Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Robust Risk Management of Retail Energy Service

Providers in Midterm Electricity Energy Markets


under Unstructured Uncertainty
Jamshid Aghaei 1; Mansour Charwand 2; Mohsen Gitizadeh 3; and Alireza Heidari 4
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Retailers purchase electric energy from the wholesale market and sell it to the customers at regulated fixed prices in deregulated
power systems. A retailer has multiple choices for electricity procurement, such as spot market, bilateral contracts, and a self-generating
facility. However, the retailer faces challenges such as the pool price and electricity demand uncertainty. To meet these challenges, this paper
proposes a midterm framework based on the information gap decision theory (IGDT) to evaluate the robust or opportunity strategy. In this
approach, the composition of two conflicting portfolio investment goals is considered including obtaining high expected portfolio return and
selling price designation, as well as controlling risk. A real case study is used to demonstrate the effectiveness of the proposed framework.
DOI: 10.1061/(ASCE)EY.1943-7897.0000462. © 2017 American Society of Civil Engineers.
Author keywords: Retailer; Risk; Portfolio optimization; Information gap decision theory.

Introduction stochastic programming method to specify the amounts of power


that can be purchased from the pool market and forward contracts
Retailers participate in power markets by purchasing electricity via as well as the optimum selling price provided by the retailers for
bilateral contracts with generating companies or from the electricity customers at a fixed price. The self-production facility and call op-
pool markets and then selling it to their customers. Additionally, tion strategies were also considered by Hatami et al. (2009). Garcia-
retailers have contracts with customers to supply their demand with Bertrand (2013) proposed a procedure to determine the energy
prespecified selling prices. Consequently, retailers have a mediator procurement by a retailer and the optimum price while considering
role by signing contracts with both the demand-side and the sup- an acceptance function to model the behavior of the customer.
pliers in power markets (Shang et al. 2014). Carrión et al. (2007) presented an idea, later developed by Carrión
The objective of the retailers is the maximization of profit. How- et al. (2009), in which a bilevel programming model was proposed
ever, the retailer must tackle variable pool prices, demand forecast to take into account competition among retailers and the client
errors, and the possibility of customers selecting different suppliers response to selling prices.
if the retailer cannot offer a competitive selling price. Accordingly, Kettunen et al. (2010) presented a stochastic programming
after selecting the selling price and making decisions regarding fu- model to define the retailer’s optimal contract portfolio. Also,
ture market involvement, the retailer needs to resolve its purchases Boroumand and Zachmann (2012) demonstrated that the physical
and sales in the pool markets (Mazer 2007). hedging that is supported by forward contracting and spot transac-
tions can be an efficient approach for risk management in decen-
tralized electricity markets. Charwand et al. (2014a, 2015) provided
Literature Review a multiobjective decision framework to simultaneously maximize
retailers’ profit and minimize selling prices to clients in order to
The perspective of the retailer has been addressed in technical lit- achieve the highest possible number of clients. Feuerriegel and
erature. Carrión et al. (2007) and Hatami et al. (2009) proposed a Neumanno (2014) proposed an optimization problem that mini-
mizes procurement costs of an electricity retailer in order to control
1
Associate Professor, Dept. of Electrical and Electronics Engineering, demand response (DR) usage as a result of the integration of inter-
Shiraz Univ. of Technology, 71557-13957 Shiraz, Iran (corresponding mittent resources of power generation. Zugno et al. (2013) pre-
author). E-mail: aghaei@sutech.ac.ir
2 sented a game-theoretical model for the participation of energy
Ph.D. Student, Dept. of Electrical and Electronics Engineering, Shiraz
Univ. of Technology, 71557-13957 Shiraz, Iran. E-mail: m.charvand@ retailers in electricity markets with flexible demand and real-time
sutech.ac.ir consumer prices.
3
Associate Professor, Dept. of Electrical and Electronics Engineering, Carrión et al. (2007, 2009), Hatami et al. (2009), and Kettunen
Shiraz Univ. of Technology, 71557-13957 Shiraz, Iran. E-mail: gitizadeh@ et al. (2010) quantified the risk factor using the conditional value-
sutech.ac.ir at-risk (CVaR) index; however, Lim et al. (2011) investigated the
4
Ph.D. Student, Australian Energy Research Institute and School of estimation errors associated with CVaR in the context of an opti-
Electrical Engineering and Telecommunications, Univ. of New South mization problem and showed that the portfolios obtained by using
Wales, Sydney, NSW 2052, Australia. E-mail: alireza.heidari@unsw the CVaR index are fragile and not reliable. So, the expected down-
.edu.au
Note. This manuscript was submitted on June 15, 2016; approved on
side risk index was applied by Ahmadi et al. (2013) to hedge risks
February 16, 2017; published online on May 11, 2017. Discussion period attributable to uncertainties.
open until October 11, 2017; separate discussions must be submitted for Wei et al. (2015) proposed a two-stage, two-level model for the
individual papers. This paper is part of the Journal of Energy Engineering, energy pricing faced by a retailer, in which the demand response of
© ASCE, ISSN 0733-9402. consumers is characterized by a Stackelberg game in the first stage

