}blackboard_R and \mathbb{C}blackboard_C denote the set of real and complex numbers, respectively. Upper and lowercase boldface letters denote matrices and column vectors, respectively. Sets are represented by calligraphic symbols. Given a vector 𝐚𝐚\mathbf{a}bold_a (a diagonal matrix 𝐀𝐀\mathbf{A}bold_A), its n𝑛nitalic_n-th (diagonal) entry is denoted by ansubscript𝑎𝑛a_{n}italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT (An)subscript𝐴𝑛(A_{n})( italic_A start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ). 𝐀()0succeeds𝐀succeeds-or-equals0\mathbf{A}\succ(\succeq)~{}0bold_A ≻ ( ⪰ ) 0 denotes that matrix 𝐀𝐀\mathbf{A}bold_A is positive (semi-) definite, and 𝐀()0precedes𝐀precedes-or-equals0\mathbf{A}\prec(\preceq)~{}0bold_A ≺ ( ⪯ ) 0 denotes that matrix 𝐀𝐀\mathbf{A}bold_A is negative (semi-) definite. The symbol ()superscripttop(\cdot)^{\top}( ⋅ ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT stands for transposition, and 𝟏,𝟎,𝐈10𝐈\mathbf{1},\mathbf{0},\mathbf{I}bold_1 , bold_0 , bold_I denote vectors of all ones and zeros and identity matrix with appropriate dimensions, respectively. Operators ()\Re(\cdot)roman_ℜ ( ⋅ ) and ()\Im(\cdot)roman_ℑ ( ⋅ ) extract the real and imaginary parts of a complex-valued argument, and act element-wise. With a slight abuse of notation, we use |||\cdot|| ⋅ | to denote the absolute value for real-valued arguments, the magnitude for complex-valued arguments, and the cardinality when the argument is a set. \|\cdot\|∥ ⋅ ∥ represents the Euclidean norm. Given a symmetric matrix 𝐀𝐀\mathbf{A}bold_A, λmax(𝐀)subscript𝜆𝐀\lambda_{\max}(\mathbf{A})italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( bold_A ) and λmin(𝐀)subscript𝜆𝐀\lambda_{\min}(\mathbf{A})italic_λ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( bold_A ) represent its largest and smallest eigenvalue, respectively. For any matrix 𝐁𝐁\mathbf{B}bold_B, it holds that 𝐁=λmax(𝐁𝐁)norm𝐁subscript𝜆superscript𝐁top𝐁\|\mathbf{B}\|=\sqrt{\lambda_{\max}(\mathbf{B}^{\top}\mathbf{B})}∥ bold_B ∥ = square-root start_ARG italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( bold_B start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_B ) end_ARG. The graph of a function ϕ::italic-ϕ\phi:\mathbb{R}\rightarrow\mathbb{R}italic_ϕ : blackboard_R → blackboard_R is the set of all points of the form (x,ϕ(x))𝑥italic-ϕ𝑥(x,\phi(x))( italic_x , italic_ϕ ( italic_x ) ), whereas the range of ϕitalic-ϕ\phiitalic_ϕ is the set of its possible output values.

II Grid Modeling and Problem Formulation

Consider a balanced three-phase power distribution network111Although the analysis and algorithms developed in this work consider balanced grids, contemporary research has demonstrated applicability of related approaches to unbalanced multiphase networks [24, 30]. We leave for future research a rigorous and detailed extension of our findings to encompass unbalanced grids. with N+1𝑁1N+1italic_N + 1 buses represented by its single-phase equivalent and modeled as an undirected graph 𝒢=(𝒩,)𝒢𝒩\mathcal{G}=(\mathcal{N},\mathcal{E})caligraphic_G = ( caligraphic_N , caligraphic_E ), where 𝒩={0,1,,N}𝒩01𝑁\mathcal{N}=\{0,1,\dots,N\}caligraphic_N = { 0 , 1 , … , italic_N } are associated with the electrical buses, and \mathcal{E}caligraphic_E represents the set of the electrical lines between these buses. We label the substation node as 0, and assume that it behaves as an ideal voltage source imposing the nominal voltage of 1 p.u. Define the following quantities:

  • unsubscript𝑢𝑛u_{n}\in\mathbb{C}italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ blackboard_C is the voltage phasor at bus n𝒩𝑛𝒩n\in\mathcal{N}italic_n ∈ caligraphic_N;

  • vnsubscript𝑣𝑛v_{n}\in\mathbb{R}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ blackboard_R is the voltage magnitude at bus n𝒩𝑛𝒩n\in\mathcal{N}italic_n ∈ caligraphic_N;

  • insubscript𝑖𝑛i_{n}\in\mathbb{C}italic_i start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ blackboard_C is the injected current phasor at bus n𝒩𝑛𝒩n\in\mathcal{N}italic_n ∈ caligraphic_N;

  • sn=pn+iqnsubscript𝑠𝑛subscript𝑝𝑛𝑖subscript𝑞𝑛s_{n}=p_{n}+iq_{n}\in\mathbb{C}italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_i italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ blackboard_C is the nodal complex power injection at bus n𝒩𝑛𝒩n\in\mathcal{N}italic_n ∈ caligraphic_N, where pn,qnsubscript𝑝𝑛subscript𝑞𝑛p_{n},q_{n}\in\mathbb{R}italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ∈ blackboard_R are the active and reactive powers, respectively. Powers take positive (negative) values, i.e., pn,qn0subscript𝑝𝑛subscript𝑞𝑛0p_{n},q_{n}\geq 0italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≥ 0 (pn,qn0subscript𝑝𝑛subscript𝑞𝑛0p_{n},q_{n}\leq 0italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≤ 0), when they are injected into (absorbed from) the grid.

We use vectors 𝐮,𝐢,𝐬N𝐮𝐢𝐬superscript𝑁\mathbf{u},\mathbf{i},\mathbf{s}\in\mathbb{C}^{N}bold_u , bold_i , bold_s ∈ blackboard_C start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT to collect the complex voltages, currents, and complex powers of buses 1,2,,N12𝑁1,2,\ldots,N1 , 2 , … , italic_N, and vectors 𝐯,𝐩,𝐪N𝐯𝐩𝐪superscript𝑁\mathbf{v},\mathbf{p},\mathbf{q}\in\mathbb{R}^{N}bold_v , bold_p , bold_q ∈ blackboard_R start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT to collect their voltage magnitudes, and active and reactive power injections. Denote by zesubscript𝑧𝑒z_{e}\in\mathbb{C}italic_z start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ∈ blackboard_C and by ye=ze1subscript𝑦𝑒superscriptsubscript𝑧𝑒1y_{e}=z_{e}^{-1}\in\mathbb{C}italic_y start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = italic_z start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ∈ blackboard_C the impedance and the admittance of line e=(m,n)𝑒𝑚𝑛e=(m,n)\in\mathcal{E}italic_e = ( italic_m , italic_n ) ∈ caligraphic_E, respectively. The network bus admittance matrix 𝐘(N+1)×(N+1)𝐘superscript𝑁1𝑁1\mathbf{Y}\in\mathbb{C}^{(N+1)\times(N+1)}bold_Y ∈ blackboard_C start_POSTSUPERSCRIPT ( italic_N + 1 ) × ( italic_N + 1 ) end_POSTSUPERSCRIPT is a symmetric matrix that can be expressed as 𝐘=𝐘L+diag(𝐲T)𝐘subscript𝐘𝐿diagsubscript𝐲𝑇\mathbf{Y}=\mathbf{Y}_{L}+{\rm diag}(\mathbf{y}_{T})bold_Y = bold_Y start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT + roman_diag ( bold_y start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ), where

(𝐘L)mn={y(m,n) if (m,n),mn,0 if (m,n),mn,kny(k,n) if m=n,subscriptsubscript𝐘𝐿𝑚𝑛casessubscript𝑦𝑚𝑛formulae-sequence if 𝑚𝑛𝑚𝑛0formulae-sequence if 𝑚𝑛𝑚𝑛subscript𝑘𝑛subscript𝑦𝑘𝑛 if 𝑚𝑛\displaystyle(\mathbf{Y}_{L})_{mn}=\begin{cases}-y_{(m,n)}&\text{ if }(m,n)\in% \mathcal{E},m\neq n,\\ 0&\text{ if }(m,n)\notin\mathcal{E},m\neq n,\\ \sum_{k\neq n}y_{(k,n)}&\text{ if }m=n,\end{cases}( bold_Y start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT = { start_ROW start_CELL - italic_y start_POSTSUBSCRIPT ( italic_m , italic_n ) end_POSTSUBSCRIPT end_CELL start_CELL if ( italic_m , italic_n ) ∈ caligraphic_E , italic_m ≠ italic_n , end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL if ( italic_m , italic_n ) ∉ caligraphic_E , italic_m ≠ italic_n , end_CELL end_ROW start_ROW start_CELL ∑ start_POSTSUBSCRIPT italic_k ≠ italic_n end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT ( italic_k , italic_n ) end_POSTSUBSCRIPT end_CELL start_CELL if italic_m = italic_n , end_CELL end_ROW

and the vector 𝐲Tsubscript𝐲𝑇\mathbf{y}_{T}bold_y start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT collects the shunt components of each bus. The matrix 𝐘Lsubscript𝐘𝐿\mathbf{Y}_{L}bold_Y start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT is a complex Laplacian matrix, and hence satisfies 𝐘L𝟏=𝟎subscript𝐘𝐿10\mathbf{Y}_{L}\mathbf{1}=\mathbf{0}bold_Y start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT bold_1 = bold_0. We partition the bus admittance matrix by separating the components associated with the substation and the ones associated with the other nodes, obtaining

𝐘=[y0𝐲0𝐲0𝐘~],𝐘matrixsubscript𝑦0superscriptsubscript𝐲0topsubscript𝐲0~𝐘\displaystyle\mathbf{Y}=\begin{bmatrix}y_{0}&\mathbf{y}_{0}^{\top}\\ \mathbf{y}_{0}&\tilde{\mathbf{Y}}\end{bmatrix},bold_Y = [ start_ARG start_ROW start_CELL italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL start_CELL bold_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL bold_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL start_CELL over~ start_ARG bold_Y end_ARG end_CELL end_ROW end_ARG ] ,

with y0subscript𝑦0y_{0}\in\mathbb{C}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ blackboard_C, 𝐲0Nsubscript𝐲0superscript𝑁\mathbf{y}_{0}\in\mathbb{C}^{N}bold_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ blackboard_C start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT, and 𝐘~N×N~𝐘superscript𝑁𝑁\tilde{\mathbf{Y}}\in\mathbb{C}^{N\times N}over~ start_ARG bold_Y end_ARG ∈ blackboard_C start_POSTSUPERSCRIPT italic_N × italic_N end_POSTSUPERSCRIPT. If the network is connected, then 𝐘~~𝐘\tilde{\mathbf{Y}}over~ start_ARG bold_Y end_ARG is invertible [31]. Let 𝐙~:=𝐘~1assign~𝐙superscript~𝐘1\tilde{\mathbf{Z}}:=\tilde{\mathbf{Y}}^{-1}over~ start_ARG bold_Z end_ARG := over~ start_ARG bold_Y end_ARG start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, the power flow equations are

𝐮𝐮\displaystyle\mathbf{u}bold_u =𝐙~𝐢+𝐮^,absent~𝐙𝐢^𝐮\displaystyle=\tilde{\mathbf{Z}}\mathbf{i}+\hat{\mathbf{u}},= over~ start_ARG bold_Z end_ARG bold_i + over^ start_ARG bold_u end_ARG , (1a)
u0subscript𝑢0\displaystyle u_{0}italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =1,absent1\displaystyle=1,= 1 , (1b)
i0subscript𝑖0\displaystyle i_{0}italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =𝟏𝐢,absentsuperscript1top𝐢\displaystyle=\mathbf{1}^{\top}\mathbf{i},= bold_1 start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_i , (1c)
uni¯nsubscript𝑢𝑛subscript¯𝑖𝑛\displaystyle u_{n}\bar{i}_{n}italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT over¯ start_ARG italic_i end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =pn+jqn,n0,formulae-sequenceabsentsubscript𝑝𝑛𝑗subscript𝑞𝑛𝑛0\displaystyle=p_{n}+jq_{n},\qquad n\neq 0,= italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + italic_j italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_n ≠ 0 , (1d)
vnsubscript𝑣𝑛\displaystyle v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =|un|absentsubscript𝑢𝑛\displaystyle=|u_{n}|= | italic_u start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | (1e)

where i¯nsubscript¯𝑖𝑛\bar{i}_{n}over¯ start_ARG italic_i end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denotes the complex conjugate of insubscript𝑖𝑛i_{n}italic_i start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and 𝐮^:=𝐙~𝐲0assign^𝐮~𝐙subscript𝐲0\hat{\mathbf{u}}:=\tilde{\mathbf{Z}}\mathbf{y}_{0}over^ start_ARG bold_u end_ARG := over~ start_ARG bold_Z end_ARG bold_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Eq. (1a) represents the Kirchoff equations and provides the relation between voltages and currents. Eqs. (1b) and (1c) hold because the substation is modeled as the slack bus. Eq. (1d) comes from the fact that all the nodes, except the substation, are modeled to be constant power buses.

Assume a subset 𝒞𝒩𝒞𝒩\mathcal{C}\subseteq\mathcal{N}caligraphic_C ⊆ caligraphic_N of buses hosts DERs, with |𝒞|=C𝒞𝐶|\mathcal{C}|=C| caligraphic_C | = italic_C. Every DER corresponds to a smart agent provided with some computational sensing capability, i.e., it measures its voltage magnitude. The remaining nodes constitute the set =𝒩𝒞𝒩𝒞\mathcal{L}=\mathcal{N}\setminus\mathcal{C}caligraphic_L = caligraphic_N ∖ caligraphic_C and are referred to as loads. For convenience, we partition the reactive powers and the voltage magnitudes by grou** together the nodes belonging to the load and generation sets

𝐪=[𝐪𝒞𝐪],𝐯=[𝐯𝒞𝐯].formulae-sequence𝐪superscriptmatrixsuperscriptsubscript𝐪𝒞topsuperscriptsubscript𝐪toptop𝐯superscriptmatrixsuperscriptsubscript𝐯𝒞topsuperscriptsubscript𝐯toptop\displaystyle\mathbf{q}=\begin{bmatrix}\mathbf{q}_{\mathcal{C}}^{\top}&\mathbf% {q}_{\mathcal{L}}^{\top}\end{bmatrix}^{\top},\quad\mathbf{v}=\begin{bmatrix}% \mathbf{v}_{\mathcal{C}}^{\top}&\mathbf{v}_{\mathcal{L}}^{\top}\end{bmatrix}^{% \top}.bold_q = [ start_ARG start_ROW start_CELL bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT end_CELL start_CELL bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT , bold_v = [ start_ARG start_ROW start_CELL bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT end_CELL start_CELL bold_v start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT .

Motivated by practical considerations, we assume that the grid is not endowed with a communication network that can be used by agents, loads, and the system operator to share information in real-time. As a notable consequence, load demands are not known to the system operator or to agents in real time, preventing the use of open-loop strategies for computing the power outputs. The (averaged) demands could, however, be reported on a hourly/daily basis for slow-timescale applications such as billing.

The massive deployment of DERs in DGs might induce voltage quality issues. For example, sudden generation drops could lead the voltages of a network with high penetration of renewables below desired operational limits and even close to collapse. Since DERs are able to provide ancillary services, reactive power compensation can be used to regulate voltage profiles. Ideally, one wants the DER reactive power setpoints to be the solution of an optimal reactive power flow (ORPF) problem of the form222More comprehensive OPRF problems could in principle be of interest in practical applications, e.g., considering line flows limitations as well. Although in this paper we focus on OPRF problems of the type (2), our approach can be readily applied to other ORPF formulations. Also, we restrict our attention to the case in which problem (2) admits a unique solution. When that is not the case, 𝐪𝒞(𝐩,𝐪)superscriptsubscript𝐪𝒞𝐩subscript𝐪\mathbf{q}_{\mathcal{C}}^{\star}(\mathbf{p},\mathbf{q}_{\mathcal{L}})bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ( bold_p , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT ) can be chosen among the set of minimizers.

