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

Network-Aware and Welfare-Maximizing Dynamic Pricing for Energy Sharing

Ahmed S. Alahmed, Guido Cavraro, Andrey Bernstein, and Lang Tong Ahmed S. Alahmed and Lang Tong are with the School of Electrical and Computer Engineering, Cornell University, Ithaca, NY, USA ({ASA278, LT35}@cornell.edu). Guido Cavraro and Andrey Bernstein are with the Power System Engineering Center, National Renewable Energy Laboratory, Golden, CO, USA ({GUIDO.CAVRARO, ABERNSTE}@nrel.gov). 0000-0002-4715-4379 0000-0003-0296-720X 0000-0003-4489-8388 0000-0003-3322-2681
Abstract

The proliferation of behind-the-meter (BTM) distributed energy resources (DER) within the electrical distribution network presents significant supply and demand flexibilities, but also introduces operational challenges such as voltage spikes and reverse power flows. In response, this paper proposes a network-aware dynamic pricing framework tailored for energy-sharing coalitions that aggregate small, but ubiquitous, BTM DER downstream of a distribution system operator’s (DSO) revenue meter that adopts a generic net energy metering (NEM) tariff. By formulating a Stackelberg game between the energy-sharing market leader and its prosumers, we show that the dynamic pricing policy induces the prosumers toward a network-safe operation and decentrally maximizes the energy-sharing social welfare. The dynamic pricing mechanism involves a combination of a locational ex-ante dynamic price and an ex-post allocation, both of which are functions of the energy sharing’s BTM DER. The ex-post allocation is proportionate to the price differential between the DSO NEM price and the energy-sharing locational price. Simulation results using real DER data and the IEEE 13-bus test systems illustrate the dynamic nature of network-aware pricing at each bus, and its impact on voltage.

I Introduction

While BTM DER are primarily adopted to provide prosumer services such as bill savings and backup power, they can also be leveraged, under proper consumer-centric mechanism design, to provide various grid services such as voltage control, system support during contingencies, and new capacity deferrals [1]. Harnessing the flexibility of BTM DER participation in grid services is usually challenged by the DSO’s lack of visibility and controllability on BTM DER alongside the absence of network-aware pricing mechanisms that can induce favorable prosumer behaviors.

The rising notion of energy sharing of a group of prosumers under the DSO’s tariff presents a compelling solution to optimize DER utilization, comply with dynamic grid constraints, and promote renewable energy integration. A major barrier facing the practical implementation of energy-sharing markets is the incorporation of network constraints into their pricing, and aligning the objectives of the self-interested energy-sharing prosumers with the global objective of maximizing the coalition’s welfare.

Despite the voluminous literature on energy-sharing systems’ DER control and energy pricing, network constraints are rarely considered due to the theoretical complexity they introduce. A short list of recent works on energy communities and energy sharing that neglected network constraints can be found here [2, 3, 4, 5]. Some works considered a coarse notion of network constraints by incorporating operating envelopes (OEs) at the point of common coupling between the energy-sharing system and the DSO [6, 7] that limit the export and imports between the two entities. Others considered OEs at the prosumer’s level [8, 9]. Few papers considered network-aware pricing mechanisms in distribution networks, such as [10, 11, 12] and the line of literature on distribution locational marginal prices (dLMP), e.g. [13, 14].

Our work differs from the existing literature in two important directions. Firstly, we consider network-aware pricing under a generic DSO NEM tariff constraint that charges the energy-sharing platform different prices based on its aggregate net consumption. Secondly, the dynamic network-aware pricing of a platform that is subject to the DSO’s fixed and exogenous NEM price gives rise to a market manager’s profit/deficit that needs to be re-allocated to the coalition members. We shed light on a unique re-allocation rule that makes the prosumers’ payment functions uniform, even if they are located on different buses and the network constraints are binding. Such a re-allocation rule is highly relevant when charging end-users, as it avoids ‘undue discrimination’, which is one of the key principles of rate design outlined by Bonbright [15].

In this paper, we present a network-aware and welfare-maximizing pricing policy for energy-sharing coalitions that aggregate DER downstream of a DSO’s revenue meter that charges the energy-sharing platform based on a generic NEM tariff. The pricing policy announces an ex-ante locational, threshold-based, and dynamic price to induce a collective prosumer response that decentrally maximizes the social welfare, while abiding by the network voltage constraints. An ex-post charge/reward is then used to ensure the market operator’s profit neutrality. We show that the market mechanism achieves an equilibrium to the Stackelberg game between the energy-sharing market operator and its prosumers. Although network constraints couple the decisions of the energy-sharing prosumers, which give rise to locational marginal prices (LMP), we show that by adopting a unique proportional re-allocation rule, the payment function becomes uniform for all prosumers, even if they are located at different buses in the energy-sharing network. Numerical simulations using the IEEE 13-bus test feeder and real BTM DER data shed more light on how the pricing policy influences prosumers’ response to ensure safe network operation.

This paper extends our work on Dynamic NEM (D-NEM) without OEs [3] and with OEs [9] by incorporating network constraints, which add substantial complexity, primarily due to coupling of the DER decisions across network buses.

For the column vector 𝒙\bm{x}, [𝒙]+[\bm{x}]^{+}, and [𝒙][\bm{x}]^{-} represent its positive and negative elements. The notation [x]x¯x¯[x]_{\underline{x}}^{\overline{x}} represents the projection of xx into the closed and convex set [x¯,x¯][\underline{x},\overline{x}] as per the rule [x]x¯x¯:=max{x¯,min{x,x¯}}[x]_{\underline{x}}^{\overline{x}}:=\max\{\underline{x},\min\{x,\overline{x}\}\}. The notation is also used for vectors, i.e., [𝐱]𝐱¯𝐱¯[\bm{x}]_{\underline{\bm{x}}}^{\overline{\bm{x}}}.

II Proposed Framework and Network Model

We consider the problem of designing a welfare-maximizing and network-aware pricing policy for an energy sharing system that bidirectionally transacts energy and money with the DSO under a general NEM tariff. Under NEM, the energy sharing platform imports from the DSO at an import rate if it is net-consuming, and collectively exports from the DSO at the export rate if it is net-producing. A budget-balanced market operator is responsible for announcing the market’s pricing policy. The market operator uses spatially varying pricing signals to adhere to its network’s operational constraints communicated by the DSO.111We posit that such DER aggregation schemes are informed by the DSO about their networks’ information, including OEs, line thermal limits, voltage limits, among others. We assume that the timescale of community members’ decision is equivalent to that of the NEM netting period [16], which allows us to adopt a single time step formulation.

A radial low voltage distribution network flow model is used to model the network power flow [17, 18]. Consider a radial distribution network described by 𝒢=(,)\mathcal{G}=({\cal B},\mathcal{L}), with ={1,,B}{\cal B}=\{1,\ldots,B\} as the set of energy sharing buses, excluding bus 0, and ={(i,j)}×\mathcal{L}=\{(i,j)\}\subset{\cal B}\times{\cal B} as the set of distribution lines between the buses, with i,ji,j as bus indices. The root bus 0 represents the secondary of the transformer and is referred to as the slack bus (substation bus). The natural radial network orientation is considered, with each distribution line pointing away from bus 0.

For each bus ii\in{\cal B}, denote by i\mathcal{L}_{i}\subseteq\mathcal{L} the set of lines on the unique path from buses 0 to ii, and by Zi,qiZ_{i},q_{i} the active and reactive power consumptions of bus ii, respectively. The magnitude of the complex voltage at bus ii is denoted by viv_{i}, and we denote the fixed and known voltage at the slack bus by v0v_{0}. For each line (i,j)(i,j)\in\mathcal{L}, denote by rijr_{ij} and xijx_{ij} its resistance and reactance. For each line, (i,j)(i,j)\in\mathcal{L}, denote by PijP_{ij} and QijQ_{ij} the real and reactive power from buses ii to jj, respectively. Let ij\ell_{ij} denote the squared magnitude of the complex branch current from ii to jj.

We adopt the distribution flow (DistFlow) model, introduced in [17], to model steady state power flow in a radial distribution network, as

Pij\displaystyle P_{ij} =Zj+k:(j,k)Pjk+rijij\displaystyle=-Z_{j}+\sum_{k:(j,k)\in\mathcal{L}}P_{jk}+r_{ij}\ell_{ij} (1a)
Qij\displaystyle Q_{ij} =qj+k:(j,k)Qjk+xijij\displaystyle=-q_{j}+\sum_{k:(j,k)\in\mathcal{L}}Q_{jk}+x_{ij}\ell_{ij} (1b)
vj2\displaystyle v_{j}^{2} =vi22(rijPij+xijQij)+(rij2+xij2)ij,\displaystyle=v_{i}^{2}-2\left(r_{ij}P_{ij}+x_{ij}Q_{ij}\right)+\left(r_{ij}^{2}+x_{ij}^{2}\right)\ell_{ij}, (1c)

where ij=(Pij2+Qij2)/vi2\ell_{ij}=(P^{2}_{ij}+Q_{ij}^{2})/v_{i}^{2} is the line losses, (1a)-(1b) are the active and reactive power balance equations, and (1c) is the voltage drop. We exploit a linear approximation of the DistFlow model above that ignores line losses, given that in practice ij0\ell_{ij}\approx 0 for all (i,j)(i,j)\in\mathcal{L}. Therefore, the linearized Distflow (LinDistFlow) equations are given by re-writing (1a)-(1c) to

