Skip to main content
Log in

Retail Equilibrium with Switching Consumers in Electricity Markets

  • Published:
Networks and Spatial Economics Aims and scope Submit manuscript

Abstract

The ongoing transformations of power systems worldwide pose important challenges, both economic and technical, for their appropriate planning and operation. A key approach to improve the efficiency of these systems is through demand-side management, i.e., to promote the active involvement of consumers in the system. In particular, the current trend is to conceive systems where electricity consumers can vary their load according to real-time price incentives, offered by retailing companies. Under this setting, retail competition plays an important role as inadequate prices or services may entail consumers switching to a rival retailer. In this work we consider a game theoretical model where asymmetric retailers compete in prices to increase their profits by accounting for the utility function of consumers. Consumer preferences for retailers are uncertain and distributed within a Hotelling line. We analytically characterize the equilibrium of a retailer duopoly, establishing its existence and uniqueness conditions for a wide class of utility functions. Furthermore, sensitivities of the equilibrium prices with respect to relevant model parameters are also provided. The duopoly model is extended to a network that includes multiple retailers with capacity constraints for which we perform static and dynamic numerical simulations. Results indicate that, depending on the retailer costs, loyalty rewards and initial market shares, the resulting equilibrium can range from complete competition to one in which a retailer has a leading or even a dominant position in the market, decreasing the consumers’ utility significantly. Moreover, the retailer network configuration also plays an important role in the competitiveness of the system.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  • Acquisti A, Varian H R (2005) Conditioning prices on purchase history. Mark Sci 24(3):367–381

    Article  Google Scholar 

  • Armstrong M (1998) Network interconnection in telecommunications. Econ J 108(448):545–564

    Article  Google Scholar 

  • Bu S, Yu F R (2013) A game-theoretical scheme in the smart grid with demand-side management: towards a smart cyber-physical power infrastructure. IEEE Trans Emerging Topics Comput 1(1):22–32

    Article  Google Scholar 

  • Csercsik D (2016) Competition and cooperation in a bidding model of electrical energy trade. Netw Spatial Econ 2016:1–31

    Google Scholar 

  • Chen Y, Bhardwaj P, Balasubramanian S (2014) The strategic implications of switching costs under customized pricing. Custom Needs Solut 1(3):188–199

    Article  Google Scholar 

  • Chen Y, Riordan M H (2007) Price and variety in the spokes model. Econ J 117(522):897–921

    Article  Google Scholar 

  • Di Giorgio A, Liberati F (2014) Near real time load shifting control for residential electricity prosumers under designed and market indexed pricing models. Appl Energy 128:119–132

    Article  Google Scholar 

  • Doganoglu T (2010) Switching costs, experience goods and dynamic price competition. QME 8(2):167–205

    Google Scholar 

  • Farrell J, Klemperer P (2007) Coordination and lock-in: competition with switching costs and network effects. In: Armstrong M, Porter R (eds) Handbook of industrial organization. chapter 31, vol 3. Elsevier, pp 1967–2072

  • Gabriel S A, Kiet S, Balakrishnan S (2004) A mixed integer stochastic optimization model for settlement risk in retail electric power markets. Netw Spatial Econ 4(4):323–345

    Article  Google Scholar 

  • Gallay O, Hongler M (2008) Market sharing dynamics between two service providers. Eur J Oper Res 190(1):241–254

    Article  Google Scholar 

  • García-Bertrand R (2013) Sale prices setting tool for retailers. IEEE Trans Smart Grid 4(4):2028–2035

    Article  Google Scholar 

  • He F, Yin Y, Wang J, Yang Y (2016) Sustainability SI: optimal prices of electricity at public charging stations for plug-in electric vehicles. Netw Spat Econ 16(1):131–154

    Article  Google Scholar 

  • Hotelling H (1929) Stability in competition. Econ J 39:41–57

    Article  Google Scholar 

  • Jiménez J L, Perdiguero J (2011) Does accessibility affect retail prices and competition? An empirical application. Netw Spatial Econ 11(4):677–69

    Article  Google Scholar 

  • Joskow P, Tirole J (2006) Retail electricity competition. Rand J Econ 37 (4):799–815

