Skip to main content

Advertisement

Log in

Emissions trading, point-of-regulation and facility siting choices in the electric markets

  • Original Article
  • Published:
Journal of Regulatory Economics Aims and scope Submit manuscript

Abstract

This paper examines the effects of emissions trading on firms’ capacity siting choices in the electric sector, considering three programs that are different by their point-of-regulation: source-, load and first-jurisdictional-deliverer approaches. We model each program by assuming electricity sales are through bilateral contracts, and analyze the solutions properties and economic and emissions implications. We have three central findings in this paper. First, when electricity sales in both directions (i.e., in-state to out-of-state and vice verse) are subject to an emissions cap, three programs will produce identical siting distribution. Second, the emissions leakage in the long-run could be lower than in the short-run case in which capacity expansion is not allowed. This is so because in the long-run power producers may opt for clean technologies in fear of the emission costs under a cap-and-trade program. Third, if compared to pure source-based program in which the import sales are not subject to an emissions cap, more pollution-intensive facilities are expected to build in uncapped states, where the regulation is less stringent. Our findings are therefore consistent with some empirical evidence of pollution haven hypothesis.

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.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. Effort at federal level is mostly led by the EPA, following the decision by the US Supreme Court in 2007. This gives the EPA the much needed authority to regulate GHG under the Clean Air Act (CAA). Burtraw et al. (2011) identified three possible regulatory tools that can be used by the EPA: fuel economy standard (CAFE) for mobile sources, new source review (NSR) and new source performance standards (NSPS) under the CAA ?111 for stationary sources. Other federal policies that might have an impact on the designs of the federal GHG policies from the power sector include the Cross-State Air Pollution Rule (CSAPR). However, it is unclear how these policies or regulations at the federal level are going to unfold in the future.

  2. RGGI experienced a setback in 2011 when the governor of New Jersey announced his intention to withdraw from the RGGI (2011), arguing that (a) RGGIs permit price is too low to induce behavior changes, (b) the state of New Jersey has already met its 2020 emission target, and (c) other existing state renewable policies, e.g., Renewable Portfolio Standard (RPS), have already provided significant incentive for emission reduction. The overlapping or “patching of energy and environmental policies could result in unintended consequences, but is subject to limited research. For example, Tsao et al. (2011) found that when a C&T is coexisted with a RPS, lowering the C&T cap might penalize renewable units, and increasing the RPS level could sometimes benefit coal and oil units and make natural gas units worse off. More recently, the NJ Governor has further vetoed a bill passed by both houses of the state legislature that would have forced it to the RGGI to cut carbon emissions in the northeast (Point Carbon 2012).

  3. The constituents of the WCI have evolved over time, partially reflecting the reality of regional and global political economy. WCI began in 2007 with five US states, including Arizona, California, New Mexico, Oregon, and Washington. It was latter joined by four Canadian provinces, i.e., British Columbia, Manitoba, Ontario, and Quebec, and two additional US states: Montana and Utah. However, currently, while the four Canadian provinces are still with WCI, the sole US state that remains in the WCI is California . There is no question that the reincarnation of the WCI will affect the investment decisions and emission leakage.

  4. They are carbon dioxide, methane, nitrous oxide, hydrofluorocarbons, perfluorocarbons and sulfur hexafluoride.

  5. Note that each sector is considered separately based on historical or benchmark emissions. Therefore, inclusion of fuel suppliers in 2015 will not affect electric sector.

  6. According to a study by the California Energy Commission (Alvarado and Griffin 2007), through the California Climate Action Registrys Power/Utility Reporting Protocol, it is possible for a precise identification of 56 % of imported power.

  7. For more discussion on the interactions of C&T and pool-typed power market, please see Wolak et al. (2007) and Burtraw (2008).

  8. More recently, in their Interim Decision, the CPUC recommended that ARB use a regional default emission factor of 1,100 lbs/MWh to represent unspecified electricity. This emission factor was meant to loosely approximate the most likely source of marginal generation, a less efficient gas fired generator located out-of-state and within the Western Interconnection. However, this gives rise to the ”power laundry” opportunity as power with its emission rate higher than the default rate could disguise as relatively cleaner sources when selling into the California market. Moreover, as dirty power sources likely favor pool-typed market, which allow them to sell their power at a lower default emission rate in spot power market to avoid emission costs, this might provide disincentives for the dirty base load units to engage in the forward markets, thereby making the market less reliable or more vulnerable to market power.

  9. Although the first-seller proposal was not discussed by the RGGI working group, both source- and load-based approaches were considered. In particular, the RGGI Staff Working Group evaluated the efficacy of various load-based policies for mitigating potential emissions leakage (Sherry 2007).

  10. A variant unbundles emissions certificates and power and requires the LSE to purchase an amount of emissions certificates. This is represented by two proposals: the \(\mathrm {CO}_2\text{ RC }\) proposal (Michel and Nielsen 2007) and the “Tradable Emission Attribute Certificates” (TEAC) proposal (Gillenwater and Breidenich 2007). Further discussion of CO2RC and TEAC can be found elsewhere (Hobbs 2007; Wolak et al. 2007).

  11. For example, when other sectors that produce imperfectly substitute goods that are not subject to the same emission cap.

  12. Note that the capacity expansion equilibrium model falls into the classical two-stage, open-loop modeling framework (Murphy and Smeers 2005; Ehrenmann and Smeers 2010). However, since we consider a perfectly competitive market (as a starting point), we do not have the issue of open- versus closed-loop equilibria in an imperfect-competition context as in Murphy and Smeers (2005), Garcia and Shen (2010), Bushnell and Ishii (2007).

  13. More detailed constraints can certainly be incorporated without changing the theoretical results to be presented, as long as the added constraints preserve the convexity of firms’ profit-maximization problem.

  14. Note that we use a hub-spoke type of representation of the network; that is, an arbitrary zone is designated as a hub, and all power flows from point \(A\) to point \(B\) can be considered as the flow from \(A\) to the hub, and from the hub to \(B\). Hence, there is only one region index for the variables \(y_i\)’s and \(w_i\)’s, as they refer to the power flow and the corresponding transmission charge from the hub to the zone \(i\). Such a modeling approach can significantly reduce the number of variables needed to represent power flows in a transmission network, and will not affect any of the theoretical or numerical results.

  15. Note that by ignoring such nonconvexities, we exclude the \(\mathrm {CO}_2\) emissions associated with the start-up/ramping processes of power plants. Fossil fuel plants have different emissions rates during such processes than their normal operation. While this is a modeling limitation, one remedy could be to use historical data of the number of starts of particular power plants as a proxy for their future operations, and calculate the corresponding \(\mathrm {CO}_2\) emissions. We can then treat these emissions as constants and minus them from firms’ allocated permits.