Pij\displaystyle P_{ij} =k𝒪(j)Zk,Qij=k𝒪(j)qk\displaystyle=-\sum_{k\in{\cal O}(j)}Z_{k},\quad\quad\quad Q_{ij}=-\sum_{k\in{\cal O}(j)}q_{k}
vj2\displaystyle v_{j}^{2} =vi22(rijPij+xijQij),\displaystyle=v_{i}^{2}-2\left(r_{ij}P_{ij}+x_{ij}Q_{ij}\right),

where 𝒪(j){\cal O}(j) represent the set node jj’s descendants, including node jj, i.e., 𝒪(j):={i:ji}{\cal O}(j):=\left\{i:\mathcal{L}_{j}\subseteq\mathcal{L}_{i}\right\}. This gives a solution for vi2v_{i}^{2} in terms of v02v_{0}^{2}, as

v02vi2=2jR~ijZj2jX~ijqj,v_{0}^{2}-v_{i}^{2}=-2\sum_{j\in{\cal B}}\tilde{R}_{ij}Z_{j}-2\sum_{j\in{\cal B}}\tilde{X}_{ij}q_{j},

where

R~ij:=(h,k)ijrhk,X~ij:=(h,k)ijxhk\tilde{R}_{ij}:=\sum_{(h,k)\in\mathcal{L}_{i}\cap\mathcal{L}_{j}}r_{hk},\quad\tilde{X}_{ij}:=\sum_{(h,k)\in\mathcal{L}_{i}\cap\mathcal{L}_{j}}x_{hk} (3)

The LinDistFlow can be compactly written as,

𝒗=𝑹𝒁𝑿𝒒+v02𝟏,\bm{v}=-\bm{R}\bm{Z}-\bm{X}\bm{q}+v_{0}^{2}\bm{1}, (4)

where 𝒗:=(v12,,vB2),𝒁:=(Z1,,ZB),𝒒:=(q1,,qB)\bm{v}:=(v_{1}^{2},\ldots,v_{B}^{2}),\bm{Z}:=(Z_{1},\ldots,Z_{B}),\bm{q}:=(q_{1},\ldots,q_{B}), and 𝑹:=[2R~ij]B×B\bm{R}:=[2\tilde{R}_{ij}]_{B\times B} and 𝑿:=[2X~ij]B×B\bm{X}:=[2\tilde{X}_{ij}]_{B\times B} are the resistance and reactance matrices, respectively. We treat the reactive power 𝒒\bm{q} as given constants rather than decision variables, which allows us to write (3) as

𝒗=𝑹𝒁+𝒗^,\bm{v}=-\bm{R}\bm{Z}+\bm{\hat{v}}, (5)

where 𝒗^:=𝑿𝒒+v02𝟏\bm{\hat{v}}:=-\bm{X}\bm{q}+v_{0}^{2}\bm{1}. The voltage magnitude vector above is constrained as

𝒗min𝒗𝒗max,\bm{v}_{\text{min}}\preceq\bm{v}\preceq\bm{v}_{\text{max}}, (6)

where 𝒗min:=vmin2𝟏\bm{v}_{\text{min}}:=v_{\text{min}}^{2}\bm{1} and 𝒗max:=vmax2𝟏\bm{v}_{\text{max}}:=v_{\text{max}}^{2}\bm{1}. Given that the second term in (5) is fixed, we re-write (6) to

𝒗¯𝒗𝒗¯,\bm{\underline{v}}\preceq\bm{v}\preceq\bm{\overline{v}}, (7)

where 𝒗¯:=𝒗max𝒗^\overline{\bm{v}}:=\bm{v}_{\text{max}}-\bm{\hat{v}} and 𝒗¯:=𝒗^𝒗min\underline{\bm{v}}:=\bm{\hat{v}}-\bm{v}_{\text{min}}. We will impose (7) on the operation of the energy-sharing market.

III Energy Sharing Mathematical Model

Let 𝒩:={1,,N}{\cal N}:=\{1,\ldots,N\} denote the set of energy-sharing system’s prosumers. Every prosumer nn is connected to one of the BB buses in the considered radial network through its revenue meter that measures the prosumer’s net consumption and BTM generation. Figure 1 shows an example 4-bus energy-sharing platform. We denote the set of prosumers connected to bus ii\in{\cal B} by 𝒩i{\cal N}_{i}, hence, 𝒩=i𝒩i{\cal N}=\bigcup_{i\in{\cal B}}{\cal N}_{i}. In this section, we model prosumers’ DER in §\SIII-A, and payment and surplus functions in §\SIII-B, followed by a formulation of the proposed bi-level program in §\SIII-C.

Refer to caption
Figure 1: A 4-bus energy-sharing platform. Z0,Zi,znZ_{0},Z_{i},z_{n}\in\mathbb{R} are the net consumption of the whole energy sharing platform, net consumption of bus ii, and net consumption of prosumer nn, respectively. z¯n0\overline{z}_{n}\geq 0 and z¯n0\underline{z}_{n}\leq 0 are the prosumer’s import and export OEs, respectively.

III-A DER Modeling

Prosumers’ DER consists of BTM renewable distributed generation (DG) and flexible loads (decision variables). The random renewable DG output of every prosumer n𝒩n\in{\cal N} is denoted by gn+g_{n}\in\mathbb{R}_{+}. The vector of prosumers’ DG output is denoted by 𝒈:=(g1,,gN)\bm{g}:=(g_{1},\ldots,g_{N}), and the aggregate DG output in the energy-sharing platform is defined by G0=n𝒩gnG_{0}=\sum_{n\in{\cal N}}g_{n}.

The flexible loads’ energy consumption vector is denoted by 𝒅n+K\bm{d}_{n}\in\mathbb{R}_{+}^{K}, where 𝒦:={1,,K}{\cal K}:=\{1,\ldots,K\} is the load (device) bundle. The devices are subject to their flexibility limits, as

𝒅n𝒟n:=[𝒅¯n,𝒅¯n],n𝒩,\bm{d}_{n}\in{\cal D}_{n}:=[\underline{\bm{d}}_{n},\overline{\bm{d}}_{n}],~{}~{}~{}\forall n\in{\cal N}, (8)

where 𝒅¯n\underline{\bm{d}}_{n} and 𝒅¯n\overline{\bm{d}}_{n} are the device bundle’s lower and upper consumption limits of prosumer n𝒩n\in{\cal N}, respectively.

The net consumption znz_{n}\in\mathbb{R} of each prosumer is the difference between its gross consumption and BTM generation, hence zn=𝟏𝒅ngnz_{n}=\bm{1}^{\top}\bm{d}_{n}-g_{n}.222The proposed pricing policy can be generalized to incorporate OEs with only little mathematical complication. We show this in the appendix. The aggregate energy-sharing net consumption is therefore Z0=n𝒩zn=iZiZ_{0}=\sum_{n\in{\cal N}}z_{n}=\sum_{i\in{\cal B}}Z_{i}.

III-B Payment, Surplus, and Profit Neutrality

The energy sharing operator designs a pricing policy χ\chi for its members, which specifies the payment function for each prosumer n𝒩n\in{\cal N} under χ\chi, denoted by Cnχ(zn)C^{\chi}_{n}(z_{n}). Energy-sharing prosumers are assumed to be rational and self-interested. Therefore, they schedule their DER based on surplus maximization. Prosumer surplus is given by

Snχ(𝒅n,gn):=Un(𝒅n)Cnχ(zn),zn=𝟏𝒅ngn,S^{\chi}_{n}(\bm{d}_{n},g_{n}):=U_{n}(\bm{d}_{n})-C^{\chi}_{n}(z_{n}),~{}~{}z_{n}=\bm{1}^{\top}\bm{d}_{n}-g_{n}, (9)

where for every n𝒩n\in{\cal N}, the utility function of the consumption bundle Un(𝒅n)U_{n}(\bm{d}_{n}) is assumed to be non-decreasing, additive, strictly concave, and continuously differentiable with a marginal utility function 𝑳n:=Un=(Ln1,,LnK)\bm{L}_{n}:=\nabla U_{n}=\left(L_{n1},\ldots,L_{nK}\right). Denote the inverse marginal utility vector by 𝒇n:=(fn1,,fnK)\bm{f}_{n}:=(f_{n1},\ldots,f_{nK}) with fnk:=Lnk1,n𝒩,k𝒦f_{nk}:=L^{-1}_{nk},\forall n\in{\cal N},k\in{\cal K}.

III-B1 Energy Sharing Payment

The operator transacts with the DSO under the NEM X tariff, introduced in [19], which charges the energy sharing coalition based on whether it is net-importing (Z0>0Z_{0}>0) or net-exporting (Z0<0Z_{0}<0) as

πNEM(Z0)={π+,Z00π,Z0<0,CNEM(Z0)=πNEM(Z0)Z0,\displaystyle\pi^{\mbox{\tiny NEM}}(Z_{0})=\begin{cases}\pi^{+},&Z_{0}\geq 0\\ \pi^{-},&Z_{0}<0\end{cases},C^{\mbox{\tiny NEM}}(Z_{0})=\pi^{\mbox{\tiny NEM}}(Z_{0})\cdot Z_{0}, (10)

where (π+,π)+(\pi^{+},\pi^{-})\in\mathbb{R}_{+} are the buy (retail) and sell (export) rates, respectively. We assume π+π\pi^{+}\geq\pi^{-}, in accordance with NEM practice [16], which also eliminates risk-free price arbitrage. The operator of the energy sharing regime is profit-neutral; a term we define next.

Definition 1 (Profit neutrality).

The operator is profit-neutral if its pricing achieves the following

n𝒩Cnχ(zn)=CNEM(n𝒩zn).\sum_{n\in{\cal N}}C^{\chi}_{n}(z_{n})=C^{\mbox{\tiny NEM}}(\sum_{n\in{\cal N}}z_{n}).

The challenging question we ask is how can the operator design a payment CnχC^{\chi}_{n}, for every n𝒩n\in{\cal N}, to achieve network-awareness, profit neutrality and equilibrium to the energy-sharing market, which we define next.

III-C Energy Sharing Stackelberg Game

We formulate this game as a bi-level mathematical program with the upper-level optimization being the operator’s pricing problem, and the lower-level optimizations representing prosumers’ optimal decisions.