𝐪𝒞(𝐩,𝐪):=argmin𝐪𝒞assignsuperscriptsubscript𝐪𝒞𝐩subscript𝐪subscriptsubscript𝐪𝒞\displaystyle\mathbf{q}_{\mathcal{C}}^{\star}(\mathbf{p},\mathbf{q}_{\mathcal{% L}}):=\arg\min_{\mathbf{q}_{\mathcal{C}}}\ bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ( bold_p , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT ) := roman_arg roman_min start_POSTSUBSCRIPT bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT end_POSTSUBSCRIPT f(𝐪𝒞)𝑓subscript𝐪𝒞\displaystyle~{}f(\mathbf{q}_{\mathcal{C}})italic_f ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) (2a)
s.t.formulae-sequencest\displaystyle\mathrm{s.t.}\ roman_s . roman_t . (1a)(1e)italic-(1aitalic-)italic-(1eitalic-)\displaystyle~{}\eqref{eq:nodevoltage}-\eqref{eq:vm_def}italic_( italic_) - italic_( italic_)
𝐯min𝐯(𝐪𝒞)𝐯maxsubscript𝐯𝐯subscript𝐪𝒞subscript𝐯\displaystyle~{}\mathbf{v}_{\min}\leq\mathbf{v}(\mathbf{q}_{\mathcal{C}})\leq% \mathbf{v}_{\max}bold_v start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ≤ bold_v ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) ≤ bold_v start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT (2b)
𝐪min𝐪𝒞𝐪maxsubscript𝐪subscript𝐪𝒞subscript𝐪\displaystyle~{}\mathbf{q}_{\min}\leq\mathbf{q}_{\mathcal{C}}\leq\mathbf{q}_{\max}bold_q start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ≤ bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ≤ bold_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT (2c)

where 𝐪min,𝐪maxCsubscript𝐪subscript𝐪superscript𝐶\mathbf{q}_{\min},\mathbf{q}_{\max}\in\mathbb{R}^{C}bold_q start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT , bold_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT are the minimum and maximum DERs’ reactive power injections; 𝐯min,𝐯maxNsubscript𝐯subscript𝐯superscript𝑁\mathbf{v}_{\min},\mathbf{v}_{\max}\in\mathbb{R}^{N}bold_v start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT , bold_v start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT are the desired voltage lower and upper bounds on all the network buses; and f:C:𝑓superscript𝐶f:\mathbb{R}^{C}\rightarrow\mathbb{R}italic_f : blackboard_R start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT → blackboard_R is the cost function of interest. The minimizer depends on the uncontrolled variables 𝐩𝐩\mathbf{p}bold_p and 𝐪subscript𝐪\mathbf{q}_{\mathcal{L}}bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT, which appear implicitly in the constraint (2b) via equation (1e). Also, notice that for given (re)active loads and active generation, the voltage at node n𝑛nitalic_n and the vector of the voltage magnitudes become functions exclusively of 𝐪𝒞subscript𝐪𝒞\mathbf{q}_{\mathcal{C}}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT, i.e.,

vn=vn(𝐪𝒞),𝐯=𝐯(𝐪𝒞).formulae-sequencesubscript𝑣𝑛subscript𝑣𝑛subscript𝐪𝒞𝐯𝐯subscript𝐪𝒞v_{n}=v_{n}(\mathbf{q}_{\mathcal{C}}),\quad\mathbf{v}=\mathbf{v}(\mathbf{q}_{% \mathcal{C}}).italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) , bold_v = bold_v ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) .

The complexity of solving (2) depends on the choice of the cost function f()𝑓f(\cdot)italic_f ( ⋅ ) and the non-convexity of (1). Tremendous advancements have been made to overcome the computational limitations via convex relaxations [1], linearized power flow equations [11], distributed optimization [9], and learning-based approaches [6]. However, solving (2) inevitably requires the knowledge of network-wide quantities (𝐩,𝐪)𝐩subscript𝐪(\mathbf{p},\mathbf{q}_{\mathcal{L}})( bold_p , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT ), centrally or via peer-to-peer communication. Because the necessary supporting real-time communication infrastructure is not prevalent for most distribution systems, the optimal 𝐪𝒞superscriptsubscript𝐪𝒞\mathbf{q}_{\mathcal{C}}^{\star}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT cannot be directly computed. Rather, inspired by the ongoing efforts towards designing local communication-free control rules for DERs and the recently reported success of neural-network-based surrogates for OPF, this work proposes a two-stage approach. In the first stage, termed the learning stage, historical data are used to learn functions that map voltages to (approximate) solutions of the ORPF problem (2). Specifically, for each agent n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C, we learn a function ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT

ϕn:,vnϕn(vn):subscriptitalic-ϕ𝑛formulae-sequencemaps-tosubscript𝑣𝑛subscriptitalic-ϕ𝑛subscript𝑣𝑛\phi_{n}:\mathbb{R}\rightarrow\mathbb{R},\;v_{n}\mapsto\phi_{n}(v_{n})italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT : blackboard_R → blackboard_R , italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ↦ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT )

that takes as an input the local voltage vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, and provides as an output the approximated ORPF solution. Given any voltage vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, ϕn(vn)subscriptitalic-ϕ𝑛subscript𝑣𝑛\phi_{n}(v_{n})italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) represents an approximation of the reactive power that the DER at node n𝑛nitalic_n would inject if its voltage is vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and the network is operating at a solution of (2). The graph of ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, namely, points of the form (vn,ϕn(vn))subscript𝑣𝑛subscriptitalic-ϕ𝑛subscript𝑣𝑛(v_{n},\phi_{n}(v_{n}))( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ), consists then of the ORPF solutions’ surrogates and describes desirable network configurations. The second stage, termed the control stage, aims to design local control rules that steer the network to the aforesaid desirable configurations. That is, the controller equilibrium points are determined by the ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT’s and, for this reason, the ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT’s are hereafter called equilibrium functions. We introduce first in Section III the local control scheme and derive conditions ensuring system stability. We build on these conditions in Section IV to guide the learning of {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT to promote the efficient operation of the DG.

III Provably Stable Local Volt/Var Control

Here, we propose a local Volt/Var controller and analyze its stability properties. For each n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C, given the equilibrium function ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, consider the reactive power update

qn(t+1)=qn(t)+ϵ(ϕn(vn(t))qn(t)).subscript𝑞𝑛𝑡1subscript𝑞𝑛𝑡italic-ϵsubscriptitalic-ϕ𝑛subscript𝑣𝑛𝑡subscript𝑞𝑛𝑡\displaystyle q_{n}(t+1)=q_{n}(t)+\epsilon(\phi_{n}(v_{n}(t))-q_{n}(t)).italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t + 1 ) = italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) + italic_ϵ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) ) - italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) ) . (3)

In (3), the new reactive power setpoint is chosen as a convex combination between the previous one and the equilibrium function evaluated at the current voltage. Rules such as (3) are referred to as incremental, because the updated reactive power setpoint is obtained by adding to the previous one an increment weighted by the stepsize parameter ϵ[0,1]italic-ϵ01\epsilon\in[0,1]italic_ϵ ∈ [ 0 , 1 ]. An equilibrium point of (3), denoted 𝐪𝒞superscriptsubscript𝐪𝒞\mathbf{q}_{\mathcal{C}}^{\sharp}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT, satisfies for n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C

qn=ϕn(vn)superscriptsubscript𝑞𝑛subscriptitalic-ϕ𝑛subscript𝑣𝑛\displaystyle q_{n}^{\sharp}=\phi_{n}(v_{n})italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT = italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) (4a)
vn=vn(𝐪𝒞).superscriptsubscript𝑣𝑛subscript𝑣𝑛superscriptsubscript𝐪𝒞\displaystyle v_{n}^{\sharp}=v_{n}(\mathbf{q}_{\mathcal{C}}^{\sharp}).italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT = italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) . (4b)

That is, (ϕn(vn),vn)subscriptitalic-ϕ𝑛superscriptsubscript𝑣𝑛superscriptsubscript𝑣𝑛(\phi_{n}(v_{n}^{\sharp}),v_{n}^{\sharp})( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) , italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) belongs to the graph of ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and hence is a desirable configuration. It remains now to establish under what conditions the algorithm (3) converges to an equilibrium. The following convergence analysis assumes that uncontrolled variables, namely, 𝐩𝐩\mathbf{p}bold_p and 𝐪subscript𝐪\mathbf{q}_{\mathcal{L}}bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT, assume arbitrary, but fixed in time, values.

As customary in the literature of reactive power control for DGs, e.g., see [11, 12, 15], we rely on the following linearization of the power flow equations to study the stability properties of (3). In general, voltage magnitudes are nonlinear functions of the nodal power injections. Define 𝐑~:=(𝐙~)assign~𝐑~𝐙\tilde{\mathbf{R}}:=\Re(\tilde{\mathbf{Z}})over~ start_ARG bold_R end_ARG := roman_ℜ ( over~ start_ARG bold_Z end_ARG ) and 𝐗~:=(𝐙~)N×Nassign~𝐗~𝐙superscript𝑁𝑁\tilde{\mathbf{X}}:=\Im(\tilde{\mathbf{Z}})\in\mathbb{R}^{N\times N}over~ start_ARG bold_X end_ARG := roman_ℑ ( over~ start_ARG bold_Z end_ARG ) ∈ blackboard_R start_POSTSUPERSCRIPT italic_N × italic_N end_POSTSUPERSCRIPT, by using a first-order Taylor expansion, the power flow equation can be linearized to obtain [10]

𝐯=𝐑~𝐩+𝐗~𝐪+|𝐮^|.𝐯~𝐑𝐩~𝐗𝐪^𝐮\displaystyle\mathbf{v}=\tilde{\mathbf{R}}\mathbf{p}+\tilde{\mathbf{X}}\mathbf% {q}+|\hat{\mathbf{u}}|.bold_v = over~ start_ARG bold_R end_ARG bold_p + over~ start_ARG bold_X end_ARG bold_q + | over^ start_ARG bold_u end_ARG | . (5)

Also, the matrices 𝐑~~𝐑\tilde{\mathbf{R}}over~ start_ARG bold_R end_ARG and 𝐗~~𝐗\tilde{\mathbf{X}}over~ start_ARG bold_X end_ARG can be decomposed according to the former partition, yielding

𝐑~=[𝐑𝐑𝐑𝐑],𝐗~=[𝐗𝐗𝐗𝐗],formulae-sequence~𝐑matrix𝐑subscript𝐑superscriptsubscript𝐑topsubscript𝐑~𝐗matrix𝐗subscript𝐗superscriptsubscript𝐗topsubscript𝐗\displaystyle\tilde{\mathbf{R}}=\begin{bmatrix}\mathbf{R}&\mathbf{R}_{\mathcal% {L}}\\ \mathbf{R}_{\mathcal{L}}^{\top}&\mathbf{R}_{\mathcal{L}\mathcal{L}}\end{% bmatrix},\quad\tilde{\mathbf{X}}=\begin{bmatrix}\mathbf{X}&\mathbf{X}_{% \mathcal{L}}\\ \mathbf{X}_{\mathcal{L}}^{\top}&\mathbf{X}_{\mathcal{L}\mathcal{L}}\end{% bmatrix},over~ start_ARG bold_R end_ARG = [ start_ARG start_ROW start_CELL bold_R end_CELL start_CELL bold_R start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL bold_R start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT end_CELL start_CELL bold_R start_POSTSUBSCRIPT caligraphic_L caligraphic_L end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ] , over~ start_ARG bold_X end_ARG = [ start_ARG start_ROW start_CELL bold_X end_CELL start_CELL bold_X start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL bold_X start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT end_CELL start_CELL bold_X start_POSTSUBSCRIPT caligraphic_L caligraphic_L end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ] ,

with 𝐑,𝐗0succeeds𝐑𝐗0\mathbf{R},\mathbf{X}\succ 0bold_R , bold_X ≻ 0 [15]. From (5), voltage magnitudes become functions exclusively of 𝐪𝒞subscript𝐪𝒞\mathbf{q}_{\mathcal{C}}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT:

𝐯(𝐪𝒞)=[𝐗𝐗]𝐪𝒞+𝐯^,𝐯subscript𝐪𝒞matrix𝐗superscriptsubscript𝐗topsubscript𝐪𝒞^𝐯\displaystyle\mathbf{v}(\mathbf{q}_{\mathcal{C}})=\begin{bmatrix}\mathbf{X}\\ \mathbf{X}_{\mathcal{L}}^{\top}\end{bmatrix}\mathbf{q}_{\mathcal{C}}+\hat{% \mathbf{v}},bold_v ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) = [ start_ARG start_ROW start_CELL bold_X end_CELL end_ROW start_ROW start_CELL bold_X start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT + over^ start_ARG bold_v end_ARG , (6)

where

𝐯^^𝐯\displaystyle\hat{\mathbf{v}}over^ start_ARG bold_v end_ARG :=[𝐯^𝒞𝐯^]=[𝐗𝐗]𝐪+𝐑~𝐩+|𝐮^|.assignabsentmatrixsubscript^𝐯𝒞subscript^𝐯matrixsubscript𝐗subscript𝐗subscript𝐪~𝐑𝐩^𝐮\displaystyle:=\begin{bmatrix}\hat{\mathbf{v}}_{\mathcal{C}}\\ \hat{\mathbf{v}}_{\mathcal{L}}\end{bmatrix}=\begin{bmatrix}\mathbf{X}_{% \mathcal{L}}\\ \mathbf{X}_{\mathcal{L}\mathcal{L}}\end{bmatrix}\mathbf{q}_{\mathcal{L}}+% \tilde{\mathbf{R}}\mathbf{p}+|\hat{\mathbf{u}}|.:= [ start_ARG start_ROW start_CELL over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ] = [ start_ARG start_ROW start_CELL bold_X start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL bold_X start_POSTSUBSCRIPT caligraphic_L caligraphic_L end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ] bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT + over~ start_ARG bold_R end_ARG bold_p + | over^ start_ARG bold_u end_ARG | . (7)

Collecting the {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT in the vector-valued function ϕbold-italic-ϕ\boldsymbol{\phi}bold_italic_ϕ, and adopting the linearization (6), the system dynamics can be described in compact form as

𝐪𝒞(t+1)=𝐪𝒞(t)+ϵ(ϕ(𝐯𝒞(t))𝐪𝒞(t)),subscript𝐪𝒞𝑡1subscript𝐪𝒞𝑡italic-ϵbold-italic-ϕsubscript𝐯𝒞𝑡subscript𝐪𝒞𝑡\displaystyle\mathbf{q}_{\mathcal{C}}(t+1)=\mathbf{q}_{\mathcal{C}}(t)+% \epsilon(\boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}(t))-\mathbf{q}_{\mathcal{C% }}(t)),bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t + 1 ) = bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) + italic_ϵ ( bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) ) - bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) ) , (8a)
𝐯𝒞(t+1)=𝐗𝐪𝒞(t+1)+𝐯^𝒞,subscript𝐯𝒞𝑡1subscript𝐗𝐪𝒞𝑡1subscript^𝐯𝒞\displaystyle\mathbf{v}_{\mathcal{C}}(t+1)=\mathbf{X}\mathbf{q}_{\mathcal{C}}(% t+1)+\hat{\mathbf{v}}_{\mathcal{C}},bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t + 1 ) = bold_Xq start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t + 1 ) + over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT , (8b)

where it is tacitly assumed that, at the timescale of the above iterates, the exogenous variables (𝐩,𝐪)𝐩subscript𝐪(\mathbf{p},\mathbf{q}_{\mathcal{L}})( bold_p , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT ) remain constant, resulting in a constant term 𝐯^𝒞subscript^𝐯𝒞\hat{\mathbf{v}}_{\mathcal{C}}over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT from (7). Let (𝐪𝒞,𝐯𝒞)superscriptsubscript𝐪𝒞superscriptsubscript𝐯𝒞(\mathbf{q}_{\mathcal{C}}^{\sharp},\mathbf{v}_{\mathcal{C}}^{\sharp})( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) an equilibrium point of (8). By definition, it must satisfy

𝐪𝒞superscriptsubscript𝐪𝒞\displaystyle\mathbf{q}_{\mathcal{C}}^{\sharp}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT =ϕ(𝐯𝒞),absentbold-italic-ϕsuperscriptsubscript𝐯𝒞\displaystyle=\boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}^{\sharp}),= bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) , (9a)
𝐯𝒞superscriptsubscript𝐯𝒞\displaystyle\mathbf{v}_{\mathcal{C}}^{\sharp}bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT =𝐗𝐪𝒞+𝐯^𝒞.absentsuperscriptsubscript𝐗𝐪𝒞subscript^𝐯𝒞\displaystyle=\mathbf{X}\mathbf{q}_{\mathcal{C}}^{\sharp}+\hat{\mathbf{v}}_{% \mathcal{C}}.= bold_Xq start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT + over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT . (9b)

In particular, it holds that qn=ϕn(vn)superscriptsubscript𝑞𝑛subscriptitalic-ϕ𝑛superscriptsubscript𝑣𝑛q_{n}^{\sharp}=\phi_{n}(v_{n}^{\sharp})italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT = italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) for each n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C, see (4), showing that the functions {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT describe all the possible equilibrium points of (8a).

Define the feasible set of reactive power injections 𝒬:=×n𝒞𝒬n\mathcal{Q}:=\times_{n\in\mathcal{C}}\mathcal{Q}_{n}caligraphic_Q := × start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT caligraphic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, with 𝒬n={qn:qmin,nqnqmax,n}subscript𝒬𝑛conditional-setsubscript𝑞𝑛subscript𝑞𝑛subscript𝑞𝑛subscript𝑞𝑛\mathcal{Q}_{n}=\{q_{n}:q_{\min,n}\leq q_{n}\leq q_{\max,n}\}caligraphic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = { italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT : italic_q start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT ≤ italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≤ italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT }. To ensure the stability of the proposed local algorithm, we require each ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT to meet the following properties:

  1. C1)

    ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is Lipschitz, i.e., there exists Ln<subscript𝐿𝑛L_{n}<\inftyitalic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT < ∞ such that |ϕn(v)ϕn(v)|Ln|vv|subscriptitalic-ϕ𝑛𝑣subscriptitalic-ϕ𝑛superscript𝑣subscript𝐿𝑛𝑣superscript𝑣|\phi_{n}(v)-\phi_{n}(v^{\prime})|\leq L_{n}|v-v^{\prime}|| italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v ) - italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | ≤ italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | italic_v - italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT |, for all v,v𝑣superscript𝑣v,v^{\prime}\in\mathbb{R}italic_v , italic_v start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ blackboard_R;

  2. C2)

    ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is nonincreasing in vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT;

  3. C3)

    range(ϕn)𝒬nrangesubscriptitalic-ϕ𝑛subscript𝒬𝑛\text{range}(\phi_{n})\subseteq\mathcal{Q}_{n}range ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ⊆ caligraphic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, i.e., ϕn:𝒬n:subscriptitalic-ϕ𝑛subscript𝒬𝑛\phi_{n}:\mathbb{R}\rightarrow\mathcal{Q}_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT : blackboard_R → caligraphic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT.