    Article  Google Scholar 

  • Kirschen D S (2003) Demand-side view of electricity markets. IEEE Trans Power Syst 18(2):520–527

    Article  Google Scholar 

  • Laffont J -J, Rey P, Tirole J (1998) Network competition: I. Overview and nondiscriminatory pricing. RAND J Econ 29(1):1–37

    Article  Google Scholar 

  • Leyffer S, Munson T (2010) Solving multi-leader-common-follower games. Optim Methods Softw 25:601–623

    Article  Google Scholar 

  • Mallet P, Granstrom P O, Hallberg P, Lorenz G, Mandatova P (2014) Power to the People!: European perspectives on the future of electric distribution. IEEE Power Energy Mag 12(2):51–64

    Article  Google Scholar 

  • Müller M, Sensfuß F, Wietschel M (2007) Simulation of current pricing-tendencies in the German electricity market for private consumption. Energy Policy 35(8):4283–4294

    Article  Google Scholar 

  • Neto P A, Friesz T L, Han K (2016) Electric power network oligopoly as a dynamic Stackelberg game. Netw Spatial Econ 16(4):1211–1241

    Article  Google Scholar 

  • Oliveira F S, Ruiz C, Conejo A J (2013) Contract design and supply chain coordination in the electricity industry. Eur J Oper Res 227(3):527–537

    Article  Google Scholar 

  • OMIE (2016) Electric spanish market operator. http://www.omie.es/en/inicio

  • Qian L P, Zhang Y J A, Huang J, Wu Y (2013) Demand response management via real-time electricity price control in smart grids. IEEE J Selected Areas Commun 31(7):1268–1280

    Article  Google Scholar 

  • Rhodes A (2014) Re-examining the effects of switching costs. Econ Theory 57 (1):161–194

    Article  Google Scholar 

  • Shaffer G, Zhang Z J (1995) Competitive coupon targeting. Market Sci 14 (4):395:1986–1998

    Article  Google Scholar 

  • Somaini P, Einav L (2013) A model of market power in customer markets. J Ind Econ 61(4):938–986

    Article  Google Scholar 

  • U.S. Department of Energy (2006) Benefits of demand response in electricity markets and recommendations for achieving them. [Online] Available: http://energy.gov/oe/downloads/benefits-demand-response-electricity-markets-and-recommendations-achieving-them-report http://energy.gov/oe/downloads/benefits-demand-response-electricity-markets-and-recommendations-achieving-them-report

  • U.S. Goverment (2015) Energy policy act of 2005, Publ. L. 109-58. [Online] Available: http://www.gpo.gov/fdsys/pkg/PLAW-109publ58/pdf/PLAW-109publ58.pdf

  • Waterson M (2003) The role of consumers in competition and competition policy. Int J Ind Organ 21(2):129–150

    Article  Google Scholar 

  • Wei W, Liu F, Mei S (2015) Energy pricing and dispatch for smart grid retailers under demand response and market price uncertainty. IEEE Trans Smart Grid 6(3):1364–1374

    Article  Google Scholar 

  • Wilson C M, Price C W (2010) Do consumers switch to the best supplier? Oxford Economic Papers, gpq006

  • Xia Y (2011) Competitive strategies and market segmentation for suppliers with substitutable products. European J Oper Res 210(2):194–203

    Article  Google Scholar 

  • Zerrahn A, Huppmann D (2017) Network expansion to mitigate market power. Netw Spatial Econ 17(2):611–644

    Article  Google Scholar 

Download references

Acknowledgments

The authors gratefully acknowledge financial support from the Spanish government through project MTM2013-44902-P

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to C. Ruiz.

Appendices

Appendix

Proof of statements

Proof of Proposition 2

From (8) and (10) we can check that, under condition (11), it holdsthat

$$\theta_{ij}^{\ast}> 1 \text{ and }\ \theta_{ji}^{\ast} \leq 0, $$

implying\({F}_{ij}^{\ast } = 1\) and\({F}_{ji}^{\ast } = 0\),with satisfies Definition 2 of a dominant position of retailer\(R_i\).

