Skip to main content
Erschienen in: Zeitschrift für Energiewirtschaft 3/2010

Open Access 01.09.2010

Cost Recovery in Congested Electricity Networks

verfasst von: Guido Pepermans, Bert Willems

Erschienen in: Zeitschrift für Energiewirtschaft | Ausgabe 3/2010

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Large scale investments in European electricity networks are foreseen in the next decade. Pricing the network at marginal cost will not be sufficient to pay for those investments as the network is a natural monopoly. This paper derives numerically the socially optimal transmission prices for cost recovery, taking into account that electricity networks are often congested, while allowing for market power in generation. The model is illustrated with a Stackelberg game for the Belgian electricity market.
Hinweise
A 2004 working paper version of this paper was titled ‘Ramsey Pricing in a Congested Network with Market Power in Generation: A Numerical Illustration for Belgium’. The authors are indebted to Ronnie Belmans, Severin Borenstein, Jim Bushnell, Bruno De Borger, Ben Hobbs, Shmuel Oren, Stef Proost, Yves Smeers, Patrick Van Cayseele, Frank Verboven and Jian Yao for their valuable comments. Any remaining errors are the responsibility of the authors. Bert Willems is the recipient of a Marie Curie Intra European Fellowship (PIEF-GA-2008-221085).

1 Introduction

Significant investments in the high voltage network will be necessary in the next decade. Two forces drive those investment needs: Cross-border transmission lines are essential to further integrate the European energy market as current cross-border connections are chronically congested and the integration of renewable energy sources in the network will require significant network upgrades to deal with intermittency problems. The cost of those network upgrades are significant and need to be recovered from network users. The International Energy Agency (IEA 2006, p. 148) estimated the cost of electricity transmission investments to be $159 billion for Europe. Given the stringent renewable energy and CO2 targets, today’s estimates are likely to be larger.
As electricity networks are natural monopolies, charging consumers a price equal to their marginal impact on the network will not suffice to cover the full network cost. Indeed, there are large scale economies and providing network services to one extra user will be cheaper than the average cost for providing the existing users.
This paper derives how network users should be charged for network usage, taking into account those economies of scales and that (parts of) the network might be congested. Our paper takes a social planner’s viewpoint and focuses on short run efficiency of network operation.
The main message of the paper is how cost recovery and efficient network usage can be combined. Energy producers and consumers pay an injection charge and a take-off charge respectively. Efficient network usage requires that no additional transaction specific charges are imposed, i.e. once a connection charge is paid, grid users can freely use the network as long as the network is uncongested. For cost recovery, consumers and generators should pay a network connection charge which is inversely proportional to the demand and supply elasticity respectively, i.e. the charge should be independent of how they use the network. Hence the paper derives a generalization of the marginal nodal prices (Schweppe et al. 1988).
We then extend the discussion to the case where generation firms have (local) market power. In that case the network operator might want to “subsidize” those firms to increase production and/or to reduce grid congestion. This might improve overall welfare.1
The interaction of cost recovery, congestion management and (local) market power are hard to disentangle in an analytical model. Therefore, the model is illustrated numerically for the Belgian electricity market. We model a Stackelberg game in which in stage 1 the network operator sets transmission charges and in stage 2 generators compete while taking transmission charges as given. Three scenarios are studied: a reference case in which we assume a network operator without budget constraint (i.e. the cost recovery issue is absent) and perfect competition in generation. The second scenario adds a budget constraint, while the last scenario drops the assumption of perfect competition and assumes Cournot competition.
The next section reviews the literature on transmission pricing and modeling market power in the energy markets. Sections 3 and 4 describe the structure of the model and the data. Section 5 discusses the simulation results and, finally, Sect. 6 concludes.

2 Literature Review

A large part of the literature on pricing electricity networks has focused on ensuring the efficient usage of the network but has neglected cost recovery.

2.1 Nodal Spot Prices

The seminal result of Schweppe et al. (1988) is that optimal transmission prices should be zero if the network is not congested. With congestion, transmission prices should be set according to the peak load pricing rule (see also Hsu 1997). Transmission prices should reflect the opportunity cost of using the transmission line. These are the so-called congestion charges or marginal nodal congestion charges.
Green (2004) compares the welfare effects of two other pricing rules with this first best benchmark. He compares a system of nodal prices, uniform prices and a hybrid, in which generators face nodal prices but consumers face a uniform price. He finds that moving from uniform prices to optimal nodal prices could raise welfare by 1.5% of the generators’ revenues. Moreover, doing so would also provide investment signals. Our approach differs from Green’s approach as we only use nodal prices, while considering the impact of a budget constraint. A study similar to Green’s was made by Bjorndal (2000), who compared the efficiency of different types of zonal pricing schemes.

2.2 Cost Recovery

There is a lively debate on how network costs, in particular the cost of network upgrades, should be allocated to the users of the grid. For instance, the integration of massive amounts of renewable energy requires investments in undersea cables to connect wind farms, the procurement of additional reserve capacity, more flexible transformers etc. The costs of such upgrades are not always allocated to the final users. Knight et al. (2005) give an overview of the current practices in several EU Member States. Most of these mechanisms allocate the cost of the existing network to all users and allocate the (average or marginal) costs of upgrades to the users of the extensions. Some proposed allocation methods suggest that firms should only pay for small network upgrades that are directly related with their usage (for instance the local transformer), while other methods define upgrade costs in a very broad sense. Firms would then have to pay ‘deep connection charges’, that is, they would need to pay for upgrade costs caused in the entire network.
Swider et al. (2008) argue that grid connection costs of renewable energy should be socialized. A related discussion is how network cost should be allocated between countries when there are imports, exports and transit flows (Olmos Camacho and Pérez-Arriaga 2007). One of the problems that arise is that contractual obligations and transit flows do not take into account the physical properties of the underlying network. Non-contracted loop-flows, i.e. flows which are the result of energy trades, do not only affect the importing and exporting country, but affect third countries as well (Daxhelet and Smeers 2005). In contrast with current practices in Europe, our model acknowledges the physical properties of the network, and fully takes into account loop flow.
The operation of the transmission network is a natural monopoly, featured by decreasing average costs as transmission output increases.2 Setting the price of network services equal to the marginal cost would generate insufficient revenue to cover costs. This is a well known problem in markets with increasing returns to scale, which is often neglected in the electricity literature and which is where the paper contributes. In order to cover costs, the network operator is required to increase transmission charges. A welfare maximizing network operator will do this in a deadweight loss minimizing way, i.e. using Ramsey prices (Ramsey 1927). Transmission tariffs are set inversely proportional to the demand and supply elasticities at the different nodes, meaning that the network operator will not only charge the users of a particular transmission line, but all users in the network.
Note that a truly welfare maximizing pricing mechanism would imply using a two-part tariff. However, we assume that the network operator does not do this, despite the fact that in many countries, it is common practice to charge a transmission fee with a fixed and a variable component. We have two motivations for this choice. First, the linear tariff assumption can be interpreted as the case where a two-part tariff is being used of which the fixed charge is insufficient to fully recover fixed transmission costs. In that case, the remainder of the costs needs to be covered by the usage charge. Second, in most cases the ‘fixed’ part is not completely fixed. It depends on actual network usage and therefore still creates a deadweight loss.3 In the absence of distortion-free instruments, it then remains optimal to use a form of Ramsey prices.
Our paper is most in line with, who also study cost recovery. They look at three models where costs are allocated proportional to different measures of network usage. They present simulation results for the Spanish market. Contrary to our paper, they assume that generation firms are perfectly competitive and that consumers do not react to the additional transmission charges used to cover costs. Advantages and disadvantages of the different methods are highlighted but no global welfare standard is used to compare different scenarios.