The role of conditions C1) – C3) will be clear from the following formal results proved in the Appendix. The first one characterizes the equilibrium points of (8).

Proposition III.1.

(Feasibility of the reactive power update and uniqueness of the equilibrium): Let {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT satisfy the conditions C1) – C3), and assume 𝐪C(0)𝒬subscript𝐪𝐶0𝒬\mathbf{q}_{C}(0)\in\mathcal{Q}bold_q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( 0 ) ∈ caligraphic_Q. The reactive power update (3) is feasible, i.e., 𝐪C(t)𝒬subscript𝐪𝐶𝑡𝒬\mathbf{q}_{C}(t)\in\mathcal{Q}bold_q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( italic_t ) ∈ caligraphic_Q, t0𝑡0t\geq 0italic_t ≥ 0. Moreover, the system (8) has an unique equilibrium point (𝐪𝒞,𝐯𝒞)superscriptsubscript𝐪𝒞superscriptsubscript𝐯𝒞(\mathbf{q}_{\mathcal{C}}^{\sharp},\mathbf{v}_{\mathcal{C}}^{\sharp})( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ).

The next result characterizes the stability properties of the equilibrium point identified in Proposition III.1.

Proposition III.2.

(Global asymptotic stability of the control rule (3)): Let {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT satisfy the conditions C1) – C3), and assume 𝐪C(0)𝒬subscript𝐪𝐶0𝒬\mathbf{q}_{C}(0)\in\mathcal{Q}bold_q start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( 0 ) ∈ caligraphic_Q. Define L=maxn𝒞Ln𝐿subscript𝑛𝒞subscript𝐿𝑛L=\max_{n\in\mathcal{C}}L_{n}italic_L = roman_max start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. If ϵitalic-ϵ\epsilonitalic_ϵ is such that

0<ϵ<min{1,2(𝐗L+1)2}0italic-ϵ12superscriptnorm𝐗𝐿120<\epsilon<\min\Big{\{}1,\frac{2}{(\|\mathbf{X}\|L+1)^{2}}\Big{\}}0 < italic_ϵ < roman_min { 1 , divide start_ARG 2 end_ARG start_ARG ( ∥ bold_X ∥ italic_L + 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG } (10)

then the equilibrium of (8) is globally asymptotically stable.

Proposition III.2 indicates that, as long as the ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT’s meet the conditions C1) – C3), one can always find ϵ>0italic-ϵ0\epsilon>0italic_ϵ > 0 so that (𝐪𝒞,𝐯𝒞)subscript𝐪𝒞subscript𝐯𝒞(\mathbf{q}_{\mathcal{C}},\mathbf{v}_{\mathcal{C}})( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) converges to the unique equilibrium point (𝐪𝒞,𝐯𝒞)superscriptsubscript𝐪𝒞superscriptsubscript𝐯𝒞(\mathbf{q}_{\mathcal{C}}^{\sharp},\mathbf{v}_{\mathcal{C}}^{\sharp})( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) under the reactive power update rule (3). Note that, in condition (10), 𝐗norm𝐗\|\mathbf{X}\|∥ bold_X ∥ is fully determined by the DG and L𝐿Litalic_L can be computed once the functions {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT have been selected. Because 𝐪subscript𝐪\mathbf{q}_{\mathcal{L}}bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT and 𝐩𝐩\mathbf{p}bold_p are fixed, the convergence of 𝐪𝒞subscript𝐪𝒞\mathbf{q}_{\mathcal{C}}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT leads, cf. (6), also to the global asymptotic convergence of 𝐯𝐯\mathbf{v}bold_v. Finally, we note that the proposed reactive power update rule (3) is a generalized version of the local control scheme proposed in [10] which considers only linear functions (instead of arbitrary nonlinear {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT satisfying C1) – C3)).

Remark III.3.

(Global vs. local asymptotic stability of the equilibrium): In our previous work [29], the equilibrium point of (8) under the reactive power update rule (3) is shown to be locally asymptotically stable if 0<ϵ<min{1,2𝐗L+1}0italic-ϵ12norm𝐗𝐿10<\epsilon<\min\{1,\frac{2}{\|\mathbf{X}\|L+1}\}0 < italic_ϵ < roman_min { 1 , divide start_ARG 2 end_ARG start_ARG ∥ bold_X ∥ italic_L + 1 end_ARG }. The previous claim roughly implies that if 𝐪𝒞(0)subscript𝐪𝒞0\mathbf{q}_{\mathcal{C}}(0)bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( 0 ) is close enough to 𝐪𝒞superscriptsubscript𝐪𝒞\mathbf{q}_{\mathcal{C}}^{\sharp}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT, then it converges to 𝐪𝒞superscriptsubscript𝐪𝒞\mathbf{q}_{\mathcal{C}}^{\sharp}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT. Our result here extends the stability properties from local to global, at the cost of reducing the selection range of ϵitalic-ϵ\epsilonitalic_ϵ, as 2𝐗L+1>2(𝐗L+1)22norm𝐗𝐿12superscriptnorm𝐗𝐿12\frac{2}{\|\mathbf{X}\|L+1}>\frac{2}{(\|\mathbf{X}\|L+1)^{2}}divide start_ARG 2 end_ARG start_ARG ∥ bold_X ∥ italic_L + 1 end_ARG > divide start_ARG 2 end_ARG start_ARG ( ∥ bold_X ∥ italic_L + 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. \bullet

Remark III.4.

(Non-incremental vs. incremental reactive power update rules): Many works in the literature consider local Volt/Var control schemes of the form [13, 14, 15]

qn(t+1)=φn(vn(t)).subscript𝑞𝑛𝑡1subscript𝜑𝑛subscript𝑣𝑛𝑡q_{n}(t+1)=\varphi_{n}(v_{n}(t)).italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t + 1 ) = italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) ) . (11)

Reactive power update rules such as (11), where the new reactive power setpoints are determined based on the local voltage without explicitly exploiting a memory of past setpoints, are referred to as non-incremental [32]. The equilibrium points of (11) satisfy

qnsubscript𝑞𝑛\displaystyle q_{n}italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =φn(vn),absentsubscript𝜑𝑛subscript𝑣𝑛\displaystyle=\varphi_{n}(v_{n}),= italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ,
vnsubscript𝑣𝑛\displaystyle v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT =vn(𝐪𝒞),absentsubscript𝑣𝑛subscript𝐪𝒞\displaystyle=v_{n}(\mathbf{q}_{\mathcal{C}}),= italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) ,

i.e., φn(vn)subscript𝜑𝑛subscript𝑣𝑛\varphi_{n}(v_{n})italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) plays the double role of the control function and the equilibrium function. Thus, even the control rule

qn(t+1)subscript𝑞𝑛𝑡1\displaystyle q_{n}(t+1)italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t + 1 ) =ϕn(vn(t)).absentsubscriptitalic-ϕ𝑛subscript𝑣𝑛𝑡\displaystyle=\phi_{n}(v_{n}(t)).= italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) ) . (12)

looks appealing for our framework because its equilibria are determined by ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, this rule corresponds to ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1 in (3). From the proof of Proposition III.2, one can show that the algorithm (12) is globally asymptotically stable if the equilibrium functions meet not only C1) – C3) but also

𝐗L<21.norm𝐗𝐿21\displaystyle\|\mathbf{X}\|L<\sqrt{2}-1.∥ bold_X ∥ italic_L < square-root start_ARG 2 end_ARG - 1 . (13)

The former condition bounds the slope of the {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT and appears often in the literature [29, 16, 10]. This means that, to ensure the stability of (11), we need to restrict the search space of potential candidates of {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT, thus risking a degradation in system performance in terms of the optimality gap at the equilibrium. The aforementioned restriction motivates adopting an incremental control like (3). \bullet

IV Learning Equilibrium Functions For Efficient Network Operation

Having established the conditions on equilibrium functions for system stability, here we lay out a data-driven approach to synthesize the functions {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT to improve operational efficiency of the DG. Specifically, our goal is to learn local equilibrium functions {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT under which the system equilibrium 𝐪𝒞(𝐩,𝐪)superscriptsubscript𝐪𝒞𝐩subscript𝐪\mathbf{q}_{\mathcal{C}}^{\sharp}(\mathbf{p},\mathbf{q}_{\mathcal{L}})bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ( bold_p , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT ) is as close as possible to the ORPF problem solution 𝐪𝒞(𝐩,𝐪)superscriptsubscript𝐪𝒞𝐩subscript𝐪\mathbf{q}_{\mathcal{C}}^{\star}(\mathbf{p},\mathbf{q}_{\mathcal{L}})bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ( bold_p , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT ). The learning process consists of the following steps. First, given that the solution of (2) depends on (𝐩,𝐪)𝐩subscript𝐪(\mathbf{p},\mathbf{q}_{\mathcal{L}})( bold_p , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT ), we build a set {(𝐩k,𝐪k)}k=1Ksuperscriptsubscriptsuperscript𝐩𝑘superscriptsubscript𝐪𝑘𝑘1𝐾\{(\mathbf{p}^{k},\mathbf{q}_{\mathcal{L}}^{k})\}_{k=1}^{K}{ ( bold_p start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT of K𝐾Kitalic_K load-generation scenarios. One can obtain the aforementioned scenarios via random sampling from assumed probability distributions, historical data, or from forecasted conditions for a look-ahead period. Second, we solve the ORPF problem (2) for these K𝐾Kitalic_K scenarios to obtain a labeled data set of corresponding minimizers 𝒟={(𝐪𝒞,k,𝐯𝒞,k)}k=1K𝒟superscriptsubscriptsuperscriptsubscript𝐪𝒞𝑘superscriptsubscript𝐯𝒞𝑘𝑘1𝐾\mathcal{D}=\{(\mathbf{q}_{\mathcal{C},k}^{\star},\mathbf{v}_{\mathcal{C},k}^{% \star})\}_{k=1}^{K}caligraphic_D = { ( bold_q start_POSTSUBSCRIPT caligraphic_C , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT, where the parametric dependencies are omitted for notational ease. Third, the entries for these minimizers are then separated for each n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C to obtain data sets of the form 𝒟n={(vn,k,qn,k)}k=1Ksubscript𝒟𝑛superscriptsubscriptsuperscriptsubscript𝑣𝑛𝑘superscriptsubscript𝑞𝑛𝑘𝑘1𝐾\mathcal{D}_{n}=\{(v_{n,k}^{\star},q_{n,k}^{\star})\}_{k=1}^{K}caligraphic_D start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = { ( italic_v start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT , italic_q start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT, and each equilibrium function ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is then trained by solving

minϕnsubscriptsubscriptitalic-ϕ𝑛\displaystyle\min_{\phi_{n}}\ roman_min start_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT k=1K|qn,kϕn(vn,k)|2superscriptsubscript𝑘1𝐾superscriptsuperscriptsubscript𝑞𝑛𝑘subscriptitalic-ϕ𝑛superscriptsubscript𝑣𝑛𝑘2\displaystyle~{}\sum_{k=1}^{K}|q_{n,k}^{\star}-\phi_{n}(v_{n,k}^{\star})|^{2}∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT | italic_q start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT - italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (14)
s.t.formulae-sequencest\displaystyle\mathrm{s.t.}\ roman_s . roman_t . ϕnmeets the conditions C1) – C3).subscriptitalic-ϕ𝑛meets the conditions C1) – C3)\displaystyle~{}\phi_{n}~{}\text{meets the conditions C1) -- C3)}.italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT meets the conditions C1) – C3) .

Typical approaches to solve (14) involves restricting the search space for function ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT via convenient parameterization leading to approaches such as polynomial regression and neural network approximation methods. Here, we adopt the latter. Enforcing the properties C1) – C3) is, in general, not trivial and depends on many aspects, e.g., the number of considered layers and the used activation functions. In particular, designing monotone neural networks might require additional considerations. Approaches to do so include, for instance, structure-based [33], gradient-constrained [34], and verification-based [35] methods. In the following, we provide a single-hidden-layer neural network design framework that achieves C1) – C3) and uses the Rectified Linear Unit (ReLU) activation function

ReLU(x)=max(0,x).ReLU𝑥0𝑥{\rm ReLU}(x)=\max(0,x).roman_ReLU ( italic_x ) = roman_max ( 0 , italic_x ) .

Note however that, in principle, any continuous and monotonic activation function could be used in our framework, e.g., the Sigmoid or the Tanh333Sigmoid(x)=11+ex,Tanh(x)=exexex+exformulae-sequenceSigmoid𝑥11superscript𝑒𝑥Tanh𝑥superscript𝑒𝑥superscript𝑒𝑥superscript𝑒𝑥superscript𝑒𝑥{\rm Sigmoid}(x)=\frac{1}{1+e^{-x}},~{}{\rm Tanh}(x)=\frac{e^{x}-e^{-x}}{e^{x}% +e^{-x}}roman_Sigmoid ( italic_x ) = divide start_ARG 1 end_ARG start_ARG 1 + italic_e start_POSTSUPERSCRIPT - italic_x end_POSTSUPERSCRIPT end_ARG , roman_Tanh ( italic_x ) = divide start_ARG italic_e start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT - italic_e start_POSTSUPERSCRIPT - italic_x end_POSTSUPERSCRIPT end_ARG start_ARG italic_e start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT - italic_x end_POSTSUPERSCRIPT end_ARG..

Each equilibrium function ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is then modeled using a single-hidden-layer neural network 𝖭(x)𝖭𝑥\mathsf{N}(x)sansserif_N ( italic_x ) as

ϕn(x)=qmax,nReLU(\displaystyle\phi_{n}(x)=q_{\max,n}-{\rm ReLU}(italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT - roman_ReLU ( qmax,n𝖭(x))\displaystyle q_{\max,n}-\mathsf{N}(x))italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT - sansserif_N ( italic_x ) )
+ReLU(qmin,n𝖭(x)),ReLUsubscript𝑞𝑛𝖭𝑥\displaystyle+{\rm ReLU}(q_{\min,n}-\mathsf{N}(x)),+ roman_ReLU ( italic_q start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT - sansserif_N ( italic_x ) ) , (15)

where

𝖭(x)=h=1HwhReLU(xbh)+β,𝖭𝑥superscriptsubscript1𝐻subscript𝑤ReLU𝑥subscript𝑏𝛽\displaystyle\mathsf{N}(x)=\sum_{h=1}^{H}w_{h}{\rm ReLU}(x-b_{h})+\beta,sansserif_N ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_h = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_ReLU ( italic_x - italic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) + italic_β , (16)

with whsubscript𝑤w_{h}italic_w start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT and bhsubscript𝑏b_{h}italic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT being the weight and bias of the hhitalic_h-th neuron unit, respectively; β𝛽\betaitalic_β being an additional bias term applied in the output layer; and H𝐻Hitalic_H is the number of neuron units in the hidden layer. Note that conditions C1) and C3) for ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT are automatically met because of the Lipschitzness of the ReLU function and because the output of the neural network ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is constrained to the set 𝒬nsubscript𝒬𝑛\mathcal{Q}_{n}caligraphic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, cf. (IV). Condition C2) is instead encoded by the next result.

Refer to caption
Figure 1: Flowchart of the proposed data-driven control framework.
Proposition IV.1.

(Universal approximation of Lipschitz nonincreasing function using ReLU activation function): Consider the single-hidden-layer neural network (16), and reorder the neuron units such that b1b2bHsubscript𝑏1subscript𝑏2subscript𝑏𝐻b_{1}\leq b_{2}\leq\cdots\leq b_{H}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ ⋯ ≤ italic_b start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT. 𝖭𝖭\mathsf{N}sansserif_N is non-increasing if and only if

j=1Jwj0,J{1,2,,H}.formulae-sequencesuperscriptsubscript𝑗1𝐽subscript𝑤𝑗0for-all𝐽12𝐻\displaystyle\sum_{j=1}^{J}w_{j}\leq 0,\quad\forall J\in\{1,2,\dots,H\}.∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≤ 0 , ∀ italic_J ∈ { 1 , 2 , … , italic_H } . (17)

Furthermore, for any Lipschitz non-increasing function g::𝑔g:\mathbb{R}\rightarrow\mathbb{R}italic_g : blackboard_R → blackboard_R and given any compact domain 𝒳𝒳\mathcal{X}\in\mathbb{R}caligraphic_X ∈ blackboard_R and η>0𝜂0\eta>0italic_η > 0, there exist H𝐻Hitalic_H, whsubscript𝑤w_{h}italic_w start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, bhsubscript𝑏b_{h}italic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, and β𝛽\betaitalic_β such that |𝖭(x)g(x)|η𝖭𝑥𝑔𝑥𝜂|\mathsf{N}(x)-g(x)|\leq\eta| sansserif_N ( italic_x ) - italic_g ( italic_x ) | ≤ italic_η for all x𝒳𝑥𝒳x\in\mathcal{X}italic_x ∈ caligraphic_X.