  16. In a perfectly competitive market, the ownership of generators are not important in the sense that it will not affect the market equilibrium in any way. Hence, the firm indices associated with \(x\)’s can be viewed as generation capacity types instead.

  17. Note that a similar result to Proposition 2 appears in Ehrenmann and Smeers (2010). However, the proof in Ehrenmann and Smeers (2010) does not carry over to the models presented in this paper. This is so because a key step in their proof requires the uniqueness of the Lagrange multiplier. With the presence of transmission constraints and \(\mathrm {CO}_2\) emissions constraints in our model, which do not appear in Ehrenmann and Smeers (2010), the multiplier may not be unique (such as in a degenerate case caused by the various constraints). Hence, our proof for Proposition 2 is more general as it certainly applies to the case where there are no transmission or emissions constraints. Even if there are nonlinear constraints in (19) (such as those incurred by transmission constraints with transmission losses), as long as the feasible region of (19) is convex with a given \(x\in X\), and the convex region has a nonempty interior, Proposition 2 is still valid. In addition, the proof of Proposition 2 (as detailed in the Appendix D) can be easily extended to incorporate stochasticity, such as stochastic demand or generators’ availability, as long as the sample spaces of all of the random variables are countable.

  18. An additional producer can be modeled to represent as a new entry. However, since the market is perfectly competitive, the aggregate capacity would be the same.

  19. For instance, the calculation of the emission leakage under the pure source-based approach goes as follows: \(\varDelta E^{\mathcal{N }}=1579 -2021 = -441\) and \(\varDelta E^{\mathcal{N }^C} = (191 + 104 + 106 + 233 + 154)- 1550 =-763)\). Therefore, \(leakage = (1-(-441)/(-763))\times 100~\% = 42.2~\%\)

  20. In reality, the magnitude of leakage depends on the assumptions of load growth and choice of emission cap. We deliberately fixed these two parameters in our analysis because we are interested in the size of the leakage when considering capacity expansion.

  21. To what extent the \(\mathrm {CO}_{2}\) costs will be passed onto the electricity prices is a function of various factors, including demand and supply elasticities, market structure,, etc. and is subject to some research (please see Chen et al. 2008).

  22. This does not mean that we do not consider retirement or mothball decisions of certain generation capacities. In our model settings, since we do not consider costs other than investment and production (i.e., we do not consider fixed operation and management costs for plants when they do not generate), firms will

    simply not operate a plant if it is not profitable to generate. Numerically, if a plant generates in a market equilibrium without \(\mathrm {CO}_2\) regulations while producing nothing in an equilibrium with \(\mathrm {CO}_2\) regulations, this implies that the plant would be mothballed or retired when carrying costs of the plant are considered.

  23. A function \(f(x)\) defined on \(X\) is said to be coercive over \(X\) if for any sequence \(\{x_k\} \subset X\) that \(||x_k|| \rightarrow \infty \), \(f(x_k) \rightarrow -\infty \).

References

  • Alvarado, A., & Griffin, K. (2007). Revised methodology to estimate the generation resource mix of California electricity imports—Update to the May 2006 staff paper. Technical Reports, California Energy Commission.

  • Bertsekas, D. P. (1999). Nonlinear programming (2nd ed.). Belmont, MA: Athena Scientific.

    Google Scholar 

  • Burtraw, D. (2007). State efforts to cap the commons: Regulating sources or consumers. Technical report. Washington, DC: Resources for the Future.

  • Burtraw, D. (2008). Regulating \(\text{ CO }_2\) in electricity markets: Sources or consumers? Climate Policy, 8(6), 588–606.

    Article  Google Scholar 

  • Burtraw, D., & Palmer, K. (2005). Allocation of \(\text{ CO }_2\) emissions allowances in the regional greenhouse gas cap-and-trade program. Technical Report. Washington, D.C.: Resources for the Future.

  • Burtraw, D., Fraas, A., & Richardson, N. (2011). Policy monitor—Greenhouse gas regulation under the Clean Air Act: A guide for economists. Review of Environmental Economics and Policy, 5(2), 293–313.

    Article  Google Scholar 

  • Bushnell, J. (2007). The implementation of California AB32 and its impact on wholesale electricity markets, Technical Report WP-170. Berkeley: University of California Energy Institute.

  • Bushnell, J., & Chen, Y. (2012). Allocation and leakage in cap-and-trade markets for \(\text{ CO }_2\). Resources and Energy Economics, 34(3), 647–668.

    Article  Google Scholar 

  • Bushnell J., & Ishii J. (2007) An equilibrium model of investment in restructured electricity markets. Center for the Study of Energy Markets (CSEM) Working Paper 164.

  • California Air Resources Board. (2008). AB32 fact sheet: California Global Warming Solutions Act. http://www.arb.ca.gov/cc/factsheets/ab32factsheet.pdf. Accessed Dec 2008.

  • California Energy Commission. (2006). 2006 Net system power report. Technical Report CEC-300-2007-007. Sacramento: California Energy Commission.

  • California Energy Commission. (2007a). Low-carbon fuel standard. http://www.energy.ca.gov/low_carbon_fuel_standard/index.html. Accessed Dec 2007.

  • California Energy Commission. (2007b). Revised methodology to estimate the generation resources mix of California electricity imports. Technical Report CEC-700-2007-007, California Energy Commission.

  • California Energy Commission. (2012a). SB 1368 Emission Performance Standards. http://www.energy.ca.gov/emission_standards/. Accessed Dec 2012.

  • California Energy Commission. (2012b). Survey of Utility Resource Planning and Procurement Practices for Application to Long-Term Procurement Planning in California. http://www.cpuc.ca.gov/NR/rdonlyres/029611EA-D7C7-4ACC-84D6-D6BA8515723A/0/ConsultantsReportonUtilityPlanningPracticesandAppendices09172008.pdf. Accessed Dec 2012.

  • Chen, Y., Sijm, J., Hobbs, B. F., & Lise, W. (2008). Implications of \(\text{ CO }_2\) emissions trading for short-run electricity market outcomes in northwest Europe. Journal of Regulatory Economics, 34(3), 251–281.

    Article  Google Scholar 

  • Chen, Y., Liu, A. L., & Hobbs, B. F. (2011). Economic and emissions implications of load-based, source-based and first-seller emissions trading programs under California AB32. Operations Research, 59(3), 696–712.

    Article  Google Scholar 

  • Condliffe, S., & Morgan, S. (2009). The effects of air quality regulations on the location decisions of pollution-intensive manufacturing plants. Journal of Regulatory Economics, 36(1), 83–93.