2.3 Congestion and Market Power

In energy markets, generators often have (local) market power, for instance in a so called load pocket.4 In situations where firms have local market power, the network operator might find it optimal to subsidize production as this might create additional benefits in the market, for instance by reducing congestion. Note that, as a consequence, the network operator will have to increase transmission chargers for other network users in order to balance his budget. Thus, subsidizing production is not necessarily welfare improving as it comes at a cost. This also happens in practice. A network operator, realizing that some firms have market power, will often sign ‘must run’ contracts with those firms, offering them long term contract at a preferential rate of return on their extra production.
The contribution of our paper is to allow for such types of (local) market power and to derive optimal transmission prices in that case. The results are similar in spirit to those found by Buchanan (1969) and Barnett (1980). They study the optimal taxation of externalities when there is market power in the output market. Typically the Pigouvian taxes, which are supposed to correct the externality, will be lower than under perfect competition due to the additional market failure (market power). Our paper is different as our regulator faces a budget constraint.

2.4 Modeling Market Power

Before describing the model itself, we briefly review the relevant literature of imperfect competition in generation. Two approaches have been developed in the specialized literature to model imperfect competition in such a multi-good market: the multi-unit auction and the supply function equilibrium approach.5 One of the major drawbacks of both approaches is that the spatial structure of the electricity market, and therefore the impact of transmission constraints, is often omitted. Both are quite difficult to apply in a market with transmission constraints. One notable exception is the work of Hobbs et al. (2000) who restrict themselves to linear supply functions. Most researchers therefore opt for some kind of Cournot market, while dropping some of the multi-good aspects of the actual market.6
In this paper generators behave à la Cournot in the energy market but are price takers in the transmission market. This is inspired by the models of Smeers and Wei (1997) and Wei and Smeers (1999). Both papers consider Cournot behavior in generation, but Smeers and Wei (1997) assume that generators perceive transmission prices as being set on the basis of congestion pricing, whereas Wei and Smeers (1999) assume regulated pricing as the basis for transmission pricing. The current model differs from Wei and Smeers (1999) in that the transmission charges are set by the network operator rather than by a regulatory rule. We assume that the welfare maximizing transmission firm also takes into account market imperfections in electricity generation when setting its transmission prices.
The electricity market is modeled as a Stackelberg game. Mathematically, this gives rise to an Mathematical Program with Equilibrium Constraints (Luo et al. 1996). Such models are becoming more and more common in the simulation of the electricity markets, but often the timing of the game is reversed: generators bid in stage 1 and the network operator balances the network in stage 2. See for instance (Ehrenmann 2004; Gabriel and Leuthold 2010; Hobbs et al. 2000 and Neuhoff et al. 2005)
In order to solve the numerical optimization problem, we relax the complementarity conditions of the second stage and use several starting values to be confident that we are close to the global optimum. For a comparison of different methods to solve MPEC problems see Fletcher and Leyffer (2002).

2.5 Long term efficiency

In our paper we do not study the long term incentives for investments in generation capacity. We restrict ourselves to short run efficiency. Cossent et al. (2009) and Joskow and Tirole (2005) discus some of the challenges we face regarding the long term integration of renewable energy into the network. We need a regulatory scheme that gives both long and short term incentives to investors to invest in new generation capacity as well as to the network operators to upgrade networks. The coordination of investments in generation and transmission assets is still an unsolved problem.
Nevertheless, we hope that the qualitative results of our paper will hold also in the long run, if the short term demand and supply elasticities are replaced with their long term equivalents. This remains for further study. Dijk et al. (2009) look at the impact of network pricing on the dynamic efficiency of investments in the electricity sector in a stochastic two stage oligopolistic model. It is shown that social and private incentives are not aligned and that additional regulatory instruments are required. Their model takes congestion into account, but cost recovery is neglected.

3 The Model

Define the sets F and G as the sets of generation firms and generation plants. Let G f be the set of generation plants owned by generation firm fF. With I being the set of network nodes, G i denotes the generation plants at node iI, and G fi the generation plants at node i owned by firm f. Furthermore, let A be the set of transmission lines in the network.
For notational simplicity, the model will be further described as if it concerned a one period model, i.e. a model that does not distinguish between peak and off-peak periods. However, the numerical simulations in Sect. 5 differentiate between peak and off-peak demand in a 4-period model.
The model distinguishes three types of players: consumers, generation firms and the network operator.
Consumers are price takers. At node i, they consume s i units of electricity. Their inverse demand for electricity, denoted as p i (s i ), is downward sloping and concave. Consumer prices include compensation for both the generation and the transmission of electricity.
Generation firmfF maximizes profits, while acting as a price taker in transmission. At node i, it owns the generation plants gG fi .
Electricity generation in plant g is q g and the generation cost is C g (q g ). Total generation costs are convex, with fixed generation costs suppressed to zero. The generation capacity of plant g is labeled \(\bar{q}_{g}\). Output should be nonnegative and cannot exceed available generation capacity. Therefore, we have \(0\le q_{g}\le\bar{q}_{g}\).
The network operator or transmission company maximizes social welfare and sets a nodal transmission charge \(\tau_{i}^{c}\) for consumers and \(\tau_{i}^{p}\) for generators.7 This is the per unit amount generators and consumers have to pay for injecting and taking power from the grid, respectively. Note that these charges can be different. For instance, a generator who generates electricity in node i and sells electricity in node j will pay\(\tau_{i}^{p}+\tau_{j}^{c}\). However, the optimal transmission charges are not uniquely defined.
First, take a node i at which no consumers are connected. For that node, the consumer transmission price \(\tau_{i}^{c}\) does not play a role and it can safely be set equal to zero. The same is true for nodes without generation. Here, \(\tau_{i}^{p}\) is not uniquely defined and the charge is set equal to zero.
Second, we can without loss of generality set one of the charges equal to zero because it is only the sum of the consumer and generation transmission charge, and not its composition that is important. The network operator has therefore one degree of freedom in setting the transmission charge components. This can easily be checked by noting that one can uniformly increase all generation tariffs with t and decrease all consumer tariffs with t without changing the sum of the charges. We can therefore arbitrarily fix one transmission price in one node equal to zero. This is done for the consumption charge in the swing node, i.e. \(\tau_{i}^{c}=0\) with i=swing node.
Finally, note that the model implicitly assumes that the charges need to be paid for all consumption and generation, even if generation and consumption are located at the same node. A generator in node i who sells electricity locally does not use the transmission network, but will have to pay a transmission payment \(\tau_{ii}=\tau_{i}^{c}+\tau_{i}^{p}\). We will call this charge the price wedge in node i as it creates a wedge between the consumer’s price and the generator’s price.
The model has two stages. In the first stage, the transmission operator sets transmission prices. In the second stage, generation firms play a Cournot game in which transmission prices and their competitor’s quantities are assumed as given. The next subsection describes the second stage of the game.

