...

Inverse Optimization for the Recovery of Market Structure from Market Outcomes:

by user

on
Category: Documents
50

views

Report

Comments

Transcript

Inverse Optimization for the Recovery of Market Structure from Market Outcomes:
Inverse Optimization for the Recovery of Market Structure
from Market Outcomes:
An Application to the MISO Electricity Market
Draft Version
John Birge, Ali Hortaçsu, & Michael Pavlin
June 26, 2014
Abstract
We propose an inverse optimization based methodology to determine market structure from the
locational pricing of a commodity. The methodology requires that the market optimally allocates goods
and that locational prices correspond to shadow prices of this optimization problem. As a case-in-point,
we study locational marginal price based electricity markets where prices are determined using the results
of a centralized optimization for clearing the market. We apply the inverse optimization methodology
to outcome data from the Midcontinent ISO electricity market and uncover transmission constraints
explaining the price variation. We also discuss analytical applications such as identifying missing data
and the residual demand derivative. To broaden the scope of applications, assumptions sucient to
justify the methodology for competitive markets are described.
1
Introduction
In electricity and many other commodity markets, price varies substantially depending on location. For
example, during the month of October 2012 in the Midcontinent ISO (MISO) day-ahead electricity market,
the average hourly price was $25.50 but the average hourly dierence between the high and low price was
$81.28. The two panels in Figure 1 illustrate both the price variation at a particular time (histogram in
Panel (a)) and the price volatility over time for each market participant as dierent active transmission
constraints change the market conditions (time-series plot in Panel(b)). Analyses of these markets would
benet from a structural model explaining price formation. However, building this model requires a precise
1
400
10
200
100
Frequency
300
distance from hourly mean ($)
5
unit
u3726
u3956
0
u6661
u4424
u6766
0
−5
−20
0
20
40
60
−10
80
0
Day Ahead LMP 2012−10−13 3:00:00
(a) Histogram of node prices.
25
50
75
100
hours since 2012−10−01 12:00:00
(b) Time-series of deviation of price received from mean price
for 5 generators.
Figure 1: Price variation in Midcontinent ISO day-ahead market.
model of the frictions generating these pricing disparities. In particular, to understand how electricity prices
are formed, we need to know how local generation contributes to congestion throughout the market footprint.
These operational details are typically private information.1 However, accurate and detailed pricing data
is often readily available. In the case of electricity markets which use the dominant locational marginal
pricing scheme, the allocation and prices correspond to the value of primal and dual variables in the economic
dispatch problem. The economic dispatch problem is typically a linear program which determines the optimal
allocation of production and consumption given market bids. The correspondance between market outcomes
and the optimal solution of a partially specied optimization program promotes an inverse optimization
based methodology. We propose a novel inverse optimization based recovery algorithm where unobservable
market structure, which corresponds to parameters of the feasible region of the economic dispatch problem,
is uncovered using observed prices and allocations.
Inverse optimization may be dened as retrieving parameters of incompletely specied mathematical
programming formulation from a partially observable optimal solution. The standard inverse optimization
problem discussed in Zhang and Liu [1996] and elsewhere is to retrieve unknown parameters of the objective
function given an optimal solution. Our problem is novel; we are searching for parameters dening the
feasible space of possible allocations. Using knowledge of the economic dispatch problem, we are able to
1 In cases such as electricity and natural gas, detailed information is unavailable due to the Critical Infrastructure Information
Act of 2002.
2
prespecify a particular form to the optimization; we also take advantage of a partial solution to the dual
problem. Like Zhang and Liu [1996], we leverage necessary conditions for optimality to rst constrain the
missing parameters and motivate an algorithm to uncover them.
We present the methodology in the context of an electricity market using locational marginal prices
(LMP) and present a numerical example. An LMP is a price at a particular point in the network and may
be dened as the cost of purveying one additional megawatt/hour (MWh) of energy at that particular point.
LMPs reect behavior of neighboring and far ung participants in the network. LMP based market design
has largely come to dominate the distribution of electricity in North America where about 60% of total
production is in markets using LMPs [Sahni et al., 2012]. The testing of our methodology will focus on the
Midcontinent Independent System Operator (MISO) which is a LMP based market spanning eleven states
and the Canadian province of Manitoba. Other LMP based markets include the California CAISO market,
the Texas ERCOT market, and the PJM Interconnection, which is the largest North American market and
serves over 60 million customers along the eastern seaboard.
From an analytical standpoint there are several benets to acquiring a structural model for these important markets. First, estimation of hidden variables such as local market power may be performed without
the endogeneity problems which plague standard econometric techniques. Second, counterfactuals may be
evaluated. For instance, an accurate model of the pricing mechanism in an electricity market allows estimation of the value of new production capacity investment which takes into account the impact of market
power on revenues. An ancillary benet of the inverse optimization methodology is that identication of the
market structure can be performed with minimal dataidentication requires only one more data sample
than there are active constraints.
While we develop this methodology in the context of electricity markets, there is the potential to apply
inverse optimization based market structure recovery whenever two structural assumptions hold: the market
results in an ecient allocation and pricing corresponds to the value of dual variables of the ecient allocation
problem. We discuss in Section 5.1 how this methodology may be extended to competitive markets for
commodities other than electricity when capacity issues are the root cause of spatial price variation. These
markets may include oil and gas and other commodities where transportation channels are inexible. In
such markets, transactions can often not be completely observed by regulators and the methodology may be
useful to regulators as well as market participants. The methodology can provide a means for regulators to
make inferences about market conditions and to perform counterfactual policy analysis.
As a proof of concept, we apply the inverse optimization methodology to real data from the MISO dayahead electricity market. We take advantage of high granularity data which includes precise prices and
allocations throughout the market footprint as well as prices for transmission lines to derive the market
3
structure. Using the derived market structure and also published bid functions for all market participants,
prices and allocations may be predicted and compared to the real market outcomes.
The results show that the methodology is eective at recovering hidden parameters determining how
transmission resources are utilized and the relative losses from dierent areas in the market. When these
parameters are used to reconstruct prices, they are very eective in explaining price variation. In many cases
we nd near perfect correlation between predicted and actual prices. However, the methodology results in
somewhat larger absolute price error of about 12%. These errors appear to be systemic eects likely due to
limitations in the data, such as missing supply and demand data. For instance, import and export data is
not available at high granularity. In Section 5.3 we discuss how the methodology may be leveraged as an
analytical tool for detecting and identifying unobservable supply and demand.
From an operational point of view the most important benet of acquiring this structural information
is likely the ability to evaluate local market power. Standard market concentration indices such as the
Herndahl-Hirschman Index are of limited value in electricity markets because transmission constraints
partially isolate submarkets resulting in frequent changes in market structure [c.f. Karthikeyan et al., 2013,
Borenstein and Bushnell, 1999]. A more principled approach, outlined in Lee et al. [2010] and Lee et al. [2011]
may be of more value to both participating rm and market observer alike. These papers propose using
the transmission constrained residual demand derivative which provides a precise picture of the incentives of
individual rms to deviate from competitive market bids into the electricity market. An ecient calculation
of this quantity for a lossless electricity market is given in Xu and Baldick [2007] and relies on bids and the
coecients describing the contributions of market participants to active congestion constraints. We adapt
this calculation to a market with transmission line losses in Section 5 and show substantial variation in the
distribution of market power in real empirical data from the MISO market.
Following a discussion of the relationship of this work to the operations research, energy, and economics