© ASCE 04017030-1 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


[as done by Jia and Tong (2016) and Yu and Hong (2016)] and a of the IGDT for distribution network operators (DNOs) to choose
risk-aversive energy dispatch model is modeled by a linear optimi- a supply resource. Charwand and Moshavash (2014b) proposed an
zation in the second stage. Jia and Tong (2016) examined the ef- IGDT-based technique under pool price uncertainty to evaluate the
fects of integrating renewable energy sources and local storage. robust strategy for a retailer. However, Charwand and Moshavash
Also, Yu and Hong (2016) proposed an algorithm based on (2014b) did not consider the uncertainties associated with the retail-
real-time pricing for achieving optimal load control of devices er’s demand. Kharrati et al. (2016) proposed an IGDT-based
in a facility by forming a virtual electricity-trading process. Wei approach to solve a retailer’s midterm planning problem. They
et al. (2015) proposed a risk-aversive decision model for a retailer modeled the uncertainty of pool prices with an IGDT model
to offer a nondominated bidding strategy when the retailer intends and considered demand uncertainty via a scenario generation
to maximize daily profit using a two-step robust linear program- method (using Monte Carlo simulation with 10 scenarios). The
ming algorithm. Considering an optimal investment problem for Karush-Kuhn-Tucker (KKT) optimality conditions were used to
a retailer in the electricity market, He et al. (2015) provided the convert the problem into a robust optimization problem. Nojavan
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

optimal solution that maximizes the weighted sum of the expected et al. (2015b) proposed an IGDT-based model to obtain the optimal
return and the variance of the wealth. From the short-term sched- bidding strategy for a retailer that should be submitted to the pool
uling point of view, a robust optimization approach for a retailer market.
considering the effect of demand response programs was developed Considering a midterm horizon, this paper proposes a risk-based
by Nojavan et al. (2015a). model decision for the retailer based on information gap theory.
The IGDT models reveal plausible solutions at times when the in-
formation is difficult to access and the system is exposed to pivotal
IGDT Background and Stochastic Optimization changes regardless of current conditions. By analyzing the results,
Decision-making problems in electricity markets are plagued with the proposed method can help the retailer to determine its con-
uncertainty. Stochastic programming is one of the methods to deal tracting and selling strategy. The similarities and differences among
with uncertainty and can provide an accurate modeling framework the robust optimization (RO) technique, IGDT, and min-max regret
in which the problems of decision making under uncertainty are optimization can be found in detail in the Appendix I.
properly formulated. Stochastic programming relies on the knowl-
edge of the stochastic processes describing the uncertain parame- Contributions
ters, e.g., the pool price. Once the uncertain parameters are modeled
using the stochastic processes, it is possible to formulate a math- The main contribution of this paper in comparison with the conven-
ematical programming problem that takes into account the uncer- tional stochastic programming approach is that the proposed ap-
tainty of these parameters. For structured modeling purposes, each proach only needs a deterministic uncertainty set, instead of a
uncertain parameter is modeled by a set of finite outcomes or sce- hard-to-obtain probability distribution function with uncertain data.
narios, where each outcome represents a plausible realization of the By optimizing the robustness strategy instead of maximizing the
uncertain parameters with the respective probability of occurrence. retailer’s profit, the proposed model ensures a minimum allowable
The number of outcomes is needed to properly represent an uncer- profit using the forecasted data. In addition to the pool price un-
tain parameter. In some approaches, the uncertainty of a parameter certainty, since the modeling of the load uncertainty is very impor-
is modeled using probability and possibility theories based on the tant in the midterm planning, comparing with the works of
Charwand and Moshavash (2014b) and Kharrati et al. (2016),
distribution functions to model and measure the trading risk. Fur-
the uncertainties associated with the retailer’s demand have been
thermore, fuzzy logic–based models use prespecified membership
considered in this paper using the IGDT method. Also, in addition
functions to model uncertainty. Similarly, Monte Carlo simulation–
to pool market and bilateral contracts, a generating facility is con-
based approaches need to guess a probability density function for
sidered as a procurement option. The problem has been solved
the uncertain variable to generate the scenarios. Making such
as a multiobjective problem to generate Pareto-optimal solutions.
assumption can sometimes lead to nonprotective conclusions, as
Table 1 shows the main differences of this work with respect to the
further discussed by Ivatloo et al. (2013) and Gabriel et al.
previous ones in detail.
(2004). IGDT methods have practical advantages over scenario-
based methods. In the case of retailer planning, an IGDT-based
model determines optimal schedules in order to achieve a target Paper Organization
profit, whereas scenario-based methods find optimal schedules The rest of the paper is structured as follows: The section “Problem
based on the limited number of possible price scenarios. In addi- Formulation” deals with the decision-making problem formula-
tion, unlike scenario-based models, IGDT-based methods guaran- tion confronted by the retailer. The section “IGDT Analysis” pres-
tee a predefined level of profit. An IGDT model, on the other hand, ents the proposed IGDT-based solution approach. The section
neither requires a particular assumption about the nature of the “Case Study” copes with the performance of the proposed solu-
uncertain parameter nor enforces any preassumption on the size tion implementation. Finally, conclusions are drawn in the section
of the uncertainty (unstructured uncertainty). In fact, the different “Conclusion and Future Work.”
scenario generation methods can result in different solutions, espe-
cially with stochastic methods (Ben-Haim 2006). Also, in the sce-
nario generation methods and the fuzzy-based approaches, the Problem Formulation
probability density function and the specific membership function
of the uncertain parameters should be extracted, respectively. The main goal of the proposed method is to decide on four oper-
The IGDT method has already been used in the optimal bidding ation variables for the retailer in the midterm horizon, namely,
strategy of a generation company (Ivatloo et al. 2013; Kazemi et al. bilateral contracting decisions, self-production, grid purchasing
2015), hydrothermal scheduling (Charwand et al. 2016), combined and selling price under pool price, and demand uncertainty. It is
heat and power (CHP) unit commitment (Aghaei et al. 2016), and assumed that the risk premiums are included in the bilateral con-
energy purchasing strategies for large customers (Zare et al. 2010, tracts. Accordingly, the problem formulation has been extracted in
2011). Also, Soroudi and Ehsan (2013b) presented the application the following subsections.