3.1 The Second Stage

Each firm f observes the transmission charges \(\tau_{i}^{p}\) and \(\tau_{i}^{c}\) as set by the network operator and plays a Cournot game. A firm f collects revenue by selling s fi units of electricity at node i at the per unit price p i . Firms also set the production level q g (gG f ) at each of their plants. Their competitor’s sales in node i, denoted by \(\tilde{s}_{-fi}\), are taken as given. Apart from generation costs, firms also pay a transmission cost \(\tau_{i}^{p}\) for injecting electricity to the network at node i, and \(\tau_{i}^{c}\) for the delivery of electricity to node i. This results in the following profit function for generation firm f,
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ1_HTML.gif
(1)
The nodal price p i that is received by generator f depends on the total sales in that node, i.e.
$$\begin{array}{@{}l}p_{i}=p_{i}(s_{i}),\\[2mm]s_{i}=s_{fi}+\tilde{s}_{-fi}\end{array}$$
where the tilde indicates that the variable is considered as given. In (1), the first term reflects revenues from electricity sales net of transmission charges paid at the consumption nodes. The second term reflects generation costs and transmission charges to put the electricity on the network. Summarizing, we have the following maximization problem for a generator:8
$$\begin{array}{@{}l@{\ }l}\displaystyle \mathop{\mathrm{Max}}\limits_{s_{fi},q_{g}(g\in G_{f})}&\displaystyle\Pi_{f}^{Gen}=\sum_{i\in I}(p_{i}-\tau_{i}^{c})\cdot s_{fi}\\[4mm]&\displaystyle\phantom{\Pi_{f}^{Gen}=}{}-\sum_{i\in I}\sum_{g\in G_{fi}}[C_{g}(q_{g})+\tau_{i}^{p}q_{g}]\\[4mm]\mbox{s.t.}&0\le q_{g}\le\bar{q}_{g}\quad(\underline{\mu}_{g},\bar{\mu}_{g})\ \forall g\in G_{f},\\[4mm]&\displaystyle\sum_{i\in I}s_{fi}=\sum_{g\in G_{f}}q_{g}\quad (\lambda_{f}^{p}),\\[4mm]&s_{i}=s_{fi}+\tilde{s}_{-fi}\quad\forall i\in I.\end{array}$$
(2)
As noted before, the first constraints reflect generation capacity constraints. The second constraint represents the energy balance at the firm level, i.e. total output should equal total sales. The last constraint represents demand. This constraint has no multiplier as it is substituted into the objective function and the other constraints before derivatives are taken.
The following first order conditions are then derived:
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ3_HTML.gif
(3)
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ4_HTML.gif
(4)
These are the standard first-order conditions for profit maximization, i.e. as long as generation constraints are not binding, marginal revenue equals marginal cost in all market segments. The Lagrange multiplier of the energy balance constraint \(\lambda_{f}^{p}\), is the value of energy in the network for generation firm f. This value is different for every firm.
Cost minimization requires that each firm equalizes the sum of the marginal cost and the generation charge at all generation plants. Profit maximization requires that marginal revenues net of consumption charges are equalized.
Note that each firm’s reaction function with respect to the sales s fi and the transmission charges, \(\tau_{i}^{c}\) and \(\tau_{i}^{p}\) can be derived from (2), (3) and (4).
The multipliers \(\underline{\mu}_{g}\) and \(\bar{\mu}_{g}\) are positive, and satisfy the complementarity conditions:
$$\begin{array}{@{}l}\underline{\mu}_{g}\ge0,\qquad \underline{\mu}_{g}\cdot q_{g}=0,\\[2mm]\bar{\mu}_{g}\ge0,\qquad \bar{\mu}_{g}\cdot(\overline{q}_{g}-q_{g})=0.\end{array}$$
(5)

3.1.1 Electricity Transmission

The model captures the technical features of the electricity system, especially at the level of electricity transport. Electricity transport is subject to physical constraints. These constraints have an impact on the power flow through the network and therefore potentially also on the pricing of transmission services. In this paper we concentrate on active power and we adopt a simplified DC flow model without losses.9
Each line aA in the network is characterized by a transmission capacity Q a . Denoting the flow over the line a as Q a , we must have
$$Q_{a}\le\bar{Q}_{a}\quad\forall a\in A.$$
(6)
Transmission must not be larger than the available transmission capacity.
The flow over a line a depends on the injection and the extraction of electrical energy in all the nodes of the network, except in the swing node. The flow on line a is equal to:
$$\sum_{i\in I/\{\mathrm{swing\ node}\}}\theta_{a,i}(q_{i}-s_{i})=Q_{a}$$
(7)
with s i ,q i the total consumption and generation in node i, respectively:
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ8_HTML.gif
(8)
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ9_HTML.gif
(9)
The flow over a transmission line is a linear function of the net injections at all nodes. The variables θ a,i are the power transmission distribution factors (PTDFs). The factors θ a,i describe how much a transaction from node i to the swing bus will change the flow on line a. For example, a transaction of size F from node i to node j can be decomposed as a transaction from node i to the swing node and a (negative) transaction from the swing node to nodej. The impact of this on line a then equals F×(θ a,iθ a,j). The PTDFs are determined by the physical properties of all the lines and the layout of the network.
Equations (6)–(9) describe the transmission possibilities of the network, i.e. they define the production feasibility set of the network operator.

3.1.2 Security of Supply

The network operator also needs to secure the supply of electricity. A minimal requirement for this is that, if unexpectedly a line goes out of service the remaining lines should still be able to transport all supplied electricity. This is the “n−1” rule. The network operator will check that for all contingencies kK the network is still capable of accommodating all flows.
For instance, if during contingency k the line αA breaks down, then the set of the remaining lines A\{a} should be able to transport the power over the network. After a contingency, the flows redistribute themselves over the network, and these new flows should still be feasible given the thermal constraints of the remaining lines.
Taking into account the security of supply for all contingencies K, the following equation needs to be added to the network equations (6)–(9):
$$\sum_{i\in I/\{\mathrm{swing\ node}\}}\theta_{a,i}^{k}(q_{i}-s_{i})\le\bar{Q}_{a}\quad\forall a\in A,\ \forall k\in K.$$
(10)
For notational simplicity, we will assume that the set of contingencies K also includes the case where all lines are available, such that we can drop equations (6) and (7). Equations (8)–(10) describe the feasible set for transmissions on the security constrained network.

3.2 The First Stage

The network operator maximizes welfare subject to a budget constraint. He sets the consumption and generation transmission charges (\(\tau_{i}^{c}\) and \(\tau_{i}^{p}\)), which can be differentiated over the nodes. It is assumed that the cost of providing transmission services is separable into operating costs and capacity costs. In the present model, operating costs and network losses are neglected. Therefore, only the capacity costs B remain.10
The profit of the network operator is then equal to:
$$\Pi^{tr}=\sum_{i\in I}(\tau_{i}^{c}s_{i}+\tau_{i}^{p}q_{i})-B.$$
(11)
The first term between brackets is the revenue of selling transmission services to consumers at node i. The second term is the revenue of selling transmission to the generators at node i. The last term represents capacity costs. By assumption, capacity costs are fixed.
The network operator maximizes
$$W=\sum_{i\in I}\int_{0}^{s_{i}}p_{i}(t)dt-\sum_{g\in G}C_{g}(q_{g})$$
(12)
subject to the energy balance at the firm level,
$$\sum_{i\in I}s_{fi}=\sum_{g\in G_{f}}q_{g}\quad \forall f\in F$$
(13)
the Cournot behavior (Sales-Production),
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ14_HTML.gif
(14)
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ15_HTML.gif
(15)
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ16_HTML.gif
(16)
the network equations
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ17_HTML.gif
(17)
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ18_HTML.gif
(18)
https://static-content.springer.com/image/art%3A10.1007%2Fs12398-010-0021-1/MediaObjects/12398_2010_21_Equ19_HTML.gif
(19)
and the budget constraint:
$$\sum_{i\in I}(\tau_{i}^{c}s_{i}+\tau_{i}^{p}q_{i})-B=\Pi^{tr}\ge0.$$
(20)

4 Data and Calibration

Before continuing with the simulations, we discuss the data that were used as an input for the model. Also, the calibration procedure will be described. The choice of the technical features of the transmission grid and of the available generation plants is inspired by the Belgian electricity system.