For retailer \(R_i\), from (4)we get \(\pi ^{\ast }_i(p_i,p_j)= 1\) andthus the optimal utility becomes

$$ u_i^r(p_i,p_j) = (p_i - c_i)v^{\ast}(p_i) $$
(34)

and attains its maximum for a price\(p_i\) thatsatisfies

$$ (p_i - c_i) v^{\prime}(p_i) + v(p_i) = 0. $$
(35)

For retailer \(R_j\),because \(\pi ^{\ast }_j(p_j,p_i)= 0\),the optimal utility vanishes and thus the profit regardless of the price\(p_j\).□

Proof of Proposition 3

From (8) and (10), we can check that under condition (13), wehave

$$\theta_{ij}^{\ast} \geq 1 \text{ and }\ 0<\theta_{ji}^{\ast} < 1, $$

implying\({F}_{ij}^{\ast } = 1\) and\({F}_{ji}^{\ast } =\theta _{ji}^{\ast }\).

Hence, from (4) we get, \(\pi ^{\ast }_i(p_i,p_j)=\pi _{ij}+(1-\theta _{ji}^{\ast })\pi _{ji}\) forretailer \(R_i\) and\(\pi ^{\ast }_j(p_1,p_2)=\pi _{ji}\ \theta _{ji}^{\ast }\) forretailer \(R_j\).The respective utilities become

$$\begin{array}{@{}rcl@{}} u_i^r(p_i,p_j) &=& (p_i - c_i)v^{\ast}(p_i)(\pi_{ij}+(1-\theta_{ji}^{\ast})\pi_{ji}) \end{array} $$
(36)
$$\begin{array}{@{}rcl@{}} u_j^r(p_i,p_j) &=& (p_j - c_j)v^{\ast}(p_j)(\pi_{ji}\theta_{ji}^{\ast}). \end{array} $$
(37)

By considering (8) and (10), the first order conditions associated to maximize (36) and (37), with respectto \(p_i\) and\(p_j\),respectively, yields

$$\begin{array}{@{}rcl@{}} &&\left( v(p_i) + (p_i - c_i) v^{\prime}(p_i) \right) \left( 2k\pi_{ij} + \pi_{ji} (\varphi(p_i) - \varphi(p_j) + k - \gamma_j) \right)\\ &&+ \pi_{ji} (p_i - c_i) v(p_i) \varphi^{\prime}(p_i) = 0\\ &&\left( v(p_j) + (p_j - c_j) v^{\prime}(p_j) \right) \left( 2k\pi_{ji} - \pi_{ji} (\varphi(p_i) - \varphi(p_j) + k - \gamma_j) \right)\\ &&+ \pi_{ji} (p_j - c_j) v(p_j) \varphi^{\prime}(p_j) = 0 . \end{array} $$

Using the definition (12) the above conditions can be rewritten as

$$\begin{array}{@{}rcl@{}} {\Gamma} (p_i;c_i) &=& - \varphi(p_j) + \frac{1 + \pi_{ij}}{\pi_{ji}} k - \gamma_j\\ {\Gamma} (p_j;c_j) &=& - \varphi(p_i) + k + \gamma_j . \end{array} $$

Proof of Proposition 4

Under condition (16) we have

$$0<\theta_{ij}^{\ast} < 1 \text{ and }\ 0<\theta_{ji}^{\ast} < 1, $$

implying\({F}_{ij}^{\ast } =\theta _{ij}^{\ast }\) and\({F}_{ji}^{\ast } =\theta _{ji}^{\ast }\).

By using (4) we get \(\pi ^{\ast }_i(p_i,p_j)=\theta _{ij}^{\ast }\ \pi _{ij}+(1-\theta _{ji}^{\ast })\pi _{ji}\),and \(\pi ^{\ast }_j(p_i,p_j)=\theta _{ji}^{\ast }\ \pi _{ji}+(1-\theta _{ij}^{\ast })\pi _{ij}\).Thus, the profit function for each retailer becomes

$$\begin{array}{@{}rcl@{}} u_i^r(p_i,p_j) &=& (p_i - c_i)v^{\ast}(p_i)(\theta_{ij}^{\ast}\ \pi_{ij}+(1-\theta_{ji}^{\ast})\pi_{ji}) \end{array} $$
(38)
$$\begin{array}{@{}rcl@{}} u_j^r(p_i,p_j) &=& (p_j - c_j)v^{\ast}(p_j)(\theta_{ji}^{\ast}\ \pi_{ji}+(1-\theta_{ij}^{\ast})\pi_{ij}). \end{array} $$
(39)