    Article  Google Scholar 

  • Considine, T. J. (1999). Economies of scale and asset values in power production. The Electricity Journal, 12(10), 37–42.

    Article  Google Scholar 

  • Convery, F. J., & Redmond, L. (2007). Market and price developments in the European Union emissions trading schemes. Review of Environmental Economics and Policy, 1, 88–111.

    Article  Google Scholar 

  • Demailly, D., & Quirion, P. (2006). \(\text{ CO }_2\) abatement, competitiveness and leakage in the European Cement Industry under the EU ETS: grandfathering versus output-based allocation. Climate Policy, 6(1), 103–113.

  • Dinda, S. (2004). Environmental Kuznets curve hypothesis: A survey. Ecological Economics, 49, 431–455.

    Article  Google Scholar 

  • Ehrenmann, A., & Smeers, Y. (2010). Stochastic equilibrium models for generation capacity expansion. Cambridge Working Papers in Economics.

  • EIA. (2009). National energy modeling system: Electricity module. http://www.eia.doe.gov/oiaf/aeo/assumption/pdf/electricity.pdf. Accessed Dec 2009.

  • EIA. (2010). Updated capital cost estimates for electricity generation plants. http://www.eia.gov/oiaf/beck_plantcosts/. Accessed Dec 2010.

  • Eskeland, G. S., & Harrison, A. E. (2003). Moving to greener pastures? Multinational and the pollution haven hypothesis. Journal of Development Economics, 1, 1–23.

    Article  Google Scholar 

  • Fiacco, A. V., & Kyparisis, J. (1986). Convexity and concavity properties of the optimal value function in parametric nonlinear programming. Journal of Optimization Theory and Applications, 48(1), 95–126.

    Google Scholar 

  • Garcia, A., & Shen, Z. (2010). Equilibrum capacity expansion under stochastic demand growth. Operations Research, 58(1), 30–42.

    Article  Google Scholar 

  • Gillenwater, M., & Breidenich, C. (2007). Internalizing carbon costs in electricity markets: Using certificates in a load-based emissions trading scheme. Technical Report, Science, Technology and Environmental Policy Program. Princeton, NJ: Woodrow Wilson School of Public and International Affairs, Princeton University.

  • Henderson, J. V. (1996). Effects of air quality regulation. American Economics Review, 86(4), 789–831.

    Google Scholar 

  • Hobbs, B. F. (2001). Linear complementarity models of Nash–Cournot competition in bilateral and poolco power markets. IEEE Transactions on Power System, 16(2), 194–202.

    Article  Google Scholar 

  • Hobbs, B. F. (2007). Theoretical equivalence of \(\text{ CO }_2\text{ RC }\) and source-based \(\text{ CO }_2\) trading. Baltimore: Department of Geography & Environmental Engineering, The Johns Hopkins University.

  • Market Advisory Committee. (2007). Market Advisory Committee to the California Air Resources Board. Recommendation for desgining a greenhouse gas cap-and-trade system for California. Sacramento: California Air Resources Board.

  • Michel, S., & Nielsen, J. (2007). An alternative load-based \(\text{ CO }_2\) cap & trade for the West. Western Resource Advocate.

  • Millimet, D. L., & List, J. A. (2004). The case of the missing pollution haven hypothesis. Journal of Regulatory Economics, 26(3), 239–262.

    Article  Google Scholar 

  • Murphy, F., & Smeers, Y. (2005). Generation capacity expansion in imperfectly competitive restructured electricity markets. Operations Research, 53(4), 646–661.

    Article  Google Scholar 

  • Pineau, P. O., Rasata, H., & Zaccour, G. (2011). A dynamic oligopolistic electricity market with interdependent market segments. The Energy Journal, 32(4), 183–217.

    Article  Google Scholar 

  • Point Carbon. (2011). Carbon Market North America June, 2011. http://www.pointcarbon.com/news/cmna/1.1545983. Accessed June 2012.

  • Point Carbon. (2012). 2012 Carbon Market North America. http://www.pointcarbon.com/polopoly_fs/1.1944458!CMNA20120727.pdf. Accessed Oct 2012.

  • Regional Greenhouse Gas Initiative. (2008). An Initiative of the Northeast and Mid Atlantic States of U.S. http://www.rggi.org/. Accessed Dec 2008.

  • Rockafellar, R. T. (1982). Lagrange multipliers and subderivatives of optimal value functions in nonlinear programming. In D. C. Sorensen & R. J. B. Wets (Eds.), Nondifferential and variational techniques in optimization (pp. 28–66). Amsterdam: North- Holland.

    Chapter  Google Scholar 

  • Rockafellar, R. T. (1997). Convex analysis. Princeton: Princeton University Press.

    Google Scholar 

  • Ruth, M., Gabriel, S. A., Palmer, K., Burtraw, D., Paul, A., Chen, Y., et al. (2008). Economic and energy impacts from Maryland’s potential participation in the Regional Greenhouse Gas Initiative: A study commissioned by the Maryland Department of the Environment. Energy Policy, 36, 2279–2289.

    Article  Google Scholar 

  • Sherry, C. (2007). Monitoring potential emissions leakage under the regional greenhouse gas initiative: Possible lessons for load-based cap-and-trade programs. http://www.rggi.org/docs/il_report_final_3_14_07.pdf. Accessed Dec 2007.

  • Stoft, S. (2002). Power system economics. Piscataway: Wiley InterScience.

    Book  Google Scholar 

  • Tsao, C. C., Campbell, J. E., & Chen, Y. (2011). When renewable portfolio standards meet cap-and-trade regulations in the power sector: Market interactions, profit implications, and policy redundancy. Energy Policy, 39(7), 3966–3974.

    Article  Google Scholar 

  • Van Horn, A., & Remedios, E. (2008). A comparison of three cap-and-trade market designs and incentives for new technologies to reduce greenhouse gases. The Electricity Journal, 21(2), 51–62.

    Article  Google Scholar 

  • Wei, J. Y., & Smeers, Y. (1999). Spatial oligopolistic electricity models with Cournot generators and regulated transmission price. Operations Research, 47(1), 102–112.

    Article  Google Scholar 

  • Western Climate Initiative. (2008). Western Climate Initiative. http://www.westernclimateinitiative.org/. Accessed Dec 2008.

  • Wolak, F., Bushnell, J., & Hobbs, B. F. (2007). Opinion on load-based and source-based trading of carbon dioxide in California. Technical Report, Market Surveillance Committee of the California ISO.