4.1 The Network

Figure 1 shows the network that has been modeled. It consists of 55 nodes and 92 lines and includes all the Belgian 380 kV and 220 kV transmission lines, but also some 380 kV lines in The Netherlands and France because they are important for the flows inside the Belgian network. The full lines on the graph are 380 kV lines, the dotted lines are 220 kV lines. The line between Gouy and Avelgem represents several lines of the 110 kV network that connect both nodes.11
We impose the n−1 rule for all Belgian 380 kV lines, except for some loose ends. Imposing the n−1 rule for these latter lines makes no sense when we do not include lines at lower voltage levels that can accommodate the flows in case of a failure of the 380 kV line. Also, the n−1 rule is not imposed for the interconnections with France and The Netherlands and for the lines within these countries, because sufficient or adequate information is lacking.

4.2 Electricity Generation

The total generation capacity connected to the grid is 14475 MW. Of this capacity, approximately 1070 MW consists of smaller generation plants which are not included in the model. These are mainly combined heat and power generation units (970 MW), and some small hydro units (90 MW). We assume that in any time period, 50% of these plants produce electricity. The remaining production capacity (13405 MW), is spread over 51 generation units, which are modeled in the paper based upon data of the year 2002.12 Each generator is assumed to have constant marginal production costs C g .
In the simulations, we assume three Cournot players, having a market share in generation capacity of 43%, 34% and 23%, respectively.13
Each player maximizes profit, taking into account plant characteristics. Generation decisions are described by the first order conditions (3) and the complementarity conditions (5). The player’s complementarity conditions are non-linear, which makes the optimization problem non convex.
Three plants are pumped storage plants, i.e. they can store energy in the form of a water reservoir. When generation costs are low, these plants consume electricity and pump water to a higher level. When generation costs are high, the reservoir is emptied and electricity is produced. The underlying decision process is not modeled in this paper. We assume that these plants generate electricity during peak periods at a marginal cost of €13 per MWh, and we count them as part of the consumption side during the off-peak periods.14

4.3 Electricity Demand

The model has been calibrated on the basis of Belgian data for electricity demand in 2002.15 In that year the average demand was 9.52 GW. Total yearly demand in Belgium is 83.4 TWh per Year. Figure 2 presents a histogram of demand in Belgium. The histogram is based on periodical observations with a length of 15 minutes. The highest and lowest observed demand levels were 13.7 GW and 5.8 GW, respectively.
Obviously, the demand for electricity is not constant over time and in order to take this into account, the numerical one-period model has been extended to a 4-period model. Currently, the model does not take full account of the potential links that could exist between these four periods. Taking these links into account would certainly enrich the model and this is seen as a priority for further research.
At least four of such potential links can be identified. First, cross-substitution can take place between time periods. For example, demand for electricity during the night will not only depend on the price in the night, but also on the price that is charged during the day. In this model it is assumed that these cross-substitution effects are zero. There is thus no intertemporal substitution.
Second, consumption and generation decision for the pumped storage plants create a link between periods and can be endogenized.
Third, intertemporal production constraints exist for generation, because generators can increase or decrease output only at a certain speed (ramping constraints). Starting up and shutting down generators is costly and requires time. Finally, the transmission firm is maximizing welfare and is subject to a budget constraint. This constraint creates an intertemporal link as the marginal welfare cost of obtaining revenue should now be equal over time periods.16 In this paper, only this last effect is taken into account.

4.4 Network Operator

The network operator has total costs of B=€649M per year (Source: Annual report ELIA, 2002). Capital costs are about 50% of the total costs, the other 50% being operating costs, such as wages and network maintenance costs. Wages and network maintenance costs are not directly related to the amount of MW transported over a line, they are inherent to the existence of that line. Therefore, as we could not find a more detailed description of the cost function of the network operator, we assume all costs to be fixed. Network losses are neglected in the model. Clearly, these would depend on the actual use of the network. With a total electricity demand of 83.4 TWh in 2002, the average cost of the network operator is €7.78 per MWh.

4.5 Calibration

The calibration of the model involves three steps. Each of these three steps is described below.

4.5.1 Fixing Periodic Aggregate Demand and the Length of Each Period

The first step is to decide about the level of electricity demand in each of the four periods and about the length of each period in a standard year. This has been done on the basis of the data presented in Fig. 2. This figure shows how often a certain demand level occurs in the Belgian market. We will consider 4 periods with average demand levels fixed at 8, 10, 11.5 and 12.5 GW. The length of each time period is then set such that the cumulative distribution function of the 4 periods approximates the observed cumulative distribution function (Table 1). As 500 MW of this demand is provided by small generators, the demand level as seen by the generators in our model is fixed 500 MW lower. Thus, the demand levels used to calibrate the demand functions are 7.5, 9.5, 11 and 12 GW.
Table 1
Calibration of the 4 time periods
Period
Observed demand (GW)
Period length
Model demand (GW)
Reference price (€ per MWh)
(hr)
(%)
1
12.5
208
2.4%
12.0
45.2
2
11.5
1759
20.1%
11.0
37.9
3
10.0
3410
38.9%
9.5
35.6
4
8.0
3383
38.6%
7.5
27.0
Average
9.6
8760
100.0%
9.1
33.0

4.5.2 Fixing a Reference Price for Each Period

Given the periodic electricity demand derived in the first step, we minimize the production costs to supply this demand. Here, it is assumed that pumped water storage can only be used in periods one and two. In periods three and four, pumped storage plants pump water into a reservoir.
Via this procedure, we obtain the marginal production cost for each period. The obtained values are increased with the average costs of the network operator (€7.78 per MWh) to obtain a reference price for each period (Table 1).

4.5.3 Fixing Periodic Electricity Demand in the Consumption Nodes

In the third step, we derive for each node a linear demand function. The price elasticity of demand is assumed to be −0.2 in all nodes and all periods.17 Total demand is distributed proportionally over the different nodes on the basis of the demand data in Van Roy (2001) and the reference prices are calculated in step 2. This information is sufficient to derive for each consumption node the parameters of the linear demand function.

4.6 Transit

The model also takes into account that the Belgian grid is used for relatively large transit flows. These flows are generally directed from France to The Netherlands and, as a first approximation, we assume an exogenous transit flow of 1000 MW from the south to the north. This transit is assumed to occur in all periods. The foreign generation and load nodes are summarized in Table 2.
Table 2
Exogenous generation levels at the foreign nodes (Negative numbers are loads)
The Netherlands
Node
 
Maasbracht
−731 MW
Geertruidenberg
−368 MW
Borssele
99 MW
TOTAL
−1000 MW
France
Node
 
Avelin
543 MW
Lonny
34 MW
Moulaine
423 MW
TOTAL
1000 MW

5 Simulation Results

This section discusses the results of three simulation runs. Section 5.1 discusses the first best case, with perfect competition in generation and no budget constraint for the network operator. The following subsections subsequently drop one of these assumptions: Sect. 5.2 adds a budget constraint for the network operator, Sect. 5.3 adds imperfect competition in generation.