Denote the consumption policy of the nnth prosumer, given the pricing policy χ\chi, by ψn,χ\psi_{n,\chi}. Formally,

ψn,χ:+𝒟n,gnCnχψn,χ(gn),\psi_{n,\chi}:\mathbb{R}_{+}\rightarrow{\cal D}_{n},\ g_{n}\stackrel{{\scriptstyle C^{\chi}_{n}}}{{\mapsto}}\psi_{n,\chi}(g_{n}),

with 𝝍χ:={ψ1,χ,,ψN,χ}\bm{\psi}_{\chi}:=\{\psi_{1,\chi},\ldots,\psi_{N,\chi}\} as the vector of prosumers’ policies. The operator strives to design a network-aware and welfare-maximizing pricing policy χ𝝍\chi^{\sharp}_{\bm{\psi}} (given 𝝍\bm{\psi}), where χ𝝍:+NN,𝒈𝑪χ:=(C1χ,,CNχ),\chi^{\sharp}_{\bm{\psi}}:\mathbb{R}_{+}^{N}\rightarrow\mathbb{R}^{N},\ \bm{g}\mapsto\bm{C}^{\chi}:=(C^{\chi}_{1},\ldots,C^{\chi}_{N}), and the welfare is defined as the sum of total prosumers’ surplus, as

Wχ,𝝍χ:=n𝒩Snχ(ψn,χ(gn),gn),W^{\chi,\bm{\psi}_{\chi}}:=\sum_{n\in{\cal N}}S_{n}^{\chi}(\psi_{n,\chi}(g_{n}),g_{n}),

The bi-level program can be compactly formulated as

maximize𝑪()\displaystyle\underset{\bm{C}(\cdot)}{\operatorname{maximize}} (Wχ𝝍=n𝒩Un(𝒅nψχ)CNEM(Z0ψχ))\displaystyle\Bigg{(}W^{\chi_{\bm{\psi}}}=\sum_{n\in{\cal N}}U_{n}(\bm{d}^{\psi^{\sharp}_{\chi}}_{n})-C^{\mbox{\tiny NEM}}(Z_{0}^{\psi^{\sharp}_{\chi}})\Bigg{)} (11a)
subject to n𝒩Cnχ(znψχ)=CNEM(Z0ψχ)\displaystyle\quad\quad\sum_{n\in{\cal N}}C^{\chi}_{n}(z_{n}^{\psi^{\sharp}_{\chi}})=C^{\mbox{\tiny NEM}}(Z_{0}^{\psi^{\sharp}_{\chi}}) (11b)
Z0ψχ=n𝒩(𝟏𝒅nψχgn)\displaystyle\quad\quad Z_{0}^{\psi^{\sharp}_{\chi}}=\sum_{n\in{\cal N}}\big{(}\bm{1}^{\top}\bm{d}^{\psi^{\sharp}_{\chi}}_{n}-g_{n}\big{)} (11c)
(𝜼¯,𝜼¯)\displaystyle(\underline{\bm{\eta}},\overline{\bm{\eta}}) 𝒗¯𝑹𝒁ψχ𝒗¯\displaystyle\quad\quad\underline{\bm{v}}\preceq-\bm{R}\bm{Z}^{\psi^{\sharp}_{\chi}}\preceq\overline{\bm{v}} (11d)
for all n𝒩\displaystyle\quad\quad\text{for all }n\in{\cal N} (11e)
𝒅nψχ:=argmax𝒅n𝒟nSnχ(𝒅n,gn):=Un(𝒅n)Cnχ(zn)\displaystyle\bm{d}^{\psi^{\sharp}_{\chi}}_{n}:=\underset{\bm{d}_{n}\in{\cal D}_{n}}{\operatorname{argmax}}~{}S^{\chi}_{n}(\bm{d}_{n},g_{n}):=U_{n}(\bm{d}_{n})-C^{\chi}_{n}(z_{n}) (11f)
subject tozn=𝟏𝒅ngn,\displaystyle\hskip 36.98866pt\text{subject to}~{}~{}z_{n}=\bm{1}^{\top}\bm{d}_{n}-g_{n}, (11g)

where

𝒁ψχ:=(n𝒩1𝟏𝒅nψχgn,,n𝒩B𝟏𝒅nψχgn).\bm{Z}^{\psi^{\sharp}_{\chi}}:=(\sum_{n\in{\cal N}_{1}}\bm{1}^{\top}\bm{d}^{\psi^{\sharp}_{\chi}}_{n}-g_{n},\ldots,\sum_{n\in{\cal N}_{B}}\bm{1}^{\top}\bm{d}^{\psi^{\sharp}_{\chi}}_{n}-g_{n}).

In the following, we will assume that problem (11) is feasible, i.e., a solution meeting all the constraints exists.

The program in (11) defines the Stackelberg strategy. Specifically, (χ,𝝍\chi^{\ast},\bm{\psi}^{\ast}) is a Stackelberg equilibrium since (a) for all χ𝒳\chi\in{\cal X} and n𝒩n\in{\cal N}, Snχ(ψn(gn),gn)Snχ(ψn(gn),gn)S^{\chi}_{n}(\psi^{\ast}_{n}(g_{n}),g_{n})\geq S^{\chi}_{n}(\psi_{n}(g_{n}),g_{n}) for all 𝝍Ψ\bm{\psi}\in\Psi; (b) for all 𝝍Ψ,Wχ,𝝍nSnχ(ψn(gn),gn)\bm{\psi}\in\Psi,W^{\chi^{\ast},\bm{\psi}^{\ast}}\geq\sum_{n}S_{n}^{\chi}(\psi^{\ast}_{n}(g_{n}),g_{n}).

IV Network-Aware Pricing and Equilibrium

At the beginning of each pricing period, the operator communicates the price to each prosumer. Given the price, prosumers simultaneously move to solve their own surplus maximization problem. At the end of the netting period, and given the resulting Z0Z_{0}, the DSO charges the energy sharing operator based on the NEM X tariff in (10). We propose the network-aware pricing policy and delineate its structure in §\SIV-A, followed by solving the optimal response of prosumers in §\SIV-B. We discuss the operator’s profit/deficit redistribution in §\SIV-C and §\SIV-E. In §\SIV-D, we establish the market equilibrium result.

IV-A Network-Aware Dynamic Pricing

The operator uses the renewable DG vector 𝒈\bm{g} to dynamically set the price taking into account network constraints. That is, the dynamic price is used to satisfy network constraints in a decentralized way by internalizing them into prosumers’ private decisions.

Network-aware pricing policy 1.

For every bus ii\in{\cal B}, the operator charges the prosumers based on a two-part pricing

χ:𝒈Cnχ(zn)=πi(𝒈)ex-ante priceznAnex-post allocation,n𝒩i,\chi^{\ast}:\bm{g}\mapsto C^{\chi^{\ast}}_{n}(z_{n})=\underbrace{\pi^{\ast}_{i}(\bm{g})}_{\text{ex-ante price}}\cdot z_{n}-\underbrace{A_{n}^{\ast}}_{\text{ex-post allocation}},\forall n\in{\cal N}_{i}, (12)

where the ex-ante bus price πi(𝐠)\pi^{\ast}_{i}(\bm{g}) abides by a two-threshold policy with thresholds

σ1(𝒈)\displaystyle\sigma_{1}(\bm{g}) =in𝒩i𝟏[𝒇n(𝟏χi+(𝒈))]𝒅¯n𝒅¯n,\displaystyle=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}\bm{1}^{\top}{[\bm{f}_{n}(\bm{1}\chi_{i}^{+}(\bm{g}))]}_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}}, (13)
σ2(𝒈)\displaystyle\sigma_{2}(\bm{g}) =in𝒩i𝟏[𝒇n(𝟏χi(𝒈))]𝒅¯n𝒅¯nσ1(𝒈),\displaystyle=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}\bm{1}^{\top}{[\bm{f}_{n}(\bm{1}\chi_{i}^{-}(\bm{g}))]}_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}}\geq\sigma_{1}(\bm{g}),

as

πi(𝒈)={χi+(𝒈),G0<σ1(𝒈)χiz(𝒈),G0[σ1(𝒈),σ2(𝒈)]χi(𝒈),G0>σ2(𝒈),\pi^{\ast}_{i}(\bm{g})=\begin{cases}\chi^{+}_{i}(\bm{g})&,G_{0}<\sigma_{1}(\bm{g})\\ \chi^{z}_{i}(\bm{g})&,G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})]\\ \chi^{-}_{i}(\bm{g})&,G_{0}>\sigma_{2}(\bm{g}),\end{cases}\quad (14)

and the price χiκ\chi_{i}^{\kappa}, where κ:={+,,z}\kappa:=\{+,-,z\}, is given by

χiκ=πκj=1BRji(η¯jη¯j)\chi_{i}^{\kappa}=\pi^{\kappa}-\sum_{j=1}^{B}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j}) (15)

where η¯j\overline{\eta}^{\ast}_{j} and η¯j\underline{\eta}^{\ast}_{j} are the dual variables of the upper and lower voltage limits in (11d), respectively, and the price πz:=μ\pi^{z}:=\mu^{\ast} is the solution of

in𝒩i𝟏[𝒇n(𝟏μ𝟏j=1BRji(η¯jη¯j))]𝒅¯n𝒅¯n=G0.\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}\bm{1}^{\top}[\bm{f}_{n}(\bm{1}\mu-\bm{1}\sum_{j=1}^{B}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j}))]_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}}=G_{0}. (16)

The two pricing policy parts are composed of a locational dynamic price that is announced ex-ante and a charge (reward) that is distributed ex-post. For every bus ii\in{\cal B}, the prosumer’s ex-post charge/reward is denoted by AnA_{n}^{\ast}, which we delineate in §\SIV-C and §\SIV-E.

The locational ex-ante price πi(𝒈)\pi^{\ast}_{i}(\bm{g}) for every ii\in{\cal B} is used to induce a collective prosumer response at each bus so that the network constraints are satisfied and the energy sharing social welfare is maximized. The energy-sharing price has a similar structure to the celebrated LMP in wholesale markets [20] in the sense that it takes into account demand, generation, location, and network physical limits. Also, like congestionless LMP, the energy-sharing price is uniform across all buses if the network constraints are nonbinding, as described in (15).

Similar to D-NEM without network constraints [3], the price obeys a two-threshold policy and it is a monotonically decreasing function of the system’s renewables 𝒈\bm{g}. As shown in (15), the thresholds partition G0G_{0}, and the price at each bus is the D-NEM price adjusted by the shadow prices of violating voltage limits. When G0[σ1(𝒈),σ2(𝒈)]G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})] the platform is energy-balanced, and the price χiz(𝒈)\chi^{z}_{i}(\bm{g}) is the sum of the dual variables for energy balance and voltage limits.