literatures, we present the electricity market model and ecient allocation problem (Section 2). In Section
3 we present the recovery methodology. We then discuss the application of this methodology to electricity
markets and the MISO market in particular in Section 4. In Section 5 we consider the broader context by
discussing missing data, competitive markets, and several applications. First, we show that the methodology
may be applied to competitive markets with capacitated shared resources. The analytical applications
include demonstrating a simple methodology for uncovering unobservable supply and discussion of how the
structural model can be used to evaluate market power and the value of capacity investments.
4
1.1
Literature and Positioning
This paper promotes the analysis of spatial price variation using tools from optimization theory. We provide
a specic application to electricity markets which use locational marginal prices. Locational marginal prices
assign a price at each location in the market which corresponds to the marginal cost of supplying that
particular location. Hogan et al. [1996] rst proposed the use of mathematical programming to determine
electricity prices; since then, locational marginal prices have come to dominate North American electricity
markets. While there is some evidence that the locational prices are succeeding at improving market eciency
[c.f. Wolak, 2011], there are potential drawbacks such as increased market power as discussed in Cardell
et al. [1997]. As identied by Sahni et al. [2012], locational priced markets also present much higher analytical
hurdles both for participants and observers which a methodology like ours may contribute to mitigating.
The methodology we introduce to uncover the market structure is a type of inverse optimization. Inverse
optimization algorithms uncover unknown parameters of a mathematical programming formulation from
observations of the optimal solution. Since the important paper of Zhang and Liu [1996], a breadth of research
has focussed on the following inference problem: given an observable feasible space and an optimal solution
to the optimization problem, determine unobservable parameters of the objective function. This problem has
been addressed for both linear programming problems [Yang and Zhang, 2007, Ahuja and Orlin, 2001, Zhang
and Liu, 1996, 1999] as well as various non-linear mathematical programming paradigms [Schaefer, 2009,
Wang, 2009, Iyengar and Kang, 2005]. Our problem diers particularly in that the unobservable parameters
of interest dene the feasible region and we assume that relevant parts of both the primal (allocation) and
dual (prices) solutions are observable. In the standard inverse optimization problem, the solution is not fully
identied. The common practice is to search for norm minimizing values for the missing parameters. In our
case, the solution is fully identied and with sucient data may be determined by solving a system of linear
equations.
Several clear themes link this work to studies in the economic literature which examine failure of the `law
of one price' in regional trade. Pioneered by the works of Engel and Rogers [1996] and [Mccallum, 1995],
these studies have focussed on the goal of understanding how distance and border eects contribute to price
dierences for physically identical goods. Recent treatments of the problem have incorporated sophisticated
models of market power [Cosar et al., 2012] and have tested these results with non-aggregated microdata
[Broda and Weinstein, 2008]. The work is characterized by a common underlying model of the contribution
of distribution infrastructure to variation in prices: transportation frictions (i.e., borders and distance) are
separable, observable, and contribute to price variation in a manner which is consistent across samples.
This model may be appropriate for prices of hard goods such as wind turbines as studied in Cosar et al.
5
[2012]. However, the electricity produced by these turbines diers in a fundamental manner: the preferred
and lowest cost distribution method is via a capacitated technology. Adding new transmission lines requires
fullling time consuming regulatory requirements and providing large capital investments. Capacity issues
which make it more dicult to close supply-demand gaps are compounded by seasonal demand uctuations
and the inability to relocate production. Natural resources are tied to a specic geography. This applies to
natural gas and oil as well as renewable electricity fuels such as hydro and wind.
Which transportation link is at capacity varies sample to sample; as a result, these frictions contribute
in dierent manners to the overall price. In electricity markets, the variation in eects is particularly acute.
For example in the MISO market, in 2012, 1295 unique transmission linesthe relevant resourcewere at
capacity. Of these 89% were at capacity in no more than 1% of the hours and no constraint was active in
more than 30% of the hours. Ecient use of data is important for deriving estimates from multiple market
states. Our results show that identication of the pricing mechanism requires only one more sample than
there are links at capacity.
In natural gas and other energy markets, theoretical models have been proposed which incorporate
capacitated technologies into equilibrium models as explanatory of geographic price dierences [Mudrageda
and Murphy, 2008, Cremer et al., 2003]. We believe that in addition to applications to electricity markets,
this paper may be taken more generally as an extention of this line of theoretical research equating price
variation to market wide capacity constraints. In Section 5 we discuss how our methodology may be applied
to the inference of market frictions when a centralized institutional structure is lacking.
2
Model and Preliminary Analysis
We focus on a one-shot electricity market facilitated by a centralized market maker. In North American electricity markets, the facilitator would be either an independent service operator (ISO) or regional transmission
organization (RTO). The model is appropriate to either forward (day-ahead) or spot (real-time) energy markets. In both of these markets, the participants submit bids while the facilitator solves an ecient allocation
problem under the assumption that these bids accurately reect the participants' preferences, i.e., if the bids
are honest, the facilitator maximizes social welfare. A bid from a participant consists of a supply function
and bounds on the interval they are willing to supply. The facilitator also takes into account how nodes are
connected by transmission resources, the capacity of these transmission resources, and other characteristics
determining how electricity ows through the network.
The market features a set of generators. Load serving entities are located at buses connected by transmission lines. Generators and load serving entities correspond to the set of market participants; buses correspond
6
to locations. The standard model of line load used by electricity markets is a linear approximation of the
true AC power ows referred to as the DC model. Details of the linearization are discussed in Appendix
A. In the linear model, the net production at a particular bus will have a proportional aect on ow over
a particular line. In eect, the transmission lines are shared transmission resources whose utilizations are
determined by locational parameters. Other than increasing and decreasing production and consumption at
particular locations in the network, the facilitator and participants are not able to inuence the routing of
electricity as they might be able in other network settings such as a data network.
2.1
Market Model
The initial model we present is specialized to electricty markets. In Section 5.1, we discuss application of
the model and methodology to other commodity markets. The electricity market model consists of a set of
participants who produce or consume electricity which is distributed through shared transmission resources.
Participants:
The market features a set of participants Q = {1..Q} where participant j produces xj MWh
of electricty. Consumption is interpreted as negative production, i.e., xj < 0.
Transmission Resources:
The market is connected via a set of transmission links, R = {1..R}. Each
of these links may be interpreted as a shared resource with a xed capacity. Link r has capacity dr . This
endowment is utilized heterogenously by participants when they produce or consume electricity.
Each participant produces or consumes electricity at one of a set of possible locations, L = {1..T }.
Variable bi denotes the net production at location i. The location determines the rate at which a participant
utilizes each resource. In particular, for each location i and resource r, there is a parameter Dri , such that, if
bi units are produced at location i, then Dri bi units of resource r are consumed. Allowing b to be the vector
of net production for each location, b is feasible if Db ≤ d. We let l(j) be the location of participant j and
P
q(i) be the set of participants operating at location i. So, net production at location i is bi = j∈q(i) xj . In
the context of an electricity market, a location would be equivalent to an electricity bus though alternative
interpretations are possible in other settings.
Transmission lines are imperfect and some proportion of production is lost in the form of heat. We
use an ane model to accomodate this line loss which associates a parameter δi equal to the proportion of
production lost from each location. This notion of loss is compatible with the MISO electricity market where