© ASCE 04017030-2 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


Table 1. Taxonomy of the IGDT Solution in the Electricity Market Strategies
Uncertain parameter(s) Planning Strategy Base
Reference Decision maker in IGDT method horizon determination objective
Kazemi et al. (2015) Genco Pool price Short-term Selling energy Maximum profit
Charwand et al. (2016) Hydrothermal Demand Short-term Selling energy Minimum cost
Aghaei et al. (2016) CHP Pool price Short-term Selling energy Maximum profit
Zare et al. (2010, 2011) Large consumers Pool price Midterm Purchasing Minimum cost
Soroudi and Ehsan (2013b) Distribution network operators Pool price and demand Midterm Purchasing Minimum cost
Charwand and Moshavash Retailer Pool price Midterm Purchasing and Maximum profit
(2014b) selling price
Nojavan et al. (2015a) Retailer Pool price Short-term Bidding Minimum cost
This work Retailer Pool price and demand Midterm Purchasing and selling price Maximum profit
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

Energy Supply Resources for the Retailer facilities and the sum of the traded electricity in the pool market
should be nonnegative during the time horizon.
Bilateral Contract
The constant power price transactions during a given period of time
Supplying Demand
can be included in the bilateral contracts that the retailer imple-
ments to reduce the risk of the volatility of the pool price. Pur- It is assumed that the client’s behavior is dependent upon the selling
chased energy via the bilateral contracts is formulated as follows: price, which is shown by ρðeÞ. Therefore, higher values of ρðeÞ will
cause the client to reduce its demand from the retailer. According to
X
B X
T
Carrión et al. (2007), this behavior is modeled using a stepwise
CB ¼ γðb; tÞPB ðb; tÞkðtÞ ð1Þ
price-quota curve (Fig. 1), which shows that a competitive price
b¼1 t¼1
should be paid for a determined amount of electricity. In formulas
B ðbÞZðbÞ ≤ PB ðb; tÞ ≤ PB ðbÞZðbÞ;
Pmin ∀ t ∈ T; ∀b∈B
max
X
I

ð2Þ Dðe; tÞ ¼ D̄ðe; i; tÞYðe; iÞ; ∀ t ∈ T; ∀e∈E ð8Þ


i¼1
Eq. (1) presents the cost of purchased energy from the bilateral
contract and Eq. (2) expresses the maximum and minimum amount X
I