5.1 First Best

In the first best the network operator does not face a budget constraint and generators are competing competitively, but the transmission capacity is limited. Table 3 shows the simulation results for a representative hour in each period. It gives total welfare, the surpluses for the economic agents, the network operator costs, the generation level and the multiplier of the budget constraint (which in this case is zero by definition). Period 1 represents peak demand, periods 2 and 3 have intermediate demand, and period 4 has off-peak demand. The column “Average” gives the values for an average hour over the course of the year, taking into account the length of each period.
Table 3
First best, results for a representative hour
First best
Average
Period 1 2,4%
Period 2 20.1%
Period 3 38.9
Period 4 38.6%
Length of the period
%
Welfare
(k€ per hr)
908.6
1647.5
1256.3
1003.4
586.9
Consumer surplus
(k€ per hr)
772.4
1301.8
1031.7
837.1
539.8
Producer surplus
(k€ per hr)
170.8
305.2
230.6
194.1
107.9
Profit network operator
(k€ per hr)
−34.5
40.5
−5.9
−27.8
−60.8
Revenue network operator
(k€ per hr)
39.5
114.6
68.2
46.3
13.3
Fixed cost network operator
(k€ per hr)
74.1
74.1
74.1
74.1
74.1
Multiplier budget constraint
(€ per €)
Total generation
(MWh per hr)
9108
11657
10880
9403
7734
Average price
(€ per MWh)
32.2
48.0
37.9
35.5
22.4
Minimum price
(€ per MWh)
18.5
26.1
23.2
17.6
15.4
Maximum price
(€ per MWh)
76.4
121.7
91.7
92.6
46.8
Table 4 shows aggregate values for all time periods and for the full year. Consumption in the low and the high demand period are 7734 MW and 11657 MW, respectively. Total annual production is 79791 GWh.18
Table 4
First best, aggregate results
First best
Sum
Period 1 208
Period 2 1,759
Period 3 3,410
Period 4 3,383
Length of the period
hrs
Welfare
(M€)
7959
342
2210
3422
1986
Consumer surplus
(M€)
6766
270
1815
2855
1826
Producer surplus
(M€)
1496
63
406
662
365
Profit network operator
(M€)
−303
8
−10
−95
−206
Revenue network operator
(M€)
346
24
120
158
45
Fixed cost network operator
(M€)
649
15
130
253
251
Multiplier budget constraint
(€ per k€)
Total generation
(GWh)
79790
2420
19138
32067
26165
In Table 4, the indicated prices are wholesale prices, covering generation as well as transmission. If all transmission charges would be set equal to zero, then in all periods the network capacity would be insufficient to satisfy the demand for transmission. Thus, congestion is an issue in the four periods, and network use must be charged in order to solve capacity problems. A welfare maximizing network operator will set charges such that distortions are minimized. The best way to do this is to tax the effective use of the network, but not the ‘intra-nodal’ trade, i.e. the network operator will set the price wedge equal to zero, i.e. τ ii =0 \((\tau_{i}^{c}=-\tau_{i}^{p})\). The reason for this is simple: setting a positive price wedge \(\tau_{ii}=\tau_{i}^{c}+\tau_{i}^{p}>0\), increases the distortion in the local market at node i, but only has an indirect effect on the network flows that cause the congestion. Note that this only makes sense for nodes at which both generators and consumers are connected. If not, the price wedge does not play a role.
These congestion charges allow the network operator to collect a revenue equal to €13 300 per hour in the low demand period and €114 600 per hour in the peak demand period. Aggregated over the four periods, the network operator would cover 53% of its fixed costs by charging these congestion charges. Note that hourly congestion revenue is mainly collected in the periods 1, 2 and 3. (See Table 6.) Congestion is low in period 4, so congestion revenue is rather low in that period.
The congestion charges are node specific. They depend on how much consumption in a node affects the flows on the congested lines. In fact, the network operator uses the standard nodal pricing model. As a consequence, consumers at different nodes will pay different prices for electricity. In the low demand period, prices range between €15 and €47 per MWh. In the peak period, the price range is €26 to €122 per MWh.
Note that at first sight, one would expect that at times of low demand, the network is more likely to be uncongested. This is however not necessarily the case as in periods of low demand only base-load plants are running and, therefore, the ‘average distance’ between consumption and generation nodes can increase compared to periods of peak demand. As a result the network is used more.19 In the case of a contingency, flows are rerouted to a greater extent, as there is often no generation to provide the electricity locally.

5.2 Second Best

In the second best case, the network operator is faced with a budget constraint. Congestion charges now need to be increased above their first best level in order to obtain sufficient revenue to cover the remaining 47% of the network operator’s cost.
These transmission prices will result in increased wholesale prices and thus reduced demand in all periods. Aggregate demand (and generation) decreases by 1.5% to 78,592 GWh. Prices are now in the range of €18 and €47 per MWh in off-peak hours and of €30 and €123 in peak hours.
On average, the network operator needs to collect €74 100 per hour. Table 5 shows that the network operator collects €168 100 per hour during the peak period, and €34 800 per hour during base load periods. He makes sure that marginal deadweight loss of collecting revenue is equal in all the four time periods. Collecting €1 000 of extra revenue creates a deadweight loss of €17.3.
Table 5
Second best, results for a representative hour
Second best
Average
Period 1 2,4%
Period 2 20.1%
Period 3 38.9
Period 4 38.6%
Length of the period
%
Welfare
(k€ per hr)
908.3
1647.1
1256.0
1003.0
586.7
Consumer surplus
(k€ per hr)
749.0
1270.5
1001.4
811.6
522.7
Producer surplus
(k€ per hr)
159.3
282.6
214.1
179.1
103.3
Profit network operator
(k€ per hr)
0.0
94.0
40.5
12.4
−39.3
Revenue network operator
(k€ per hr)
74.1
168.1
114.6
86.5
34.8
Fixed cost network operator
(k€ per hr)
74.1
74.1
74.1
74.1
74.1
Multiplier budget constraint
(€ per k€)
17.3
17.3
17.3
17.3
17.3
Total generation
(MWh per hr)
8971
11521
10722
9262
7612
Average price
(€ per MWh)
34.8
50.8
40.7
38.3
24.7
Minimum price
(€ per MWh)
21.7
29.7
27.0
21.0
17.9
Maximum price
(€ per MWh)
76.8
122.8
93.0
92.6
46.8
Table 6
Second best, aggregate results
Second best
Average
Period 1 208
Period 2 1,759
Period 3 3,410
Period 4 3,383
Length of the period
hrs
Welfare
(M€)
7957
342
2209
3421
1985
Consumer surplus
(M€)
6561
264
1761
2768
1768
Producer surplus
(M€)
1395
59
377
611
349
Profit network operator
(M€)
0
20
71
42
−133
Revenue network operator
(M€)
649
35
202
295
118
Fixed cost network operator
(M€)
649
15
130
253
251
Multiplier budget constraint
(€per k€)
17.3
17.3
17.3
17.3
17.3
Total generation
(GWh)
78590
2392
18861
31585
25753
The network operator increases transmission tariffs, in order to cover his costs. The solution to this problem is known as Ramsey pricing. The basic idea is that prices are increased in a way that minimizes distortions, which amounts to applying price increases that are inversely proportional to the demand elasticities.
The use of Ramsey prices has two effects. On the one hand, the higher transmission prices will decrease the total demand for transmission. On the other hand, the pattern of the flows over the network will change. In the first best, transmission quantities depend on the price levels and the marginal production costs. In the second best, transmission quantities will also depend on the demand and supply elasticities.