a similar loss factor is associated with each producer [Ma et al., 2009a]. The total system production must
P
P
equal total consumption plus losses:
i∈L bi +
i∈L δi bi = 0 . Note that participants that are co-located
may trade electricity amongst themselves without loss or impact of transmission capacities; however, positive
7
or negative net production at each location may impact loss from the network and consumes transmission
resources.
Market Mechanism:
Production and consumption bids are described by a real valued supply function Sj
to each participant j . The function Sj is dened over an interval [`j , uj ], where Sj (xj ) is the announced total
cost to participant j of producing xj MWh. In the case of consumption, xj < 0, Sj (xj ) is the announced
negative of the value to participant j of consuming −xj MWh. While participants will typically behave as
either producer `j ≥ 0 or consumer uj ≤ 0, there is nothing which precludes participants from behaving as
both. For instance, the owner of a partially charged battery operating in an electricity market may wish to
P
either buy or sell electricity depending on price. The total system welfare is − j∈Q Sj (xj ).
The facilitator solves the ecient allocation program which determines the ecient production and consumption from each participant under the assumption that bids are an accurate reection of true preferences:
min
x
X
(1)
Sj (xj )
j∈Q
s.t bi =
X
∀i ∈ L,
xj
(2)
j∈q(i)
X
(1 + δi )bi = 0,
(3)
i∈L
Db ≤ d,
(4)
x ≥ `,
(5)
x ≤ u.
(6)
In the above model, prices may be assigned to each location. The ecient allocation program has linear
constraints but typically also has quadratic relationships and integer variables due to non-linear supply
functions. The linearized version that we focus on here is, however, still used for computing prices as the
marginal impacts from changing constraints.
2.2
Preliminary Analysis:
Program 1-6 implies a series of optimality conditions gleaned from standard constrained optimization theory
(see, e.g., Karush-Kuhn-Tucker conditions in Nocedal and Wright [1999]). Consider dual variables π, φ, ρ, α, γ
corresponding to constraints 2-6. We refer to the vector −π as the locational prices and the vector −ρ as
the resource prices. Then, the necessary conditions for optimality are:
8
Stationarity:
X
Dki ρk + πi + (1 + δi )φ = 0,
Primal feasibility:
∀i ∈ B,
X
bi =
k
xj , ∀i ∈ L.,
j∈q(i)
(7)
−πl(j) + α − γ = S 0 (xj ),
(9)
X
j ∈ P.
(8)
Dual feasibility:
ρ, α, γ ≤ 0.
(1 + δi )bi = 0,
(10)
Db ≤ d,
(11)
x ≥ `,
(12)
x ≤ u.
(13)
i∈L
Complementary slackness:
(14)
ρ(Db − d) = 0,
(15)
γ(x − `) = 0,
(16)
α(u − x) = 0.
(17)
Note that these are necessary conditions for optimality under very general continuity condtions [c.f. Nocedal
and Wright, 1999] and will hold even with piecewise and non-convex supply functions with the corresponding
linearization.
With the supply equilibrium constraint in Equation 3, the program is overspecied. To avoid this problem,
it is standard in electrical engineering to select a reference location r which is assumed to not aect ow on
the lines. The rth column of matrix D, which contains the parameters for the utilization of each resource,
corresponds to the reference location and is a zero vector; the parameters for other locations will compensate.
These is a dierent matrix D for each reference location though the optimization problem remains equivalent.
The following observations allows the unobserved dual variable φ to be removed.
Proposition 1.
In an optimal solution of Program 1,D and φ have the following properties:
1. Buses with zero coecients in all active D constraints have the same lmp.
2. φ is equal to the lmp of the reference location corrected by the loss.
φ=
9
−πr
.
(1 + δr )
(18)
3
Recovery of Market Structure from Outcomes
In this section, we discuss the identication of utilization parameters for resources which are scarce in at
least one data sample, i.e., the link is at full capacity and has a non-zero resource price. Since resource
prices are identically equal to zero for any transmission resource with surplus capacity, parameters for these
resources cannot be identied; hence, we exclude these resources from the recovery procedure. We refer to
resources with a positive resource price as active resources. By convention, each row of D corresponds to a
resource which is active in at least one sample.
Consider a set of noise-free data samples S = {1..s}, where a sample j = {−π j , −ρj , x} species observed
commodity, resource prices, and allocations. Let K be the set of active resources in the data samples. The
identity φ =
−πr
(1+δr )
from Proposition1 allows φ to be replaced in Equation 7. Rather than determining δi
, the relative productivity of a location i to the reference location, qi =
1+δi
1+δr ,
will be explored. With the
relative productivity vector q , an equivalent program to the Ecient Allocation program can be formulated
P
by replacing the constraint i∈L (1 + δi )bi = 0 with the equation:
X
(19)
qi bi = 0.
i∈L
The following theorem details how and when the unknown parameters are identied.
Theorem 2.
For an arbitrarily selected reference location r, consider equations:
X
Dki ρjk = qi πrj − πij
, ∀i ∈ L, j ∈ S,
(20)
∀i ∈ L, j ∈ S,
(21)
k∈K
X
xjp = bji ,
p∈q(i)
X
(22)
qi bi = 0;
i∈L
then, D and q are identied if the above system of equations has rank (|K| + 1)|L|. The remaining unknown
parameters of the primal program, d, may be recovered from Dk bj = dk .
Proof. Equation 20 is simply Equation 7 updated with φ = qi πr as stated in Proposition 1. The system of
equations contains (|K| + 1)|L| variables which are the elements of matrix D and q , requiring the system to
have this rank for a unique solution. Dk bj = dk follows directly from the primal formulation. Dk is identied
as just stated, and bj is specied from data allowing d to be identied.
10
Corollary 3.
Consider a set of resource constraints K active in a set of samples S . Then, consider the
bipartite constraint graph G dened by nodes K ∪ S and edges (k, s) if constraint k is active in sample s.
The coecients D for constraints K and the relative productivities q are identied if and only if for each
connected component in G with constraints Kc and samples S c , the rank of
X
Dki ρjk = qi πrj − πij ,
∀i ∈ L, j ∈ S c ,
(23)
k∈K
is greater or equal to |Kc ||L| and, for one connected component, the rank of these equations is greater or
equal to (|Kc | + 1)|L|.
Estimation:
Error may be incorporated into this model with the additive residual for each of Equations
20. Assuming a normal error distribution and that each equation has error with zero mean and common
variance, the least-squares solution is an appropriate estimator when the problem is overspecied (rank
is greater than (|K| + 1)|L|). We use this single step estimator in the application section. Through the
remainder of the text, we refer to this estimation procedure as the IO recovery algorithm.
3.1
Numerical Example
The example is illustrated in Figure 2. There are four locations (buses), L = {1..4}, connected into a single
cycle by lines {1 − 2, 2 − 3, 3 − 4, 1 − 4}. We assume no transmission loss (i.e., q = [1, 1, 1, 1]). Flow may
go in either direction on these links (i.e., y may be negative). The links have reactance parameters z = 1
except for the link 1-4 where z = 2. The topology and reactances contribute to the matrix D. The precise
derivation is described in Appendix A. There is a single capacity constraint on line 2-3 which limits the
absolute ow to 50MW.
The positive and negative ows over lines 1 − 2, 2 − 3, 3 − 4, and 1 − 4 are respectively in rows 1,5; 2,6;
3,7 and 4,8 with d containing the corresponding limits. As discussed, only capacity constraints that are
active in a sample are relevant; so, we may ignore resources with innite capacity. In this market, there is
a single price-taking demand participant at Bus 4 requiring 300MW. This may be supplied from a low price
producer at Location 1 or a high price producer at Location 3.
Location is important in determining resource utilizations. If 100 additional MW are requested at Location 4, depending on whether it is supplied from Location 1 or Location 3, ow over the transmission line 2-3
will either increase by 40MW or decrease by 20MW. The following shows that this can dramatically aect
prices throughout the network.
Table 1, describes two bidding scenario which are identical except for demand at Location 4. Scenario 1
11
Participant 1
Min:0 MW
Max:500 MW
Cost:$20/MWh
z=1
1
2
Limit: 50MW
z=2
Participant 4
Min:-300 MW
Max:-300 MW
Participant 2
Min:0 MW
Max:0 MW
z=1
z=1
4
3
Participant 3
Min:0 MW
Max:500 MW
Cost:$30/MWh
Figure 2: Electricity Network in Numerical Example for Sample 2.
Participant
1
3
4
Scenario 1
S1 (x1 ) = 20x1 , `1 = 0, u1 = 500
S3 (x3 ) = 30x3 , `3 = 0, u3 = 500
S4 (x4 ) = 0, `3 = 100, u3 = 100
Scenario 2
S1 (x1 ) = 20x1 , `1 = 0, u1 = 500
S3 (x3 ) = 30x3 , `3 = 0, u3 = 500
S4 (x4 ) = 0, `3 = 300, u3 = 300
Table 1: Bidding scenarios for the numerical example
is the low demand scenario; Scenario 2 is the high demand scenario. We will illustrate the ecient allocation
program and use the outputs as data samples to demonstrate the IO recovery algorithm.
The optimal ow problem for Scenario 1 is:
max20x1 + 30x2
(24)
s.t. x = b,
(25)
b1 + b2 + b3 + b4 = 0,
(26)
0.4b1 + 0.6b2 − 0.2b3 ≤ 50,
(27)
−0.4b1 − 0.6b2 + 0.2b3 ≤ 50,
(28)
x ≥ [0, 0, 0, −100],
(29)
x ≤ [500, 0, 500, −100].
(30)
12
Solving this program results in:
x1 =[100, 0, 0, −100],
−π 1 =[20, 20, 20, 20],
−ρ1 =[0],
φ1 =20.
In Scenario 2, the production bounds in ecient allocation program change to:
x ≥ [0, 0, 0, −300],
(31)
x ≤ [500, 0, 500, −300].
(32)
Solving the program for Scenario 2 results in:
x2 =[183.3, 0, 116.7, −300],
−π 2 =[20, 16.7, 30, 26.7],
−ρ2 =[16.7],
φ2 =26.7.