of purchased energy. ρðeÞ ¼ ρðe; iÞ; ∀e∈E ð9Þ


i¼1
Pool Trading
The retailer takes part in the pool market in order to satisfy a certain ρ̄ðe; i − 1ÞYðe; iÞ ≤ ρðe; iÞ ≤ ρ̄ðe; iÞYðe; iÞ; ∀ e ∈ E; ∀i∈I
amount of its load. However, the retailer can either purchase or ð10Þ
sell energy if profitable. The net transaction cost in the pool is
as follows: X
I
XT Yðe; iÞ ¼ 1; ∀e∈E ð11Þ
CP ¼ λðtÞPP ðtÞ ð3Þ i¼1
t¼1
Eqs. (8)–(11) model a step function relationship between de-
mand of each client group and time slot, Dðe; tÞ, and selling price,
Self-Production
ρðeÞ, in which by increasing the offering price, the supplied de-
A piecewise linear function has been considered as the energy op-
mand would be decreased. ρðeÞ is the selling price settled by
eration cost model for the generating facilities as done by Ahmadi
the retailer for the client group e. To make output easier and more
et al. (2013). Eq. (4) shows the energy operation cost of the avail-
obvious than ρðe; iÞ, this variable is used.
able generating units
X
T X
G D(e, t)
CG ¼ πðgÞEG ðg; tÞ ð4Þ (MWh)
t¼1 g¼1

D(e,1, t)
0 ≤ EG ðg; tÞ ≤ Emax
G ðgÞ − EG ðg − 1Þ;
max
∀ t ∈ T; g ¼ 2; : : : ; G
ð5Þ D(e, 2, t)

0 ≤ EG ð1; tÞ ≤ Emax
G ð1Þ; ∀t∈T ð6Þ
….
X
G
EG ðg; tÞ þ PP ðtÞ ≥ 0; ∀t∈T ð7Þ D(e, I, t)
g¼1
ρ(e,1) ρ(e, 2) ρ(e, I) ρ(e)
Constraints (5)–(7) are relevant to the operation of the generat- ($ / MWh)
ρ(e, 0) ρ(e,1) ρ(e, 2) ρR(l, I)
ing facilities and express the maximum amount of selling energy in
the pool market by the generating units’ capacity. It is expressed by
Fig. 1. Price-quota curve
constraint (7) that the electricity produced by the generating

© ASCE 04017030-3 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


Based on the stepwise price-quota curve, the revenue gained
from providing electricity to the customers is calculated as follows:
X
I
Rðe; tÞ ¼ D̄ðe; i; tÞρðe; iÞ; ∀ t ∈ T; ∀e∈E ð12Þ
i¼1
Fig. 2. Price band
Constraint (13) indicates the electricity balance for the retailer in
each period of each scenario as follows:
X  and a related joint variance–covariance matrix. However, in this
XE X I B XG
paper the unstructured uncertain parameters (pool price and de-
D̄ðe;i;tÞYðe;iÞ ¼ PB ðb;tÞkðtÞ þ EG ðg;tÞ þ PP ðtÞ ;
mand) are supposed (without any distribution function). Accord-
e¼1 i¼1 b¼1 g¼1
ingly, the envelope bound is used.
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