The thresholds and locational prices can be computed while preserving prosumers’ privacy. The operator do not need the functional form of prosumers’ utilities or marginal utilities but rather asks the prosumers to submit a value for every device kk at a given price.

IV-B Optimal Prosumer Decisions

After the network-aware price is announced, consumers simultaneously move to solve their own surplus maximization problem by determining their optimal decision policy ψn,χ:+𝒟n,gnCnχ𝒅nψ:=ψn(gn),n𝒩\psi^{\ast}_{n,\chi^{\ast}}:\mathbb{R}_{+}\rightarrow{\cal D}_{n},\ g_{n}\stackrel{{\scriptstyle{C^{\chi^{\ast}}_{n}}}}{{\mapsto}}\bm{d}_{n}^{\psi^{\ast}}:=\psi^{\ast}_{n}(g_{n}),\forall n\in{\cal N}. Therefore, from the surplus definition in (9), each prosumer solves

𝒅nψ=\displaystyle\bm{d}_{n}^{\psi^{\ast}}= argmax𝒅n𝒟nSnχ(𝒅n,gn):=Un(𝒅n)πi(𝒈)zn\displaystyle\underset{\bm{d}_{n}\in{\cal D}_{n}}{\operatorname{argmax}}~{}~{}S_{n}^{\chi^{\ast}}(\bm{d}_{n},g_{n}):=U_{n}\left(\bm{d}_{n}\right)-\pi^{\ast}_{i}(\bm{g})\cdot z_{n}
subject to zn=𝟏𝒅nrn,\displaystyle\text{subject to }\hskip 5.69046ptz_{n}=\bm{1}^{\top}\bm{d}_{n}-r_{n}, (17)

where AnA^{\ast}_{n} was omitted because it is announced after consumption decisions are exercised.

Lemma 1 (Prosumer optimal consumption).

Under every bus ii\in{\cal B}, given the pricing policy χ\chi^{\ast}, the prosumer’s optimal consumption is

𝒅nψ(πi)=[𝒇n(𝟏πi)]𝒅¯n𝒅¯n,n𝒩i.\displaystyle\bm{d}^{\psi^{\ast}}_{n}(\pi^{\ast}_{i})=[\bm{f}_{n}(\bm{1}\pi^{\ast}_{i})]_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}},~{}~{}\forall n\in{\cal N}_{i}. (18)

By definition, the aggregate net consumption is

Z0ψ(𝝅(𝒈))=in𝒩i{𝒅nψ(χi+)gn,G0<σ1(𝒈)𝒅nψ(χiz)gn,G0[σ1(𝒈),σ2(𝒈)]𝒅nψ(χi)gn,G0>σ2(𝒈),\displaystyle Z_{0}^{\psi^{\ast}}(\bm{\pi}^{\ast}(\bm{g}))=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}\begin{cases}\bm{d}^{\psi^{\ast}}_{n}(\chi^{+}_{i})-g_{n}&,G_{0}<\sigma_{1}(\bm{g})\\ \bm{d}^{\psi^{\ast}}_{n}(\chi^{z}_{i})-g_{n}&,G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})]\\ \bm{d}^{\psi^{\ast}}_{n}(\chi^{-}_{i})-g_{n}&,G_{0}>\sigma_{2}(\bm{g}),\end{cases} (19)

where 𝛑:=(π1,,πB)\bm{\pi}^{\ast}:=(\pi^{\ast}_{1},\ldots,\pi^{\ast}_{B}), and Z0ψ(𝛑(𝐠))>0Z_{0}^{\psi^{\ast}}(\bm{\pi}^{\ast}(\bm{g}))>0 if G0<σ1(𝐠)G_{0}<\sigma_{1}(\bm{g}), Z0ψ(𝛑(𝐠))=0Z_{0}^{\psi^{\ast}}(\bm{\pi}^{\ast}(\bm{g}))=0 if G0[σ1(𝐠),σ2(𝐠)]G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})], and Z0ψ(𝛑(𝐠))<0Z_{0}^{\psi^{\ast}}(\bm{\pi}^{\ast}(\bm{g}))<0 if G0>σ2(𝐠)G_{0}>\sigma_{2}(\bm{g}).

Proof.

We drop the prosumer subscript nn for brevity. The objective in (17) is strictly concave and differentiable. The Lagrangian function of the surplus maximization problem, for a prosumer under bus ii, is

(𝒅,𝜸¯,𝜸¯)=πi(𝒈)zU(𝒅)+𝜸¯(𝒅𝒅¯)𝜸¯(𝒅𝒅¯),\mathscr{L}(\bm{d},\overline{\bm{\gamma}},\underline{\bm{\gamma}})=\pi^{\ast}_{i}(\bm{g})\cdot z-U\left(\bm{d}\right)+\overline{\bm{\gamma}}^{\top}(\bm{d}-\overline{\bm{d}})-\underline{\bm{\gamma}}^{\top}(\bm{d}-\underline{\bm{d}}),

where 𝜸¯+K\overline{\bm{\gamma}}\in\mathbb{R}_{+}^{K} and 𝜸¯+K\underline{\bm{\gamma}}\in\mathbb{R}_{+}^{K} are the Lagrangian multipliers of the upper and lower consumption limits. From the KKT conditions we have

𝒅=𝟏πi(𝒈)𝑳(𝒅ψ)+𝜸¯𝜸¯=𝟎,\nabla_{\bm{d}}\mathscr{L}=\bm{1}\pi^{\ast}_{i}(\bm{g})-\bm{L}(\bm{d}^{\psi^{\ast}})+\overline{\bm{\gamma}}-\underline{\bm{\gamma}}=\bm{0},

therefore, for each device k𝒦k\in{\cal K}, we have

dkψ\displaystyle d_{k}^{\psi^{\ast}} ={fk(πi),γ¯k=γ¯k=0d¯k,γ¯k>0,γ¯k=0d¯k,γ¯k=0,γ¯k>0\displaystyle=\begin{cases}f_{k}(\pi^{\ast}_{i})&,\overline{\gamma}_{k}=\underline{\gamma}_{k}=0\\ \overline{d}_{k}&,\overline{\gamma}_{k}>0,\underline{\gamma}_{k}=0\\ \underline{d}_{k}&,\overline{\gamma}_{k}=0,\underline{\gamma}_{k}>0\end{cases}
=:[fk(πi)]d¯kd¯k,\displaystyle=:[f_{k}(\pi^{\ast}_{i})]_{\underline{d}_{k}}^{\overline{d}_{k}},

where fk:=Lk1f_{k}:=L^{-1}_{k}.

Give the aggregate net consumption definition Z0=n𝒩(𝟏𝒅ngn)Z_{0}=\sum_{n\in{\cal N}}(\bm{1}^{\top}\bm{d}_{n}-g_{n}) and the dynamic price in (14), one can easily get (19). Finally, from (13), we can re-formulate (19) as

Z0ψ(𝝅(𝒈))\displaystyle Z_{0}^{\psi^{\ast}}(\bm{\pi}^{\ast}(\bm{g})) ={σ1(𝒈)G0,G0<σ1(𝒈)0,G0[σ1(𝒈),σ2(𝒈)]σ2(𝒈)G0,G0>σ2(𝒈),\displaystyle=\begin{cases}\sigma_{1}(\bm{g})-G_{0}&,G_{0}<\sigma_{1}(\bm{g})\\ 0&,G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})]\\ \sigma_{2}(\bm{g})-G_{0}&,G_{0}>\sigma_{2}(\bm{g}),\end{cases}

which proves the sign of Z0ψ(𝝅(𝒈))Z_{0}^{\psi^{\ast}}(\bm{\pi}^{\ast}(\bm{g})) under each piece. ∎

IV-C Ex-Post Allocation

Unlike the ex-ante price, the ex-post allocation is distributed after the prosumers schedule their DER. The operator may choose to accrue the ex-post charge amount of each prosumer to be distributed after multiple netting periods rather than at every netting period. The ex-post fee is essentially levied to achieve profit neutrality. After the price is announced and the transaction with the DSO is settled, the profit/deficit that the operator accumulates A(𝒈):=in𝒩iAnA^{\ast}(\bm{g}):=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}A^{\ast}_{n} is, using Def.1,

in𝒩iCnχ(zn)CNEM(n𝒩zn)\displaystyle\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}C^{\chi}_{n}(z_{n})-C^{\mbox{\tiny NEM}}(\sum_{n\in{\cal N}}z_{n}) =0(10),(12)\displaystyle=0\stackrel{{\scriptstyle\text{(\ref{eq:Pcommunity}),(\ref{eq:PricingPolicy})}}}{{\Longrightarrow}}
in𝒩i(πi(𝒈)znAnπNEM(Z0)zn)\displaystyle\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}(\pi^{\ast}_{i}(\bm{g})\cdot z_{n}-A^{\ast}_{n}-\pi^{\mbox{\tiny NEM}}(Z_{0})\cdot z_{n}) =0\displaystyle=0
in𝒩i(πi(𝒈)znπNEM(Z0)zn)\displaystyle\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}(\pi^{\ast}_{i}(\bm{g})\cdot z_{n}-\pi^{\mbox{\tiny NEM}}(Z_{0})\cdot z_{n}) =in𝒩iAn\displaystyle=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}A^{\ast}_{n}
in𝒩i(πi(𝒈)πNEM(Z0))zn\displaystyle\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}(\pi^{\ast}_{i}(\bm{g})-\pi^{\mbox{\tiny NEM}}(Z_{0}))\cdot z_{n} =A(𝒈).\displaystyle=A^{\ast}(\bm{g}).

One can see that the larger the differential between the energy sharing price and NEM price (πi(𝒈)πNEM(Z0),i\pi^{\ast}_{i}(\bm{g})-\pi^{\mbox{\tiny NEM}}(Z_{0}),\forall i\in{\cal B}), the larger the profit/deficit. Note that if the network constraints are non-binding, i.e., η¯i=η¯i=0,i\overline{\eta}^{\ast}_{i}=\underline{\eta}^{\ast}_{i}=0,\forall i\in{\cal B}, then A(𝒈)=0A^{\ast}(\bm{g})=0, and the pricing policy becomes one-part; see D-NEM in [3].