In Scenario 2, transmission resource 2-3 is at capacity resulting in price variation throughout the network.
φ2 is equal to the price at bus 4, the reference bus, in this lossless system.
Switching to the IO recovery algorithm, assume that there are two data samples (x1 , −π 1 , −ρ1 ) and
(x2 , −π 2 , −ρ2 ).
13
Equations 7 provide:
−D11 20 − 20 + φ1 = 0,
−D12 20 − 20 + φ1 = 0,
−D13 20 − 20 + φ1 = 0,
−D14 20 − 20 + φ1 = 0,
−D11 16.7 − 20 + φ2 = 0,
−D12 16.7 − 16.7 + φ2 = 0,
−D13 16.7 − 30 + φ2 = 0,
−D14 16.7 − 26.7 + φ2 = 0.
The next step is to select the reference location and substitute −qi πrs for φs in each constraint after which
the system of equations can be solved. Selecting r = 4 recovers the original D = [0.4, 0.6, −0.2, 0] and
q = [1, 1, 1, 1]. An alternative selection such as r = 1 results in D = [0, 0.2, −0.6, −0.4] which is equivalent
P
once b∈L b = 0 is taken into account.
4
Application to the MISO Electricity Market
MISO provides day-ahead energy, real-time energy and ancillary services markets for 11 American states and
a Canadian province. We perform a case study applying the IO recovery algorithm to the MISO day-ahead
energy market in March of 2011. This market uses locational marginal pricing and ts the model described
in Section 2.1.
In the day-ahead energy market participants include generators (supply), load-serving entities (demand),
and strictly nancial players who participate on both sides of the market. We will continue to refer to
market submissions on both sides of the market as bids though MISO refers to supply and demand bids
respectively as oers and bids. Regardless of type, participants are treated simlarly and submit marginal
cost functions as non-decreasing piecewise linear functions. In the case of generators, these may have up to
10 regions and may be either step functions or continuous. For other participants, only a step function with
up to nine components may be submitted. The resulting supply functions, which return the total cost, are
either piecewise linear or piecewise quadratic such that the Ecient Allocation program is a mixed integer
quadratic program (QIP).
14
4.1
Empirics
The goal of the emipirical section is to show that in the context of a real market with noisy data, the IO
recovery algorithm generates informative parameters regarding the relative utilization of key resources and
the relative loss throughout the network. To show this result, we rst recover the parameters and then test
their accuracy by using them to generate new prices from bids and by comparing those prices to the prices
generated by MISO for that same set of bids.
We focus our attention on several cases taken from the day ahead market in March of 2011. For each of
these examples we do the following:
1. Recovery: Using the IO recovery algorithm recover constraints K using a set of identifying samples SK .
2. In-sample testing: (IS) Take bids and oers for s ∈ SK determine prices using Program 1 and the
estimated constraints.
3. Out-of-sample testing: (OS) Take bids and oers for s ∈
/ SK where active constraints are a subset of
K and determine prices using Program 1.
4. Well-out-of-sample testing: (WOS) Take bids and oers for s ∈
/ SK where active constraints are not a
subset of K and determine prices using Program 1.
5. Comparison to benchmarks: To test the model quality, we compare the accuracy to prices generated
with two benchmark algorithms. The rst benchmark (NC) is a model where we assume that there is
no congestion, performing an unconstrained uniform price auction with the observed bids. The second
benchmark (UC) produces prices using a naive data-centric approach: assume prices remain unchanged
from the rst sample.
We have selected three cases (Test1-Test3) from March of 2011 where respectively 1, 2, and 3 unique constraints are active and identied. For each case, |K| + 1 IS samples are selected and in all cases yield a full
rank system of linear equations. OS and WOS samples were also selected from March of 2011. For each
example, three OS samples were selected closest in date to the example IS. WOS samples are common to all
examples. The four WOS samples were randomly selected satisfying the conditions that (1) they occured in
March of 2011, (2) the active set is not a subset of the example active sets, and (3) they respectively have
1, 3, 5, and 7 active constraints. All tests have the same WOS set. The hours and active constraints are
reported in Table 2. The procedure also assumes that we have an accurate map from participant to location.
This was accomplished by clustering participants and locations with persistent identical prices. They are
assumed to remain constant over the course of the study.
15
ID
Hour
Test1:
Active Resources
ID
Hour
Active Resources
Test2:
In sample
In sample
IS-1-0
2011-03-26 23:00:00
680
IS-2-0
2011-03-26 23:00:00
680
IS-1-1
2011-03-26 02:00:00
680
IS-2-1
2011-03-27 00:00:00
680:1473
IS-2-2
2011-03-26 02:00:00
680
Out of sample
Out of sample
OS-1-1
2011-03-26 03:00:00
680
OS-2-1
2011-03-26 03:00:00
680
OS-1-2
2011-03-27 11:00:00
680
OS-2-2
2011-03-27 01:00:00
680:1473
IS-3-0
2011-03-18 12:00:00
100:1526:1530
WOS-?-0
2011-03-18 21:00:00
1531
IS-3-1
2011-03-18 13:00:00
100:1526:1530
WOS-?-1
2011-03-20 22:00:00
100:680:1530
IS-3-2
2011-03-18 14:00:00
100:1526:1530
WOS-?-2
2011-03-13 11:00:00
3:307:1257:1496:1504
IS-3-3
2011-03-18 18:00:00
100:1526:1530
WOS-?-3
2011-03-20 14:00:00
100:307:1329:1358:1473:1530:1537
OS-3-1
2011-03-18 19:00:00
100:1530
OS-3-2
2011-03-18 20:00:00
100:1526:1530
Test3:
All Tests:
In sample
Well out of sample
Out of sample
Table 2: Samples for each test case
Determining Market Outcomes:
To determine the predicted LMPs and MW allocations, one may solve
the QIP using the complete supply function bids in Program 1. Then, x the binary variables and resolve
the model to determine the shadow prices and associated LMPs. This would correspond to the sequence
of operations for MISO which rst solves the Unit Commitment problem and then solves the continuous
Economic Dispatch problem to determine prices. This process is computationally challenging due to the
large number of binary variables. To speed up the process we essentially allow MISO to solve the Unit
Commitment by taking advantage of our knowledge of the true allocation values and force the MW values
to be in the observed active segment of their supply function. We then solve what is an estimated version of
the ecient allocation program to determine prices and the precise allocation. Comparison with the network
when there are no transmission constraints illuminates the degree to which the transmission structure is
impacting these outcomes.
4.2
Results
Results of applying the IO pricing mechanism are summarized in Table 3. These may be contrasted with
the NC and UC benchmarks reported in Tables 4 and 5. The measures we consider are the correlation of
predicted LMPs with true LMPs, the average absolute error between the predicted and true LMPs, and the
squared error associated with the linear regression lmptrue = a · lmppred + b . We also report on the same
values when the outliers are removed. An outlier is dened as the 3% of locations with maximum absolute
LMP error.
To summarize Table 3, we nd that the prices generated using the derived constraints are very highly
16
ID
All locations and participants
corr-coef
IS-1-0
IS-1-1
OS-1-0
OS-1-1
WOS-1-0
WOS-1-1
WOS-1-2
WOS-1-3
IS-2-0
IS-2-1
IS-2-2
OS-2-0
OS-2-1
WOS-2-0
WOS-2-1
WOS-2-2
WOS-2-3
IS-3-0
IS-3-1
IS-3-2
IS-3-3
OS-3-0
OS-3-1
WOS-3-0
WOS-3-1
WOS-3-2
WOS-3-3
1.000
0.999
1.000
0.951
0.723
0.990
0.349
0.974
1.000
0.470
0.993
0.842
0.456
0.723
0.988
0.350
0.974
0.416
0.567
0.399
0.384
0.157
0.210
0.140
0.744
0.314
0.581
std-reg-err
0.020
0.042
0.026
7.023
9.832
0.334
16.542
1.048
0.010
107.298
0.311
11.519
109.181
9.830
0.338
16.432
0.926
687.254
180.706
669.760
4 119.915
26 105.651
28 562.426
9 407.944
21.780
205.694
105.566
avg-abs-err
2.522
1.606
0.976
7.555
5.909
2.471
13.150
1.605
2.672
2.391
1.747
1.156
2.126
5.910
2.379
13.159
1.783
16.285
10.037
7.053
12.192
27.242
23.414
14.741
2.684
14.314
2.348
Outliers removed
mw avg-abs-err
12.588
10.778
10.363
14.457
16.013
9.473
16.522
8.022
12.703
11.789
10.841
10.157
12.482
15.706
9.505
16.861
8.065
14.777
16.094
14.075
14.216
13.372
13.661
16.742
10.480
18.432
8.744
corr-coef
1.000
0.999
1.000
0.951
0.744
0.991
0.392
0.982
1.000
0.987
0.999
0.996
0.993
0.744
0.989
0.392
0.982
0.978
0.978
0.988
0.971
0.946
0.895
0.710
0.984
0.369
0.987
std-reg-err
0.015
0.034
0.020
6.292
8.487
0.290
15.433
0.705
0.007
0.844
0.041
0.188
0.436
8.486
0.290
15.329
0.635
1.796
1.589
0.764
2.039
8.151
11.000
17.789
0.765
17.289
0.818
avg-abs-err
2.501
1.564
0.929
7.169
5.679
2.301
12.860
1.465
2.662
2.075
1.690
1.008
1.830
5.679
2.192
12.870
1.641
13.296
8.443
3.830
5.373
12.121
7.694
5.303
2.092
12.883
1.035
mw avg-abs-err
4.001
4.157
4.044
3.727
4.693
4.125
4.559
3.451
4.076
3.734
4.265
4.064
4.238
4.520
4.138
4.651
3.503
4.246
4.360
3.851
3.863
3.491
3.625
5.114
4.196
5.535
3.501
Table 3: Price and MW predictions from using IO recovery algorithm
correlated with the true prices and error can largely be removed through a linear shift. In the case of Tests
2 and 3, there are a small number of outliers which degrade the summary statistics. When outliers are
removed the delity of both Test 2 and 3 improves dramatically. For example, for IS-3-0, removing the
outliers improves correlation from 0.416 to 0.978 and the standard error is reduced from 687.254 to 1.796.
We expand in Section on the distribution of error in our detailed discussion of Test 3 below. Below, we will
discuss in depth how this error is systemic and appears to reect missing generation data in a congested