∀t∈T ð13Þ The aim of employing the IGDT method is to assist the decision
maker (DM) in setting the decision variables to the values in such a
Accordingly, the retailer’s objective function ½Jðd; xÞ is as way to avoid the risk of not obtaining the minimum necessities
follows: while there are uncertainties of uncontrollable parameters. The per-
T X
X E X I formance requirements that explain the requirement from the
Max profit∶ ρðe; iÞD̄ðe; i; tÞ − λðtÞPP ðtÞ objective function can be shown in terms of profit function. The
t¼1 e¼1 i¼1 robustness and opportunity functions are used to evaluate these
X
B X
G  necessities.
− γðb; tÞPB ðb; tÞkðtÞ − πðgÞEG ðg; tÞ ð14Þ An info-gap robustness function, μ̂ðd; J c Þ, would be interpreted
b¼1 g¼1 as the maximum value that μ takes in such a way that the minimum
required or desired profit of the problem indicated by J c, is met,
i.e.
n o
IGDT Analysis μ̂ðd; J c Þ ¼ max μ∶ min Jðd; xÞ ≥ J c ð16Þ
μ ~
x∈Uðμ;xÞ
The common problems associated with decision making are when
the reliable data to make the base decisions is spare or when the In fact, this function expresses the risk-aversion aspect of the
problem parameters (i.e., pool price and demand in this paper) are retailer strategy. Jðd; xÞ is the retailer’s objective function (profit).
poorly understood. In this regard, the IGDT is developed as a This variable is a function of the decision variables (d) and uncer-
powerful tool to make decisions under unstructured uncertainty tain parameters (x). The opportunity function, β̂ðd; J w Þ, indicates
without needing the membership or probability density functions. the possibility of getting a higher profit derived from desired
Ben-Haim (2006) gave a complete review on uncertainty modeling variations of uncertain parameters. This function is a kind of invio-
approaches for power system studies that makes sense of the lability against windfall profit and shows the risk-taking possibility
strengths and weaknesses of different methods. Methods were of an optimal strategy for a retailer. In other words, in the com-
classified into different categories, including probabilistic, possibil- petitive market environment, knowing the behavior of the other
istic, joint possibilistic-probabilistic, IGDT, robust optimization participants can increase the benefit of each participant. Conse-
(RO), and interval analysis. IGDT measures the information gap quently, the strategy of each participant is a kind of secreting
between the predicted and actual values of an uncertain parameter for other participants. For that reason, all market players should
to provide a robust strategy against losses and an opportunistic make their strategy with lack of knowledge about other competi-
strategy to windfall benefits, without sacrificing the performance tors. Accordingly, the info-gap model can provide this opportunity
requirements. to have a desired level of profit in the presence of uncertainties.
The corresponding formulation can be expressed by the following
minimization problem:
Process Model and Performance Requirements n o
The prior information about the uncertain variable is included in the β̂ðd; J w Þ ¼ min μ∶ max Jðd; xÞ ≥ J w ð17Þ
μ ~
x∈Uðμ;xÞ
IGDT model of the uncertainty. The gap between unknown and
known values as a function of known parameters (i.e., the fore- The opportunity function [denoted here by β̂ðd; J w Þ] is defined
casted value of the uncertain parameter, standard deviation, etc.) to evaluate the possibility of achieving a higher profit resulting
can be expressed by the uncertainty model. The IGDT consists from desired variations of the uncertain parameter. J w is the tar-
of several models to present the uncertainty of parameters. In this geted higher profit that may be achieved if the uncertain parameter
paper, the envelope-bound model is taken from the work of Zare
et al. (2010), as follows: favorably deviates from the forecast value by at least β̂. In fact,
   β̂ðd; J w Þ is the minimum value of the decision that can endure
 x − x~  and still obtain the high profit denoted by J w.
~
Uðμ; xÞ ¼    ≤ μ; μ ≥ 0 ð15Þ
x~ 
IGDT Model for the Retailer
where μ = uncertainty horizon of parameter x and x~ = predicted
value of x. An envelope-bound model (Fig. 2) is used in which Considering a midterm horizon, an IGDT-based model has been
the magnitude of deviation is proportional to the forecasted value. developed for a retailer. To evaluate the robustness of decisions,
Another uncertainty model is the weighted mean-square error a risk-averse model is proposed against conditions of low demand
(WMSE) (Ben-Haim 2006; Zare et al. 2011). The WMSE model, and high pool prices, in which decisions are robust against low
unlike the envelope-bound model, assumes that the pool prices profits. Furthermore, the retailer is able to analyze its selling price
have normal distribution functions with estimated average values and portfolio strategy from the results of the proposed method.

© ASCE 04017030-4 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


Robust Strategy μ̂ðd; J c Þ ¼ maxðα or ζÞ
T X
A risk-averse retailer tends to schedule so that it is not exposed to
X E X I
losses or low profit attributable to undesired deviations in pool pri- subject to ~
ρðe; iÞð1 − ζÞDðe; i; tÞ
ces and demand from the forecasted value. Thus, the robust func- t¼1 e¼1 i¼1
tion is stated as X
B
n o ~
− ð1 þ αÞλðtÞP P ðtÞ − γðb; tÞPB ðb; tÞkðtÞ
μ̂ðd; J c Þ ¼ max μ∶ min Jðd; xÞ ≥ J c ¼ ð1 − σÞJ 0 ð18Þ b¼1
μ ~
x∈Uðμ;xÞ
X
G 
− πðgÞEG ðg; tÞ ≥ ð1 − σÞJ 0
J 0 is considered the maximum profit on the basis of the fore- g¼1
casted prices and demand, i.e., the DM is performed merely based Eqs: ð1Þ–ð13Þ ð20Þ
on the forecasted prices and risk is neglected; σ indicates a profit
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