5.3 Strategic Behavior of Generators

The last simulation looks at how imperfect competition influences the second best model. Now, all three limitations are present: transmission constraints, market power in generation and a budget constraint for the grid operator. Market power in generation is modeled as Cournot competition. Table 7 presents the results aggregated over the four periods. The results do not come as a surprise. Ceteris paribus, less competition reduces welfare.
Table 7
Aggregate results of the 4 scenarios
Scenario
 
First best
Second best
Cournot
Welfare
(M€ per yr)
7959
7957
7310
Consumer surplus
(M€ per yr)
6766
6561
3778
Producer surplus
(M€ per yr)
1496
1395
3532
Profit network operator
(M€ per yr)
−303
0
0
Revenue network operator
(M€ per yr)
346
649
649
Fixed cost network operator
(M€ per yr)
649
649
649
Multiplier budget constraint
(€ per k€)
17.3
30.6
Total generation
(GWh)
79,790
78,590
59,691
Average price
(€ per MWh)
32.2
34.8
75.7
Minimum price
(€ per MWh)
18.5
21.7
68.7
Maximum price
(€ per MWh)
76.4
76.8
92.1
The multiplier of the budget constraint measures the net cost of giving one Euro to the network operator. The effect is about twice as large with Cournot as with perfect competition. There are two reasons for this. As total demand drops by almost 30% from perfect competition (= the second best) to the Cournot model, the network operator needs to increase the transmission tariffs with 30% in order to obtain the same revenue. At the same time, transmission prices create a larger deadweight loss, as there is already a deadweight loss due to the strategic behavior of the Cournot players.
The average price for electricity increases from €34.8 per MWh under perfect competition to €75.7 per MWh in the Cournot setting.
In an oligopoly with 3 players, welfare is about 8% lower than with perfect competition. Consumer surplus drops by 42%, producers’ surplus more than doubles.
In the Cournot model, the location of the firms in the grid might be important in determining the market power of the generators. If firms have geographically dispersed production capacities, the effect of congestion might be much smaller than when firms are geographically concentrated.
Also the ownership structure of generation capacity might affect the market outcome. If all firms own a diverse portfolio with base load and peak load plants, then each firm will have an incentive to withhold some production capacity at the margin, as it will increase the price it receives for the infra-marginal plants. In the opposite case where one firm only owns base load plants and another firm only peak load plants, there are fewer incentives to reduce production. These effects of location and ownership remain a topic for further research.

6 Conclusions

This paper looks at the socially optimal transmission prices in a congested network with imperfect competition in electricity generation and a network operator facing a binding budget constraint. It shows that generators and consumers have to pay different transmission prices at the social optimum. These differences reflect the fact that the network operator needs to collect revenues and that the generation sector is not competitive.
The model in this paper allows imperfect competition in generation, but assumes that generators are price takers in the transmission market. The network operator is a Stackelberg leader who sets transmission price before generators decide about generation and sales. The model is illustrated with three simulation runs: a first best scenario (limited transmission capacity), a second best scenario (a binding budget constraint), and a second best scenario with Cournot competition in generation. The parameterization of the model is inspired by the technical characteristics of the Belgian electricity system. It includes the Belgian high voltage transmission grid and the lines in France and the Netherlands which are important for the Belgian network. The network is presented as a linearized DC-load flow model. Transmission is limited by the thermal constraints of the lines and n−1 security constraints are imposed.
The model provides a framework that can help policy makers to design transmission tariffs. It shows how prices should be set in response to changes in market power. The model links some of the standard regulation literature on Ramsey pricing with the electricity literature on optimal pricing of transmission networks. The qualitative results that come out of the model can be very informative, but some reservation is at place if one would consider implementing such a pricing scheme. There are a number of reasons for this.
We assume that generators are price takers in the transmission market, while, in practice, generators might abuse their locational market power. Other types of models are needed if such market behavior exists. (See for example Borenstein et al. (2000).)
The model neglects entry in generation, and only derives short run optimal prices. These prices might not give the right long run incentives for investing in new generation capacity.
The network operator is assumed to have perfect information about generation costs and demand. In practice, this information is not readily available. Any mechanism to allocate transmission capacity will have to take into account this information asymmetry.
The network operator might not be maximizing welfare, but rather profits or the interests of some political pressure groups. This is the reason why the network operator is regulated and is limited in setting transmission prices. A companion paper studies the strategic behavior of the network operator (Pepermans and Willems 2006).

Open Access

This article is distributed under the terms of the Creative Commons Attribution Noncommercial License which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited.
Open AccessThis is an open access article distributed under the terms of the Creative Commons Attribution Noncommercial License (https://​creativecommons.​org/​licenses/​by-nc/​2.​0), which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited.

Unsere Produktempfehlungen

Zeitschrift für Energiewirtschaft

Zeitschrift für Energiewirtschaft (ZfE) ist eine unabhängige Fachzeitschrift mit aktuellen Themen zu Energiewirtschaft, Energiepolitik und Energierecht. JETZT BESTELLEN

Fußnoten
1
In practice such payments might take the form of must-run contracts.
 
2
This output includes the transport of electricity, but also other functions, such as the provision of reliable electricity supply (operation of reserve power and balancing markets) and of qualitative power (e.g. voltage level, frequency stability).
 
3
The fixed part often depends on the maximal usage of the network over a certain time period, or the average use of the network over a time period. As the fixed part is a function of the actual use of the network, users will take it into account when they make their decision about how much transmission they want to use. They will transport less electricity than the welfare optimum, and deadweight loss is created.
 
4
A load pocket is an area where there is insufficient transmission capability to reliably supply 100% of the electric load without relying on generation capacity that is physically located within that area. It is the result of high concentrations of intensive power use inevitable in a big city and limitations, known as constraints, on the transmission system that limit the ability of load to be served by generating resources located remotely. This definition was taken from ‘http://​www.​uspowergen.​com/​2008/​02/​27/​what-is-a-load-pocket/​.
 
5
See for instance the models of von der Fehr and Harbord (1993) for the multi-unit action approach. The supply function equilibrium concept is based on Klemperer and Meyer (1989). Holmberg et al. (2008) show that a multi-unit discrete supply function equilibrium might converge to the continuous supply function concept.
 
6
This choice is supported by an empirical study of Wolak and Patrick (2001) who suggest that Cournot competition is an appropriate representation of the electricity generation market. Willems et al. (2009) compare supply function equilibria and Cournot models and show that there are not significant differences between both models when one looks at the available market data. Other studies using Cournot competition are Borenstein et al. (1999, 2000), Borenstein and Bushnell (1999), Cardell et al. (1997), Hogan (1997), Oren (1997) and Stoft (1997).
 
7
Within the set of linear price structures, this is the most general assumption. It encompasses a number of ‘price structure’ options as special cases. For example, only charging consumption, only charging generation, a separate but uniform tariff for generation and consumption and one uniform tariff for both generation and consumption as the most extreme case.
 
8
Note that in this formulation nodal sales of the generators can become negative. Generators can act like arbitrageurs that buy electricity in one region and sell it in another. They will, however, still take into account the effect on the marginal revenue in both regions. Therefore, with a limited number of firms, not all price differences will be arbitraged away. As the number of generation firms increases, arbitrage will improve.
 