region of the network. In Section 5 we discuss how the recovered model may be used to infer the magnitude
and location of the unobservable data.
Table 4 shows the same outcome of the UC benchmark when transmission constraints are removed from
the ecient allocation problem. Price dierences produced in this model are then exclusively a result of
the recovered loss coecients. Without the transmission constraints, the correlation between predicted
and observed prices decreases for in-sample and out-of-sample tests with the exception of Test 3. The UC
benchmark does not produce systemic outliers; removing outliers produces little improvement. Excluding
outliers, IO recovery algorithm performs better than the UC benchmark with respect to average absolute
pricing error on all instances and with respect to correlation on all but four instances. The four instances
which show higher correlation when the UC benchmark is used are well-out-of-sample tests. This may be
explained by loss parameters but not transmission constraints being consistent across the in-sample and
well-out-of-sample tests.
17
ID
All locations and participants
corr-coef
IS-1-0
IS-1-1
OS-1-0
OS-1-1
WOS-1-0
WOS-1-1
WOS-1-2
WOS-1-3
IS-2-0
IS-2-1
IS-2-2
OS-2-0
OS-2-1
WOS-2-0
WOS-2-1
WOS-2-2
WOS-2-3
IS-3-0
IS-3-1
IS-3-2
IS-3-3
OS-3-0
OS-3-1
WOS-3-0
WOS-3-1
WOS-3-2
WOS-3-3
0.893
0.872
0.871
0.947
0.839
0.854
0.470
0.891
0.893
0.900
0.872
0.871
0.897
0.839
0.854
0.470
0.891
0.749
0.761
0.719
0.749
0.714
0.812
0.856
0.552
0.614
0.612
std-reg-err
0.386
0.456
0.460
0.200
0.563
0.512
1.488
0.389
0.386
0.362
0.456
0.460
0.371
0.563
0.512
1.484
0.389
0.407
0.389
0.446
0.408
0.455
0.316
0.247
0.638
0.578
0.574
avg-abs-err
5.891
5.578
5.414
12.449
8.112
4.512
15.192
4.778
5.891
5.992
5.578
5.414
5.506
8.112
4.512
15.221
4.778
18.133
12.337
8.340
9.486
17.475
11.629
8.813
5.306
15.922
5.613
Outliers removed
mw avg-abs-err
20.187
16.412
16.445
26.323
21.059
13.475
21.887
12.025
20.057
18.213
16.294
16.331
18.471
20.922
13.351
21.808
11.905
23.058
22.299
21.076
20.565
21.693
20.557
23.029
15.660
24.198
14.142
corr-coef
0.894
0.892
0.897
0.944
0.839
0.890
0.482
0.925
0.894
0.899
0.892
0.897
0.894
0.839
0.890
0.482
0.925
0.728
0.737
0.703
0.725
0.686
0.785
0.863
0.579
0.604
0.632
std-reg-err
0.362
0.383
0.371
0.196
0.545
0.402
1.411
0.278
0.362
0.345
0.383
0.371
0.357
0.545
0.402
1.407
0.278
0.354
0.342
0.402
0.373
0.400
0.288
0.204
0.627
0.554
0.554
avg-abs-err
5.741
5.436
5.272
12.338
8.036
4.262
14.958
4.600
5.741
5.844
5.436
5.272
5.354
8.036
4.262
14.987
4.600
17.948
12.190
8.197
9.231
17.228
11.486
8.736
5.088
15.686
5.462
mw avg-abs-err
5.622
5.451
5.409
7.242
5.857
4.929
6.151
4.216
5.622
5.183
5.451
5.409
5.686
5.857
4.929
6.158
4.196
5.800
5.312
4.925
4.831
4.956
4.903
6.179
5.105
6.784
4.337
Table 4: Benchmark 1 NC: no transmission constraints
Table 4 shows the same outcome of the NC benchmark when the rst hour's price and allocation are used
to predict prices in the other hours. The hours used for the prediction are respectively IS-1-0, IS-2-0, and
IS-3-0. Broadly, this predictor does not suer from large outliers and as a result does well on the in-sample
and out-of-sample tests. The IO mechanism is competitive with the NC on these IS and OS tests and of
course uses recovered market structure as opposed to market outputs to construct prices.
Figures 3 and 4 show the relationship between predicted and true prices for in-sample and out-of-sample
conditions for Tests 1 and 2. In both cases there is high correlation between the predicted and true prices. In
the case of Test 2, Figure 4-b shows that error in the out-of-sample test largely results from several outliers
which are priced very high and are the major contributors to the error.
Figure 5 compares recovered coecients for constraints 680 and 1473 using dierent sets of samples to
generate them. A log scale is used on the graph which accentuates error in the smaller coecients. There
is strong correlation between recovered coecients, particularly at higher coecient values (top right corner
of the plots).
The prices output for Test 3 are shown in Figure 6. Unlike the other examples, in this case there is a set
of locations which appear to be systematically overpriced by the model. In the case of Test 3 the demand is
larger resulting in higher levels of congestion and this set of nodes is isolated from the rest of the market.
In Section 5.3 we discuss how this may be an artifact of missing supply and how the structural model may
be used to uncover this unobservable supply.
18
ID
All locations and participants
corr-coef
std-reg-err
avg-abs-err
Outliers removed
mw avg-abs-err
corr-coef
std-reg-err
avg-abs-err
mw avg-abs-err
IS-1-0
IS-1-1
OS-1-0
OS-1-1
WOS-1-0
WOS-1-1
WOS-1-2
WOS-1-3
0.999
0.999
0.968
0.717
0.993
0.342
0.973
0.060
0.066
1.957
14.868
0.437
27.040
1.621
2.270
2.756
7.036
3.464
6.058
9.795
4.172
10.811
11.414
16.217
23.258
19.450
19.041
19.724
0.999
0.999
0.967
0.739
0.993
0.387
0.978
0.046
0.051
1.827
12.764
0.394
24.698
1.311
2.192
2.679
6.875
3.198
5.976
9.458
4.070
6.436
7.015
9.437
13.435
12.894
12.251
12.978
0.999
0.999
0.999
0.994
0.717
0.993
0.342
0.973
0.082
0.060
0.066
0.373
14.868
0.437
27.040
1.621
0.423
2.270
2.756
0.615
3.464
6.058
9.795
4.172
8.072
10.811
11.414
11.069
23.258
19.450
19.041
19.724
0.999
0.999
0.999
0.995
0.739
0.993
0.387
0.978
0.072
0.046
0.051
0.281
12.764
0.394
24.698
1.311
0.388
2.192
2.679
0.565
3.198
5.976
9.458
4.070
3.423
6.436
7.015
5.350
13.435
12.894
12.251
12.978
0.999
0.998
0.965
0.997
0.990
0.802
0.947
0.441
0.964
0.045
0.091
1.457
0.138
0.405
7.622
2.201
17.190
1.514
5.837
9.847
8.650
0.785
6.503
9.332
17.981
4.129
16.038
2.024
4.260
8.680
7.106
7.254
11.865
35.288
17.009
31.741
0.999
0.998
0.963
0.998
0.990
0.828
0.953
0.514
0.970
0.045
0.086
1.461
0.070
0.404
6.516
1.880
15.869
1.251
5.796
9.796
8.595
0.715
6.433
9.170
17.837
3.898
15.876
0.828
1.981
4.620
3.952
3.837
5.901
22.146
9.188
18.925
IS-2-0
IS-2-1
IS-2-2
OS-2-0
OS-2-1
WOS-2-0
WOS-2-1
WOS-2-2
WOS-2-3
IS-3-0
IS-3-1
IS-3-2
IS-3-3
OS-3-0
OS-3-1
WOS-3-0
WOS-3-1
WOS-3-2
WOS-3-3
Table 5: Benchmark 2 UC: uses rst IS sample prices and allocations for each test
(a) In Sample
(b) Out of Sample
Figure 3: Predicted vs. True LMPs, predictions using Test 1 with one active constraint
19
(a) In Sample
(b) Out of Sample
Figure 4: Predicted vs. True LMPs, predictions using Test 2 with two active constraint
(a) Constraint 680,
x-axis:IS-1-0,IS-1-1
y-axis:OS-1-0,OS-1-1
(b) Constraint 680,
x-axis:IS-2-0,IS-2-1,IS-2-2
y-axis:IS-2-0,OS-2-1,OS-2-2
(c) Constraint 1473,
x-axis:IS-2-0,IS-2-1,IS-2-2
y-axis:IS-2-0,OS-2-1,OS-2-2
Figure 5: Recovered D coecients from dierent sample sets
(a) In Sample
(b) Out of Sample
Figure 6: Predicted vs. True LMPs, predictions using Test 3 with three active constraint
20
The results indicate that the recovered market structure is quite accurate: the prices are very highly
correlated with the true prices and removing the constraints reduces quality except where they should not
be included (the well-out-of-sample tests). However, we are not able to precisely predict prices which tend
to be shifted either as a whole or in bunches of similar locations. Missing supply and demand was mentioned
above. MISO allows participants to also participate in both the ancillary services market and energy markets.
A subset of participants also partaking in selling regulatory volumes may also be a cause of error (Ma et al.
[2009b]).
5
Applications and Extensions
5.1
Application to Competitive Markets
In this section we show that a competitive market where variation in prices is explained through competition
for a set of resources with xed capacities is amenable to the methodology developed for electricity markets.
We present sucient assumptions for an equilibrium in such a market to satisfy the twin market operations
and pricing requirements, i.e., any equilibrium results in an ecient allocation with prices corresponding to
the dual solution. The IO recovery algorithm can then be directly applied.
In some likely applications the frictions driving price dierences may also be due to a capacitated transportation resource and the relative location of producers and consumers. For instance, pipeline capacity may
explain large locational price dierences in natural gas markets. The central assumption is that empirically
observed commodity prices correspond to optimal dual values of the production constraints in the optimal
solution, i.e., standard necessary conditions for optimality provide a set of linear equations where variables
are the resource utilization parameters. This pricing assumption may be justiable in decentralized markets
which are in competitive equilibrium. The relationship between dual prices in mathematical programming
models and competitive equilibria have long been well known (see, e.g., Samuelson [1952] and Takayama and
Judge [1964]).
The competitive market model augments the electricity market model described in Section 2.1 by also
considering the set of resource holders. We assume that each resource is held by several arbitrarily small
agents each with marginal cost of zero. Then, consider the set of prices (−π, −ρ) where a producer receives
revenue −πl(i) xi and resource holders in total produce θj of resource j and receive revenue −ρj θj where
P
θj = i Dji bi . For simplicity we assume that the market is lossless.
The notion of a competitive equilibrium identied here requires (1) that agents are prot maximizing given
prices and production constraints, (2) the market allocation is feasible, and (3) no arbitrage opportunities
21
exist. We implement (3) by asserting that no agents can protably purchase the commodity at location
a, pay the resource costs for that location, and receive the price for the reference location. The following