deviation index that is used to model the minimum desired profit by Solving the optimization problem stated in Eq. (20) leads to the
the retailer. In a risk-averse strategy, the objective of the IGDT robust strategy based on critical profit, J c . This means that the sol-
model is to maximize the uncertainty parameters (α and ζ) while ution ensures a minimum profit of J c provided that all absolute
the required performance is met relative forecast errors are less than μ̂, which is the maximized tol-
erable error. Obviously, if the market prices are increased, the
μ̂ðd; J c Þ ¼ maxðα or ζÞ cost of buying energy increases as well. If the production (self-
X T X E X
I generation plus bilateral contracts) is greater than the predicted
subject to Min ρðe; iÞDðe; i; tÞ − λðtÞPP ðtÞ load, and if the market price is high, the retailer is willing to sell
t¼1 e¼1 i¼1 its excess energy to the market (if there is a buyer). On the other
X
B X
G  hand, if the market price is low, the retailer is obliged to reduce its
− γðb; tÞPB ðb; tÞkðtÞ − πðgÞEG ðg; tÞ ≥ ð1 − σÞJ 0 production facilities to avoid losses.
b¼1 g¼1 As stated, Jðd; xÞ is the retailer’s objective function (profit).
~ ≤ λðtÞ ≤ ð1 þ αÞλðtÞ
ð1 − αÞλðtÞ ~ This variable is a function of the decision variables (d) and uncer-
tain parameters (x). So, J c (desired profit) is a controllable param-
~
ð1 − ζÞDðe; ~
i; tÞ ≤ Dðe; i; tÞ ≤ ð1 þ ζÞDðe; i; tÞ eter via σ [profit deviation index, Eq. (18)]. So, μ is the robust
Eqs: ð1Þ–ð13Þ ð19Þ objective function (variable) and μ (α or ζ) cannot be set to zero
or another value.
The risk level of the retailer’s planning can be controlled in this Opportunistic Strategy
model through a confidence level, J c ¼ ð1 − σÞJ 0 , which is deter- Electricity markets may experience unexpected price falls and de-
mined by the retailer according to its risk management policy. mand spikes. A risk-seeking retailer tends to take advantage of
There are various midterm strategies that the retailer can select ac- such desirable variations utilizing an opportunity function. Using
cording to its market situation, competitors’ status, and position. Eq. (17), the opportunity function can be written in mathematical
These strategies include the (1) costs strategy, (2) differentiation form as
strategy, and (3) focus strategy. In the cost strategy (or invasive
n o
strategy), retailers are looking to increase their market share and
β̂ðd; J w Þ ¼ min μ∶ max Jðd; xÞ ≥ J w ¼ ð1 þ τ ÞJ 0 ð21Þ
cost superiority over other competing companies. Important tools μ ~
x∈Uðμ;xÞ
in implementing this policy are the price reduction of energy sell-
ing, new and exciting products, and advertising and marketing in- where J w = target profit that the retailer expects to get, if there are
vestment, as well as increments in efficiency. In the differentiation price falls and demand spikes, which is a factor of the expected
strategy, the retailer offers unique services that will create added profit J 0 through τ . Target profit J w is usually greater than J c .
value for clients. In the focus strategy (defensive strategy), the The opportunity function can be stated using a similar formulation
retailer focuses on a small and narrow part of the retail market.
It offers differentiated services and makes added value for its clients β̂ðd; J c Þ ¼ minðα or ζÞ
in order to benefit from their loyalty and to provide services to X T XE X
I
them. Small companies often choose to take a defensive strategy
subject to Max ρðe; iÞDðe; i; tÞ − λðtÞPP ðtÞ
because of their limited competitive facilities and resources. This t¼1 e¼1 i¼1
is acceptable in the short term. With increasing competition, defen-
X
B X
G 
sive strategies cause clients to decrease in the future and the market − γðb; tÞPB ðb; tÞkðtÞ − πðgÞEG ðg; tÞ ≥ ð1 þ τ ÞJ 0
to shrink unless the retailer uses other competitive market advan- b¼1 g¼1
tages. So, the determining factor (σ) depends entirely on the retail-
~ ≤ λðtÞ ≤ ð1 þ αÞλðtÞ
ð1 − αÞλðtÞ ~
er’s risk management policy.
The pool price and demand are related to each other. But this ~
ð1 − ζÞDðe; ~
i; tÞ ≤ Dðe; i; tÞ ≤ ð1 þ ζÞDðe; i; tÞ
correlation must be modeled before formulation (in the forecasting
process). In IGDT, it is assumed that there is an initial forecast of Eqs: ð1Þ–ð13Þ ð22Þ
pool price and demand (without making assumptions about their
structure). By the given uncertainty μ, the maximum profit seems likely to
With the given uncertainty μ, the minimum profit in Eq. (19) take place for the lowest price and greatest demand permitted by the
seems likely to take place for the highest price and the lowest de- IGDT model at the horizon of the uncertainty μ that are equal to
mand at the horizon of uncertainty μ that is equal to ð1 þ αÞλðtÞ ~ ~ and ð1 þ ζÞDðe;
ð1 − αÞλðtÞ ~ i; tÞ, respectively. Thus, as illustrated
~
and ð1 − ζÞDðe; i; tÞ. Thus, the aforementioned problem can be ex- in Eq. (23), the opportunity model would be stated in a simplified
pressed in the form of the following equation: form as a single-objective optimization problem

© ASCE 04017030-5 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


Table 2. Computational Size of the Problem profit for the retailer. This procedure is also followed in converting
Items Number Eqs. (17)–(23) in which the price is lower than the forecasted value.
The above description is true in the case of the worst-case scenario
Binary variables EI þ B
for the load.
Real variables ðG þ B þ 1ÞT þ EI þ 2
Constraints 2B þ 2TðG þ 1Þ þ Eð2I þ 1Þ þ 2
The optimization solution algorithm can be found in detail in
Appendix II.