There might not be unique way to re-allocate the operator’s profit/deficit A(𝒈)A^{\ast}(\bm{g}). A profit-sharing coalitional game can be established to fairly re-allocate the operator’s profit/deficit. In §\SIV-E, we propose a proportional allocation rule that makes the payment function uniform for all prosumers.

IV-D Stackelberg Equilibrium

Under the solution (χ,ψ)(\chi^{\ast},\psi^{\ast}), with A(𝒈)A^{\ast}(\bm{g}) as in (IV-C) ,the operator is profit-neutral.

We show next that the network-aware pricing achieves a Nash equilibrium to the leader-follower game in $\$III-C.

Theorem 1.

The solution (χ,ψ\chi^{\ast},\psi^{\ast}) is a Stackelberg equilibrium that also achieves social optimality, i.e.,

(𝒅1ψ,,𝒅Nψ)=argmax(𝒅1,,𝒅N)\displaystyle(\bm{d}_{1}^{\psi^{\ast}},\ldots,\bm{d}_{N}^{\psi^{\ast}})=\underset{(\bm{d}_{1},\ldots,\bm{d}_{N})}{\rm argmax} in𝒩iUn(𝒅n)CNEM(Z0)\displaystyle~{}~{}\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}U_{n}(\bm{d}_{n})-C^{\mbox{\tiny NEM}}(Z_{0})
subject to Z0=n𝒩(𝟏𝒅ngn)\displaystyle~{}~{}Z_{0}=\sum_{n\in{\cal N}}\big{(}\bm{1}^{\top}\bm{d}_{n}-g_{n}\big{)}
𝒅n𝒟nn𝒩\displaystyle\hskip 11.38092pt\bm{d}_{n}\in{\cal D}_{n}~{}\forall n\in{\cal N}
𝒗¯𝑹𝒁𝒗¯.\displaystyle\hskip 11.38092pt\underline{\bm{v}}\preceq-\bm{R}\bm{Z}\preceq\overline{\bm{v}}.
Proof.

See the appendix. ∎

The proof of Theorem 1 solves an upper bound of (11) that relaxes the profit-neutrality constraint (11b).

IV-E Energy Sharing Payment Uniformity

We propose here a unique way to allocate the operator’s profit/deficit A(𝒈)A^{\ast}(\bm{g}). For every bus ii\in{\cal B}, the allocation to every prosumer is given by

An(𝒈)=(πi(𝒈)πNEM(n𝒩zn))zn,n𝒩i,A_{n}^{\ast}(\bm{g})=\left(\pi^{\ast}_{i}(\bm{g})-\pi^{\mbox{\tiny NEM}}(\sum_{n\in{\cal N}}z_{n})\right)\cdot z_{n},\forall n\in{\cal N}_{i}, (20)

which has three favourable features. First, it redistributes the profit/deficit proportionally to the prosumers based on how far their energy-sharing price from the DSO’s NEM price, which reflects how much they paid (got paid) for voltage correction. Second, it makes prosumer payment functions Cnχn𝒩C_{n}^{\chi^{\ast}}\forall n\in{\cal N} uniform. Indeed, plugging (20) into (12) cancels out the locational dynamic price πi(𝒈)\pi^{\ast}_{i}(\bm{g}), and yields a simple, uniform payment function that charges customers based on the NEM price, i.e., for every bus ii\in{\cal B},

Cnχ(zn)=πNEM(Z0)zn,n𝒩i.C_{n}^{\chi^{\ast}}(z_{n})=\pi^{\mbox{\tiny NEM}}(Z_{0})\cdot z_{n},\forall n\in{\cal N}_{i}.

Third, unlike the computationally expensive coalitional-game-based profit allocation schemes such as the Shapley value, the allocation rule in (20) is straightforward and directly correlates the allocation to the energy-sharing price and the prosumer’s net consumption. The decentralization argument may not hold under the allocation in (20), as it compensates prosumers explicitly based on their own net consumption, which may influence their consumption decisions resulting in deviations from the welfare-maximizing decisions. It might be, however, too difficult for prosumers to anticipate if the operator performs the re-allocation at every multiple netting periods rather than at every single netting period.

V Numerical Study

Refer to caption
Figure 2: The IEEE 13-bus test feeder.
Refer to caption
Figure 3: Summary of the numerical tests on the four considered scenarios. The lower panel reports the ex-ante energy prices obtained after solving the energy sharing platform optimization problem (11). The upper panel shows the cumulative power demand at each bus obtained after the energy sharing operator dispatched the energy prices. The middle panel reports the resulting bus voltage magnitudes.

Our network-aware market mechanism was validated on the IEEE 13-bus feeder converted to a single-phase equivalent [21], see Figure 2. Bus 1 is the substation and represents the network slack bus. Buses 2 to 13 instead host 23 prosumers. For every n𝒩n\in{\cal N}, the following utility function was chosen

Un(dn)={αndn12βndn2,0dnαnβnαn22βn,dn>αnβn,U_{n}(d_{n})=\left\{\begin{array}[]{ll}\alpha_{n}d_{n}-\frac{1}{2}\beta_{n}d_{n}^{2},&0\leq d_{n}\leq\frac{\alpha_{n}}{\beta_{n}}\\ \frac{\alpha_{n}^{2}}{2\beta_{n}},&d_{n}>\frac{\alpha_{n}}{\beta_{n}},\end{array}\right. (21)

where the parameters αn,βn\alpha_{n},\beta_{n} were learned and calibrated using historical retail prices333The retail prices were taken from Data.AustinTexas.gov historical residential rates in Austin, TX. and consumptions444For the historical consumption data, we used pre-2018 PecanStreet data for households in Austin, TX. and by assuming an elasticity of 0.21 taken from [22] (see appendix D in [16]). The minimum demand was set to 𝒅¯n=𝟎{\underline{\bm{d}}_{n}}=\bm{0} for every n𝒩n\in{\cal N}, whereas the maximum demands 𝒅¯n{\overline{\bm{d}}_{n}} and DER generations were obtained using data from the PecanStreet dataset. We set 𝒗min=0.95\bm{v}_{\text{min}}=0.95 p.u. and 𝒗max=1.05\bm{v}_{\text{max}}=1.05 p.u.

In our simulations, we considered four scenarios, described in the following, that differ in the DER generation levels. For each scenario, we used the exact AC power flow solver to obtain bus voltages, whereas we solved the optimization problems relying on the power flow equation linearization (4). The results are shown in Figure 3.

Scenario 1: the DER generation here is zero for each prosumer. Hence, G0=0G_{0}=0 and G0<σ1(𝒈)G_{0}<\sigma_{1}(\bm{g}). The energy-sharing system is importing energy. In this case, the energy sharing optimization problem solutions are such that η¯i0\underline{\eta}^{\ast}_{i}\neq 0, i.e., some voltages are on the lower bound 𝒗min\bm{v}_{\text{min}}. The resulting prices are in general higher than π+\pi^{+}.

Scenario 2: the DER generation G0G_{0} is non-zero but still not enough to cover the demand, i.e., G0<σ1(𝐠)G_{0}<\sigma_{1}(\bm{g}). Hence, the energy-sharing system is importing energy. However, the optimum demands are such that all the voltages are within the desired bounds and the energy prices are equal π+\pi^{+}.

Scenario 3: the DER generation was further increased in this scenario and σ1(𝒈)G0σ2(𝒈)\sigma_{1}(\bm{g})\leq G_{0}\leq\sigma_{2}(\bm{g}). That is, the energy-sharing platform did not exchange active power with the external network. The energy sharing platform optimization problem provides an energy price within π+\pi^{+} and π\pi^{-}; voltage limits are satisfied at the optimal consumption.

Scenario 4: here, we increased the generation until G0σ2(𝒈)G_{0}\geq\sigma_{2}(\bm{g}). The platform exports power to the grid. The energy sharing optimization problem solution is such that the voltages in some locations are exactly 𝒗max\bm{v}_{\text{max}} and the Lagrange multipliers vector 𝜼¯\overline{\bm{\eta}}^{\ast} is different from zero. The energy prices are smaller than π\pi^{-} and close to zero, i.e., consumption is incentivized to take full advantage of generation.

Some observations are in order. In general, we observe that increasing the DER generation G0G_{0} results in the decrease of energy prices. The energy prices can in principle be greater than π+\pi^{+}, see Scenario 1. This is to ensure that the voltage constraints are satisfied by decreasing the power demand. Finally, we note a slight difference between the true and the expected (i.e., the ones computed by the energy-sharing platform optimization problem) voltage magnitudes. Indeed, we see that the voltages in Scenario 4 are all strictly lower than 𝒗max\bm{v}_{\text{max}} even though we obtained η¯i0\overline{\eta}^{\ast}_{i}\neq 0, see the middle panel of Figure 3. This can be explained by the fact that (11) was solved relying on the linearized equations (4) rather than on the true power flow equations. Note, however, that using the true equation would result in a nonconvex energy sharing optimization problem possibly displaying multiple local minima.

VI Conclusion

In this work, we propose a network-aware and welfare-maximizing market mechanism for energy-sharing coalitions that aggregate small but ubiquitous BTM DER downstream of a DSO’s revenue meter, charging the energy-sharing systems using a generic NEM tariff. The proposed pricing policy has ex-ante and ex-post pricing components. The ex-ante locational and threshold-based price decreases as the energy-sharing generation-to-demand ratio increases. The price is used to induce a collective prosumer reaction that decentrally maximizes social welfare while being network-cognizant. On the other hand, the ex-post charge/reward is used to enforce the market operator’s profit-neutrality condition. We show that the market mechanism achieves an equilibrium to the Stackelberg game between the operator and its prosumers. We also show that a unique proportional rule to re-allocate the operator’s profit/deficit can make the payment function of all energy-sharing prosumers uniform, even when the network constraints are binding. Our simulation results leverage real DER data on an IEEE 13-bus test feeder system to show how the dynamic pricing drives the energy sharing’s flexible consumption to abide by the network voltage limits.