  • Yao, J., Adler, I., & Oren, S. S. (2008). Modeling and computing two-settlement oligopolistic equilibrium in a congested electricity network. Operations Research, 56(1), 34–47.

    Article  Google Scholar 

Download references

Acknowledgments

The work of the first author is partially supported by the research funding from the Graduate Research Council (GRC), University of California, Merced; the second author is partially supported by National Science Foundation Grant CMMI-1234057. We would also like to thank the editor Prof. Crew for his helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yihsu Chen.

Appendices

Appendix A notation summary

A summary of notation is provided here for the ease of reading the modeling sections. In general we use script capital letters for sets, capital letters for parameters, lower-case letters for variables and indices and greek lower-case letters for dual variables. Also we use bold letters to represent vectors of appropriate dimensions.

Sets, Indices and Dimensions

\(i, j \in \mathcal{N }\)

Set of zones in a network, including both power producers and LSEs; \(|\mathcal{N }| = N\)

\(\mathcal{N }^C \subset \mathcal{N }\)

Set of zones that are under the emissions cap

\(\mathcal{N }^{NC} \subset \mathcal{N }\)

Set of zones that are not under the emissions cap; \(\mathcal{N }^{NC} = \mathcal{N }\backslash \mathcal{N }^C\)

\(l \in \mathcal{A }\)

Set of (directional) transmission links in the network \(\mathcal{N }\); \(|\mathcal{A }| = A\)

\(f\in \mathcal{F }\)

Set of power producers; \(|\mathcal{F }| = F\).

Parameters

\(P_j^0\)

Price intercept of the inverse demand function in zone \(j\) ($/MW(h))

\(Q_j^0\)

Quantity intercept of the inverse demand function in zone \(j\) ($/MW(h))

\({\overline{E}}^{^P}_f/{\overline{E}}^{^{LSE}}_j\)

Initial free allocation of \(\mathrm {CO}_2\) allowances to power producer \(f\) or LSE \(j\) (tons)

\(E_{fi}\)

Emission rate of firm \(f\) in zone \(i\) (tons/MW(h))

\({\overline{E}}\)

System-wide \(\mathrm {CO}_2\) cap (tons)

\(\alpha _{fi}\)

Per unit cost of capacity expansion ($/MW)

\(K_{fi}\)

Firm \(f\)’s existing capacity at location \(i\) (MW); \(K_{fi} \ge 0,\ \forall f\in \mathcal{F },\ i\in \mathcal{N }\)

\(C_{fi}(\cdot )\)

Firm \(f\)’s production cost function in zone \(i\) ($/MW(h))

\(PTDF_{li}\)

The \((l, i)\)-th element of the power transmission-distribution factor matrix

\(T_l\)

Transmission capacity bound on line \(l\); \(T_l > 0 \ \forall l\in \mathcal{A }\)

\({{1}\!\!\mathrm{l}}_{\left[ {\cdot } \right] }\)

An indicator function—if the condition in the brackets is true, \({{1}\!\!\mathrm{l}}_{\left[ {\cdot } \right] } = 1\); else, \({{1}\!\!\mathrm{l}}_{\left[ {\cdot } \right] } = 0\).

Variables

\(d_{j}\)

Electricity demand in zone \(j\) (MW(h))

\(s_{fij}\)

LSE \(j\)’s purchase of power from firm \(f\) in zone \(i\) (MW(h)). (Note that there is

 

assumed to be just one LSE per zone; more general assumptions are possible,

 

but would not change the conclusions of our analysis.)

\(s_{fj}\)

Power firm \(f\)’s sale of electricity to LSE in zone \(j\) (MW(h))

\(g_{fij}\)

Electricity sold to LSE \(j\), produced by firm \(f\) in zone \(i\) (MW(h))

\(x_{fi}\)

New capacity addition by firm \(f\) at location \(i\) (MW)

\(y_j\)

MWs transmitted from the hub to zone \(j\) (MW)

\(w_j\)

Fees charged for transmitting electricity from the hub to zone \(j\) ($/MW)

\({\tilde{p}}_{j}\)

The price of electricity paid by consumers to an LSE in zone \(j\) ($/MW(h))

\(p_{fij}\)

The price of electricity produced by generator \((f,i)\) sold to LSE \(j\) ($/MW(h))

\(p^{CO_2}\)

\(\mathrm {CO}_2\) allowance price ($/ton)

Appendix B Theoretical properties of the open-loop models

In this Appendix we provide theoretical results for the three equilibrium models introduced in Sect. 3—Model Mo-SB, Model Mo-LB and FJD). More specifically, we show equilibrium existence and uniqueness.

Proposition 3

Given the inverse demand functions in (12), the investment cost functions in (13) and the production cost functions in (14), there exists an equilibrium to Model Mo-SB, Model Mo-LB and Model FJD, respectively.

Proof

If consumers’ highest willingness-to-pay for electricity is less than (or equal to) the lowest-cost generation unit’s marginal cost, (that is, \(\max _{f\in {\mathcal{F }},\ i, j\in {\mathcal{N }}}\{P^0_j - a_{fi}\} \le 0 \)), then it is straightforward to verify that \(\mathbf 0 \) (a vector of all 0’s with a proper dimension) is an equilibrium to all the three models. If \(\max _{f\in {\mathcal{F }},\ i, j\in {\mathcal{N }}}\{P^0_j - a_{fi}\} > 0 \), then let every variable be 0 except for the capacity expansion variables \(x_{fi}\) and the Lagrange multipliers associated with the generation capacity constraints (denoted as \(\rho _{fi}\)). For the nonzero variables, we set that \(\rho ^{\circ }_{fi} = \max _{j\in {\mathcal{N }}}\{P^0_j - a_{fi}, 0\}\), and \(x^{\circ }_{fi} = \rho _{fi}/\alpha _{fi}\), \(\forall \ f\in {\mathcal{F }}\) and \(i \in {\mathcal{N }}\). Then the tuple \((\mathbf 0 ,\mathbf x ^{\circ }, \varvec{\rho }^{\circ })\) is feasible to each of the mixed linear complementarity problem resulted from Model Mo-SB, Mo-LB and FJD. The rest of the proof follows exactly as that in Chen et al. (2011), Theorem 1. \(\square \)

The following proposition is a straightforward extension of Theorem 2 in Chen et al. (2011), and the proof is hence omitted.