β̂ðd; J c Þ ¼ minðα or ζÞ Case Study


XT X E X I
subject to ~
ρðe; iÞð1 þ ζÞDðe; i; tÞ The proposed model has been implemented in a case study where
t¼1 e¼1 i¼1 contractual period was assumed to be 1 month. The opportunity
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

X
B problems and robustness are considered as a mixed integer nonlin-
~
− ð1 − αÞλðtÞP P ðtÞ − γðb; tÞPB ðb; tÞkðtÞ ear programming (MINLP) problem. The MINLP optimization
b¼1 problem is then solved in GAMS software by a DICOPT solver.
X
G  The size of the problem is presented in Table 2 in terms of variables
− πðgÞEG ðg; tÞ ≥ ð1 þ τ ÞJ 0 and constraints.
g¼1

Eqs: ð1Þ–ð13Þ ð23Þ Data


A stepwise price-quota curve including 100 steps of the load pro-
Therefore, a higher profit of J w would be obtained provided that vided by the retailer has been assumed [as done by Carrión et al.
the future price and demand desirably deviate from the forecasted (2007)], considering one type of costumer. Demand is divided into
values by β̂, denoting the minimum needed desirable variations of three levels including peak (P), shoulder (S), and valley (V).
price that make J w obtainable. Twelve bilateral contracts are provided as the input of the simula-
Regarding Eq. (16) [also Eqs. (18) and (19)], the worst-case sce- tion. The best estimated load profiles of the retailer (MWh) and
nario for the market price (and load) is considered. As mentioned price data ($=MWh) are presented in Fig. 3.
previously, to model the uncertainty of parameters, the envelope-
bound model is selected with the structure depicted in Fig. 2.
Robust Strategy
In this structure, the variable upper and lower bounds,
~
i.e., ð1 − αÞλðtÞ ~
and ð1 þ αÞλðtÞ, are considered for the price. In this section three different robust strategies are tested against
Based on this band, as much as the market price is low [e.g., the (1) price uncertainty, (2) demand uncertainty, and (3) simultaneous
forecasted value is higher than the real value (lower bound)], the price and demand uncertainty.
cost of purchasing electricity would be decreased and vice versa. Having taken into consideration the price data, the proposed
In the optimization problem, as the two boundaries can be inves- problem, i.e., the profit maximization problem, is simulated based
tigated, therefore Eq. (16) [or Eq. (19)] is considered for the on Eq. (14). The value of the expected optimal profit, J 0 , is
worst-case situations in which the market price is higher than $2,528,400. Afterward, Case 1 (ζ ¼ 0) of the optimization problem
the forecasted value (lower profit from the expected value). Thus, Eq. (20) is solved considering different values for σ, resulting in
the method is seeking the solution to be robust against the price different critical profits J c . For instance, if σ ¼ 0.43, the value
increasing. Converting Eqs. (19) and (20) shows that the higher of the critical profit would be ensured at J c ¼ $1,769,880, provided
market price (higher than the forecasted value) can lead to a lower that none of the deviations is greater than μ̂ ¼ 37%. This means
$/MWh
MWh

3050 60
Demand price

2850 50

2650 40

2450 30

2250 20

2050 10

1850 0
1 6 11 16 21 26 31 36 41 46 51 56 61 66 71 76 81
Period

Fig. 3. Predicted pool price and load profile

© ASCE 04017030-6 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


0.45 Robustness value, (d,Jc)

0.4

0.35

0.3

0.25

0.2

0.15
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

0.1

0.05

0
1,517,040 1,643,460 1,769,880 1,896,300 2,022,720 2,149,140 2,275,560 2,401,980 2,503,116
Profit level, Jc($)