VII Acknowledgment

The work of Ahmed S. Alahmed and Lang Tong was supported in part by the National Science Foundation under Award 2218110 and the Power Systems Engineering Research Center (PSERC) under Research Project M-46. The work was authored in part 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 is provided by the U.S. DOE Office of Energy Efficiency and Renewable Energy Building Technologies Office, United States. 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.

References

  • [1] P. De Martini, L. Gallagher, E. Takayesu, R. Hanley, and P. Henneaux, “Unlocking consumer der potential: Consumer-centric approaches for grid services,” IEEE Power and Energy Magazine, vol. 20, no. 4, pp. 76–84, 2022.
  • [2] L. Han, T. Morstyn, and M. McCulloch, “Incentivizing prosumer coalitions with energy management using cooperative game theory,” IEEE Trans. Power Systems, vol. 34, no. 1, 2019.
  • [3] A. S. Alahmed and L. Tong, “Dynamic net metering for energy communities,” IEEE Transactions on Energy Markets, Policy and Regulation, pp. 1–12, 2024.
  • [4] Y. Yang, G. Hu, and C. J. Spanos, “Optimal sharing and fair cost allocation of community energy storage,” IEEE Trans. Smart Grid, vol. 12, no. 5, Sep. 2021.
  • [5] Y. Chen, S. Mei, F. Zhou, S. H. Low, W. Wei, and F. Liu, “An Energy Sharing Game with Generalized Demand Bidding: Model and Properties,” IEEE Transactions on Smart Grid, vol. 11, no. 3, pp. 2055–2066, 2020.
  • [6] N. Vespermann, T. Hamacher, and J. Kazempour, “Access economy for storage in energy communities,” IEEE Transactions on Power Systems, vol. 36, no. 3, pp. 2234–2250, 2021.
  • [7] A. Fleischhacker, C. Corinaldesi, G. Lettner, H. Auer, and A. Botterud, “Stabilizing energy communities through energy pricing or PV expansion,” IEEE Transactions on Smart Grid, vol. 13, no. 1, pp. 728–737, 2022.
  • [8] M. I. Azim, G. Lankeshwara, W. Tushar, R. Sharma, M. R. Alam, T. K. Saha, M. Khorasany, and R. Razzaghi, “Dynamic operating envelope-enabled P2P trading to maximize financial returns of prosumers,” IEEE Transactions on Smart Grid, vol. 15, no. 2, pp. 1978–1990, 2024.
  • [9] A. S. Alahmed, G. Cavraro, A. Bernstein, and L. Tong, “Operating-envelopes-aware decentralized welfare maximization for energy communities,” in 2023 59th Annual Allerton Conference on Communication, Control, and Computing (Allerton), 2023, pp. 1–8.
  • [10] Y. Chen, C. Zhao, S. H. Low, and A. Wierman, “An energy sharing mechanism considering network constraints and market power limitation,” IEEE Transactions on Smart Grid, vol. 14, no. 2, pp. 1027–1041, 2023.
  • [11] C. P. Mediwaththe and L. Blackhall, “Network-Aware Demand-Side Management Framework with A Community Energy Storage System Considering Voltage Constraints,” IEEE Transactions on Power Systems, vol. 36, no. 2, pp. 1229–1238, 2021.
  • [12] N. Li, “A market mechanism for electric distribution networks,” Proceedings of the IEEE Conference on Decision and Control, vol. 54rd IEEE, no. IEEE CDC, pp. 2276–2282, 2015.
  • [13] L. Bai, J. Wang, C. Wang, C. Chen, and F. Li, “Distribution locational marginal pricing (DLMP) for congestion management and voltage support,” IEEE Transactions on Power Systems, vol. 33, no. 4, pp. 4061–4073, 2018.
  • [14] A. Papavasiliou, “Analysis of distribution locational marginal prices,” IEEE Transactions on Smart Grid, vol. 9, no. 5, pp. 4872–4882, 2018.
  • [15] J. Bonbright, A. Danielsen, D. Kamerschen, and J. Legler, Principles of Public Utility Rates.   Public Utilities Reports, 1988.
  • [16] A. S. Alahmed and L. Tong, “Integrating distributed energy resources: Optimal prosumer decisions and impacts of net metering tariffs,” SIGENERGY Energy Inform. Rev., vol. 2, no. 2, Aug. 2022.
  • [17] M. Baran and F. Wu, “Optimal sizing of capacitors placed on a radial distribution system,” IEEE Transactions on Power Delivery, vol. 4, no. 1, pp. 735–743, 1989.
  • [18] ——, “Network reconfiguration in distribution systems for loss reduction and load balancing,” IEEE Transactions on Power Delivery, vol. 4, no. 2, pp. 1401–1407, 1989.
  • [19] A. S. Alahmed and L. Tong, “On net energy metering X: Optimal prosumer decisions, social welfare, and cross-subsidies,” IEEE Trans. Smart Grid, vol. 14, no. 02, 2023.
  • [20] F. Schweppe, M. Caramanis, R. Tabors, and R. Bohn, Spot Pricing of Electricity.   Springer US, 1988.
  • [21] G. Cavraro and V. Kekatos, “Inverter probing for power distribution network topology processing,” IEEE Transactions on Control of Network Systems, vol. 6, no. 3, pp. 980–992, 2019.
  • [22] A. Asadinejad, A. Rahimpour, K. Tomsovic, H. Qi, and C. fei Chen, “Evaluation of residential customer elasticity for incentive based demand response programs,” Electric Power Systems Research, 2018.

[Incorporating Operating Envelopes] Here, we present the pricing policy under OEs at the prosumer’s revenue meter, as shown in Fig.1.

OEs limit the net consumption of every prosumer n𝒩n\in{\cal N}, as

zn𝒵n:=[z¯n,z¯n],z_{n}\in\mathcal{Z}_{n}:=[\underline{z}_{n},\overline{z}_{n}], (22)

where z¯n0\underline{z}_{n}\leq 0 and z¯n0\overline{z}_{n}\geq 0 are the export and import envelopes at the prosumers’ meters, respectively. From the analysis in [9], the network-aware pricing policy generalizes as in the following policy.

Network-aware pricing policy 2.

For every bus ii\in{\cal B}, the pricing policy charges the prosumers based on a two-part pricing

χ:𝒈Cnχ(zn)=πi(𝒈)ex-ante priceznAnex-post allocation,n𝒩i,\chi^{\ast}:\bm{g}\mapsto C^{\chi^{\ast}}_{n}(z_{n})=\underbrace{\pi^{\ast}_{i}(\bm{g})}_{\text{ex-ante price}}\cdot z_{n}-\underbrace{A_{n}^{\ast}}_{\text{ex-post allocation}},\forall n\in{\cal N}_{i},

where the ex-ante bus price πi(𝐠)\pi^{\ast}_{i}(\bm{g}) abides by a two-threshold policy with thresholds

σ1(𝒈)\displaystyle\sigma_{1}(\bm{g}) =in𝒩i[𝟏[𝒇n(𝟏χi+)]𝒅¯n𝒅¯n]z¯n+gnz¯n+gn,\displaystyle=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}\left[\bm{1}^{\top}{[\bm{f}_{n}(\bm{1}\chi_{i}^{+})]}_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}}\right]^{\overline{z}_{n}+g_{n}}_{\underline{z}_{n}+g_{n}},
σ2(𝒈)\displaystyle\sigma_{2}(\bm{g}) =in𝒩i[𝟏[𝒇n(𝟏χi)]𝒅¯n𝒅¯n]z¯n+gnz¯n+gnσ1(𝒈),\displaystyle=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}\left[\bm{1}^{\top}{[\bm{f}_{n}(\bm{1}\chi_{i}^{-})]}_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}}\right]^{\overline{z}_{n}+g_{n}}_{\underline{z}_{n}+g_{n}}\geq\sigma_{1}(\bm{g}),

as

πi(𝒈)={χi+(𝒈),G0<σ1(𝒈)χiz(𝒈),G0[σ1(𝒈),σ2(𝒈)]χi(𝒈),G0>σ2(𝒈),\pi^{\ast}_{i}(\bm{g})=\begin{cases}\chi^{+}_{i}(\bm{g})&,G_{0}<\sigma_{1}(\bm{g})\\ \chi^{z}_{i}(\bm{g})&,G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})]\\ \chi^{-}_{i}(\bm{g})&,G_{0}>\sigma_{2}(\bm{g}),\end{cases}\quad (23)

and the price χiκ\chi_{i}^{\kappa}, where κ:={+,,z}\kappa:=\{+,-,z\}, is given by

χiκ=πκjRji(η¯jη¯j)\chi_{i}^{\kappa}=\pi^{\kappa}-\sum_{j\in{\cal B}}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j}) (24)

where η¯j\overline{\eta}^{\ast}_{j} and η¯j\underline{\eta}^{\ast}_{j} are the dual variables of the upper and lower voltage limits in (11d), respectively, and the price πz:=μ\pi^{z}:=\mu^{\ast} is the solution of

in𝒩i[𝟏[𝒇n(𝟏μ𝟏jRji(η¯jη¯j))]𝒅¯n𝒅¯n]z¯n+gnz¯n+gn=G0.\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}\left[\bm{1}^{\top}[\bm{f}_{n}(\bm{1}\mu-\bm{1}\sum_{j\in{\cal B}}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j}))]_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}}\right]^{\overline{z}_{n}+g_{n}}_{\underline{z}_{n}+g_{n}}=G_{0}.

For every bus ii\in{\cal B}, the prosumer’s ex-post charge/reward is denoted by AnA_{n}^{\ast}, which we delineate in §\SIV-C and §\SIV-E.

[Mathematical Proofs]

-A Lemma 2 and Proof of Lemma 2

Lemma 2 (Maximum welfare under centralized operation).