Its proof can be found in the Appendix, and note that the monotonicity of 𝖭𝖭\mathsf{N}sansserif_N remains after the projection (IV). This design imposes weight constraints (17) and output constraints (IV) on a single-hidden-layer neural network (16) so that it meets the conditions C1) – C3) by construction. Leveraging the universal approximation property, the optimization problem (14) is then equivalent to the parameterized formulation

min𝐰,𝐛,βsubscript𝐰𝐛𝛽\displaystyle\min_{\mathbf{w},\mathbf{b},\beta}\ roman_min start_POSTSUBSCRIPT bold_w , bold_b , italic_β end_POSTSUBSCRIPT k=1K|qn,kϕn(vn,k)|2superscriptsubscript𝑘1𝐾superscriptsuperscriptsubscript𝑞𝑛𝑘subscriptitalic-ϕ𝑛superscriptsubscript𝑣𝑛𝑘2\displaystyle~{}\sum_{k=1}^{K}|q_{n,k}^{\star}-\phi_{n}(v_{n,k}^{\star})|^{2}∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT | italic_q start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT - italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (18)
s.t.formulae-sequencest\displaystyle\mathrm{s.t.}\ roman_s . roman_t . (IV),(16),(17)italic-(IVitalic-)italic-(16italic-)italic-(17italic-)\displaystyle\eqref{eq:enc_c3},~{}\eqref{eq:single-layer-NN},~{}\eqref{eq:enc_% c2}italic_( italic_) , italic_( italic_) , italic_( italic_)

which can be solved to local optima using suitable renditions of (stochastic) gradient descent prevalent for neural network training. Also, exploiting the fact that the ReLU function is used as an activation function, the Lipschitz constant Lnsubscript𝐿𝑛L_{n}italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT of each ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT can be easily computed, see (17), as

Ln=maxJ{1,..,H}|j=1Jwj|.\displaystyle L_{n}=\max_{J\in\{1,..,H\}}\Big{|}\sum_{j=1}^{J}w_{j}\Big{|}.italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = roman_max start_POSTSUBSCRIPT italic_J ∈ { 1 , . . , italic_H } end_POSTSUBSCRIPT | ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | .

The two stages of the proposed framework are summarized in Algorithm 1. Fig. 1 provides a flowchart to illustrate the overall scheme.

Algorithm 1 Local Volt/Var Control Framework
1:
2:Historical data in the form of K𝐾Kitalic_K load-generation scenarios {(𝐩k,𝐪k)}k=1Ksuperscriptsubscriptsuperscript𝐩𝑘superscriptsubscript𝐪𝑘𝑘1𝐾\{(\mathbf{p}^{k},\mathbf{q}_{\mathcal{L}}^{k})\}_{k=1}^{K}{ ( bold_p start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT , bold_q start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT
3:Solve the ORPF problem (2) for each load-generation scenario, obtain the optimal reactive powers {𝐪𝒞,k}k=1Ksuperscriptsubscriptsubscriptsuperscript𝐪𝑘𝒞𝑘1𝐾\{\mathbf{q}^{\star,k}_{\mathcal{C}}\}_{k=1}^{K}{ bold_q start_POSTSUPERSCRIPT ⋆ , italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT
4:Compute the optimal voltages {𝐯,k}k=1Ksuperscriptsubscriptsuperscript𝐯𝑘𝑘1𝐾\{\mathbf{v}^{\star,k}\}_{k=1}^{K}{ bold_v start_POSTSUPERSCRIPT ⋆ , italic_k end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT by solving the power flow equations (1) with the power injections {(𝐩k,𝐪k,𝐪𝒞,k)}k=1Ksuperscriptsubscriptsuperscript𝐩𝑘subscriptsuperscript𝐪𝑘subscriptsuperscript𝐪𝑘𝒞𝑘1𝐾\{(\mathbf{p}^{k},\mathbf{q}^{k}_{\mathcal{L}},\mathbf{q}^{\star,k}_{\mathcal{% C}})\}_{k=1}^{K}{ ( bold_p start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT , bold_q start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_L end_POSTSUBSCRIPT , bold_q start_POSTSUPERSCRIPT ⋆ , italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT
5:Build the data set of optimal reactive powers and voltages 𝒟={(𝐪𝒞,k,𝐯𝒞,k)}k=1K𝒟superscriptsubscriptsuperscriptsubscript𝐪𝒞𝑘superscriptsubscript𝐯𝒞𝑘𝑘1𝐾\mathcal{D}=\{(\mathbf{q}_{\mathcal{C}}^{\star,k},\mathbf{v}_{\mathcal{C}}^{% \star,k})\}_{k=1}^{K}caligraphic_D = { ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ , italic_k end_POSTSUPERSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ , italic_k end_POSTSUPERSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT
6:Separate the entries of these minimizers to obtain, for each n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C, data sets of the form 𝒟n={(vn,k,qn,k)}k=1Ksubscript𝒟𝑛superscriptsubscriptsuperscriptsubscript𝑣𝑛𝑘superscriptsubscript𝑞𝑛𝑘𝑘1𝐾\mathcal{D}_{n}=\{(v_{n}^{\star,k},q_{n}^{\star,k})\}_{k=1}^{K}caligraphic_D start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = { ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ , italic_k end_POSTSUPERSCRIPT , italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ , italic_k end_POSTSUPERSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT
7:Learn the equilibrium functions {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT by solving (18)
1:
2:Learned equilibrium functions {ϕn}n𝒞subscriptsubscriptitalic-ϕ𝑛𝑛𝒞\{\phi_{n}\}_{n\in\mathcal{C}}{ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT, network parameter 𝐗𝐗\mathbf{X}bold_X, constant L=maxn𝒞Ln𝐿subscript𝑛𝒞subscript𝐿𝑛L=\max_{n\in\mathcal{C}}L_{n}italic_L = roman_max start_POSTSUBSCRIPT italic_n ∈ caligraphic_C end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, stepsize parameter ϵitalic-ϵ\epsilonitalic_ϵ selected according to (10)
3:
4:Measuring its local voltage magnitude vn(t)subscript𝑣𝑛𝑡v_{n}(t)italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t )
5:Updating the reactive power as per (3)
Remark IV.2.

(Enhancing the capability to regulate voltages when they are not within desired limits through pseudo data points): In the above exposition, the data set points are solutions to the ORPF problem (2), which are subject to the constraint (2b). That is, for each n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C, the equilibrium function ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is trained only using data points such that vmin,nvn,kvmax,nsubscript𝑣𝑛superscriptsubscript𝑣𝑛𝑘subscript𝑣𝑛v_{\min,n}\leq v_{n,k}^{\star}\leq v_{\max,n}italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT ≤ italic_v start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ≤ italic_v start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT, i.e., not when the voltages exceed the limits. Nevertheless, in practical implementation, a DG might experience load-generation scenarios in which (2) is infeasible and the voltages do not meet the desired constraints. Engineering considerations suggest that in such cases the available reactive power capability should be maximally used to alleviate the voltage violations as much as possible. Namely, for each n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C, if vn<vmin,nsubscript𝑣𝑛subscript𝑣𝑛v_{n}<v_{\min,n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT < italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT (vn>vmax,nsubscript𝑣𝑛subscript𝑣𝑛v_{n}>v_{\max,n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT > italic_v start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT), then qn=qmax,nsubscript𝑞𝑛subscript𝑞𝑛q_{n}=q_{\max,n}italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT (qn=qmin,n)subscript𝑞𝑛subscript𝑞𝑛(q_{n}=q_{\min,n})( italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT ), see [10, 14]. To ensure that the learned function ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT meets this condition, we can add a certain number of additional pseudo data points to the data set, e.g., K¯¯𝐾\underline{K}under¯ start_ARG italic_K end_ARG points of the form {(v¯n,k,qmax,n)}k=1K¯superscriptsubscriptsubscript¯𝑣𝑛𝑘subscript𝑞𝑛𝑘1¯𝐾\{(\underline{v}_{n,k},q_{\max,n})\}_{k=1}^{\underline{K}}{ ( under¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT under¯ start_ARG italic_K end_ARG end_POSTSUPERSCRIPT, and K¯¯𝐾\overline{K}over¯ start_ARG italic_K end_ARG points of the form {(v¯n,k,qmin,n)}k=1K¯superscriptsubscriptsubscript¯𝑣𝑛𝑘subscript𝑞𝑛𝑘1¯𝐾\{(\overline{v}_{n,k},q_{\min,n})\}_{k=1}^{\overline{K}}{ ( over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT over¯ start_ARG italic_K end_ARG end_POSTSUPERSCRIPT, with v¯n,kvmin,n,v¯n,kvmax,nformulae-sequencesubscript¯𝑣𝑛𝑘subscript𝑣𝑛subscript¯𝑣𝑛𝑘subscript𝑣𝑛\underline{v}_{n,k}\leq v_{\min,n},\overline{v}_{n,k}\geq v_{\max,n}under¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT ≤ italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT , over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT ≥ italic_v start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT. These points could be uniformly spaced or randomly sampled. Here we adopt the former method, illustrated in Fig. 2. \bullet

Refer to caption
(a) Without pseudo data points
Refer to caption
(b) With pseudo data points
Figure 2: An illustration of the role of the pseudo data points for DER at node n𝑛nitalic_n. Blue and orange points respectively represent true and pseudo data points, while the dark red curves are instances of learned equilibrium functions. Adding pseudo data points helps the equilibrium functions reach maximum reactive power compensation capability when voltage exceeds the limits.
Remark IV.3.

(Deep neural network parameterization of non-increasing function via gradient penalization): Though the single-hidden-layer neural network parameterization of (16) achieves universal approximation, it requires the “width” of the neural networks, i.e., H𝐻Hitalic_H to be sufficiently large. Instead, in certain situations, deeper neural networks could achieve better approximations than the shallower ones, even if they are much narrower [36]. On the other hand, the structure-based restrictions enforced on the neural networks to guarantee monotonicity could in some cases restrict expressibility and lead to unsatisfactory approximation results [35]. To overcome these challenges, we describe a deep neural network parameterization approach by incorporating the monotonicity requirement as a penalization in the cost function of the learning process. Suppose for each n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C, ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is parameterized by a deep neural network, and denote by dϕn(vn)𝑑subscriptitalic-ϕ𝑛subscript𝑣𝑛d\phi_{n}(v_{n})\in\mathbb{R}italic_d italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ∈ blackboard_R the (sub)gradient of ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT with respect to vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. The cost function in (14) is then replaced by

k=1K|qn,kϕn(vn,k)|2+γnmax(0,dϕn(vn,k)),superscriptsubscript𝑘1𝐾superscriptsuperscriptsubscript𝑞𝑛𝑘subscriptitalic-ϕ𝑛superscriptsubscript𝑣𝑛𝑘2subscript𝛾𝑛0𝑑subscriptitalic-ϕ𝑛superscriptsubscript𝑣𝑛𝑘\displaystyle\sum_{k=1}^{K}|q_{n,k}^{\star}-\phi_{n}(v_{n,k}^{\star})|^{2}+% \gamma_{n}\max(0,\;d\phi_{n}(v_{n,k}^{\star})),∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT | italic_q start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT - italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT roman_max ( 0 , italic_d italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ) ) ,

where γn>0subscript𝛾𝑛0\gamma_{n}>0italic_γ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT > 0 is a tuning parameter. During implementation, one can gradually increase γnsubscript𝛾𝑛\gamma_{n}italic_γ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT until ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is verified to be non-increasing. \bullet

Remark IV.4.

(On the comparison with existing reinforcement learning-based approaches): Recent literature has also investigated reinforcement learning approaches for learning stability guaranteed local Volt/Var controllers, e.g., [21, 22]. However, due to the lack of communication in the training phase, the cost function that the whole system seeks to minimize in such settings can only be separable, i.e., the summation of all the local cost functions at each node. Therefore, these approaches generally cannot cope with cost functions such as power losses which shows coupling among nodes. In contrast, since our approach only uses off-line collected data, any type of cost function could, in principle, be considered when solving the ORPF problem (2). \bullet

V Numerical Tests

Refer to caption
Figure 3: The IEEE 37-bus feeder.

We conduct case studies on the IEEE 37-bus feeder. We omit regulators, incorporate five solar generators, and convert it to its single-phase equivalent, see Fig. 3. The feeder has 25 buses with non-zero load, and the five solar generators are the DERs participating in reactive power compensation.

For our experiments, we use minute-based load and solar generation data retrieved from the Pecan Street data set (June 1, 2018) [37]. The first 75 nonzero load buses from the data set are aggregated every 3 loads and normalized to obtain 25 load profiles. Similarly, we obtain 5 solar generation profiles for the active power of DERs. The normalized load profiles for the 24-hour period are scaled so that the demand peak is 1.65 times the nominal load. We synthesize the reactive loads by scaling the active demand to match the power factors of the IEEE 37-bus feeder. Fig. 4 shows the total demand and solar generation across the feeder.

To evaluate the effectiveness of our control framework, it is compared with the standard linear droop control from [10, 14] which dictates

qn(t+1)=ϱn(vn(t))subscript𝑞𝑛𝑡1subscriptitalic-ϱ𝑛subscript𝑣𝑛𝑡\displaystyle q_{n}(t+1)=\varrho_{n}(v_{n}(t))italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t + 1 ) = italic_ϱ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) )
ϱn(vn):={qmax,nvn(t)vmin,n,qmin,nvn(t)vmax,n,cn(vn(t)vmin,n)+qmax,notherwise,assignsubscriptitalic-ϱ𝑛subscript𝑣𝑛casessubscript𝑞𝑛subscript𝑣𝑛𝑡subscript𝑣𝑛subscript𝑞𝑛subscript𝑣𝑛𝑡subscript𝑣𝑛subscript𝑐𝑛subscript𝑣𝑛𝑡subscript𝑣𝑛subscript𝑞𝑛otherwise\displaystyle\varrho_{n}(v_{n}):=\begin{cases}q_{\max,n}&v_{n}(t)\leq v_{\min,% n},\\ q_{\min,n}&v_{n}(t)\geq v_{\max,n},\\ -c_{n}(v_{n}(t)\!-\!v_{\min,n})\!+\!q_{\max,n}&\text{otherwise},\end{cases}italic_ϱ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) := { start_ROW start_CELL italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT end_CELL start_CELL italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) ≤ italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT end_CELL start_CELL italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) ≥ italic_v start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL - italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) - italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT ) + italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT end_CELL start_CELL otherwise , end_CELL end_ROW

where cn=qmax,nqmin,nvmax,nvmin,nsubscript𝑐𝑛subscript𝑞𝑛subscript𝑞𝑛subscript𝑣𝑛subscript𝑣𝑛c_{n}=\frac{q_{\max,n}-q_{\min,n}}{v_{\max,n}-v_{\min,n}}italic_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = divide start_ARG italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT - italic_q start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT end_ARG, and with the the optimized droop control design444In [28], the voltage limits for maximum reactive power provision and absorption are selected as vmin,n=0.9subscript𝑣𝑛0.9v_{\min,n}=0.9italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT = 0.9 p.u. and vmin,n=1.1subscript𝑣𝑛1.1v_{\min,n}=1.1italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT = 1.1 p.u, respectively. To ensure the DERs reach maximum reactive power compensation capability when voltage exceeds the limits, as discussed in Remark IV.2, here they are set to vmin,n=0.95subscript𝑣𝑛0.95v_{\min,n}=0.95italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT = 0.95 p.u. and vmin,n=1.05subscript𝑣𝑛1.05v_{\min,n}=1.05italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT = 1.05. from [28]

qn(t+1)=ρn(vn(t))subscript𝑞𝑛𝑡1subscript𝜌𝑛subscript𝑣𝑛𝑡\displaystyle q_{n}(t+1)=\rho_{n}(v_{n}(t))italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t + 1 ) = italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) )
ρn(vn):={qmax,nvnvmin,n,v¯min,nvnv¯min,nvmin,nqmax,nvmin,n<vn<v¯min,n,0v¯min,nvnv¯max,n,vnv¯max,nv¯max,nv¯max,nqmin,nv¯max,n<vn<vmax,n,qmin,nvnvmax,nassignsubscript𝜌𝑛subscript𝑣𝑛casessubscript𝑞𝑛subscript𝑣𝑛subscript𝑣𝑛subscript¯𝑣𝑛subscript𝑣𝑛subscript¯𝑣𝑛subscript𝑣𝑛subscript𝑞𝑛subscript𝑣𝑛subscript𝑣𝑛subscript¯𝑣𝑛0subscript¯𝑣𝑛subscript𝑣𝑛subscript¯𝑣𝑛subscript𝑣𝑛subscript¯𝑣𝑛subscript¯𝑣𝑛subscript¯𝑣𝑛subscript𝑞𝑛subscript¯𝑣𝑛subscript𝑣𝑛subscript𝑣𝑛subscript𝑞𝑛subscript𝑣𝑛subscript𝑣𝑛\displaystyle\rho_{n}(v_{n}):=\begin{cases}q_{\max,n}&v_{n}\leq v_{\min,n},\\ \frac{\bar{v}_{\min,n}-v_{n}}{\bar{v}_{\min,n}-v_{\min,n}}q_{\max,n}&v_{\min,n% }<v_{n}<\bar{v}_{\min,n},\\ 0&\bar{v}_{\min,n}\leq v_{n}\leq\bar{v}_{\max,n},\\ \frac{v_{n}-\bar{v}_{\max,n}}{\bar{v}_{\max,n}-\bar{v}_{\max,n}}q_{\min,n}&% \bar{v}_{\max,n}<v_{n}<v_{\max,n},\\ q_{\min,n}&v_{n}\geq v_{\max,n}\end{cases}italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) := { start_ROW start_CELL italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT end_CELL start_CELL italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≤ italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL divide start_ARG over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT end_ARG italic_q start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT end_CELL start_CELL italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT < italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT < over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT ≤ italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≤ over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL divide start_ARG italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT - over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT end_ARG italic_q start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT end_CELL start_CELL over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT < italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT < italic_v start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_q start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT end_CELL start_CELL italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≥ italic_v start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT end_CELL end_ROW

where v¯min,nsubscript¯𝑣𝑛\bar{v}_{\min,n}over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT and v¯max,nsubscript¯𝑣𝑛\bar{v}_{\max,n}over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT are parameters satisfying vmin,n<v¯min,nv¯max,n<vmax,nsubscript𝑣𝑛subscript¯𝑣𝑛subscript¯𝑣𝑛subscript𝑣𝑛v_{\min,n}<\bar{v}_{\min,n}\leq\bar{v}_{\max,n}<v_{\max,n}italic_v start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT < over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_min , italic_n end_POSTSUBSCRIPT ≤ over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT < italic_v start_POSTSUBSCRIPT roman_max , italic_n end_POSTSUBSCRIPT which are optimized given the data set prescribing a day-ahead forecast.