9
Such a model assumes that line resistance is small relative to reactance, that voltage magnitudes are the same at all nodes, and that voltage angles between nodes at opposite ends of a transmission line are small. Engineers often use the linearised model of the network for long term planning. See Schweppe et al. (1988).
The alternative, AC-power flow, was used in a previous version of the program, but did not give fundamentally different results.
 
10
It was noted before that B can also be interpreted as being the fraction of the transmission firms costs that needs to be collected via usage charges, i.e. net of the revenues collected via the standing charge of a two-part tariff.
 
11
Network data was kindly provided by Peter Van Roy and Konrad Purchala of the K.U.Leuven Electrical Engineering Department. More detailed information on origin and destination, voltage level, admittance, thermal capacity… is available upon request.
 
12
Some of the data was kindly provided by Leonardo Meeus and Kris Voorspools of the Departments of Electrical and Mechanical Engineering, respectively. Data was also taken from several editions of the BFE statistical yearbook and the annual report of Electrabel. Data are available from the authors upon request.
 
13
One firm received all nuclear power plants, while the remaining power plants were allocated to two competitors of more or less equal size. Both these firms received a mix of generation technologies. Joint ownership of generation plants was excluded and only one generation firm is allowed per generation node.
 
14
A better modeling of the pumped storage plants would require taking into account the capacity constraint of the water reservoir, and to make the generation and consumption decisions endogenous.
 
15
The network of one part of Luxembourg forms an integral part of the Belgian network. Demand levels for that part are included in the model here.
 
16
Note that there would be no (binding) budget constraint if the network operator would be a profit maximizer.
 
17
The own price elasticity is based on estimates found in the literature. Lijesen (2007) provides a survey of studies that estimate short term price elasticities. These estimates range from −0.071 to −1.113. Koichiro (2010) estimates the price response in the Californian energy market. He compares electricity demand of consumers in the same neighborhoods facing different type of prices because they belonged to different utilities. He obtains marginal price elasticities of −0.18 to −0.20 during the California electricity crisis in 2000, and in the range −0.135 to −0.211 for the full time period (1999 to 2006).
 
18
This is lower than the 83.4 TWh that was used to calibrate the demand functions. The reason for this is that we neglected the network constraints when we were calibrating. The network constraints reduce demand for electricity.
 
19
Of course, this depends on the location of the base-load plants in the network. If base-load plants are not distributed homogenously in the network congestion is larger.
 