theorem shows that these conditions are sucient to ensure that the identifying equations are sound and
that the IO recovery algorithm is applicable.
Theorem 4.
Consider prices (−π, −ρ) and production (x, θ). If this system is in a competitive equilibrium,
then Equations 7-17 hold.
Proof. φ may be interpreted as the marginal production cost of the lowest cost unconstrained production.
α and γ may be interpreted as the dierence between the received price and the marginal production cost
for each participant.
CE⇒Primal Feasibility (Equations 9-13): Requirement (2) is a statement of primal feasibility (Equations
9-13).
CE⇒Dual Feasibility (Equation 14) : ρ ≤ 0 is a statement of non-negative resource prices which is
consistent with our statement that the marginal resource production cost is zero for all agents.
CE⇒Complementary Slackness (Equations 15-17): Equation 15 is equivalent to ρ = 0 ⇐ Db < d and
ρ < 0 ⇒ Db = d; these must hold since any price greater than zero would allow a resource holder to
protably increase production if there is surplus capacity. Equation 16 is equivalent to γ = 0 ⇐ x > `
and γ < 0 ⇒ x = `; with Equation 8 negative γ implies pricing below the marginal cost in which case
revenue is maximized by producing the minimum volume `. Equation 17 is equivalent to α = 0 ⇐ x < u and
α < 0 ⇒ x = u; with Equation 8 negative α implies pricing above the marginal cost in which case revenue
is maximized by producing the maximum volume u.
CE⇒Stationarity (Equations 7 and 8): Equation 7 corresponds to eliminating the arbitrage opportunity.
Equation 8 holds by assumption.
5.2
Transmission Constrained Market Power Variation
The transmission constraint parameters are useful for measuring the distribution of market power through
the market footprint. As noted in [Borenstein and Bushnell, 1999] and [Karthikeyan et al., 2013] and others,
market power can vary substantially in an electricity and this variation can be both over the footprint and
over time. This spatial and temporal variation depends on the impact of a particular participant on active
constraints and the ability of particular participants to make up for production shortfalls which is dened
by the loss parameters. To capture this variation, Lee et al. [2010] and Lee et al. [2011] argue that useful
market power measures can be constructed by using the residual demand derivative for each participant in
the market. The residual demand derivative denes the demand surplus that a particular participant would
22
(a) Active Constraint 680
(b) Active Constraints 680, 1473 (c) Active Constraints 100,1526,1530
Figure 7: Distribution of Residual Demand Derivative for Samples with Diering Active Constraints
observe for a marginal increase in price. In the transmission constrained electricity market this depends
both on the transmission constraint contribution factors D and the loss parameters δ .The residual demand
derivative is a precise reection of the current market state and without some form of aggregation, does not
necessarily reect the a priori beliefs of participants. These beliefs may provide a more accurate explanation
for the structure of a particular bid. For instance, only the realized set, and no highly likely alternative sets,
of active constraints are taken into account. Xu and Baldick [2007] show that in a lossless locationally priced
electricity market, this realized residual demand derivative can be eciently calculated for each location
given bids and the constraint factors of active constraints. To adapt this analyses to our model requires only
a minor modication.In short, replacing the lossless energy balance equation with Constraint 3 and carrying
through the same derivation as in Xu and Baldick [2007], the residual demand derivative for a particular
participant i located at the reference bus becomes:
RDD(i) = −δ̄iT Λi δ̄i + δ̄iT Λi D̄i (D̄iT Λi D̄i )−1 D̄iT Λi δ̄i
where, δ̄i and D̄i are respectively the participant loss vector and the participant active constraint contribution
matrix and relevant entries for i removed (i.e. δ̄i is Q − 1 dimension vector D̄i is a Q − 1 × Q − 1 dimension
matrix). With b(j) denoting the bus at which participant j is located and for convenience allowing i = Q,
(δ̄i )j = δb(j) and (D̄i )rj = Drb(j) .Λi is a diagonal matrix with entries corresponding to Sj00 (xj ) of each
participant except i (i.e. Λi is a Q − 1 × Q − 1 dimension matrix). To calculate the residual demand
derivative for a participant i not at the reference bus, δ and D can easily be refactored so that b(i) is the
reference bus by dividing δ by δb(i) and zeroing the b(i)th row of D using δ .
We have implemented this calculation for each of the market samples considered in Section 4. Histograms
showing the distribution of residual demand derivatives for each participant are shown in Figure 7. This
gure shows an outcome for each of the three test sets with one, two or three active constraints. Panels (a)
23
(a) Coecients of Constraint 100
(b) Coecients of Constraint 1526 (c) Coecients of Constraint 1530
Figure 8: Relationship between residual demand derivative and active constraint coecients for example
IS-3-0
and (b) of Figure 7 show very similar ranges of market power though the less constrained example in Panel
(a) appears to show a slightly more uniform distribution of residual demand derivatives. Panel (c) on the
other hand is quite distinct. While the bulk of the participants face similar residual demand derivatives in
the -20 to 0 range, there is a long left tail where participants face substantially higher magnitude residual
demand derivatives. Figure 8 shows the relationship between the residual demand derivative and constraint
coecients. There is little visible correlation in panels (a) and (b) between the residual demand derivative
and coecients of constraints 100 and 1526. On the other hand Figure 8(c) shows that the residual demand
derivative is highly correlated with coecients for constraint 1530.
5.3
Analytical Applications
In this paper we have derived and demonstrated a novel structure recovery methodology. Structural approaches provide an advantage over purely statistical approaches by allowing analysis of hidden parameters
and counterfactual analysis. Recovery of parameters of capacity lines, D and d, is an example of a hidden
parameter not accessible through standard statistical methods. Relevant examples that have not yet been
explored include uncovering unobservable supply and demand, estimating market power, and estimating the
value of capacity investment. We will demonstrate a naive version of uncovering unobservable supply and
demand and discuss implementation of the other analyses.
While bids and outcomes are publicly reported for participants in most electricity markets, there are
supply and demand sources in electricity markets that are not reported. These may include imports and
exports and grandfathered bilateral transactions. In practice, these injections and withdrawals will not
be decision variables in the optimal ow program but they nonetheless aect transmission constraints and
thereby indirectly impact prices. As a result, when new prices are generated without this missing data, the
lack of supply and demand in certain regions may bias prices. This type of error would appear as a systemic
24
(a) Original predicted LMPs
(b) Predicted LMPs with 61MW at Location 1811
Figure 9: Predicted vs. True LMPs for Test 3 with and without absolute error minimizing additional supply
bias in the price for a set of locations with similar coecients for one or more transmission constraints,
which is consistent with the type of error observed in our predictions of MISO prices. This suggests a
simple moment matching algorithm where supply is added at overpriced nodes and demand is increased at
underpriced nodes until an error measure such as the absolute pricing error of the LMPs is minimized.
Since Test 3 from Section 4 very clearly shows this type of bias we use this as a case study. The relation
between predicted and actual MW for Test 3 in-sample IS-3-0 is depicted in Figure 9 Panel (a). A collection
of outliers appears in the top right hand corner of this gure which are severely overpriced by the recovered
model resulting in an absolute pricing error of $16.285. These outliers correspond to a set of pricing nodes
in northern Wisconsin and Michigan. We have aggregate information on day ahead import plan and in this
hour, 1511 MWh were planned for import from the IESO market (the Ontario energy market) and 1600
MWh from the PJM market. It is plausible that a substantial amount of energy may have moved into the
MISO market from Ontario resulting in this descrepancy. Under this assumption, we applied a naive version
of the moment matching technique by adding a supply of uncharged MWs of energy at location 1811 (MISO
pricing node UPPC.AZ) which is one of the nodes in the cluster of outliers. The absolute error is reduced to
$14.74 when 61MW are added. The resulting prices from the absolute error minimizer are shown in Figure
9.
Recovered market structure allows estimates of the value of a capacity investment to be made which
take into account market power. If there is substantial demand elasticity at the location, revenue may be
much less than if the new capacity was sold at the previously observed prices. The revenue needs to take
into account the price eect of the injection of capacity at that particular point in the network. Using the
recovered model, the eect of the new capacity on the local price can be recovered by rerunning the allocation
program with the new bids. More sophisticated analyses may consider strategic factors and also modify the
25
IS-3-0
15.6
15.5
15.4