Refer to caption
Figure 4: Minute-based data for the total (feeder-wise) solar power generation and active power demand.

V-A The Learning Stage

We considered an optimization problem of the form (2) where the cost is set to

f(𝐪𝒞)𝑓subscript𝐪𝒞\displaystyle f(\mathbf{q}_{\mathcal{C}})italic_f ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) =α𝐯(𝐪𝒞)𝟏+(1α)(𝐪𝐑~𝐪+𝐩𝐑~𝐩),absent𝛼subscriptnorm𝐯subscript𝐪𝒞11𝛼subscriptsuperscript𝐪top~𝐑𝐪superscript𝐩top~𝐑𝐩\displaystyle=\alpha\underbrace{\|\mathbf{v}(\mathbf{q}_{\mathcal{C}})-\mathbf% {1}\|}_{\text{\char 172}}+(1-\alpha)\underbrace{(\mathbf{q}^{\top}\tilde{% \mathbf{R}}\mathbf{q}+\mathbf{p}^{\top}\tilde{\mathbf{R}}\mathbf{p})}_{\text{% \char 173}},= italic_α under⏟ start_ARG ∥ bold_v ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) - bold_1 ∥ end_ARG start_POSTSUBSCRIPT ① end_POSTSUBSCRIPT + ( 1 - italic_α ) under⏟ start_ARG ( bold_q start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG bold_R end_ARG bold_q + bold_p start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT over~ start_ARG bold_R end_ARG bold_p ) end_ARG start_POSTSUBSCRIPT ② end_POSTSUBSCRIPT ,

where ① and ② aims to minimize the voltage deviations and power losses [10], respectively, and the parameter α𝛼\alphaitalic_α trades-off those two objectives. We assume the 5 DERs have uniform generation capabilities, precisely, 𝐪max=0.4×𝟏subscript𝐪0.41\mathbf{q}_{\max}=0.4\times\mathbf{1}bold_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 0.4 × bold_1 MVAR and 𝐪min=𝐪maxsubscript𝐪subscript𝐪\mathbf{q}_{\min}=-\mathbf{q}_{\max}bold_q start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = - bold_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. The voltage limit vectors are set to 𝐯max=1.05×𝟏subscript𝐯1.051\mathbf{v}_{\max}=1.05\times\mathbf{1}bold_v start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1.05 × bold_1 p.u. and 𝐯min=0.95×𝟏subscript𝐯0.951\mathbf{v}_{\min}=0.95\times\mathbf{1}bold_v start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0.95 × bold_1 p.u.

The data set for the learning process is built using the aforesaid power demands and generations obtained from the Pecan Street data, which are intended as day-ahead forecasts. We use the CVX toolbox [38] to solve the ORPF problem (2) with linearized power flow (5). Note however that, one can use any other power flow models to solve the ORPF problem. We add the pseudo data points to the obtained data set as described in Remark IV.2 with K¯=K¯=700¯𝐾¯𝐾700\underline{K}=\overline{K}=700under¯ start_ARG italic_K end_ARG = over¯ start_ARG italic_K end_ARG = 700, which results in a total of 2840284028402840 data points for each DER. We implement the neural network approach according to Proposition IV.1 using TensorFlow 2.7.0 and conduct the training process in Google Colab with a single TPU with 32 GB memory. The number of episodes and the number of neurons H𝐻Hitalic_H are 2000 and 1000, respectively, and the neural networks are trained using the Adam optimizer [39] with the learning rate initialized at 0.01 and decays every 500 steps with a base of 0.5.

Fig. 5 plots the solutions to the ORPF problem (2) for all data profiles, the equilibrium function ϕ32subscriptitalic-ϕ32\phi_{32}italic_ϕ start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT learned with and without pseudo points, the standard droop function ϱ32subscriptitalic-ϱ32\varrho_{32}italic_ϱ start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT, and the optimized droop function ρ32subscript𝜌32\rho_{32}italic_ρ start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT for the DER at node 32 with α=13𝛼13\alpha=\frac{1}{3}italic_α = divide start_ARG 1 end_ARG start_ARG 3 end_ARG. In contrast to the case in which no pseudo points are added in the learning process, the learned equilibrium function with pseudo points reaches maximum reactive power compensation capability when voltage exceeds the limits. We further summarize in Table I the average loss for the whole training data set using the learned equilibrium functions and optimized/standard droop control functions, i.e.,

k=1K𝐪𝒞,k(𝐯𝒞,k)2KC,subscriptsuperscript𝐾𝑘1superscriptnormsuperscriptsubscript𝐪𝒞𝑘superscriptsubscript𝐯𝒞𝑘2𝐾𝐶\displaystyle\frac{\sum^{K}_{k=1}\|\mathbf{q}_{\mathcal{C},k}^{\star}-\square(% \mathbf{v}_{\mathcal{C},k}^{\star})\|^{2}}{K\cdot C},divide start_ARG ∑ start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT ∥ bold_q start_POSTSUBSCRIPT caligraphic_C , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT - □ ( bold_v start_POSTSUBSCRIPT caligraphic_C , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ) ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_K ⋅ italic_C end_ARG , (19)

where \square is ϕbold-italic-ϕ\boldsymbol{\phi}bold_italic_ϕ for the data-based method, 𝝆𝝆\boldsymbol{\rho}bold_italic_ρ for the optimized droop control, and ϱbold-italic-ϱ\boldsymbol{\varrho}bold_italic_ϱ for the standard droop control. The results illustrate the enhanced optimality of the learned equilibrium functions in approximating ORPF solutions compared to benchmarks.

TABLE I: Average loss values for all data profiles
α𝛼\alphaitalic_α Learned equil. func. Opt. droop func. Std. droop func.
0 0.00900.0090\mathbf{0.0090}bold_0.0090 0.0305 0.0438
1/3 0.00600.0060\mathbf{0.0060}bold_0.0060 0.0171 0.0395
1/2 0.00710.0071\mathbf{0.0071}bold_0.0071 0.0429 0.0662
2/3 0.01750.0175\mathbf{0.0175}bold_0.0175 0.0724 0.0852
1 0.03770.0377\mathbf{0.0377}bold_0.0377 0.0878 0.0965

Next, we illustrate the advantage of using the incremental algorithm. Recall that to guarantee the convergence of the non-incremental algorithm, i.e., ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1, one needs to further enforce an additional slope constraint (13) on the learned equilibrium functions, cf. Remark III.4. Fig. 6 shows that this additional slope constraint leads to larger approximation errors of the learned equilibrium functions in fitting the data set (we do not consider the pseudo points during learning here for fairness), and thus degrades the optimality of system performance.

Refer to caption
Figure 5: The orange solid and red dashed curves are, respectively, the learned equilibrium functions with and without considering pseudo data points for the DER at node 32. The blue dash-dotted and green dotted curves are, respectively, the optimized linear droop control function [28] and standard linear droop function [10, 14]. The comparison between the orange solid and red dashed curves illustrates the role of pseudo data points in learning equilibrium functions. The former reaches the maximum reactive power compensation capability when the voltage exceeds the limits, whereas the latter does not.
Refer to caption
Figure 6: Comparison of average training losses (MSE) for all DERs with and and without the additional slope constraint (13) on the equilibrium functions as described in Remark III.4. The mean and standard deviations are evaluated based on five random seeds. This additional slope constraint leads to a larger approximation error of the learned equilibrium functions in fitting the data points.

V-B The Control Stage

We run the following simulations using the learned equilibrium functions for the case α=13𝛼13\alpha=\frac{1}{3}italic_α = divide start_ARG 1 end_ARG start_ARG 3 end_ARG with the pseudo data points considered and assume that 𝐪𝒞(0)=𝟎subscript𝐪𝒞00\mathbf{q}_{\mathcal{C}}(0)=\mathbf{0}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( 0 ) = bold_0. Matpower [40] was used to solve the power flow equation. First, we verify the convergence properties of the proposed reactive power update rule (3) stated in Proposition III.2. Consider the scenario where the load-generation profiles are fixed, Fig. 7 reports the evolution of the DERs’ reactive power setpoints using load-generation profiles of the 695-th minute and considers 120 iterations of (3). For ϵ=0.369italic-ϵ0.369\epsilon=0.369italic_ϵ = 0.369, the reactive power setpoint trajectories converge to their final values, cf. Fig. 7(a), whereas the case ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1 fails, cf. Fig. 7(b). This is consistent with the sufficient condition

0<ϵ<min{1,2(𝐗L+1)2}=0.36910italic-ϵ12superscriptnorm𝐗𝐿120.36910<\epsilon<\min\Big{\{}1,\frac{2}{(\|\mathbf{X}\|L+1)^{2}}\Big{\}}=0.36910 < italic_ϵ < roman_min { 1 , divide start_ARG 2 end_ARG start_ARG ( ∥ bold_X ∥ italic_L + 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG } = 0.3691

derived in Proposition III.2.

Next, we test the proposed data-based control method in a scenario where the load-generation profiles are time varying. Specifically, we obtain load-generation profiles by randomly perturbing (5%) the consumption data used to learn the equilibrium functions. This can be interpreted as having the data from the data set prescribing a day-ahead forecast, whereas their random perturbation acts as the true realization of the load-generation scenarios. These loads and generations are minute-based and we consider 120 iterations of (3) per minute with ϵ=0.369italic-ϵ0.369\epsilon=0.369italic_ϵ = 0.369. Fig. 8 compares the evolution of the maximum/minimum voltages under the proposed data-based control method, the optimized droop control method, the ORPF solutions, and the case where no control action is taken. One can observe that, in contrast to the proposed data-based method, the optimized droop control method induces instability issues within 12:00 and 16:00 causing voltages to oscillate. We further compare in Fig. 9 the distances between the actual reactive power setpoints and ORPF solutions, i.e., 𝐪𝒞𝐪𝒞normsubscript𝐪𝒞superscriptsubscript𝐪𝒞\|\mathbf{q}_{\mathcal{C}}-\mathbf{q}_{\mathcal{C}}^{\star}\|∥ bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ∥. Compared to the benchmarks, the reactive power setpoints during evolution under the proposed data-based method remains much closer to the ORPF solutions, thus leading to significantly improved optimality. To further illustrate the effectiveness and advantages of the proposed data-based control method, Table II summarizes the comparison results of the proposed data-based control method against the standard and optimized linear droop control methods, as well as the case where no control action is taken for different values of α𝛼\alphaitalic_α. We quantify the performance by the average of the distances of the actual and optimal reactive power setpoints, i.e., average of 𝐪𝒞(t)𝐪𝒞(t)normsubscript𝐪𝒞𝑡superscriptsubscript𝐪𝒞𝑡\|\mathbf{q}_{\mathcal{C}}(t)-\mathbf{q}_{\mathcal{C}}^{\star}(t)\|∥ bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) - bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ( italic_t ) ∥ for the entire day. It can be observed that the proposed data-based control method outperforms the benchmark methods in all cases.

Refer to caption
(a) ϵ=0.369italic-ϵ0.369\epsilon=0.369italic_ϵ = 0.369
Refer to caption
(b) ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1
Figure 7: Evolution of reactive power setpoints under the proposed reactive power update rule (3) with (a) ϵ=0.369italic-ϵ0.369\epsilon=0.369italic_ϵ = 0.369 and (b) ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1, where we use the power data profiles of the 695-th minute and consider 120 iterations. This verifies the sufficient condition 0<ϵ<min{1,2(𝐗L+1)2}=0.36910italic-ϵ12superscriptnorm𝐗𝐿120.36910<\epsilon<\min\{1,\frac{2}{(\|\mathbf{X}\|L+1)^{2}}\}=0.36910 < italic_ϵ < roman_min { 1 , divide start_ARG 2 end_ARG start_ARG ( ∥ bold_X ∥ italic_L + 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG } = 0.3691 in Proposition III.2 to ensure global asymptotic stability.
Refer to caption
(a) Maximum Voltage
Refer to caption
(b) Minimum Voltage
Figure 8: Evolution of the maximum (top) and minimum (bottom) voltages of the IEEE 37-bus network under the proposed data-based, ORPF solutions, optimized linear droop, standard linear droop, and no control methods. For all minute-based data profiles, the ORPF problem is feasible and thus 𝐪𝒞superscriptsubscript𝐪𝒞\mathbf{q}_{\mathcal{C}}^{\star}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT always exists. The optimized droop control induces voltage instability issues, causing voltages oscillations during 12:00 and 16:00, while the proposed data-based method guarantees the convergence of voltages for every minute-based data profile.
Refer to caption
Figure 9: Evolution of the distance between actual reactive power setpoints and ORPF solutions of the IEEE 37-bus network under the proposed data-based, optimized linear droop, standard linear droop, and no control methods.
TABLE II: Average distances between actual reactive power setpoints and ORPF solutions for entire day
α𝛼\alphaitalic_α 0 1/3 1/2 2/3 1
Data-based 0.11850.1185\mathbf{0.1185}bold_0.1185 0.09850.0985\mathbf{0.0985}bold_0.0985 0.07840.0784\mathbf{0.0784}bold_0.0784 0.11150.1115\mathbf{0.1115}bold_0.1115 0.16450.1645\mathbf{0.1645}bold_0.1645
Opt. Droop 0.2786 0.2474 0.3728 0.4410 0.4854
Std. Droop 0.2886 0.3311 0.4076 0.4699 0.5047
No Control 0.3160 0.5081 0.6842 0.8029 0.8358

V-C Robustness to Voltage Measurement Noise

Here, we test the robustness of the proposed data-based method against the voltage measurement noise. Similar to Section V-B, we use the learned equilibrium functions to run the simulations, but add random perturbations to the measurement of the local voltages. Table III summarizes the distances between actual reactive power setpoints and the ORPF solutions under different voltage measurement noise levels. Specifically, 0.5% perturbation to the voltage measurement corresponds to a common level of precision among smart meters in the United States [41], whereas the case of 1% perturbation represents the biggest error allowed in power systems. Table III indicates that, even in the latter case, the proposed data-based method still significantly outperforms the benchmark methods without measurement noise.

TABLE III: Average distances under measurement noise
Noise α𝛼\alphaitalic_α 0 1/3 1/2 2/3 1
0.2% 0.1210 0.1017 0.0936 0.1291 0.1739
0.5% 0.1317 0.1176 0.1378 0.1862 0.2237
1.0% 0.1571 0.1553 0.2127 0.2762 0.3129

V-D Discussion

Our simulation results above validate the improved performance of the proposed data-based method compared to the linear droop control method for different control goals. In fact, apart from considering the minimization of voltage deviations and power losses, our framework allows the users to consider any other type of cost functions, depending on specific control goals, to learn purely local controllers that steer system operating points to approximated ORPF solutions. However, as shown in Table I, different cost functions could result in different optimality gaps between the proposed data-based method and the ORPF approach. Since the data set we construct only maps the local voltage to the local optimal reactive power setpoint, it is possible that one fixed voltage corresponds to multiple optimal reactive power setpoints. On the other hand, it could also be that the optimal solution pairs are not as close to the non-increasing shape as we require the equilibrium functions to be. We refer to these phenomena as data inconsistency. We note that different selections of cost function significantly influence the data inconsistency, and thus leads to very different optimality gaps. For example, as Table I suggests, the data becomes significantly more inconsistent when the minimization of the voltage deviations takes a more important role in the cost function. As part of our follow-up work, we plan to include other available local information to alleviate the data inconsistency challenge, e.g., prevailing (re)active power injections as additional inputs of the equilibrium function. Another important observation is that, although the ORPF approach strictly guarantees that the voltages are within limits, our approach does not. For instance, in Fig. 8, the voltage nadir during evolution under the proposed data-based method slightly violates the voltage limits. The reason is that when α𝛼\alphaitalic_α is relatively small, many of the optimal solutions given by the ORPF problem lie on the boundary of the voltages limits. Since the local surrogates only provide approximations of the optimal solutions, the actual converged voltages can easily go out of limits in such situations. On the other hand, as pointed out in [17], purely local control strategies generally have no guarantee on desired regulation, in the sense that the equilibrium 𝐪𝒞superscriptsubscript𝐪𝒞\mathbf{q}_{\mathcal{C}}^{\sharp}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT of (8) could result in a 𝐯(𝐪𝒞)[𝐯min,𝐯max]𝐯superscriptsubscript𝐪𝒞subscript𝐯subscript𝐯\mathbf{v}(\mathbf{q}_{\mathcal{C}}^{\sharp})\notin[\mathbf{v}_{\min},\mathbf{% v}_{\max}]bold_v ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) ∉ [ bold_v start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT , bold_v start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ], even if there indeed exists 𝐪𝒞subscript𝐪𝒞\mathbf{q}_{\mathcal{C}}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT such that 𝐯(𝐪𝒞)[𝐯min,𝐯max]𝐯subscript𝐪𝒞subscript𝐯subscript𝐯\mathbf{v}(\mathbf{q}_{\mathcal{C}})\in[\mathbf{v}_{\min},\mathbf{v}_{\max}]bold_v ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) ∈ [ bold_v start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT , bold_v start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ].