Proposition 4

Let a vector \(\mathbf z \) represent the variables (\(\mathbf{x,d, s,g,w,y, } \ {\tilde{\mathbf{p}}},\mathbf{p, p^{CO_2}}\)) in one of the equilibrium models and \(\varvec{\zeta }\) represent the (vector of) dual variables corresponding to the constraints. Assume that the system-wide \(\mathrm {CO}_2\) cap among the three models is the same; namely,

$$\begin{aligned} \displaystyle {\overline{E}}^{SB} = {\overline{E}}^{LB} = {\overline{E}}^{FJD}. \end{aligned}$$
(24)

Further assume that the assumptions in Proposition 3 hold. Then if \((\mathbf z , \varvec{\zeta })\) solves the mixed complementarity problem (MiCP) derived from Model Mo-SB, Mo-LB or FJD, it solves the MiCP derived from the other two models.

Appendix C Proof of Lemma 1

Lemma 1

The inequalities (20)–(22) are equivalent.

Proof

Since we consider an electricity market with bilateral contracts; that is, \(s_{fij} = g_{fij}, \forall f, i, j\), we can replace the \(s_{f ij}\)’s in (21) and (22) with \(g_{fij}\)’s. Let \({\mathcal{N }} \times {\mathcal{N }}\) denote the Cartesian product of the set of node indices. It can be partitioned into the following four sets.

$$\begin{aligned} \mathcal{I }_1 := {\mathcal{N }}^C \times {\mathcal{N }}^C,\ \mathcal{I }_2 := {\mathcal{N }}^C \times {\mathcal{N }}^{NC},\ \mathcal{I }_3 := {\mathcal{N }}^{NC} \times {\mathcal{N }}^C,\ \mathcal{I }_4 := {\mathcal{N }}^{NC} \times {\mathcal{N }}^{NC}.\nonumber \\ \end{aligned}$$
(25)

For any \((i, j)\in {\mathcal{N }}\times {\mathcal{N }}\), it must be in one and only one of the above four sets. Using the partitioned sets, we can rewrite (20)–(22) as follows.

$$\begin{aligned} \text{ Modified } \text{ SB: } \displaystyle \sum \limits _{f\in {\mathcal{F }}}\left( \underset{(i, j) \in \mathcal{I }_1\cup \mathcal{I }_2\cup \mathcal{I }_3}{\sum \limits }E_{fi}g_{fij}\right) \le {\overline{E}} \\ \text{ Modified } \text{ LB: } \displaystyle \sum \limits _{f\in {\mathcal{F }}}\left( \sum \limits _{(i, j) \in \mathcal{I }_1\cup \mathcal{I }_3\cup \mathcal{I }_3}E_{fi}g_{fij}\right) \le {\overline{E}} \\ \text{ FJD: } \displaystyle \sum \limits _{f\in {\mathcal{F }}}\left( \sum \limits _{(i, j) \in \mathcal{I }_3\cup \mathcal{I }_1\cup \mathcal{I }_2}E_{fi}g_{fij}\right) \le {\overline{E}}. \end{aligned}$$

Apparently the above three inequalities are equivalent. \(\square \)

Appendix D Proof of Proposition 2

To prove Proposition 2, we first discuss solution existence of the two-stage problem (16). Let \(Z(x)\) denote the feasible region of the 2nd-stage problem (i.e., Problem (19)) with respect to a given \(x\in X\), where \(X\) is the Cartesian product of all firms’ feasible capacity expansion sets \(X_f\) (that is, \( X = \varPi _{f\in {\mathcal{F }}} X_f\)). To rule out uninteresting cases, we make the blanket assumption that \(X\ne \emptyset \). Even with this assumption, \(Z(x)\) may still be empty with a given \(x \in X\). In this case, by convention we assign \(\mathcal{Q }(x)\) to be \(-\infty \). For \(X\), note that throughout the paper we only consider the case where \(X\subset \mathfrak R ^{FN}_+\) (namely, all the \(x_{fi}\)’s are nonnegative).Footnote 22 Then it is easy to check that \(\mathbf 0 \) (the vector of zeros of a proper dimension) is always in \(Z(x)\) for any \(x\in X\). Hence, \(\mathcal{Q }(x) > -\infty \), \(\forall x\in X\). With the \(P_j(\cdot )\) and \(C_{fi}(\cdot )\) functions defined in (12) and (14), respectively, the objective function \(Q(\mathbf z )\) is convex quadratic and can be easily seen to be coercive over \(Z(x)\).Footnote 23 Since \(Z(x)\) is always nonempty and closed (as all the functions in the constraint set of (19) are continuous) for each \(x\in X\), together with the coerciveness \(Q(\mathbf z )\), an optimal solution of the 2nd-stage problem (19) (denoted as \(\mathbf z ^*\)) exists for each \(x\in X\). (See the existence result in Bertsekas (1999), Proposition A.8.) Consequently, \(\mathcal{Q }(x)\) is finite for each \(x\in X\). Then with the investment cost functions given in (13), \(\varPi (x)\) is also coercive over \(X\). With a further assumption of \(X\) being a closed set, we have the following existence result regarding to the two-stage problem (16).

Lemma 2

Let the inverse demand functions, the investment cost functions, and the production cost functions be given as in (12)–(14), respectively. Let \(X\subset \mathfrak R ^{FN}_+\) be nonempty and closed. Then an optimal solution of (16) exists. \(\square \)

Note that in showing the solution existence, we do not require the set \(X\) to be bounded. Though capacity expansions are bounded by engineering or other constraints in reality, modeling-wise not requiring explicit bounds on the expansion variables is nonetheless more convenient.

We then present several results regarding to the properties of the optimal value function \(\mathcal{Q }(x)\) in the two-stage problem (16).

Lemma 3

Given the conditions in Lemma 2 and that \(X\) is a convex set, \(\mathcal{Q }(x)\) is concave on \(X\).

Proof

With the given inverse demand functions and cost functions, \(Q(z)\) is easily seen to be concave. In addition, \(Z(x)\), the parameterized feasible region that can be viewed as a set-valued mapping, is apparently a convex mapping as all the constraints defining the region are linear. Then the concavity of \(\mathcal{Q }(x)\) on \(X\) follows Proposition 2.1 in Fiacco and Kyparisis (1986). \(\square \)

The optimal value function \(\mathcal{Q }(x)\) is in general not differentiable. With the concavity of \(\mathcal{Q }(x)\), however, it has subgradients on \(X\), and the collection of the subgradients is referred to as the subdifferential, denoted as \(\partial \mathcal{Q }(x)\). Given an \(x\in X\), by the discussion for Lemma 2, an optimal solution to Problem (19) exists. Let it be denoted as \({\tilde{z}}(x)\) (or simply \({\tilde{z}}\)). The optimal solution may not be unique though, such as in the case of linear production cost functions. Let \(Z^*(x)\) denote the set of optimal solutions to Problem (19) with a given \(x\). Since the constraints in (19) are all linear, the Linear Constraint Qualification (LCQ) holds for the entire set \(Z(x)\). Then for each \({\tilde{z}}\in Z^*(x)\), there exist corresponding Lagrange multipliers. Let \(\mathcal P ({\tilde{z}}; x)\) denote the set of Lagrange multipliers corresponding to \({\tilde{z}}\) and associated with the capacity constraints in (19). To establish the relationship between the equilibrium model and the optimization model, it is key to show the relationship between \(\partial \mathcal{Q }(x)\) and \(\mathcal P ({\tilde{z}}; x)\), which is formally stated in the following lemma.