Avg Absolute Error
15.3
15.2
15.1
15
14.9
14.8
14.7
14.6
0
20
40
60
80
100
120
MW at location 1811
Figure 10: Absolute error in price predictions as supply is increased at Location 1811
bidding strategies of other agents to ensure they remain rational.
6
Conclusion
This paper develops an inverse optimization based method for recovering market structure from allocation
and pricing data. The methodology takes advantage of a long history of equating market eciency and
optimization. It is justied when the allocation and locational prices correspond to the primal and dual
of the optimal assignment problem. The IO recovery algorithm implements the methodology in a context
specialized for electricity markets where prices are determined by competition for transmission resources and
line loss. The IO recovery algorithm recovers utilizations of transmission resources from the locational prices
alone and with allocation information the capacities of these resources are also determined.
In the context of a case study of the MISO electricity market, we studied the application of this methodology. Our ndings show that the recovered market structure does a good but imperfect job at reconstructing
market prices. We hypothesize that the imperfections are primarily due to missing allocation data. In
extensions, we discuss naive methods to estimate missing supply and demand for regions in the market.
Markets where transportation links between regions are capacitated are challenging for managers because
market conditions can shift dramatically depending on which constraints are active. The methodology
put forward in this paper can help managers determine their bidding and operational strategies by using
structural models to make market inferences and perform counterfactuals. Similarly, this may provide a
mechanism for regulators to make inferences about markets where they cannot observe all transactions but
have some knowledge of prices and allocation volumes. In the applications and extensions we have discussed
estimation of missing missing market data, estimation of market power and analysis of investment decisions
as particularly worthy of further study. With further renement of these and related techniques, structure
recovery methodologies based on inverse optimization may be a useful analytical tool for participants in
26
institutionalized markets like electricity markets as well as decentralized but ecient markets exhibiting
locational price variation.
References
Ravindra K Ahuja and James B Orlin. Inverse Optimization. Operations Research, 49(5):771783, 2001.
Severin Borenstein and James Bushnell. Market Power in Electricity Markets: Beyond Concentration Measures. Energy Journal, 20(4):6588, 1999.
Christian Broda and David E Weinstein. Understanding International Price Dierences Using Barcode Data
Christian Broda. NBER Working Paper, (14017), 2008.
Judith B Cardell, Carrie Cullen Hitt, and William W Hogan. Market power and strategic interaction in
electricity networks. Resource and Energy Economics, 19(1):109137, 1997.
A Kerem Cosar, Paul L E Grieco, and Felix Tintelnot. Borders , Geography , and Oligopoly: Evidence from
the Wind Turbine Industry. 2012.
Helmuth Cremer, Farid Gasmi, and Jean-Jacques Laont. Access to Pipelines in Competitive Gas Markets.
Journal of Regulatory Economics, 24(1):533, 2003.
Charles Engel and John H. Rogers. How wide is the border? The American Economic Review, 86(5):pp.
11121125, 1996.
William W. Hogan, E. Grant Read, and Brendan J. Ring. Using Mathematical Programming for Electricity
Spot Pricing. pages 209221, 1996.
Garud Iyengar and Wanmo Kang. Inverse conic programming with applications. Operations Research Letters,
33(3):319330, May 2005. ISSN 01676377. doi: 10.1016/j.orl.2004.04.007.
SP Karthikeyan, IJ Raglend, and DP Kothari. A review on market power in deregulated electricity market.
International Journal of Electrical Power & Energy Systems, 48:139147, June 2013.
Yen-yu Lee, Jin Hur, Ross Baldick, and Salvador Pineda. New Indices of Market Power. IEEE Transactions
on Power Systems, 26(681-689):19, 2010.
YY Lee, Ross Baldick, and Jin Hur. Firm-based measurements of market power in transmission-constrained
electricity markets. Power Systems, IEEE Transactions, 26(4):19621970, 2011.
27
Xingwang Ma, Haili Song, Mingguo Hong, Jie Wan, Yonghong Chen, and Eugene Zak. The securityconstrained commitment and dispatch for midwest iso day-ahead co-optimized energy and ancillary service
market. In Power & Energy Society General Meeting, 2009. PES'09. IEEE, pages 18. IEEE, 2009a.
Xingwang Ma, Haili Song, Mingguo Hong, Jie Wan, Yonghong Chen, and Eugene Zak. The securityconstrained commitment and dispatch for midwest iso day-ahead co-optimized energy and ancillary service
market. In Power & Energy Society General Meeting, 2009. PES'09. IEEE, pages 18. IEEE, 2009b.
John Mccallum. National Borders Matter : Canada-U . S . Regional Trade Patterns. American Economic
Review, 85(3):615623, 1995.
M. Mudrageda and F. H. Murphy. An Economic Equilibrium Model of the Market for Marine Transportation
Services in Petroleum Products. Operations Research, 56(2):278285, March 2008.
Jorge Nocedal and Stephen J. Wright. Numerical optimization. 2nd edition, 1999. ISBN 9780387303031.
Mandhir Sahni, Richard Jones, and Yunzhi Cheng. Beyond the Crystal Ball. IEEE Power & Energy
Magazine, 10(4):3542, 2012.
Paul A Samuelson. Spatial Price Equilibrium and Linear Programming. American Economic Review, 42(3):
283303, 1952.
Andrew J. Schaefer. Inverse integer programming. Optimization Letters, 3(4):483489, June 2009. ISSN
1862-4472. doi: 10.1007/s11590-009-0131-z.
B. Stott, J. Jardim, and O. Alsac. DC Power Flow Revisited. IEEE Transactions on Power Systems, 24(3):
12901300, August 2009.
T Takayama and G Judge. Spatial Equilibrium and Quadratic Programming. Journal of Farm Economics,
46(1):6793, 1964.
Lizhi Wang. Cutting plane algorithms for the inverse mixed integer linear programming problem. Operations
Research Letters, 37(2):114116, March 2009. ISSN 01676377. doi: 10.1016/j.orl.2008.12.001.
Frank A Wolak. Measuring the benets of greater spatial granularity in short-term pricing in wholesale
electricity markets. American Economic Review, 101(3):247, 2011.
Allen J Wood and Bruce F Wollenberg. Power generation, operation, and control. Wiley-Interscience, 2012.
Lin Xu and Ross Baldick. Transmission-constrained residual demand derivative in electricity markets. Power
Systems, IEEE Transactions on, 22(4):15631573, 2007.
28
Xiaoguang Yang and Jianzhong Zhang. Partial inverse assignment problems under norm. Operations Research
Letters, 35(1):2328, January 2007. ISSN 01676377. doi: 10.1016/j.orl.2005.12.003.
Jianzhong Zhang and Zhenhong Liu. Calculating some inverse linear programming problems. Journal of
Computational and Applied Mathematics, 72:261273, 1996.
Jianzhong Zhang and Zhenhong Liu. A further study on inverse linear programming problems. Journal of
Computational and Applied Mathematics, 106:345359, 1999.
A
Electricity distribution as a linear allocation problem
The standard assumptions underlying the DC approximation to AC power ow are (1) an assumption of no
line loss, (2) bus voltage magnitudes are unit, and (3) voltage angle dierences across branches are small
(the implications of these assumptions are discussed in Stott et al. [2009]). The standard formulation for the
power ows (c.f. Wood and Wollenberg [2012]) uses voltage angles θi for each node i:
X
bi = 0,
(33)
Bθ = b,
(34)
θr = 0,
(35)
y(i,j) =
1
(θi − θj ),
z(i,j)
∀(i, j) ∈ L.
(36)
Node r is the reference bus which is selected arbitrarily and set to zero so that the problem is well dened.
−1
B is determined by per unit line reactances z(i,j) where for i 6= j , Bij = z(i,j)
if branch (i, j) exists and
P
1
zero otherwise and Bii = (g,h)∈Ei z(g,h)
where Ei is the set of links involving bus i. To put this system of
equations into a more useful form for our analysis, we dene the matrix Bf as an |L| × |N | matrix where
row k corresponds to edge (i, j), (Bf )ki = 1/zk , (Bf )kj = −1/zk , and for h ∈
/ {i, j}, (Bf )kh = 0 (zk is the
line reactance of line k ). Then letting matrix D = Bf B , y = Db, and the constraints on production at the
buses may be succinctly represented as:
X
bi = 0,
(37)
|Db| ≤ d,
(38)
where the second constraint is from the capacity d(i,j) of link (i, j).
29
In keeping with the MISO's optimization model (described in Ma et al. [2009a]), we consider losses
approximated by linear functions of the injections into the network. The loss model, L-LIN, considers loss
as a linear function of net production and consumption at each node. Positive loss parameters δi+ and δi−
are assigned to each node i :
X
bi +
X
δi− b+
i +
X
δi− b−
i = 0,
(39)
(40)
|Db| ≤ d.
There are more principled ways of incorporating loss which would also adjust matrices B and Bf to
accurately account for ow changes from loss. However, since the goal is to infer the product of these
matrices from actual data, these details are not too concerning.
A.1
Flows in Numerical Example
This is a numerical example of the construction of the DC ows from a description of the network. We
assume no loss in this example. We have four nodes {1..4} connected into a single cycle by links L =
{1 − 2, 2 − 3, 3 − 4, 1 − 4} (let these links have respective indices 1..4 and be referred to as ek ). Flow may go
in either direction on these links (y may be negative) and have reactance z = 1 except for the link 1-4 where
z = 2.
Participant 1
Min:0 MW
Max:500 MW
Cost:$20/MWh
1
z=1
2
Limit: 50MW
z=2
Participant 4
Min:-300 MW
Max:-300 MW
Participant 2
Min:0 MW
Max:0 MW
z=1
4
z=1
3
Connectivity may be described in the following adjacency matrix:
30
Participant 3
Min:0 MW
Max:500 MW
Cost:$30/MWh