VI Conclusions

We have presented a data-driven framework to design local Volt/Var controllers capable of steering a power distribution network towards efficient network configurations. Building on the idea of learning local surrogates that map local voltages to reactive power setpoints that approximate the ORPF solution, we have proposed a local control update scheme and identified conditions on surrogates and control parameters so that the reactive power point converges in a global asymptotic sense. By constructing a labeled data set of ORPF solutions with different load and generation profiles, we have trained neural networks whose resulting parameterized functions meet the conditions on surrogates by design to fit the data set. We have shown in AC power flow simulation tests that the proposed framework guarantees the voltage stability and significantly reduces the operation cost compared to prevalent local control approaches. Future research directions include considering the regulation of legacy devices and unbalanced DGs, enhancing data consistency by making use of other local information in building the data set, reducing the optimality gap during the learning process, and extending the proposed framework to a more general scenario where we take advantage of communication among neighboring agents.

Appendix A Technical Lemmas

Lemma A.1.

(Bauer and Fike Theorem [42, Corollary 6.3.4]): Let 𝐀,𝐄n×n𝐀𝐄superscript𝑛𝑛\mathbf{A},\mathbf{E}\in\mathbb{R}^{n\times n}bold_A , bold_E ∈ blackboard_R start_POSTSUPERSCRIPT italic_n × italic_n end_POSTSUPERSCRIPT with 𝐀𝐀\mathbf{A}bold_A normal. If λ^^𝜆\hat{\lambda}over^ start_ARG italic_λ end_ARG is an eigenvalue of 𝐀+𝐄𝐀𝐄\mathbf{A}+\mathbf{E}bold_A + bold_E, then there exists an eigenvalue λ𝜆\lambdaitalic_λ of 𝐀𝐀\mathbf{A}bold_A such that |λ^λ|𝐄^𝜆𝜆norm𝐄|\hat{\lambda}-\lambda|\leq\|\mathbf{E}\|| over^ start_ARG italic_λ end_ARG - italic_λ | ≤ ∥ bold_E ∥.

Lemma A.2.

(Positive semidefiniteness of 𝐗𝐌𝐗𝐌\mathbf{X}\mathbf{M}bold_XM and upper bound of 𝐗𝐌norm𝐗𝐌\|\mathbf{X}\mathbf{M}\|∥ bold_XM ∥): The matrix 𝐗𝐌𝐗𝐌\mathbf{X}\mathbf{M}bold_XM is positive semidefinite. Moreover, it holds

𝐗𝐌𝐗L.norm𝐗𝐌norm𝐗𝐿\displaystyle\|\mathbf{X}\mathbf{M}\|\leq\|\mathbf{X}\|L.∥ bold_XM ∥ ≤ ∥ bold_X ∥ italic_L .
Proof.

Let (λi,𝝃i)subscript𝜆𝑖subscript𝝃𝑖(\lambda_{i},\boldsymbol{\xi}_{i})( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , bold_italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) be a left eigenpair for 𝐗𝐌𝐗𝐌\mathbf{X}\mathbf{M}bold_XM. Then, (λi,𝝃i𝐗12)subscript𝜆𝑖subscript𝝃𝑖superscript𝐗12(\lambda_{i},\boldsymbol{\xi}_{i}\mathbf{X}^{\frac{1}{2}})( italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , bold_italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT bold_X start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ) is a left eigenpair for the symmetric matrix 𝐗12𝐌𝐗120succeeds-or-equalssuperscript𝐗12superscript𝐌𝐗120\mathbf{X}^{\frac{1}{2}}\mathbf{M}\mathbf{X}^{\frac{1}{2}}\succeq 0bold_X start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT bold_MX start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ⪰ 0. Indeed,

𝝃i𝐗12𝐗12𝐌𝐗12subscript𝝃𝑖superscript𝐗12superscript𝐗12superscript𝐌𝐗12\displaystyle\boldsymbol{\xi}_{i}\mathbf{X}^{\frac{1}{2}}\mathbf{X}^{\frac{1}{% 2}}\mathbf{M}\mathbf{X}^{\frac{1}{2}}bold_italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT bold_X start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT bold_X start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT bold_MX start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT =𝝃i𝐗𝐌𝐗12=λi𝝃i𝐗12.absentsubscript𝝃𝑖superscript𝐗𝐌𝐗12subscript𝜆𝑖subscript𝝃𝑖superscript𝐗12\displaystyle=\boldsymbol{\xi}_{i}\mathbf{X}\mathbf{M}\mathbf{X}^{\frac{1}{2}}% =\lambda_{i}\boldsymbol{\xi}_{i}\mathbf{X}^{\frac{1}{2}}.= bold_italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT bold_XMX start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT = italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT bold_italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT bold_X start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT .

Therefore, 𝐗𝐌𝐗𝐌\mathbf{X}\mathbf{M}bold_XM is positive semidefinite as well. Also,

𝐗𝐌𝐗𝐌𝐗L,norm𝐗𝐌norm𝐗norm𝐌norm𝐗𝐿\displaystyle\|\mathbf{X}\mathbf{M}\|\leq\|\mathbf{X}\|\|\mathbf{M}\|\leq\|% \mathbf{X}\|L,∥ bold_XM ∥ ≤ ∥ bold_X ∥ ∥ bold_M ∥ ≤ ∥ bold_X ∥ italic_L ,

which completes the proof. ∎

Appendix B Proofs of Propositions

Proof of Proposition III.1.

First, we show by induction the feasibility of the reactive power update (3). The initial power injection qn(0)subscript𝑞𝑛0q_{n}(0)italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( 0 ) belongs to 𝒬nsubscript𝒬𝑛\mathcal{Q}_{n}caligraphic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT by hypothesis. Assume now that qn(t)𝒬nsubscript𝑞𝑛𝑡subscript𝒬𝑛q_{n}(t)\in\mathcal{Q}_{n}italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) ∈ caligraphic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and C3) holds. Then qn(t+1)subscript𝑞𝑛𝑡1q_{n}(t+1)italic_q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t + 1 ) is the convex combination of two elements of 𝒬nsubscript𝒬𝑛\mathcal{Q}_{n}caligraphic_Q start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. Next, we show that (8) has an unique equilibrium point. From (9), the equilibrium exists if 𝐪𝒞=𝐡(𝐪𝒞)subscript𝐪𝒞𝐡subscript𝐪𝒞\mathbf{q}_{\mathcal{C}}=\mathbf{h}(\mathbf{q}_{\mathcal{C}})bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT = bold_h ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) has a solution, where 𝐡:𝒬𝒬:𝐡𝒬𝒬\mathbf{h}:\mathcal{Q}\rightarrow\mathcal{Q}bold_h : caligraphic_Q → caligraphic_Q is a continuous vector function with 𝐡(𝐪𝒞)=ϕ(𝐗𝐪𝒞+𝐯^𝒞)𝐡subscript𝐪𝒞bold-italic-ϕsubscript𝐗𝐪𝒞subscript^𝐯𝒞\mathbf{h}(\mathbf{q}_{\mathcal{C}})=\boldsymbol{\phi}(\mathbf{X}\mathbf{q}_{% \mathcal{C}}+\hat{\mathbf{v}}_{\mathcal{C}})bold_h ( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) = bold_italic_ϕ ( bold_Xq start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT + over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ). Since 𝒬𝒬\mathcal{Q}caligraphic_Q is convex and compact, according to Brouwer’s Fixed Point Theorem [43, Corollary 6.6], such a solution exists. Finally, to show uniqueness, we reason by contradiction. Assume both (𝐪𝒞,𝐯𝒞)superscriptsubscript𝐪𝒞superscriptsubscript𝐯𝒞(\mathbf{q}_{\mathcal{C}}^{\sharp},\mathbf{v}_{\mathcal{C}}^{\sharp})( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) and (𝐪𝒞,𝐯𝒞)superscriptsubscript𝐪𝒞superscriptsubscript𝐯𝒞(\mathbf{q}_{\mathcal{C}}^{\natural},\mathbf{v}_{\mathcal{C}}^{\natural})( bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT ) are equilibrium points for (8) with 𝐪𝒞𝐪𝒞superscriptsubscript𝐪𝒞superscriptsubscript𝐪𝒞\mathbf{q}_{\mathcal{C}}^{\sharp}\neq\mathbf{q}_{\mathcal{C}}^{\natural}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ≠ bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT. From (9a),

𝐪𝒞𝐪𝒞superscriptsubscript𝐪𝒞superscriptsubscript𝐪𝒞\displaystyle\mathbf{q}_{\mathcal{C}}^{\natural}-\mathbf{q}_{\mathcal{C}}^{\sharp}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT - bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT =ϕ(𝐯𝒞)ϕ(𝐯𝒞)=𝐃(𝐯𝒞𝐯𝒞),absentbold-italic-ϕsuperscriptsubscript𝐯𝒞bold-italic-ϕsuperscriptsubscript𝐯𝒞𝐃superscriptsubscript𝐯𝒞superscriptsubscript𝐯𝒞\displaystyle=\boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}^{\natural})-% \boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}^{\sharp})=\mathbf{D}(\mathbf{v}_{% \mathcal{C}}^{\natural}-\mathbf{v}_{\mathcal{C}}^{\sharp}),= bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT ) - bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) = bold_D ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) , (20)

where 𝐃C×C𝐃superscript𝐶𝐶\mathbf{D}\in\mathbb{R}^{C\times C}bold_D ∈ blackboard_R start_POSTSUPERSCRIPT italic_C × italic_C end_POSTSUPERSCRIPT is a diagonal matrix with

Dn={ϕn(vn)ϕn(vn)vnvnvnvn,0vn=vn.subscript𝐷𝑛casessubscriptitalic-ϕ𝑛superscriptsubscript𝑣𝑛subscriptitalic-ϕ𝑛superscriptsubscript𝑣𝑛superscriptsubscript𝑣𝑛superscriptsubscript𝑣𝑛superscriptsubscript𝑣𝑛superscriptsubscript𝑣𝑛0superscriptsubscript𝑣𝑛superscriptsubscript𝑣𝑛\displaystyle D_{n}=\begin{cases}\frac{\phi_{n}(v_{n}^{\natural})-\phi_{n}(v_{% n}^{\sharp})}{v_{n}^{\natural}-v_{n}^{\sharp}}&v_{n}^{\natural}\neq v_{n}^{% \sharp},\\ \hfil 0&v_{n}^{\natural}=v_{n}^{\sharp}.\end{cases}italic_D start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = { start_ROW start_CELL divide start_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT ) - italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT - italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT end_ARG end_CELL start_CELL italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT ≠ italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT , end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT = italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT . end_CELL end_ROW

From C2), ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is nonincreasing in vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT for all n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C. Hence, Dn0,n𝒞formulae-sequencesubscript𝐷𝑛0for-all𝑛𝒞D_{n}\leq 0,\forall n\in\mathcal{C}italic_D start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≤ 0 , ∀ italic_n ∈ caligraphic_C, and 𝐃0precedes-or-equals𝐃0\mathbf{D}\preceq 0bold_D ⪯ 0. On the other hand, (9b) yields

𝐪𝒞𝐪𝒞=𝐗1(𝐯𝒞𝐯𝒞).superscriptsubscript𝐪𝒞superscriptsubscript𝐪𝒞superscript𝐗1superscriptsubscript𝐯𝒞superscriptsubscript𝐯𝒞\mathbf{q}_{\mathcal{C}}^{\natural}-\mathbf{q}_{\mathcal{C}}^{\sharp}=\mathbf{% X}^{-1}(\mathbf{v}_{\mathcal{C}}^{\natural}-\mathbf{v}_{\mathcal{C}}^{\sharp}).bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT - bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT = bold_X start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) .

Then, it follows that

(𝐗1𝐃)(𝐯𝒞𝐯𝒞)=𝟎.superscript𝐗1𝐃superscriptsubscript𝐯𝒞superscriptsubscript𝐯𝒞0(\mathbf{X}^{-1}-\mathbf{D})(\mathbf{v}_{\mathcal{C}}^{\natural}-\mathbf{v}_{% \mathcal{C}}^{\sharp})=\mathbf{0}.( bold_X start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - bold_D ) ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT ) = bold_0 .

Since 𝐗0succeeds𝐗0\mathbf{X}\succ 0bold_X ≻ 0, it holds that 𝐗10succeedssuperscript𝐗10\mathbf{X}^{-1}\succ 0bold_X start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ≻ 0, 𝐗1𝐃0succeedssuperscript𝐗1𝐃0\mathbf{X}^{-1}-\mathbf{D}\succ 0bold_X start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - bold_D ≻ 0. As a consequence, 𝐯𝒞𝐯𝒞=𝟎superscriptsubscript𝐯𝒞superscriptsubscript𝐯𝒞0\mathbf{v}_{\mathcal{C}}^{\natural}-\mathbf{v}_{\mathcal{C}}^{\sharp}=\mathbf{0}bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT = bold_0 and 𝐪𝒞𝐪𝒞=𝟎superscriptsubscript𝐪𝒞superscriptsubscript𝐪𝒞0\mathbf{q}_{\mathcal{C}}^{\natural}-\mathbf{q}_{\mathcal{C}}^{\sharp}=\mathbf{0}bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♮ end_POSTSUPERSCRIPT - bold_q start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ♯ end_POSTSUPERSCRIPT = bold_0, cf. (9a), which is a contradiction. This completes the proof. ∎

Proof of Proposition III.2.

Consider the voltage evolution under (8),

𝐯𝒞(t+1)=𝐗𝐪𝒞(t+1)+𝐯^𝒞subscript𝐯𝒞𝑡1subscript𝐗𝐪𝒞𝑡1subscript^𝐯𝒞\displaystyle\mathbf{v}_{\mathcal{C}}(t+1)=\mathbf{X}\mathbf{q}_{\mathcal{C}}(% t+1)+\hat{\mathbf{v}}_{\mathcal{C}}bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t + 1 ) = bold_Xq start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t + 1 ) + over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT
=(1ϵ)𝐗𝐪𝒞(t)+ϵ𝐗ϕ(𝐯𝒞(t))+(1ϵ)𝐯^𝒞+ϵ𝐯^𝒞absent1italic-ϵsubscript𝐗𝐪𝒞𝑡italic-ϵ𝐗bold-italic-ϕsubscript𝐯𝒞𝑡1italic-ϵsubscript^𝐯𝒞italic-ϵsubscript^𝐯𝒞\displaystyle=(1-\epsilon)\mathbf{X}\mathbf{q}_{\mathcal{C}}(t)+\epsilon% \mathbf{X}\boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}(t))+(1-\epsilon)\hat{% \mathbf{v}}_{\mathcal{C}}+\epsilon\hat{\mathbf{v}}_{\mathcal{C}}= ( 1 - italic_ϵ ) bold_Xq start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) + italic_ϵ bold_X bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) ) + ( 1 - italic_ϵ ) over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT + italic_ϵ over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT
=(1ϵ)𝐯𝒞(t)+ϵ(𝐗ϕ(𝐯𝒞(t))+𝐯^𝒞):=𝐠(𝐯𝒞(t)).absent1italic-ϵsubscript𝐯𝒞𝑡italic-ϵ𝐗bold-italic-ϕsubscript𝐯𝒞𝑡subscript^𝐯𝒞assign𝐠subscript𝐯𝒞𝑡\displaystyle=(1-\epsilon)\mathbf{v}_{\mathcal{C}}(t)+\epsilon(\mathbf{X}% \boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}(t))+\hat{\mathbf{v}}_{\mathcal{C}})% :=\mathbf{g}(\mathbf{v}_{\mathcal{C}}(t)).= ( 1 - italic_ϵ ) bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) + italic_ϵ ( bold_X bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) ) + over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) := bold_g ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ( italic_t ) ) .