Lemma 4

Given the conditions in Lemma 2, for an \(x\in X\), let \({\tilde{z}} \in Z^*(x)\). \(\forall \rho \in \mathcal{P }({\tilde{z}}; x)\), \(\rho \in \partial \mathcal{Q }(x)\). If in addition that the Magasarian-Fromovitz Constraint Qualification (MFCQ) holds at each \({\tilde{z}} \in Z^*(x)\), then \(\displaystyle \partial \mathcal{Q }(x) = \text{ conv }\{\cup _{{\tilde{z}}\in Z^*(x)}\mathcal{P }({\tilde{z}}; x)\}\), where ‘\(\mathrm {conv}\)’ denotes the convex hull of a set.

Proof

Let \({\bar{x}}\) be an arbitrary point in \(X\), and let \({\bar{z}}\) be the corresponding optimal solution. Since the constraints in (19) are all linear, the Linear Constraint Qualification (LCQ) holds for any \(z\in Z({\bar{x}})\). Hence there exist Lagrange multipliers associated with \({\bar{z}}\). For the ease of argument, we use \(h(z)=0\), a vector of functions with proper dimension, to represent all equality constraints in (19), and use \(l(z)\le 0\) to represent all inequality constraints in (19) except for the generation capacity constraint. Let \(\mathbf g \) denote the vector of \((\sum _{j\in {\mathcal{N }}}g_{fij})_{f\in {\mathcal{F }}, i\in {\mathcal{N }}}\). Further define \(\bar{v}\) and \(\bar{u}\) to be the vectors of Lagrange multipliers (with appropriate dimensions) associated with \(h({\bar{z}})\) and \(l({\bar{z}})\). (19) is easily seen to be a convex optimization problem given an \(x\in X\). Hence the strong duality of nonlinear programming holds. With the Lagrangian dual function of (19) (considering minimizing the negative of \(Q(z)\)) defined as follows

$$\begin{aligned} \theta (\rho (x), u, v) \equiv \displaystyle \inf _{z\in \mathfrak R ^{N+FN+2A}}\left\{ -Q(z) + \rho ^T(x)(\mathbf g - K - x) + u^Tl(z) + v^Th(z)\right\} \nonumber \\ \end{aligned}$$
(26)

we have the following

$$\begin{aligned} -\mathcal{Q }({\bar{x}})&= \displaystyle \theta (\rho ({\bar{x}}), {\bar{u}}, \bar{v})\nonumber \\&= \inf _{z\in \mathfrak R ^{N+FN+2A}}\left\{ -Q(z) \!+\! \rho ^T({\bar{x}})(\mathbf g \!-\! K \!-\! {\bar{x}})\!+\! {\bar{u}}^Tl(z) \!+\! \bar{v}^Th(z)\right\} \nonumber \\&= \rho ^T({\bar{x}})(x - {\bar{x}}) + \displaystyle \inf _{z\in \mathfrak R ^{N+FN+2A}}\left\{ -Q(z) + \rho ^T({\bar{x}})(\mathbf g - K - x)\right. \nonumber \\&\left. + {\bar{u}}^Tl(z) + \bar{v}^Th(z)\right\} \nonumber \\&\le \rho ({\bar{x}})^T(x - {\bar{x}}) - \mathcal{Q }(x), \end{aligned}$$
(27)

where the last inequality follows from the weak duality of nonlinear programming. By the definition of subgradients, we obtain that \(-\rho (x)\in \partial \left( - \mathcal{Q }(x)\right) \) for \(x\in X\), and hence, \(\rho (x) \in \partial \mathcal{Q }(x)\) for \(x \in X\).

To show that \(\partial \mathcal{Q }(x) = \text{ conv }\{\cup _{{\tilde{z}}\in Z^*(x)} \mathcal{P }({\tilde{z}}; x)\}\), we apply the result in Rockafellar (1982) (Corollary 2 of Theorem 2), which states that if \(\mathcal{Q }(x)\) is tame and that if the MFCQ holds at each \({\tilde{z}}\in Z^*(x)\), then

$$\begin{aligned} \partial \mathcal{Q }(x) = \text{ cl } \text{ conv }\left\{ \displaystyle \left[ \bigcup _{{\tilde{z}}\in Z^*(x)} \mathcal{P }({\tilde{z}}; x)\right] \bigcap \partial \mathcal{Q }(x) \right\} . \end{aligned}$$
(28)

As stated in Rockafellar (1982), ‘tameness’ is a local boundedness assumption on how a parameterized feasible set varies with the parameter. Let \(x^{\prime }\in X\) be a small perturbation of \(x\); i.e., for a \(\delta > 0\), \(|| x^{\prime } - x || \le \delta \). In the discussion of Lemma 2, the feasible set of Problem 16, \(Z(x)\), is always nonempty for each \(x\in X\). In addition, an optimal solution of the 2nd-stage problem (19) always exists with a given \(x \in X\). Hence, there exist finite \(\omega \) and \(\omega ^{\prime }\) such that \(\mathcal{Q }(x) = \omega \) and \(\mathcal{Q }(x^{\prime }) = \omega ^{\prime }\). Let \(\omega _0 = \min \{\omega , \omega ^{\prime }\} - \epsilon \) with \(\epsilon > 0\). Then by the coerciveness of \(Q(z)\) shown in the discussion of Lemma 2, we obtain that the level set \(\{z^{\prime }\in Z(x^{\prime }): Q(z^{\prime }) \ge \omega _0 \}\) is bounded. Hence, by Proposition 9 in Rockafellar (1982), \(\mathcal{Q }(x)\) is tame. Together with the MFCQ assumption in the lemma, the characterization of the subdifferential of \(\mathcal{Q }(x)\) in (28) holds true. By the first part of this proof, we know that \(\cup _{{\tilde{z}}\in Z^*(x)} \mathcal{P }({\tilde{z}}; x) \subset \partial \mathcal{Q }(x)\). Hence,

$$\begin{aligned} \partial \mathcal{Q }(x) = \text{ cl } \text{ conv }\left\{ \displaystyle \left[ \bigcup _{{\tilde{z}}\in Z^*(x)} \mathcal{P }({\tilde{z}}; x)\right] \right\} . \end{aligned}$$
(29)