1


 1



 0

1
1
0
1
1
1
1
0
1

1


0 

.

1 

1
This connectivity is reected in the admittance matrix:

1.5


 −1

B=

 0

−0.5
−1
−0.5
0
2
−1
−1
2
0
−1



0 

.

−1 

1.5
Selecting bus 4 as the reference bus, we set θr = θ4 = 0. Then we set B−r as B less the row and column
−1
x−r .
corresponding to the reference bus. Then B−r θ−r = x−r . B−r may be inverted such that θ−r = B−r
−1
Then dene Bθ as B−r
with an additional row and column of zeros inserted at the position corresponding
to the reference bus. In our example these matrices are:



−1
B−r
1.2
0.8
0.4
0


 1.2 0.8 0.4 
 0.8 1.2 0.6 0


 and Bθ = 
=

0.8
1.2
0.6





 0.4 0.6 0.8 0

0.4 0.6 0.8
0
0
0 0





.



Finally, the angles may be concisely represented as θ = Bθ x.
Now to get the ow on each link, consider connectivity matrix E as described in earlier documents, i.e.,
Eki = 1 if ek = (i, j) for some j , Ekj = −1 if ek = (i, j) and otherwise Eki = 0. Bf is simply matrix E row
by row multiplied by 1/zk such that Bf θ = y as in Equation 36. Then,

1


 0

E=

 0

1
−1
0
1
−1
0
1
0
0
0


1
−1





 0
0 

 and Bf = 


−1 
 0


−1
0.5
The ow over links can then be expressed as y = Bf θ = Bf Bθ x and
31
0
1
−1
0
1
0
0
0



0 

.

−1 

−0.5

0.4


 0.4

Bf Bθ = 

 0.4

0.6
−0.4
−0.2
0.6
−0.2
0.6
0.8
0.4
0.2

B
B
f θ 

Then D = 
.
−Bf Bθ

32
0



0 

.

0 

0
Fly UP