Literatur
Zurück zum Zitat Barnett AH (1980) The Pigouvian tax rule under monopoly. Am Econ Rev 70(5):1037–1041 Barnett AH (1980) The Pigouvian tax rule under monopoly. Am Econ Rev 70(5):1037–1041
Zurück zum Zitat Bjorndal M (2000) Topics on electricity transmission pricing. University of Bergen, Bergen, p 153 Bjorndal M (2000) Topics on electricity transmission pricing. University of Bergen, Bergen, p 153
Zurück zum Zitat Borenstein S, Bushnell JB (1999) An empirical analysis of the potential for market power in California’s electricity industry. J Ind Econ 47(3):285–323 Borenstein S, Bushnell JB (1999) An empirical analysis of the potential for market power in California’s electricity industry. J Ind Econ 47(3):285–323
Zurück zum Zitat Borenstein S, Bushnell JB, Knittel CR (1999) Market power in electricity markets: beyond concentration measures. Energy J 20(4):65–88 Borenstein S, Bushnell JB, Knittel CR (1999) Market power in electricity markets: beyond concentration measures. Energy J 20(4):65–88
Zurück zum Zitat Borenstein S, Bushnell JB, Stoft S (2000) The competitive effects of transmission capacity in a deregulated electricity industry. Rand J Econ 31(2):294–325 CrossRef Borenstein S, Bushnell JB, Stoft S (2000) The competitive effects of transmission capacity in a deregulated electricity industry. Rand J Econ 31(2):294–325 CrossRef
Zurück zum Zitat Buchanan JM (1969) External diseconomies, corrective taxes, and market structure. Am Econ Rev 59(1):174 Buchanan JM (1969) External diseconomies, corrective taxes, and market structure. Am Econ Rev 59(1):174
Zurück zum Zitat Cardell JB, Hitt CC, Hogan WW (1997) Market power and strategic interaction in electricity networks. Resour Energy Econ 19(1–2):109–137 CrossRef Cardell JB, Hitt CC, Hogan WW (1997) Market power and strategic interaction in electricity networks. Resour Energy Econ 19(1–2):109–137 CrossRef
Zurück zum Zitat Cossent R, Gómez T, Frías P (2009) Towards a future with large penetration of distributed generation: is the current regulation of electricity distribution ready? Regulatory recommendations under a European perspective. Energy Policy 37(3):1145–1155 Cossent R, Gómez T, Frías P (2009) Towards a future with large penetration of distributed generation: is the current regulation of electricity distribution ready? Regulatory recommendations under a European perspective. Energy Policy 37(3):1145–1155
Zurück zum Zitat Daxhelet O, Smeers Y (2005) Inter-Tso compensation mechanism, p 15 Daxhelet O, Smeers Y (2005) Inter-Tso compensation mechanism, p 15
Zurück zum Zitat Dijk J, Petropoulos G, Willems B (2009) Generation investment and access regulation in the electricity market: a real option approach, p 22 Dijk J, Petropoulos G, Willems B (2009) Generation investment and access regulation in the electricity market: a real option approach, p 22
Zurück zum Zitat Ehrenmann A (2004) Equilibrium problems with equilibrium constraints and their application to electricity markets. Fitzwilliam College, Fitzwilliam, p 154 Ehrenmann A (2004) Equilibrium problems with equilibrium constraints and their application to electricity markets. Fitzwilliam College, Fitzwilliam, p 154
Zurück zum Zitat Fletcher R, Leyffer S (2002) Numerical experience with solving Mpecs as Nlps (No NA/210) Fletcher R, Leyffer S (2002) Numerical experience with solving Mpecs as Nlps (No NA/210)
Zurück zum Zitat Gabriel SA, Leuthold FU (2010) Solving discretely-constrained Mpec problems with applications in electric power markets. Energy Econ 32(1):3–14 CrossRef Gabriel SA, Leuthold FU (2010) Solving discretely-constrained Mpec problems with applications in electric power markets. Energy Econ 32(1):3–14 CrossRef
Zurück zum Zitat Green R (2004) Electricity transmission pricing: how much does it cost to get it wrong? Cambridge, p 25 Green R (2004) Electricity transmission pricing: how much does it cost to get it wrong? Cambridge, p 25
Zurück zum Zitat Hobbs BF, Metzler CB, Pang JS (2000) Strategic gaming analysis for electric power systems: an mpec approach. IEEE Trans Power Syst 15(2):638–645 CrossRef Hobbs BF, Metzler CB, Pang JS (2000) Strategic gaming analysis for electric power systems: an mpec approach. IEEE Trans Power Syst 15(2):638–645 CrossRef
Zurück zum Zitat Hogan WW (1997) A market power model with strategic interaction in electricity networks. Energy J 18(4):107–141 Hogan WW (1997) A market power model with strategic interaction in electricity networks. Energy J 18(4):107–141
Zurück zum Zitat Holmberg P, Newbery D, Ralph D (2008) Supply function equilibria: step functions and continuous representations (No 0863). Cambridge Holmberg P, Newbery D, Ralph D (2008) Supply function equilibria: step functions and continuous representations (No 0863). Cambridge
Zurück zum Zitat Hsu M (1997) An introduction to the pricing of electric power transmission. Util Policy 6(3):257–270 CrossRef Hsu M (1997) An introduction to the pricing of electric power transmission. Util Policy 6(3):257–270 CrossRef
Zurück zum Zitat IEA (2006) World energy outlook 2006. Paris, p 601 IEA (2006) World energy outlook 2006. Paris, p 601
Zurück zum Zitat Joskow P, Tirole J (2005) Merchant transmission investment. J Ind Econ 53(2):233–264 CrossRef Joskow P, Tirole J (2005) Merchant transmission investment. J Ind Econ 53(2):233–264 CrossRef
Zurück zum Zitat Knight RC, Montez JP, Knecht F, Bouquet T (2005) Distribution generation connection charging within the European Union—review of current practises. In: Future options and European policy recommendations, p 73 Knight RC, Montez JP, Knecht F, Bouquet T (2005) Distribution generation connection charging within the European Union—review of current practises. In: Future options and European policy recommendations, p 73
Zurück zum Zitat Koichiro I (2010) How do consumers respond to nonlinear pricing? Evidence from household electricity demand. In: 15th power conference. University of California Energy Institute, Berkeley Koichiro I (2010) How do consumers respond to nonlinear pricing? Evidence from household electricity demand. In: 15th power conference. University of California Energy Institute, Berkeley
Zurück zum Zitat Lijesen MG (2007) The real-time price elasticity of electricity. Energy Econ 29(2):249–258 CrossRef Lijesen MG (2007) The real-time price elasticity of electricity. Energy Econ 29(2):249–258 CrossRef
Zurück zum Zitat Luo ZQ, Pang JS, Ralph D (1996) Mathematical programs with equilibrium constraints. Cambridge University Press, New York Luo ZQ, Pang JS, Ralph D (1996) Mathematical programs with equilibrium constraints. Cambridge University Press, New York
Zurück zum Zitat Neuhoff K, Barquin J, Boots MG, Ehrenmann A, Hobbs BF, Rijkers FAM, Vazquez M (2005) Network-constrained Cournot models of liberalized electricity markets: the devil is in the details. Energy Econ 27(3):495–525 CrossRef Neuhoff K, Barquin J, Boots MG, Ehrenmann A, Hobbs BF, Rijkers FAM, Vazquez M (2005) Network-constrained Cournot models of liberalized electricity markets: the devil is in the details. Energy Econ 27(3):495–525 CrossRef
Zurück zum Zitat Olmos Camacho L, Pérez-Arriaga IJ (2007) Comparison of several Inter-Tso compensation methods in the context of the internal electricity market of the European Union. Energy Policy 35(4):2379–2389 CrossRef Olmos Camacho L, Pérez-Arriaga IJ (2007) Comparison of several Inter-Tso compensation methods in the context of the internal electricity market of the European Union. Energy Policy 35(4):2379–2389 CrossRef
Zurück zum Zitat Oren SS (1997) Economic inefficiency of passive transmission rights in congested electricity systems with competitive generation. Energy J 18(1):63–83 Oren SS (1997) Economic inefficiency of passive transmission rights in congested electricity systems with competitive generation. Energy J 18(1):63–83
Zurück zum Zitat Pepermans G, Willems B (2006) Network unbundling, ownership structure and oligopolies: a case study for the Belgian electricity market. Compet Regul Netw Ind 1(2):231–262 Pepermans G, Willems B (2006) Network unbundling, ownership structure and oligopolies: a case study for the Belgian electricity market. Compet Regul Netw Ind 1(2):231–262
Zurück zum Zitat Ramsey FP (1927) A contribution to the theory of taxation. Econ J 37:47–61 CrossRef Ramsey FP (1927) A contribution to the theory of taxation. Econ J 37:47–61 CrossRef
Zurück zum Zitat Schweppe FC, Caramanis MC, Tabors RD, Bohn R (1988) Spot pricing of electricity. Dordrecht, Kluwer, p 355 Schweppe FC, Caramanis MC, Tabors RD, Bohn R (1988) Spot pricing of electricity. Dordrecht, Kluwer, p 355
Zurück zum Zitat Smeers Y, Wei J-Y (1997) Spatially oligopolistic model with opportunity cost pricing for transmission capacity reservations—a variational approach (No 9717). Louvain-La-Neuve, p 26 Smeers Y, Wei J-Y (1997) Spatially oligopolistic model with opportunity cost pricing for transmission capacity reservations—a variational approach (No 9717). Louvain-La-Neuve, p 26
Zurück zum Zitat Stoft SE (1997) The effect of the transmission grid on market power (No mimeo). Berkley Stoft SE (1997) The effect of the transmission grid on market power (No mimeo). Berkley
Zurück zum Zitat Swider DJ, Beurskens L, Davidson S, Twidell J, Pyrko J, Prüggler W, Auer H, Vertin K, Skema R (2008) Conditions and costs for renewables electricity grid connection: examples in Europe. Renew Energy 33(8):1832–1842 CrossRef Swider DJ, Beurskens L, Davidson S, Twidell J, Pyrko J, Prüggler W, Auer H, Vertin K, Skema R (2008) Conditions and costs for renewables electricity grid connection: examples in Europe. Renew Energy 33(8):1832–1842 CrossRef
Zurück zum Zitat Van Roy P (2001) Study of the technical economical aspects of the liberalisation of the electricity market. PhD dissertation, K.U. Leuven. ISBN 90-5682-325-6 Van Roy P (2001) Study of the technical economical aspects of the liberalisation of the electricity market. PhD dissertation, K.U. Leuven. ISBN 90-5682-325-6
Zurück zum Zitat von der Fehr N, Harbord D (1993) Spot market competition in the UK electricity industry. Econ J 103:531–546 CrossRef von der Fehr N, Harbord D (1993) Spot market competition in the UK electricity industry. Econ J 103:531–546 CrossRef
Zurück zum Zitat Wei J-Y, Smeers Y (1999) Spatial oligopolistic electricity models with Cournot generators and regulated transmission prices. Oper Res 47(1):102–112 MATHCrossRef Wei J-Y, Smeers Y (1999) Spatial oligopolistic electricity models with Cournot generators and regulated transmission prices. Oper Res 47(1):102–112 MATHCrossRef
Zurück zum Zitat Willems B, Rumiantseva I, Weigt H (2009) Cournot versus supply functions: what does the data tell us? Energy Econ 31(1):38–47 CrossRef Willems B, Rumiantseva I, Weigt H (2009) Cournot versus supply functions: what does the data tell us? Energy Econ 31(1):38–47 CrossRef
Zurück zum Zitat Wolak FA, Patrick R (2001) The impact of market rules and market structure on the price determination process in England and Wales electricity market. No NBER Working Paper 8248, Stanford, p 88 Wolak FA, Patrick R (2001) The impact of market rules and market structure on the price determination process in England and Wales electricity market. No NBER Working Paper 8248, Stanford, p 88
Metadaten
Titel
Cost Recovery in Congested Electricity Networks
verfasst von
Guido Pepermans
Bert Willems
Publikationsdatum
01.09.2010
Verlag
Vieweg Verlag
Erschienen in
Zeitschrift für Energiewirtschaft / Ausgabe 3/2010
Print ISSN: 0343-5377
Elektronische ISSN: 1866-2765
DOI
https://doi.org/10.1007/s12398-010-0021-1

Weitere Artikel der Ausgabe 3/2010

Zeitschrift für Energiewirtschaft 3/2010 Zur Ausgabe

Veranstaltungskalender

Veranstaltungen 2010