Since the functions in the optimization problem (19) are all continuously differentiable, the Lagrange multiplier set is closed, and so is \(\mathcal{P }({\tilde{z}}; x)\) for each \({\tilde{z}}\in Z^*(x)\). Further by the MFCQ assumption, \(\mathcal{P }({\tilde{z}}; x)\) is also bounded with each \({\tilde{z}}\in Z^*(x)\). Then \(\cup _{{\tilde{z}}\in Z^*(x)} \mathcal{P }({\tilde{z}}; x)\) is a closed bounded set, and the convex hull of which is closed. Hence,

$$\begin{aligned} \partial \mathcal{Q }(x) = \text{ conv }\left\{ \displaystyle \left[ \bigcup _{{\tilde{z}}\in Z^*(x)} \mathcal{P }({\tilde{z}}; x)\right] \right\} . \end{aligned}$$
(30)

\(\square \)

We need one more technical result before proving Proposition 2, which also has intuitive economic meaning.

Lemma 5

Let \(X\) be given as \(\{x\in \mathfrak R ^{FN}: x\ge 0\}\), and \(x^*\) be an optimal solution of the Problem (16). Then under the conditions of Lemma 2, for each \({\tilde{z}} \in Z^*(x^*)\) and \({\tilde{\rho }} \in \mathcal{P }({\tilde{z}}; x^*)\), \(\alpha _{fi} x_{fi}^*\ge {\tilde{\rho }}_{fi}\), \(\forall f\in {\mathcal{F }},\ i\in {\mathcal{N }}\), where \(\alpha _{fi}\)’s are the coefficients of the capacity expansion variables \(x_{fi}\)’s in (13).

Proof

By Lemma 2, an optimal solution exists to Problem (16). Let \(x^*\) be such a solution. Suppose that there exists a \({\tilde{z}}\in Z^*(x^*)\) and a \({\tilde{\rho }} \in \mathcal{P }({\tilde{z}}; x^*)\) such that for an \(h\in {\mathcal{F }}\) and \(j\in {\mathcal{N }}\), \({\tilde{\rho }}_{hj} > \alpha _{hj}x_{hj}^*\). Define a vector \({\bar{x}}\in \mathfrak R ^{FN}\) such that \({\bar{x}}_{hj} = x^*_{hj} + \epsilon \) and \({\bar{x}}_{fi} = x^*_{fi}\) for all \(h\ne f\in {\mathcal{F }}\) and \(j\ne i\in {\mathcal{N }}\), where \(\epsilon \) is an arbitrarily small positive number. By the way that the set \(X\) is defined, \({\bar{x}}\) is also in \(X\). By Lemma 4, \({\tilde{\rho }} \in \partial \mathcal{Q }{x}\). Then we have the following.

$$\begin{aligned} \varPi ({\bar{x}})&= \displaystyle - \sum \limits _{f\in {\mathcal{F }}, i \in {\mathcal{N }}} \frac{1}{2}\alpha _{fi} {\bar{x}}_{fi} + \mathcal{Q }({\bar{x}}) \nonumber \\&\ge \displaystyle - \sum \limits _{f\in {\mathcal{F }}, i \in {\mathcal{N }}} \frac{1}{2}\alpha _{fi} {\bar{x}}_{fi} + \mathcal{Q }(x^*) - {\tilde{\rho }}^T(x^* - {\bar{x}}) \nonumber \\&= \displaystyle - \sum \limits _{f\in {\mathcal{F }}, i \in {\mathcal{N }}} \frac{1}{2}\alpha _{fi} x^*_{fi} - \alpha _{hj} \left( \epsilon x^*_{hj} + \frac{\epsilon ^2}{2}\right) + \mathcal{Q }(x^*) + {\tilde{\rho }}_{hj} \epsilon \nonumber \\&= \displaystyle \varPi (x^*) + \epsilon \left( {\tilde{\rho }}_{hj} - \alpha _{hj}x_{hj}^* - \frac{1}{2}\alpha _{hj}\epsilon \right) , \end{aligned}$$
(31)

where the first inequality is implied by the definition of subgradients. Since it is assumed that \({\tilde{\rho }}_{hj} > \alpha _{hj}x_{hj}^*\), then there must exist an \(\epsilon > 0\) such that (\({\tilde{\rho }}_{hj} - \alpha _{hj}x_{hj}^* - \frac{1}{2}\alpha _{hj}\epsilon ) > 0\), which implies that \(\varPi ({\bar{x}}) > \varPi (x^*)\). This contradicts with the assumption that \(x^*\) maximizes \(\varPi (x)\) over \(X\). Hence, the claim of the lemma holds true. \(\square \)

The interpretation of Lemma 5 is that for any optimal capacity expansion decision \(x_{fi}\), its corresponding (per unit) capital cost must be greater than or equal to the marginal value of expanding that capacity.

Now we are ready to prove Proposition 2, which is restated below (in more mathematically precise statements) to ease the presentation.

Proposition 2

Assume that the set \(X\) is given as \(\{x\in \mathfrak R ^{FN}: x\ge 0\}\). Under the conditions in Proposition 1, if a tuple \((x^*_,\ d^*_,\ s^*_,\ g^*_,\ y^*_,\ \zeta ^*)\) solves Model FJD, where \(\zeta ^*\equiv (\ w^*_,\ {\tilde{p}}^*_,\ p^*_,\ p^{{CO_2}^*})\), then \(x^*\) solves the two-stage optimization problem (16) with \((d^*_,\ s^*_,\ g^*_,\ y^*) \in {\arg \max }\{Q(z): {z\in Z(x^*)}\}\). Conversely, if \(x^*\) solves (16) with \((d^*, s^*, g^*, y^*) \in {\arg \max }\{Q(z): {z\in Z(x^*)}\} \), then there exists a vector \(\zeta ^*\) such that \((x^*, d^*, g^*, y^*, \zeta ^*)\) solves Model FJD. In addition, the two-stage optimization problem (16) has a unique solution in capacity expansion decisions.