We show that, for small enough values of ϵitalic-ϵ\epsilonitalic_ϵ, the operator 𝐠:CC:𝐠superscript𝐶superscript𝐶\mathbf{g}:\mathbb{R}^{C}\to\mathbb{R}^{C}bold_g : blackboard_R start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT is a contraction, i.e.,

𝐠(𝐯𝒞)𝐠(𝐯𝒞)𝐯𝒞𝐯𝒞<1,norm𝐠subscript𝐯𝒞𝐠superscriptsubscript𝐯𝒞normsubscript𝐯𝒞superscriptsubscript𝐯𝒞1\displaystyle\frac{\|\mathbf{g}(\mathbf{v}_{\mathcal{C}})-\mathbf{g}(\mathbf{v% }_{\mathcal{C}}^{\prime})\|}{\|\mathbf{v}_{\mathcal{C}}-\mathbf{v}_{\mathcal{C% }}^{\prime}\|}<1,divide start_ARG ∥ bold_g ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) - bold_g ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ∥ end_ARG start_ARG ∥ bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∥ end_ARG < 1 , (21)

for any 𝐯𝒞,𝐯𝒞Csubscript𝐯𝒞superscriptsubscript𝐯𝒞superscript𝐶\mathbf{v}_{\mathcal{C}},\mathbf{v}_{\mathcal{C}}^{\prime}\in\mathbb{R}^{C}bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT , bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT. Indeed, define a diagonal matrix 𝐌C×C𝐌superscript𝐶𝐶\mathbf{M}\in\mathbb{R}^{C\times C}bold_M ∈ blackboard_R start_POSTSUPERSCRIPT italic_C × italic_C end_POSTSUPERSCRIPT with the n𝑛nitalic_n-th diagonal entry being

Mn={|ϕn(vn)ϕ(vn)||vnvn|vnvn,0vn=vn.subscript𝑀𝑛casessubscriptitalic-ϕ𝑛subscript𝑣𝑛italic-ϕsuperscriptsubscript𝑣𝑛subscript𝑣𝑛superscriptsubscript𝑣𝑛subscript𝑣𝑛superscriptsubscript𝑣𝑛0subscript𝑣𝑛superscriptsubscript𝑣𝑛\displaystyle M_{n}=\begin{cases}\frac{|\phi_{n}(v_{n})-\phi(v_{n}^{\prime})|}% {|v_{n}-v_{n}^{\prime}|}&v_{n}\neq v_{n}^{\prime},\\ \hfil 0&v_{n}=v_{n}^{\prime}.\end{cases}italic_M start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = { start_ROW start_CELL divide start_ARG | italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) - italic_ϕ ( italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | end_ARG start_ARG | italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | end_ARG end_CELL start_CELL italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ≠ italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT . end_CELL end_ROW

Then, it follows that

𝐠(𝐯𝒞)𝐠(𝐯𝒞)norm𝐠subscript𝐯𝒞𝐠superscriptsubscript𝐯𝒞\displaystyle\|\mathbf{g}(\mathbf{v}_{\mathcal{C}})-\mathbf{g}(\mathbf{v}_{% \mathcal{C}}^{\prime})\|∥ bold_g ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) - bold_g ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ∥
=(1ϵ)(𝐯𝒞𝐯𝒞)+ϵ𝐗(ϕ(𝐯𝒞)ϕ(𝐯𝒞))absentnorm1italic-ϵsubscript𝐯𝒞superscriptsubscript𝐯𝒞italic-ϵ𝐗bold-italic-ϕsubscript𝐯𝒞bold-italic-ϕsuperscriptsubscript𝐯𝒞\displaystyle=\big{\|}(1-\epsilon)(\mathbf{v}_{\mathcal{C}}-\mathbf{v}_{% \mathcal{C}}^{\prime})+\epsilon\mathbf{X}\big{(}\boldsymbol{\phi}(\mathbf{v}_{% \mathcal{C}})-\boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}^{\prime})\big{)}\big{\|}= ∥ ( 1 - italic_ϵ ) ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) + italic_ϵ bold_X ( bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) - bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ) ∥
=(1ϵ)sign(𝐯𝒞𝐯𝒞)|𝐯𝒞𝐯𝒞|\displaystyle=\big{\|}(1-\epsilon)\rm{sign}(\mathbf{v}_{\mathcal{C}}-\mathbf{v% }_{\mathcal{C}}^{\prime})|\mathbf{v}_{\mathcal{C}}-\mathbf{v}_{\mathcal{C}}^{% \prime}|= ∥ ( 1 - italic_ϵ ) roman_sign ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT |
ϵ𝐗sign(𝐯𝒞𝐯𝒞)|ϕ(𝐯𝒞)ϕ(𝐯𝒞)|\displaystyle\qquad\qquad\qquad-\epsilon\mathbf{X}\rm{sign}(\mathbf{v}_{% \mathcal{C}}-\mathbf{v}_{\mathcal{C}}^{\prime})|\boldsymbol{\phi}(\mathbf{v}_{% \mathcal{C}})-\boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}^{\prime})|\big{\|}- italic_ϵ bold_X roman_sign ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) - bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | ∥
=(1ϵ)|𝐯𝒞𝐯𝒞|ϵ𝐗|ϕ(𝐯𝒞)ϕ(𝐯𝒞)|absentnorm1italic-ϵsubscript𝐯𝒞superscriptsubscript𝐯𝒞italic-ϵ𝐗bold-italic-ϕsubscript𝐯𝒞bold-italic-ϕsuperscriptsubscript𝐯𝒞\displaystyle=\big{\|}(1-\epsilon)|\mathbf{v}_{\mathcal{C}}-\mathbf{v}_{% \mathcal{C}}^{\prime}|-\epsilon\mathbf{X}|\boldsymbol{\phi}(\mathbf{v}_{% \mathcal{C}})-\boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}}^{\prime})|\big{\|}= ∥ ( 1 - italic_ϵ ) | bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | - italic_ϵ bold_X | bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) - bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) | ∥
(1ϵ)𝐈ϵ𝐗𝐌𝐯𝒞𝐯𝒞,absentnorm1italic-ϵ𝐈italic-ϵ𝐗𝐌normsubscript𝐯𝒞superscriptsubscript𝐯𝒞\displaystyle\leq\|(1-\epsilon)\mathbf{I}-\epsilon\mathbf{X}\mathbf{M}\|\|% \mathbf{v}_{\mathcal{C}}-\mathbf{v}_{\mathcal{C}}^{\prime}\|,≤ ∥ ( 1 - italic_ϵ ) bold_I - italic_ϵ bold_XM ∥ ∥ bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∥ ,

where we have used in the second equality the fact that ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is nonincreasing in vnsubscript𝑣𝑛v_{n}italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT for each n𝒞𝑛𝒞n\in\mathcal{C}italic_n ∈ caligraphic_C, and thus sign(ϕ(𝐯𝒞)ϕ(𝐯𝒞))=sign(𝐯𝒞𝐯𝒞)signbold-italic-ϕsubscript𝐯𝒞bold-italic-ϕsuperscriptsubscript𝐯𝒞signsubscript𝐯𝒞superscriptsubscript𝐯𝒞\rm{sign}(\boldsymbol{\phi}(\mathbf{v}_{\mathcal{C}})-\boldsymbol{\phi}(% \mathbf{v}_{\mathcal{C}}^{\prime}))=-\rm{sign}(\mathbf{v}_{\mathcal{C}}-% \mathbf{v}_{\mathcal{C}}^{\prime})roman_sign ( bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT ) - bold_italic_ϕ ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ) = - roman_sign ( bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ).

To prove (21), it is sufficient to show that there always exists ϵitalic-ϵ\epsilonitalic_ϵ such that (1ϵ)𝐈ϵ𝐗𝐌<1norm1italic-ϵ𝐈italic-ϵ𝐗𝐌1\|(1-\epsilon)\mathbf{I}-\epsilon\mathbf{X}\mathbf{M}\|<1∥ ( 1 - italic_ϵ ) bold_I - italic_ϵ bold_XM ∥ < 1, which is equivalent to proving that λmax(𝚪)<1subscript𝜆𝚪1\lambda_{\max}(\boldsymbol{\Gamma})<1italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( bold_Γ ) < 1, where 𝚪0succeeds-or-equals𝚪0\boldsymbol{\Gamma}\succeq 0bold_Γ ⪰ 0 is

𝚪𝚪absent\displaystyle\boldsymbol{\Gamma}\triangleqbold_Γ ≜ [(1ϵ)𝐈ϵ𝐗𝐌][(1ϵ)𝐈ϵ𝐗𝐌]superscriptdelimited-[]1italic-ϵ𝐈italic-ϵ𝐗𝐌topdelimited-[]1italic-ϵ𝐈italic-ϵ𝐗𝐌\displaystyle\left[(1-\epsilon)\mathbf{I}-\epsilon\mathbf{X}\mathbf{M}\right]^% {\top}\left[(1-\epsilon)\mathbf{I}-\epsilon\mathbf{X}\mathbf{M}\right][ ( 1 - italic_ϵ ) bold_I - italic_ϵ bold_XM ] start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT [ ( 1 - italic_ϵ ) bold_I - italic_ϵ bold_XM ]
=\displaystyle== (1ϵ)2𝐈ϵ(1ϵ)(𝐌𝐗+𝐗𝐌)+ϵ2𝐌𝐗𝐗𝐌.superscript1italic-ϵ2𝐈italic-ϵ1italic-ϵsuperscript𝐌𝐗top𝐗𝐌superscriptitalic-ϵ2superscript𝐌𝐗top𝐗𝐌\displaystyle(1-\epsilon)^{2}\mathbf{I}-\epsilon(1-\epsilon)(\mathbf{M}\mathbf% {X}^{\top}+\mathbf{X}\mathbf{M})+\epsilon^{2}\mathbf{M}\mathbf{X}^{\top}% \mathbf{X}\mathbf{M}.( 1 - italic_ϵ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_I - italic_ϵ ( 1 - italic_ϵ ) ( bold_MX start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT + bold_XM ) + italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_MX start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_XM .

Rewrite 𝚪𝚪\boldsymbol{\Gamma}bold_Γ as

𝚪=𝚪absent\displaystyle\boldsymbol{\Gamma}=bold_Γ = (12ϵ)𝐈ϵ(𝐌𝐗+𝐗𝐌):=𝐀subscript12italic-ϵ𝐈italic-ϵsuperscript𝐌𝐗top𝐗𝐌assignabsent𝐀\displaystyle\underbrace{(1-2\epsilon)\mathbf{I}-\epsilon(\mathbf{M}\mathbf{X}% ^{\top}+\mathbf{X}\mathbf{M})}_{:=\mathbf{A}}under⏟ start_ARG ( 1 - 2 italic_ϵ ) bold_I - italic_ϵ ( bold_MX start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT + bold_XM ) end_ARG start_POSTSUBSCRIPT := bold_A end_POSTSUBSCRIPT
+ϵ2(𝐈+𝐗𝐌+𝐌𝐗+𝐌𝐗𝐗𝐌):=𝐄.superscriptitalic-ϵ2subscript𝐈𝐗𝐌superscript𝐌𝐗topsuperscript𝐌𝐗top𝐗𝐌assignabsent𝐄\displaystyle\qquad+\epsilon^{2}\underbrace{(\mathbf{I}+\mathbf{X}\mathbf{M}+% \mathbf{M}\mathbf{X}^{\top}+\mathbf{M}\mathbf{X}^{\top}\mathbf{X}\mathbf{M})}_% {:=\mathbf{E}}.+ italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT under⏟ start_ARG ( bold_I + bold_XM + bold_MX start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT + bold_MX start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_XM ) end_ARG start_POSTSUBSCRIPT := bold_E end_POSTSUBSCRIPT .

Note that 𝐀𝐀\mathbf{A}bold_A is symmetric. According to Lemma A.1, provided in the Appendix, it holds that 0λmax(𝚪)λmax(𝐀)+ϵ2𝐄0subscript𝜆𝚪subscript𝜆𝐀superscriptitalic-ϵ2norm𝐄0\leq\lambda_{\max}(\boldsymbol{\Gamma})\leq\lambda_{\max}(\mathbf{A})+% \epsilon^{2}\|\mathbf{E}\|0 ≤ italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( bold_Γ ) ≤ italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( bold_A ) + italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∥ bold_E ∥. Now, it remains to show that there always exists ϵitalic-ϵ\epsilonitalic_ϵ such that

λmax(𝐀)+ϵ2𝐄<1.subscript𝜆𝐀superscriptitalic-ϵ2norm𝐄1\displaystyle\lambda_{\max}(\mathbf{A})+\epsilon^{2}\|\mathbf{E}\|<1.italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ( bold_A ) + italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∥ bold_E ∥ < 1 .

Let λ¯=λmin(𝐌𝐗+𝐗𝐌)¯𝜆subscript𝜆superscript𝐌𝐗top𝐗𝐌\underline{\lambda}=\lambda_{\min}(\mathbf{M}\mathbf{X}^{\top}+\mathbf{X}% \mathbf{M})under¯ start_ARG italic_λ end_ARG = italic_λ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( bold_MX start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT + bold_XM ). According to Lemma A.2, 𝐌𝐗,𝐗𝐌0succeeds-or-equalssuperscript𝐌𝐗top𝐗𝐌0\mathbf{M}\mathbf{X}^{\top},\mathbf{X}\mathbf{M}\succeq 0bold_MX start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT , bold_XM ⪰ 0, and therefore λ¯0¯𝜆0\underline{\lambda}\geq 0under¯ start_ARG italic_λ end_ARG ≥ 0. This condition is then equivalent to

1ϵ(2+λ¯)+ϵ2𝐄<1,1italic-ϵ2¯𝜆superscriptitalic-ϵ2norm𝐄11-\epsilon(2+\underline{\lambda})+\epsilon^{2}\|\mathbf{E}\|<1,1 - italic_ϵ ( 2 + under¯ start_ARG italic_λ end_ARG ) + italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∥ bold_E ∥ < 1 ,

which yields

0<ϵ<(2+λ¯)𝐄.0italic-ϵ2¯𝜆norm𝐄0<\epsilon<\frac{(2+\underline{\lambda})}{\|\mathbf{E}\|}.0 < italic_ϵ < divide start_ARG ( 2 + under¯ start_ARG italic_λ end_ARG ) end_ARG start_ARG ∥ bold_E ∥ end_ARG .

Finally, according to Lemma A.2, it holds

𝐄norm𝐄\displaystyle\|\mathbf{E}\|∥ bold_E ∥ 1+2𝐗𝐌+𝐗𝐌2absent12norm𝐗𝐌superscriptnorm𝐗𝐌2\displaystyle\leq 1+2\|\mathbf{X}\mathbf{M}\|+\|\mathbf{X}\mathbf{M}\|^{2}≤ 1 + 2 ∥ bold_XM ∥ + ∥ bold_XM ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
1+2𝐗L+𝐗2L2=(𝐗L+1)2,absent12norm𝐗𝐿superscriptnorm𝐗2superscript𝐿2superscriptnorm𝐗𝐿12\displaystyle\leq 1+2\|\mathbf{X}\|L+\|\mathbf{X}\|^{2}L^{2}=(\|\mathbf{X}\|L+% 1)^{2},≤ 1 + 2 ∥ bold_X ∥ italic_L + ∥ bold_X ∥ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( ∥ bold_X ∥ italic_L + 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ,

and hence

2+λ¯𝐄2𝐄2(𝐗L+1)2.2¯𝜆norm𝐄2norm𝐄2superscriptnorm𝐗𝐿12\displaystyle\frac{2+\underline{\lambda}}{\|\mathbf{E}\|}\geq\frac{2}{\|% \mathbf{E}\|}\geq\frac{2}{(\|\mathbf{X}\|L+1)^{2}}.divide start_ARG 2 + under¯ start_ARG italic_λ end_ARG end_ARG start_ARG ∥ bold_E ∥ end_ARG ≥ divide start_ARG 2 end_ARG start_ARG ∥ bold_E ∥ end_ARG ≥ divide start_ARG 2 end_ARG start_ARG ( ∥ bold_X ∥ italic_L + 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG .

Considering that we require ϵ[0,1]italic-ϵ01\epsilon\in[0,1]italic_ϵ ∈ [ 0 , 1 ], (10) follows, concluding the proof. ∎

Proof of Proposition IV.1.

To show the sufficiency of (17), notice that, for any x𝑥x\in\mathbb{R}italic_x ∈ blackboard_R, if x<b1𝑥subscript𝑏1x<b_{1}italic_x < italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, then 𝖭(x)=β𝖭𝑥𝛽\mathsf{N}(x)=\betasansserif_N ( italic_x ) = italic_β; and for xb1𝑥subscript𝑏1x\geq b_{1}italic_x ≥ italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, 𝖭(x)𝖭𝑥\mathsf{N}(x)sansserif_N ( italic_x ) is divided into H𝐻Hitalic_H segments with the slope of the J𝐽Jitalic_J-th segment being j=1Jwjsuperscriptsubscript𝑗1𝐽subscript𝑤𝑗\sum_{j=1}^{J}w_{j}∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT for J{1,2,,H}𝐽12𝐻J\in\{1,2,\dots,H\}italic_J ∈ { 1 , 2 , … , italic_H }. Given (17), it follows that 𝖭𝖭\mathsf{N}sansserif_N is nonincreasing. To show the necessity of (17), suppose 𝖭𝖭\mathsf{N}sansserif_N is nonincreasing. Then, there exists a segment of the equilibrium function, e.g., the J𝐽Jitalic_J-th, J{1,2,,H}𝐽12𝐻J\in\{1,2,\dots,H\}italic_J ∈ { 1 , 2 , … , italic_H } which is increasing. Hence, j=1Jwj>0superscriptsubscript𝑗1𝐽subscript𝑤𝑗0\sum_{j=1}^{J}w_{j}>0∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT > 0 which is a contradiction.