The maximum welfare under centralized operation, that solves

W(𝒈):=maximize(𝒅1,,𝒅N)\displaystyle W^{\ast}(\bm{g}):=\underset{(\bm{d}_{1},\ldots,\bm{d}_{N})}{\rm maximize} in𝒩iUn(𝒅n)CNEM(Z0)\displaystyle~{}~{}\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}U_{n}(\bm{d}_{n})-C^{\mbox{\tiny NEM}}(Z_{0})
subject to Z0=n𝒩(𝟏𝒅ngn)\displaystyle~{}~{}Z_{0}=\sum_{n\in{\cal N}}\big{(}\bm{1}^{\top}\bm{d}_{n}-g_{n}\big{)} (25)
𝒅n𝒟nn𝒩\displaystyle\hskip 11.38092pt\bm{d}_{n}\in{\cal D}_{n}~{}\forall n\in{\cal N}
(𝜼¯,𝜼¯):\displaystyle(\underline{\bm{\eta}},\overline{\bm{\eta}}): 𝒗¯𝑹𝒁𝒗¯.\displaystyle\hskip 11.38092pt\underline{\bm{v}}\preceq-\bm{R}\bm{Z}\preceq\overline{\bm{v}}.

obeys by two-thresholds as

W(𝒈)\displaystyle W^{\ast}(\bm{g}) =in𝒩i\displaystyle=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}
{Un(𝒅n+(𝒈))π+(𝟏𝒅n+(𝒈)gn),G0<σ1(𝒈)Un(𝒅nz(𝒈)),G0[σ1(𝒈),σ2(𝒈)]Un(𝒅n(𝒈))π(𝟏𝒅n(𝒈)gn),G0>σ2(𝒈),\displaystyle\begin{cases}U_{n}(\bm{d}^{+}_{n}(\bm{g}))-\pi^{+}(\bm{1}^{\top}\bm{d}^{+}_{n}(\bm{g})-g_{n})&,G_{0}<\sigma_{1}(\bm{g})\\ U_{n}(\bm{d}^{z}_{n}(\bm{g}))&,G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})]\\ U_{n}(\bm{d}^{-}_{n}(\bm{g}))-\pi^{-}(\bm{1}^{\top}\bm{d}^{-}_{n}(\bm{g})-g_{n})&,G_{0}>\sigma_{2}(\bm{g}),\end{cases}

where the thresholds (σ1(𝐠),σ2(𝐠)\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})) are as in (13), the consumption 𝐝nκ(𝐠),n𝒩\bm{d}^{\kappa}_{n}(\bm{g}),\forall n\in{\cal N} and κ:={+,,z}\kappa:=\{+,-,z\}, is

𝒅nκ(𝒈):=[𝒇n(𝟏πκ𝟏jRji(η¯jη¯j))]𝒅¯n𝒅¯n,n𝒩i,\displaystyle\bm{d}^{\kappa}_{n}(\bm{g}):=[\bm{f}_{n}(\bm{1}\pi^{\kappa}-\bm{1}\sum_{j\in{\cal B}}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j}))]_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}},\forall n\in{\cal N}_{i},

and η¯i,η¯i0,i\overline{\eta}^{\ast}_{i},\underline{\eta}^{\ast}_{i}\geq 0,\forall i\in{\cal B} is from the KKT conditions in (27)-(28).

Proof of Lemma 2

The convex non-differentiable program in (25) is a generalization to the standalone consumer decision problem under the DSO’s NEM X regime in [19] with (a) the additional dimension of NN users located at BB buses and (b) network constraints. Therefore, (25) can be divided into three convex and differentiable programs based on the energy-sharing net consumption Z0Z_{0}, namely (𝒫Z00,𝒫Z0=0,𝒫Z00{\cal P}_{Z_{0}\geq 0},{\cal P}_{Z_{0}=0},{\cal P}_{Z_{0}\leq 0}), as

𝒫Z00:maximize(𝒅1,,𝒅N)\displaystyle{\cal P}_{Z_{0}\geq 0}:\underset{(\bm{d}_{1},\ldots,\bm{d}_{N})}{\rm maximize} in𝒩iUn(𝒅n)π+Z0\displaystyle~{}~{}\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}U_{n}(\bm{d}_{n})-\pi^{+}\cdot Z_{0}
subject to Z00\displaystyle~{}~{}Z_{0}\geq 0
𝒅n𝒟nn𝒩\displaystyle\hskip 11.38092pt\bm{d}_{n}\in{\cal D}_{n}~{}\forall n\in{\cal N}
(𝜼¯+,𝜼¯+):\displaystyle(\underline{\bm{\eta}}^{+},\overline{\bm{\eta}}^{+}): 𝒗¯𝑹𝒁𝒗¯.\displaystyle\hskip 11.38092pt\underline{\bm{v}}\preceq-\bm{R}\bm{Z}\preceq\overline{\bm{v}}.
𝒫Z0=0:maximize(𝒅1,,𝒅N)\displaystyle{\cal P}_{Z_{0}=0}:\underset{(\bm{d}_{1},\ldots,\bm{d}_{N})}{\rm maximize} in𝒩iUn(𝒅n)\displaystyle~{}~{}\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}U_{n}(\bm{d}_{n})\hskip 43.53267pt
subject to Z0=0\displaystyle~{}~{}Z_{0}=0
𝒅n𝒟nn𝒩\displaystyle\hskip 11.38092pt\bm{d}_{n}\in{\cal D}_{n}~{}\forall n\in{\cal N}
(𝜼¯,𝜼¯):\displaystyle(\underline{\bm{\eta}}^{-},\overline{\bm{\eta}}^{-}): 𝒗¯𝑹𝒁𝒗¯.\displaystyle\hskip 11.38092pt\underline{\bm{v}}\preceq-\bm{R}\bm{Z}\preceq\overline{\bm{v}}.
𝒫Z00:maximize(𝒅1,,𝒅N)\displaystyle{\cal P}_{Z_{0}\leq 0}:\underset{(\bm{d}_{1},\ldots,\bm{d}_{N})}{\rm maximize} in𝒩iUn(𝒅n)πZ0\displaystyle~{}~{}\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}U_{n}(\bm{d}_{n})-\pi^{-}\cdot Z_{0}
subject to Z00\displaystyle~{}~{}Z_{0}\leq 0
𝒅n𝒟nn𝒩\displaystyle\hskip 11.38092pt\bm{d}_{n}\in{\cal D}_{n}~{}\forall n\in{\cal N}
(𝜼¯0,𝜼¯0):\displaystyle(\underline{\bm{\eta}}^{0},\overline{\bm{\eta}}^{0}): 𝒗¯𝑹𝒁𝒗¯.\displaystyle\hskip 11.38092pt\underline{\bm{v}}\preceq-\bm{R}\bm{Z}\preceq\overline{\bm{v}}.

It has been shown in [19] that the optimal consumption policy is a two-thresholds policy on the aggregate renewables G0G_{0}. The three problems above are generalizations of [19] that incorporate multiple prosumers and buses dimension and network constraints dimension, which therefore yields, for every bus ii\in{\cal B}, the following optimal consumption vector

𝒅n(𝒈)={𝒅n+(𝒈),G0<σ1(𝒈)𝒅nz(𝒈),G0[σ1(𝒈),σ2(𝒈)]𝒅n(𝒈),G0>σ2(𝒈),,n𝒩i,\bm{d}^{\ast}_{n}(\bm{g})=\begin{cases}\bm{d}^{+}_{n}(\bm{g})&,G_{0}<\sigma_{1}(\bm{g})\\ \bm{d}^{z}_{n}(\bm{g})&,G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})]\\ \bm{d}^{-}_{n}(\bm{g})&,G_{0}>\sigma_{2}(\bm{g}),\end{cases},\quad\forall n\in{\cal N}_{i},

where the thresholds (σ1(𝒈),σ2(𝒈)\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})) are as in (13) and the vector 𝒅nκ(𝒈)\bm{d}^{\kappa}_{n}(\bm{g}), where κ:={+,,z}\kappa:=\{+,-,z\}, is given by

𝒅nκ(𝒈):=[𝒇n(𝟏πκ𝟏jRji(η¯jη¯j))]𝒅¯n𝒅¯n,n𝒩i.\displaystyle\bm{d}^{\kappa}_{n}(\bm{g}):=[\bm{f}_{n}(\bm{1}\pi^{\kappa}-\bm{1}\sum_{j\in{\cal B}}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j}))]_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}},\forall n\in{\cal N}_{i}. (26)

By generalizing the special case in [19] through incorporating the additional dimension of NN users located at BB buses and the network constraints, one can see that the price πz(𝒈)\pi^{z}(\bm{g}) is as in (16). Lastly, for every ii\in{\cal B}, the dual variable η¯i0\overline{\eta}^{\ast}_{i}\geq 0 is computed from the KKT conditions

η¯i(jRijk𝒩j(𝟏𝒅nκ(𝒈)gn))\displaystyle\overline{\eta}^{\ast}_{i}\left(-\sum_{j\in{\cal B}}R_{ij}\sum_{k\in{\cal N}_{j}}(\bm{1}^{\top}\bm{d}^{\kappa}_{n}(\bm{g})-g_{n})\right) =0\displaystyle=0 (27a)
jRijk𝒩j(𝟏𝒅nκ(𝒈)gn)v¯i\displaystyle-\sum_{j\in{\cal B}}R_{ij}\sum_{k\in{\cal N}_{j}}(\bm{1}^{\top}\bm{d}^{\kappa}_{n}(\bm{g})-g_{n})-\overline{v}_{i} 0\displaystyle\leq 0 (27b)

and η¯i0\underline{\eta}^{\ast}_{i}\geq 0 is similarly computed from

η¯i(jRijk𝒩j(𝟏𝒅nκ(𝒈)gn))\displaystyle\underline{\eta}^{\ast}_{i}\left(-\sum_{j\in{\cal B}}R_{ij}\sum_{k\in{\cal N}_{j}}(\bm{1}^{\top}\bm{d}^{\kappa}_{n}(\bm{g})-g_{n})\right) =0\displaystyle=0 (28a)
jRijk𝒩j(𝟏𝒅nκ(𝒈)gn)v¯i\displaystyle-\sum_{j\in{\cal B}}R_{ij}\sum_{k\in{\cal N}_{j}}(\bm{1}^{\top}\bm{d}^{\kappa}_{n}(\bm{g})-g_{n})-\underline{v}_{i} 0\displaystyle\geq 0 (28b)