By considering (8) and (10), the first order conditions associated to maximize (38) and (39), with respect to\(p_i\) and\(p_j\),respectively, yield

$$\begin{array}{@{}rcl@{}} &&\left( v(p_i) + (p_i - c_i) v^{\prime}(p_i) \right) \left( \varphi(p_i) - \varphi(p_j) +k+ \pi_{ij} \gamma_i - \pi_{ji} \gamma_j \right)\\ &&+ (p_i - c_i) v(p_i) \varphi^{\prime}(p_i) = 0\\ &&\left( v(p_j) + (p_j - c_j) v^{\prime}(p_j) \right) \left( \varphi(p_j) - \varphi(p_i) + k + \pi_{ji} \gamma_j - \pi_{ji} \gamma_i \right)\\ &&+ (p_j - c_j) v(p_j) \varphi^{\prime}(p_j) = 0. \end{array} $$

Using definition (12) the above optimal conditions can be rewritten as

$$\begin{array}{@{}rcl@{}} {\Gamma} (p_i;c_i) &=& -\varphi(p_j) + k + \pi_{ij} \gamma_i - \pi_{ij} \gamma_j\\ {\Gamma} (p_j;c_j) &=& -\varphi(p_i) + k - \pi_{ij} \gamma_i + \pi_{ji} \gamma_j. \end{array} $$

Proof of Theorem 1

As \({\Gamma }(c;c) = -\varphi (c)\),under Condition C.4 and from Lemma 1 there always exists a value\(p_i \in [c_i;\tau (c_i))\) satisfying(23) and \(p_j \in [c_j;\tau (c_j))\) satisfying(24), given any value for \(p_j\) or\(p_i\) respectively.Furthermore, this value is unique.

For\(q \in [-\varphi (c);\infty )\) define

$$G(q; c) \in {\Gamma}^{-1} (q;c) , \quad G(q; c) < \tau(c). $$

Lemma 1 implies G is a continuous and increasing function of q on\([-\varphi (c);\infty )\).

From this definition, a solution of

$$\begin{array}{@{}rcl@{}} p_i &=& G(-\varphi(p_j) + \kappa_i; c_i) \end{array} $$
(40)
$$\begin{array}{@{}rcl@{}} p_j &=& G(-\varphi(p_i) + \kappa_j; c_j) , \end{array} $$
(41)

is also a solution of (23)–(24).Define

$$\omega_i \equiv G(-\varphi(\tau(c_j)) + \kappa_i;c_i), \quad \omega_j \equiv G(-\varphi(\tau(c_i)) + \kappa_j;c^j) , $$

and note that for any\(p = (p_i, p_j)\) it will hold that\(p \in [c_i;\omega _i)\times [c_j;\omega _j)\). Hence, by considering theproperty that \(-\varphi \) isan increasing function, \(\omega _k < \tau (c_k)\),for \(k=\{i,j\}\),and \({\Gamma }(c;c) = -\varphi (c)\),we have

$$\begin{array}{@{}rcl@{}} && -\varphi(c_i) \leq - \varphi(c_j) + \kappa_i \leq -\varphi(p_j) + \kappa_i < -\varphi(\tau(c_j)) + \kappa_i\\ && \Rightarrow c_i \leq G(-\varphi(p_j) + \kappa_i; c_i) < \omega_i , \end{array} $$

with an equivalent bound holding for \(\omega _j\).

As a consequence, system (40)–(41) defines a fixed-point iteration\(p = {\Phi }(p)\) where\({\Phi } : [c_i;\omega _i)\times [c_j;\omega _j) \rightarrow [c_i;\omega _i)\times [c_j;\omega _j)\) andis continuous on that set. Brouwer’s fixed-point theorem then implies the existence of a fixed point for\({\Phi }\), that will also be asolution for (23)–(24). □

Proof of Theorem 2

Consider the fixed-point dynamics for the solutiondefined by (40)–(41), which we write in compact form as\(p = {\Phi }(p)\),and define the fixed-point iteration

$$ p = {\Phi}_j({\Phi}_i(p)) \equiv \psi (p) , $$
(42)

where \({\Phi } = ({\Phi }_i \ {\Phi }_j )^T\) and\(\psi : [c_j ; \tau (c_j)) \rightarrow [c_j ; \tau (c_j))\). Thisiteration is equivalent to (40)–(41), as any solution of (42) provides a solution for (40)–(41) by letting\(p_j = p\) and\(p_i = {\Phi }_i(p)\),and a solution for (42) can be obtained from (40)–(41) by letting\(p = p_j\).