Proof

By Lemma 3, the optimal value function \(\mathcal{Q }(x)\) is concave over \(X\). Together with the convex quadratic investment functions (13) and the assumption on the set \(X\), the two-stage optimization problem (16) is a convex problem. By the general result of optimality conditions of convex programming (see Theorem 28.3 in Rockafellar 1997), \(x\) is an optimal solution to (16) if and only if there exist \(\gamma \in \mathfrak R ^{FN}\) such that

$$\begin{aligned} 0&\in \varLambda ^T x - \partial \mathcal{Q }(x) - \gamma , \nonumber \\ 0&\le x \perp \gamma \ge 0, \end{aligned}$$
(32)

where \(\varLambda \in \mathfrak R ^{FN\times FN}\) is a diagonal matrix with the diagonal entries being the investment cost parameters \(\alpha _{fi}\)’s in (13). The optimality condition (32) can be equivalently stated as that for some \(\xi \in \partial \mathcal{Q }(x^*)\),

$$\begin{aligned} 0\le x \perp \varLambda ^Tx - \xi \ge 0. \end{aligned}$$
(33)

Now we first assume that \((x^*_,\ d^*_,\ s^*_,\ g^*_,\ y^*_,\ \zeta ^*)\) solves Model FJD. Since the optimization problems in Model FJD are all convex with linear constraints, the first-order conditions are both necessary and sufficient conditions of optimality. By grouping the KKT systems of problems (1), (5), (9), (10), together with the market clearing conditions (11), it is straightforward to check that the resulting mixed complementarity problem (minus the first-order condition corresponding to the variable \(x^*_f\) of the generator’s problem (9)) is exactly the KKT system of the 2nd-stage optimization problem (19), with \(x\) set at \(x^*\). Since (19) is a convex program as well with a fixed \(x\), then satisfying the KKT system is a sufficient optimality condition; and hence, \(z^* \equiv (d^*_,\ s^*_,\ g^*_,\ y^*) \in {\arg \max }\{Q(z): {z\in Z(x^*)}\}\). Let \(\rho ^*\) denote the vector of Lagrange multipliers associated with the capacity constraints in the generating firms’ optimization problem (9). Then \(\rho ^* \in \mathcal{P }(z^*; x^*)\), the set of Lagrange multipliers corresponding to \(z^*\) and associated with the capacity constraints in (19). By Lemma 4, \(\rho ^* \in \partial \mathcal{Q }(x^*)\). Then the first-order condition corresponding to the variable \(x^*_f\) of the generator’s problem (9) is exactly the same as (33), and hence, \(x^*\) is an optimal solution of (16).

Conversely, if \(x^*\) is an optimal solution to the optimization problem (16), then (33) holds for some \(\xi \in \partial \mathcal{Q }(x^*)\). By Lemma 4, we know that \(\xi \in \text{ conv }\{\cup _{{\tilde{z}}\in Z^*(x^*)} \mathcal{P }({\tilde{z}}; x^*)\}\). Let \(\mathcal M ({\tilde{z}}; x^*)\) denote the set of Lagrange multipliers associated with a \({\tilde{z}}\in Z^*(x^*)\), the optimal solution set of the 2nd-stage problem (19) with the given \(x^*\). \(\mathcal M ({\tilde{z}}; x^*)\) is convex by the well-known fact about Lagrange multiplier sets (see, for example, Bertsekas 1999). Then the set \(\mathcal{P }({\tilde{z}}; x^*)\) as a projection of \(\mathcal M ({\tilde{z}}; x^*)\) onto the \(\rho \)-space is also a convex set. By a corollary to the Carathéodory’s Theorem (Corollary 17.1.1 in Rockafellar 1997), which states that the convex hull of a collection of convex sets can be expressed by a convex combination of affinely independent points, each belonging to a different convex set in the collection, we can write \(\xi \) as follows.

$$\begin{aligned} \displaystyle \xi = \sum \limits _{m=0}^{FN} \tau _m {\tilde{\rho }}^m,\ \text{ with }\ {\tilde{\rho }}^m\in \mathcal{P }^m({\tilde{z}};x^*),\ \sum \limits _{m=0}^{FN} \tau _m = 1,\ \tau _m \ge 0,\ \forall m = 0,\ldots ,\! FN.\nonumber \\ \end{aligned}$$
(34)

By (33), with the same \({\tilde{\rho }}^m\)’s and \(\tau _m\)’s as in (34), we have that

$$\begin{aligned} 0&\le x^* \perp \varLambda ^Tx^* - \sum \limits _{m=0}^{FN} \tau _m {\tilde{\rho }}^m \ge 0 \Longrightarrow 0\\&\le x^*_{fi} \perp \sum \limits _{m=0}^{FN}\tau _m(\alpha _{fi}x_{fi}^* - {\tilde{\rho }}^m_{fi}) \ge 0,\quad \forall f \in {\mathcal{F }}, i\in {\mathcal{N }}. \end{aligned}$$

Let \(\mathcal{I }\) be an index set such that \(\mathcal{I } := \{(f, i) \in {\mathcal{F }} \times {\mathcal{N }}:\ x_{fi}^* > 0 \}\). By complementarity, for each \((h, j) \in \mathcal{I }\), \(\displaystyle \sum \nolimits _{m=0}^{FN}\tau _m(\alpha _{hj}x_{hj}^* - {\tilde{\rho }}^m_{hj}) = 0\). By Lemma 5, \(\alpha _{hj}x_{hj}^*- {\tilde{\rho }}^m_{hj} \ge 0\) for each \(m\). Hence, \(\alpha _{hj}x_{hj}^* - {\tilde{\rho }}^m_{hj} = 0\) for each \((h, j) \in \mathcal{I }\) and \(m = 0, \ldots , FN\). Then pick an arbitrary \(m\) from \(0, \ldots , FN\), we have that

$$\begin{aligned} 0\le x^* \perp \varLambda x^* - {\tilde{\rho }}^m \ge 0. \end{aligned}$$
(35)

Let \(z^*\equiv (d^*_,\ s^*_,\ g^*_,\ y^*)\) be the optimal solution of (19) corresponding to the multiplier \({\tilde{\rho }}^m\) with the given \(x^*\), and let \(\mathcal M ^m(z^*; x^*)\) be the corresponding multiplier set. Then for any \(\zeta ^* \in \mathcal M ^m(z^*; x^*)\), it is straightforward to check that the KKT system of (19), together with (35), is exactly the same as the mixed complementarity problem of Model FJD, which consists of the KKT systems of (1), (5), (9) and (10), together with the market clearing conditions (11). Hence, the tuple \((x^*_,\ z^*_,\ \zeta ^*)\) solves Model FJD.

The uniqueness of solutions to the optimization model (16) follows directly from the above equivalence result and Proposition 1 (part (b)). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Chen, Y., Liu, A.L. Emissions trading, point-of-regulation and facility siting choices in the electric markets. J Regul Econ 44, 251–286 (2013). https://doi.org/10.1007/s11149-013-9224-9

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11149-013-9224-9

Keywords

JEL Classification

Navigation