We next show the universal approximation property. Given a compact domain 𝒳𝒳\mathcal{X}caligraphic_X of the form x¯xx¯¯𝑥𝑥¯𝑥\underline{x}\leq x\leq\overline{x}under¯ start_ARG italic_x end_ARG ≤ italic_x ≤ over¯ start_ARG italic_x end_ARG, consider an equispaced partition of 𝒳𝒳\mathcal{X}caligraphic_X into H𝐻Hitalic_H intervals, with the length of each interval being s=x¯x¯H𝑠¯𝑥¯𝑥𝐻s=\frac{\overline{x}-\underline{x}}{H}italic_s = divide start_ARG over¯ start_ARG italic_x end_ARG - under¯ start_ARG italic_x end_ARG end_ARG start_ARG italic_H end_ARG. Consider the function

𝖭(x)=h=1HwhReLU(xbh)+β,𝖭𝑥superscriptsubscript1𝐻subscript𝑤ReLU𝑥subscript𝑏𝛽\displaystyle\mathsf{N}(x)=\sum_{h=1}^{H}w_{h}{\rm ReLU}(x-b_{h})+\beta,sansserif_N ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_h = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_ReLU ( italic_x - italic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) + italic_β ,

whose parameters are defined as: β=g(x¯)𝛽𝑔¯𝑥\beta=g(\underline{x})italic_β = italic_g ( under¯ start_ARG italic_x end_ARG ), bh=x¯+(h1)ssubscript𝑏¯𝑥1𝑠b_{h}=\underline{x}+(h-1)sitalic_b start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = under¯ start_ARG italic_x end_ARG + ( italic_h - 1 ) italic_s, w1=g(x¯+s)g(x¯)ssubscript𝑤1𝑔¯𝑥𝑠𝑔¯𝑥𝑠w_{1}=\frac{g(\underline{x}+s)-g(\underline{x})}{s}italic_w start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG italic_g ( under¯ start_ARG italic_x end_ARG + italic_s ) - italic_g ( under¯ start_ARG italic_x end_ARG ) end_ARG start_ARG italic_s end_ARG, w2=g(x¯+2s)g(x¯+s)sw1subscript𝑤2𝑔¯𝑥2𝑠𝑔¯𝑥𝑠𝑠subscript𝑤1w_{2}=\frac{g(\underline{x}+2s)-g(\underline{x}+s)}{s}-w_{1}italic_w start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG italic_g ( under¯ start_ARG italic_x end_ARG + 2 italic_s ) - italic_g ( under¯ start_ARG italic_x end_ARG + italic_s ) end_ARG start_ARG italic_s end_ARG - italic_w start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT,…, wh=g(x¯+hs)g(x¯+(h1)s)sk=1h1wksubscript𝑤𝑔¯𝑥𝑠𝑔¯𝑥1𝑠𝑠superscriptsubscript𝑘11subscript𝑤𝑘w_{h}=\frac{g(\underline{x}+hs)-g(\underline{x}+(h-1)s)}{s}-\sum_{k=1}^{h-1}w_% {k}italic_w start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = divide start_ARG italic_g ( under¯ start_ARG italic_x end_ARG + italic_h italic_s ) - italic_g ( under¯ start_ARG italic_x end_ARG + ( italic_h - 1 ) italic_s ) end_ARG start_ARG italic_s end_ARG - ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_h - 1 end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. Note that it holds that

k=1hwk=g(x¯+hs)g(x¯+(h1)s)s0superscriptsubscript𝑘1subscript𝑤𝑘𝑔¯𝑥𝑠𝑔¯𝑥1𝑠𝑠0\sum_{k=1}^{h}w_{k}=\frac{g(\underline{x}+hs)-g(\underline{x}+(h-1)s)}{s}\leq 0∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_h end_POSTSUPERSCRIPT italic_w start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = divide start_ARG italic_g ( under¯ start_ARG italic_x end_ARG + italic_h italic_s ) - italic_g ( under¯ start_ARG italic_x end_ARG + ( italic_h - 1 ) italic_s ) end_ARG start_ARG italic_s end_ARG ≤ 0

because g(x)𝑔𝑥g(x)italic_g ( italic_x ) is nonincreasing, for h{1,2,,H}12𝐻h\in\{1,2,\dots,H\}italic_h ∈ { 1 , 2 , … , italic_H }. Hence, inequality (17) is satisfied. It can be observed that

|𝖭(x)g(x)|Lgs𝖭𝑥𝑔𝑥subscript𝐿𝑔𝑠|\mathsf{N}(x)-g(x)|\leq L_{g}s| sansserif_N ( italic_x ) - italic_g ( italic_x ) | ≤ italic_L start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_s

where Lgsubscript𝐿𝑔L_{g}italic_L start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT is the Lipschitz constant of g𝑔gitalic_g. Hence, by setting HLg(x¯x¯)η=O(1/η)𝐻subscript𝐿𝑔¯𝑥¯𝑥𝜂𝑂1𝜂H\geq\frac{L_{g}(\overline{x}-\underline{x})}{\eta}=O(1/\eta)italic_H ≥ divide start_ARG italic_L start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( over¯ start_ARG italic_x end_ARG - under¯ start_ARG italic_x end_ARG ) end_ARG start_ARG italic_η end_ARG = italic_O ( 1 / italic_η ), it holds that |𝖭(x)g(x)|Lgsη𝖭𝑥𝑔𝑥subscript𝐿𝑔𝑠𝜂|\mathsf{N}(x)-g(x)|\leq L_{g}s\leq\eta| sansserif_N ( italic_x ) - italic_g ( italic_x ) | ≤ italic_L start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT italic_s ≤ italic_η for all x𝒳𝑥𝒳x\in\mathcal{X}italic_x ∈ caligraphic_X. This completes the proof. ∎

References

  • [1] S. H. Low, “Convex relaxation of optimal power flow - Part I: Formulations and equivalence,” IEEE Transactions on Control of Network Systems, vol. 1, no. 1, pp. 15–27, 2014.
  • [2] B. Cui and X. A. Sun, “A new voltage stability-constrained optimal power-flow model: Sufficient condition, SOCP representation, and relaxation,” IEEE Transactions on Power Systems, vol. 33, no. 5, pp. 5092–5102, 2018.
  • [3] L. Gan, N. Li, U. Topcu, and S. H. Low, “Optimal power flow in tree networks,” in IEEE Conf. on Decision and Control, (Firenze, Italy), pp. 2313–2318, Dec. 2013.
  • [4] X. Pan, T. Zhao, M. Chen, and S. Zhang, “DeepOPF: A deep neural network approach for security-constrained DC optimal power flow,” IEEE Transactions on Power Systems, vol. 36, no. 3, pp. 1725–1735, 2021.
  • [5] D. Owerko, F. Gama, and A. Ribeiro, “Optimal power flow using graph neural networks,” in IEEE Int. Conf. on Acoustics, Speech and Signal Processing, (Barcelona, Spain), pp. 5930–5934, 2020.
  • [6] M. K. Singh, S. Gupta, V. Kekatos, G. Cavraro, and A. Bernstein, “Learning to optimize power distribution grids using sensitivity-informed deep neural networks,” in IEEE Int. Conf. on Communications, Control, and Computing Technologies for Smart Grids, (Tempe, AZ, USA), Nov. 2020.
  • [7] M. K. Singh, V. Kekatos, and G. B. Giannakis, “Learning to solve the AC-OPF using sensitivity-informed deep neural networks,” IEEE Transactions on Power Systems, vol. 37, no. 4, pp. 2833–2846, 2022.
  • [8] F. Fioretto, T. W. K. Mak, and P. Van Hentenryck, “Predicting AC optimal power flows: Combining deep learning and lagrangian dual methods,” in AAAI Conference on Artificial Intelligence, (New York, USA), pp. 630–637, Feb. 2020.
  • [9] E. Dall’Anese, H. Zhu, and G. B. Giannakis, “Distributed optimal power flow for smart microgrids,” IEEE Transactions on Smart Grid, vol. 4, no. 3, pp. 1464–1475, 2013.
  • [10] G. Cavraro and R. Carli, “Local and distributed voltage control algorithms in distribution networks,” IEEE Transactions on Power Systems, vol. 33, no. 2, pp. 1420–1430, 2017.
  • [11] E. Dall’Anese and A. Simonetto, “Optimal power flow pursuit,” IEEE Transactions on Smart Grid, vol. 9, no. 2, pp. 942–952, 2018.
  • [12] G. Qu and N. Li, “Optimal distributed feedback voltage control under limited reactive power,” IEEE Transactions on Power Systems, vol. 35, no. 1, pp. 315–331, 2020.
  • [13] “IEEE standard for interconnection and interoperability of distributed energy resources with associated electric power systems interfaces,” IEEE Std 1547-2018 (Revision of IEEE Std 1547-2003), pp. 1–138, 2018.
  • [14] K. Turitsyn, P. Sulc, S. Backhaus, and M. Chertkov, “Options for control of reactive power by distributed photovoltaic generators,” Proceedings of the IEEE, vol. 99, no. 6, pp. 1063–1073, 2011.
  • [15] H. Zhu and H. J. Liu, “Fast local voltage control under limited reactive power: Optimality and stability analysis,” IEEE Transactions on Power Systems, vol. 31, no. 5, pp. 3794–3803, 2015.
  • [16] X. Zhou, M. Farivar, Z. Liu, L. Chen, and S. H. Low, “Reverse and forward engineering of local voltage control in distribution networks,” IEEE Transactions on Automatic Control, vol. 66, no. 3, pp. 1116–1128, 2021.
  • [17] S. Bolognani, R. Carli, G. Cavraro, and S. Zampieri, “On the need for communication for voltage regulation of power distribution grids,” IEEE Transactions on Control of Network Systems, vol. 6, no. 3, pp. 1111–1123, 2019.
  • [18] C. Dawson, S. Gao, and C. Fan, “Safe control with learned certificates: A survey of neural Lyapunov, barrier, and contraction methods,” arXiv preprint arXiv:2202.11762, 2022.
  • [19] W. Cui, Y. Jiang, and B. Zhang, “Reinforcement learning for optimal primary frequency control: A Lyapunov approach,” IEEE Transactions on Power Systems, 2022. To appear.
  • [20] Z. Yuan, C. Zhao, and J. Cortés, “Reinforcement learning for distributed transient frequency control with stability and safety guarantees,” Systems & Control Letters, 2022. Submitted.
  • [21] W. Cui, J. Li, and B. Zhang, “Decentralized safe reinforcement learning for inverter-based voltage control,” Electric Power Systems Research, vol. 211, p. 108609, 2022.
  • [22] Y. Shi, G. Qu, S. H. Low, A. Anandkumar, and A. Wierman, “Stability constrained reinforcement learning for real-time voltage control,” in American Control Conference, (Atlanta, GA), pp. 2715–2721, June 2022.
  • [23] C. Zhang, Y. Xu, Y. Wang, Z. Y. Dong, and R. Zhang, “Three-stage hierarchically-coordinated Voltage/Var control based on PV inverters considering distribution network voltage stability,” IEEE Transactions on Sustainable Energy, vol. 13, no. 2, pp. 868–881, 2021.
  • [24] S. Gupta, S. Chatzivasileiadis, and V. Kekatos, “Deep learning for optimal Volt/VAR control using distributed energy resources,” arXiv preprint arXiv:2211.09557, 2022.
  • [25] S. Karagiannopoulos, P. Aristidou, and G. Hug, “Data-driven local control design for active distribution grids using off-line optimal power flow and machine learning techniques,” IEEE Transactions on Smart Grid, vol. 10, no. 6, pp. 6461–6471, 2019.
  • [26] P. S. Torre and P. Hidalgo-Gonzalez, “Decentralized optimal power flow for time-varying network topologies using machine learning,” Electric Power Systems Research, vol. 212, p. 108575, 2022.
  • [27] X. Sun, J. Qiu, and J. Zhao, “Optimal local Volt/Var control for photovoltaic inverters in active distribution networks,” IEEE Transactions on Power Systems, vol. 36, no. 6, pp. 5756–5766, 2021.
  • [28] H. Ji, C. Wang, P. Li, J. Zhao, G. Song, F. Ding, and J. Wu, “A centralized-based method to determine the local voltage control strategies of distributed generator operation in active distribution networks,” Applied Energy, vol. 228, pp. 2024–2036, 2018.
  • [29] G. Cavraro, Z. Yuan, M. K. Singh, and J. Cortés, “Learning local Volt/Var controllers towards efficient network operation with stability guarantees,” in IEEE Conf. on Decision and Control, (Cancun, Mexico), pp. 5056–5061, Dec. 2022.
  • [30] J. Feng, Y. Shi, G. Qu, S. H. Low, A. Anandkumar, and A. Wierman, “Stability constrained reinforcement learning for real-time voltage control in distribution systems,” arXiv preprint arXiv:2209.07669, 2022.
  • [31] A. M. Kettner and M. Paolone, “On the properties of the compound nodal admittance matrix of polyphase power systems,” IEEE Transactions on Power Systems, vol. 34, no. 1, pp. 444–453, 2019.
  • [32] M. Farivar, X. Zhou, and L. Chen, “Local voltage control in distribution systems: An incremental control algorithm,” in IEEE Int. Conf. on Communications, Control, and Computing Technologies for Smart Grids, (Miami, FL, USA), pp. 732–737, Nov. 2015.
  • [33] H. Daniels and M. Velikova, “Monotone and partially monotone neural networks,” IEEE Transactions on Neural Networks, vol. 21, no. 6, pp. 906–917, 2010.
  • [34] A. Wehenkel and G. Louppe, “Unconstrained monotonic neural networks,” in Conference on Neural Information Processing Systems, vol. 32, (Vancouver, Canada), pp. 1543–1553, Dec. 2019.
  • [35] X. Liu, X. Han, N. Zhang, and Q. Liu, “Certified monotonic neural networks,” in Conference on Neural Information Processing Systems, vol. 33, (Vancouver, Canada), pp. 15427–15438, Dec. 2020.
  • [36] I. Safran and O. Shamir, “Depth-width tradeoffs in approximating natural functions with neural networks,” in International Conference on Machine Learning, (Sydney, Australia), pp. 2979–2987, Aug. 2017.
  • [37] “Dataport: The world’s largest energy data resource,” Pecan Street Inc, 2015. Available at https://dataport.pecanstreet.org/.
  • [38] M. Grant and S. Boyd, “CVX: Matlab software for disciplined convex programming, version 2.1,” Mar. 2014. Available at http://cvxr.com/cvx.
  • [39] D. P. Kingma and J. Ba, “Adam: A method for stochastic optimization,” in International Conference on Learning Representations, (San Diego, CA, USA), May 2015.
  • [40] R. D. Zimmerman, C. E. Murillo-Sánchez, and R. J. Thomas, “Matpower: Steady-state operations, planning and analysis tools for power systems research education,” IEEE Transactions on Power Systems, vol. 26, no. 1, pp. 12–19, 2011.
  • [41] EEI-AEIC-UTC, “Smart meters and smart meter systems: A metering industry perspective,” Washington, DC, USA, Edison Elect. Inst., White Paper, 2011.
  • [42] R. A. Horn and C. R. Johnson, Matrix Analysis. Cambridge University Press, 2012.
  • [43] K. C. Border, Fixed Point Theorems with Applications to Economics and Game Theory. Cambridge, UK: Cambridge University Press, 1985.
[Uncaptioned image] Zhenyi Yuan (Student Member, IEEE) was born in Jiangxi, China, in 1998. He is currently working toward the Ph.D. degree in mechanical engineering at the University of California, San Diego, CA, USA. Prior to that, he was enrolled in successive undergraduate and postgraduate program at the Honors School of Harbin Institute of Technology, Harbin, China, where he received the B.S. and M.S. degrees in control science and engineering in 2018 and 2020, respectively. He was also a visiting research assistant with the Department of Information Engineering of the Chinese University of Hong Kong, Hong Kong, in 2021. His research interests lie at the intersection of control, optimization and learning, with applications to smart grids and robotic systems.
[Uncaptioned image] Guido Cavraro (Member, IEEE) received the Ph.D. degree in Information Engineering from the University of Padova, Italy, in 2015. He was a visiting scholar at the California Institute for Energy and Environment (CIEE) at U.C. Berkeley in 2014. In 2015 and 2016, he was a postdoctoral associate at the Department of Information Engineering of the University of Padova. From 2016 to 2018, he was a postdoctoral associate with the Bradley Department of Electrical and Computer Engineering of Virginia Tech, USA. Currently, he is a Senior Researcher with the Power Systems Engineering Center at National Renewable Energy Laboratory, USA. His research interests include control, optimization, and estimation with applications to power systems.
[Uncaptioned image] Manish K. Singh (Member, IEEE) received the B.Tech. degree from the Indian Institute of Technology (BHU), Varanasi, India, in 2013; and the M.S. and Ph.D. degrees in electrical engineering from Virginia Tech, Blacksburg, VA, USA, in 2018 and 2021, respectively. He is currently a postdoctoral researcher with the University of Minnesota, Minneapolis, MN. During 2013-2016, he worked as an Engineer in the Smart Grid Dept. of POWERGRID, the central transmission utility of India. His research interests are focused on optimization, control, and learning techniques to develop algorithmic solutions for the operation and analysis of electric power systems as well as water and natural gas networks.
[Uncaptioned image] Jorge Cortés (Fellow, IEEE) received the Licenciatura degree in mathematics from Universidad de Zaragoza, Zaragoza, Spain, in 1997, and the Ph.D. degree in engineering mathematics from Universidad Carlos III de Madrid, Madrid, Spain, in 2001. He held postdoctoral positions with the University of Twente, Twente, The Netherlands, and the University of Illinois at Urbana-Champaign, Urbana, IL, USA. He was an Assistant Professor with the Department of Applied Mathematics and Statistics, University of California, Santa Cruz, CA, USA, from 2004 to 2007. He is a Professor in the Department of Mechanical and Aerospace Engineering, University of California, San Diego, CA, USA. He is a Fellow of IEEE, SIAM, and IFAC. His research interests include distributed control and optimization, network science, nonsmooth analysis, reasoning and decision making under uncertainty, network neuroscience, and multi-agent coordination in robotic, power, and transportation networks.