G is differentiable wrt p on \([-\varphi (c_k);\infty )\),\(k=i,j\), andwe have that

$$\begin{array}{@{}rcl@{}} G^{\prime} \left( -\varphi(p) + \kappa_i;c_i \right) & = & \frac{v (p)}{{\Gamma}^{\prime}(G(-\varphi(p) + \kappa_i; c_i); c_i)} \end{array} $$
(43)
$$\begin{array}{@{}rcl@{}} G^{\prime} \left( -\varphi({\Phi}_i(p)) + \kappa_j;c_j \right) & = & \frac{v ({\Phi}_i(p)) }{{\Gamma}^{\prime}(G(-\varphi({\Phi}_i(p)) + \kappa_j; c_j); c_j)} . \end{array} $$
(44)

Using the expression for \({\Gamma }^{\prime }\) in(22), we can write

$$ {\Gamma}^{\prime}(p;c) = v(p) (1 + \rho(p;c)) , $$
(45)

for \(\rho \) definedin (25). We can rewrite (43)–(44) as

$$\begin{array}{@{}rcl@{}} G^{\prime} \left( -\varphi(p) + \kappa_i;c_i \right) & = & \frac{v(p)}{v(G(-\varphi(p) + \kappa_i; c_i))} \frac{1}{1 + \rho(G(-\varphi(p) + \kappa_i; c_i);c_i)}\\ \end{array} $$
(46)
$$\begin{array}{@{}rcl@{}} G^{\prime} \left( -\varphi({\Phi}_i(p)) + \kappa_j;c_j \right) & = & \frac{v({\Phi}_i(p))}{v(G(-\varphi({\Phi}_i(p)) + \kappa_j; c_j))}\\ & & \null \times \frac{1}{1 + \rho(G(-\varphi({\Phi}_i(p)) + \kappa_j; c_j);c_j)} . \end{array} $$
(47)

From (42) and using (40)–(41), replacing (46)–(47) and\({\Phi }_i(p) = G(-\varphi (p) + \kappa _i; c_i)\) wehave

$$\begin{array}{@{}rcl@{}} \psi^{\prime}(p) & = & \frac{v(p)}{v(G(-\varphi({\Phi}_i(p)) + \kappa_j; c_j))}\\ & & \null \times \frac{1}{1 + \rho(G(-\varphi({\Phi}_i(p)) + \kappa_j; c_j);c_j)} \frac{1}{1 + \rho(G(-\varphi(p) + \kappa_i; c_i);c_i)} . \end{array} $$

From Condition C.3’, for all\(p \in [ \bar c_k;\tau (c_k))\) and\(k=i,j\) it holdsthat \(1+\rho (p;c_k) > \sqrt {v(\bar c)/v(\bar \tau )}\), and as wealso have \(v(\bar c) \geq v(\bar c_j) \geq v(p) \geq v(\tau (c_j)) \geq v(\bar \tau )\), it followsthat

$$\psi^{\prime}(p) < \frac{v(\bar c_j)}{v(\tau (c_j))} \frac{v(\bar \tau)}{v(\bar c)} < 1 , $$

implying\(1 - \psi ^{\prime }(p) > 0\).

From this bound, \(p - \varphi (p)\) isstriclty increasing on \([\bar c_j; \tau (c_j))\),there can only be one zero of the function in that interval and the fixed point for (42) is unique on\([\bar c_i;\tau (c_i))\times [\bar c_j;\tau (c_j) )\).□

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ruiz, C., Nogales, F.J. & Prieto, F.J. Retail Equilibrium with Switching Consumers in Electricity Markets. Netw Spat Econ 18, 145–180 (2018). https://doi.org/10.1007/s11067-018-9384-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11067-018-9384-3

Keywords

Navigation