\blacksquare

-B Proof of Theorem 1

Recall the bi-level program of operator and prosumers decisions

maximize𝑪χ()\displaystyle\underset{\bm{C}^{\chi}(\cdot)}{\operatorname{maximize}} (Wχ𝝍=n𝒩Un(𝒅nψχ)CNEM(Z0ψχ))\displaystyle\Bigg{(}W^{\chi_{\bm{\psi}}}=\sum_{n\in{\cal N}}U_{n}(\bm{d}^{\psi^{\sharp}_{\chi}}_{n})-C^{\mbox{\tiny NEM}}(Z_{0}^{\psi^{\sharp}_{\chi}})\Bigg{)} (29a)
subject to n𝒩Cnχ(znψχ)=CNEM(Z0ψχ)\displaystyle\quad\quad\sum_{n\in{\cal N}}C^{\chi}_{n}(z_{n}^{\psi^{\sharp}_{\chi}})=C^{\mbox{\tiny NEM}}(Z_{0}^{\psi^{\sharp}_{\chi}}) (29b)
Z0ψχ=n𝒩(𝟏𝒅nψχgn)\displaystyle\quad\quad Z_{0}^{\psi^{\sharp}_{\chi}}=\sum_{n\in{\cal N}}\big{(}\bm{1}^{\top}\bm{d}^{\psi^{\sharp}_{\chi}}_{n}-g_{n}\big{)} (29c)
(𝜼¯,𝜼¯)\displaystyle(\underline{\bm{\eta}},\overline{\bm{\eta}}) 𝒗¯𝑹𝒁ψχ𝒗¯\displaystyle\quad\quad\underline{\bm{v}}\preceq-\bm{R}\bm{Z}^{\psi^{\sharp}_{\chi}}\preceq\overline{\bm{v}} (29d)
for all i=1,,B,n𝒩i\displaystyle\quad\quad\text{for all }i=1,\ldots,B,n\in{\cal N}_{i} (29e)
𝒅nψχ:=argmax𝒅n𝒟nSnχ(𝒅n,gn):=Un(𝒅n)Cnχ(zn)\displaystyle\bm{d}^{\psi^{\sharp}_{\chi}}_{n}:=\underset{\bm{d}_{n}\in{\cal D}_{n}}{\operatorname{argmax}}~{}S^{\chi}_{n}(\bm{d}_{n},g_{n}):=U_{n}(\bm{d}_{n})-C^{\chi}_{n}(z_{n}) (29f)
subject tozn=𝟏𝒅ngn.\displaystyle\hskip 36.98866pt\text{subject to}~{}~{}z_{n}=\bm{1}^{\top}\bm{d}_{n}-g_{n}. (29g)

We solve a relaxed version of (29) that does not require profit-neutrality, hence (29b) is removed, and for every n𝒩n\in{\cal N}, we want to find the price πn\pi_{n} in C~nχ()=πnzn\tilde{C}_{n}^{\chi}(\cdot)=\pi_{n}\cdot z_{n}. So, (29) is reformulated to

maximize𝑪~χ()\displaystyle\underset{\tilde{\bm{C}}^{\chi}(\cdot)}{\operatorname{maximize}} (Wχ𝝍=n𝒩Un(𝒅nψχ)CNEM(Z0ψχ))\displaystyle\Bigg{(}W^{\chi_{\bm{\psi}}}=\sum_{n\in{\cal N}}U_{n}(\bm{d}^{\psi^{\sharp}_{\chi}}_{n})-C^{\mbox{\tiny NEM}}(Z_{0}^{\psi^{\sharp}_{\chi}})\Bigg{)}
subject to Z0ψχ=n𝒩(𝟏𝒅nψχgn)\displaystyle\quad\quad Z_{0}^{\psi^{\sharp}_{\chi}}=\sum_{n\in{\cal N}}\big{(}\bm{1}^{\top}\bm{d}^{\psi^{\sharp}_{\chi}}_{n}-g_{n}\big{)}
(𝜼¯,𝜼¯)\displaystyle(\underline{\bm{\eta}},\overline{\bm{\eta}}) 𝒗¯𝑹𝒁ψχ𝒗¯\displaystyle\quad\quad\underline{\bm{v}}\preceq-\bm{R}\bm{Z}^{\psi^{\sharp}_{\chi}}\preceq\overline{\bm{v}}
for all i=1,,B,n𝒩i\displaystyle\quad\quad\text{for all }i=1,\ldots,B,n\in{\cal N}_{i}
Lkn(dknψχ)π+λ¯knλ¯kn=0,k𝒦\displaystyle~{}~{}~{}L_{kn}(d_{kn}^{\psi_{\chi}^{\sharp}})-\pi+\overline{\lambda}^{\ast}_{kn}-\underline{\lambda}^{\ast}_{kn}=0,~{}~{}\forall k\in{\cal K}
d¯kndknψχ0λ¯kn0,k𝒦\displaystyle\hskip 45.52458pt\overline{d}_{kn}-d_{kn}^{\psi_{\chi}^{\sharp}}\geq 0\perp\overline{\lambda}^{\ast}_{kn}\geq 0,~{}~{}\forall k\in{\cal K}
dknψχ0λ¯kn0,k𝒦\displaystyle\hskip 66.86414ptd_{kn}^{\psi_{\chi}^{\sharp}}\geq 0\perp\underline{\lambda}^{\ast}_{kn}\geq 0,~{}~{}\forall k\in{\cal K}

where xyx\perp y means that xx and yy are perpendicular. From Lemma 1, the lower level’s KKT conditions can be replaced as the following

maximize𝑪~χ()\displaystyle\underset{\tilde{\bm{C}}^{\chi}(\cdot)}{\operatorname{maximize}} (Wχ𝝍=n𝒩Un(𝒅nψχ)CNEM(Z0ψχ))\displaystyle\Bigg{(}W^{\chi_{\bm{\psi}}}=\sum_{n\in{\cal N}}U_{n}(\bm{d}^{\psi^{\sharp}_{\chi}}_{n})-C^{\mbox{\tiny NEM}}(Z_{0}^{\psi^{\sharp}_{\chi}})\Bigg{)}
subject to Z0ψχ=n𝒩(𝟏𝒅nψχgn)\displaystyle\quad\quad Z_{0}^{\psi^{\sharp}_{\chi}}=\sum_{n\in{\cal N}}\big{(}\bm{1}^{\top}\bm{d}^{\psi^{\sharp}_{\chi}}_{n}-g_{n}\big{)}
(𝜼¯,𝜼¯)\displaystyle(\underline{\bm{\eta}},\overline{\bm{\eta}}) 𝒗¯𝑹𝒁ψχ𝒗¯\displaystyle\quad\quad\underline{\bm{v}}\preceq-\bm{R}\bm{Z}^{\psi^{\sharp}_{\chi}}\preceq\overline{\bm{v}}
for all i=1,,B,n𝒩i\displaystyle\quad\quad\text{for all }i=1,\ldots,B,n\in{\cal N}_{i}
𝒅nψχ(πn)=[𝒇n(𝟏πn)]𝒅¯n𝒅¯n.\displaystyle~{}~{}~{}\bm{d}^{\psi^{\sharp}_{\chi}}_{n}(\pi_{n})=[\bm{f}_{n}(\bm{1}\pi_{n})]_{\underline{\bm{d}}_{n}}^{\overline{\bm{d}}_{n}}.

Note that if πn\pi^{\ast}_{n} is found, we have equilibrium. The program above is similar to the central program (25) in Lemma 2, but with the prices πn,n𝒩\pi_{n},\forall n\in{\cal N} as decision variables rather than the consumptions 𝒅n,n𝒩\bm{d}_{n},\forall n\in{\cal N}. Therefore, the threshold structure in Lemma 2 holds with the prices used to compute 𝒅n,n𝒩\bm{d}^{\ast}_{n},\forall n\in{\cal N} being the optimal prices. Therefore, we have, for every bus ii\in{\cal B},

πi(𝒈)={π+jRji(η¯jη¯j),G0<σ1(𝒈)πz(𝒈)jRji(η¯jη¯j),G0[σ1(𝒈),σ2(𝒈)]πjRji(η¯jη¯j),G0>σ2(𝒈),\pi^{\ast}_{i}(\bm{g})=\begin{cases}\pi^{+}-\sum_{j\in{\cal B}}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j})&,G_{0}<\sigma_{1}(\bm{g})\\ \pi^{z}(\bm{g})-\sum_{j\in{\cal B}}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j})&,G_{0}\in[\sigma_{1}(\bm{g}),\sigma_{2}(\bm{g})]\\ \pi^{-}-\sum_{j\in{\cal B}}R_{ji}(\overline{\eta}^{\ast}_{j}-\underline{\eta}^{\ast}_{j})&,G_{0}>\sigma_{2}(\bm{g}),\end{cases}\quad

One can easily see from Lemma 1, that by plugging the equilibrium price πi(𝒈),i\pi^{\ast}_{i}(\bm{g}),\forall i\in{\cal B}, the prosumers’ optimal consumption decisions match the consumption decisions under centralized operation in Lemma 2.

Now, we note that for every bus ii\in{\cal B}, the pricing policy C~nχ=πi(𝒈)zn,n𝒩i\tilde{C}^{\chi^{\ast}}_{n}=\pi^{\ast}_{i}(\bm{g})\cdot z_{n},\forall n\in{\cal N}_{i} does not achieve profit neutrality. Therefore, for every bus ii\in{\cal B}, the pricing policy is augmented by the ex-post allocation AnA^{\ast}_{n} to become Cnχ=πi(𝒈)znAn,n𝒩iC^{\chi^{\ast}}_{n}=\pi^{\ast}_{i}(\bm{g})\cdot z_{n}-A^{\ast}_{n},\forall n\in{\cal N}_{i} with in𝒩iAn=in𝒩i(πi(𝒈)πNEM(Z0))zn\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}A^{\ast}_{n}=\sum_{i\in{\cal B}}\sum_{n\in{\cal N}_{i}}(\pi^{\ast}_{i}(\bm{g})-\pi^{\mbox{\tiny NEM}}(Z_{0}))\cdot z_{n}. One can then apply Def.1 to show that 𝑪χ\bm{C}^{\chi^{\ast}} achieves profit neutrality. \blacksquare