Fig. 4. Robustness curve (Case 1)

that if the actual prices are within this range, the profit obtained by the high robustness of the strategy taken by the retailer. Purchasing
the retailer would be at least J c ¼ $1,769,880. In the risk-neutral from the pool increases with increasing J c (Fig. 5). However, pur-
case (without any errors), the retailer should purchase 9.8% of its chase from the bilateral contracts and generation units is increased
electricity from bilateral contracts, 79.9% from the pool, and 10.3% with J c . These are such a logical result denoting that in the case of
from self-generating in which σ ¼ 0. Also, the selling price offered choosing a more robust strategy by the retailer, it is desired to pur-
by the retailer to the clients should be $72/MWh. If the retailer chase more electricity from the sources with no uncertainty in price.
takes the risk-averse strategy, the results obtained from the simu- To attract as much consumption as possible, the retailer offered a
lation of the robustness function will be helpful. lower selling price to the customers as the retailer is interested in
Fig. 4 demonstrates the results of the optimum robustness func- purchasing more from the pool (lower σ). The selling price in-
tion values μ̂ versus profit level (J c ) derived from solving the creases with a risk-averse strategy, because the percentage of the
optimization problem Eq. (20) considering σ ¼ 0 − 0.6. Fig. 3 retailer’s customers is lower. Consequently it is less profitable,
shows that σ ¼ 0 (J c ¼ $2,528,400) relates to the zero robustness and the retailer is concerned solely with a safety margin.
(μ̂ ¼ 0). In Case 2, demand deviation is only intended as the source of
The robustness parameter starts to rise as σ increases, implying uncertainty (α ¼ 0). Solving Eq. (20), the optimum decisions in
that a higher deviation from the predicted price can be endured with robust strategy are achieved in which σ is changed from 0 to
the cost of lower profit expectations. Deciding on high robustness 45%. As already defined, the value of ζ demonstrates the uncer-
by the retailer leads to a low profit. Similarly, low profit of the tainty of the demand. Fig. 6 demonstrates the robustness function
retailer demonstrates the state of being more risk averse and also values (μ̂) versus profit level (J c ) such that μ̂ reaches its maximum

90 Electricity procurement (%) Bilateral contracts Pool Self generation

80

70

60

50

40

30

20

10

0
1,517,040 1,643,460 1,769,880 1,896,300 2,022,720 2,149,140 2,275,560 2,401,980 2,503,116
Profit level, Jc($)

Fig. 5. Optimal procurement from different sources (Case 1)

© ASCE 04017030-7 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


0.5
Robustness value, (d,Jc)
0.45

0.4

0.35

0.3

0.25

0.2
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

0.15

0.1

0.05

0
2,528,400 2,401,980 2,275,560 2,149,140 2,022,720 1,896,300 1,769,880 1,643,460 1,517,040 1,390,620
Profit level, Jc($)

Fig. 6. Robustness curve (Case 2)

Table 3. Percent of Load Procurement Strategies (Case 2) Table 4 presents the solution for the load procurement strategy.
For Solution 6, the load procurement percentages are 60.0, 24.6,
σ Pool Bilateral contract Self-generation and 15.4% for the pool market, bilateral contracts, and self-
(%) μ̂ (%) (%) (%)
generation facility, respectively. In this strategy, the expected profit
0 0 86.4 3.9 9.7 would not be less than 0.8 of the maximum profit, if the price de-
45 0.464 74.6 7.2 18.2 viations are lower than 1.311 times the forecasted price and the
demand deviations are higher than 1.05 times the forecasted
demand.
value, μ̂ ¼ 0.464, in σ ¼ 0.45. Optimal procurement from different
sources is presented in Table 3.
In Case 3, the simultaneous price and demand deviation are con- Opportunistic Strategy
sidered as the sources of uncertainty. First, it is supposed that the To evaluate the profit gained under the opportunity function, in this
retailer chooses σ ¼ 0.2 (J c ¼ $2,022,720), and Eq. (20) is solved section three different strategies are examined, as follows:
for the values of σ in which the maximum value of α is achieved at • Case 4: opportunistic strategy against price uncertainty;
0.234 (ζ ¼ 0). The problem is solved as a multiobjective problem • Case 5: opportunistic strategy against demand uncertainty; and
to generate Pareto-optimal solutions as shown in Fig. 7. Taking into • Case 6: opportunistic strategy considering simultaneous de-
account the particular preference of the application, the most- mand and price uncertainties.
desired and satisfying solution is picked out by utilizing different In Case 4, the value of τ in Eq. (23) varies from 0 to 0.2 to
methodologies such as a fuzzy decision maker or analytic hierarchy discuss the profit gain considering an opportunity function. Fig. 8
process (AHP) method after the Pareto-optimal solutions are ex- depicts the opportunity function values (β̂) versus the target profits
tracted by solving the optimization problem. (J w ). More price deviations from the predicted values are needed

values
0.25

0.2

0.15

0.1

0.05

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
values

Fig. 7. Pareto front of robustness strategy for σ ¼ 0.2 (Case 3)

© ASCE 04017030-8 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


Table 4. Percent of Load Procurement Strategies to have a higher desired target profit. For instance, to get a profit,
Bilateral 20% more than J 0 (J w ¼ $3,034,080) the prices have to be at least
Solution Pool contract Self-generation 12.9% lower than the forecasted prices. Also, the selling price to
number ζ α (%) (%) (%) the clients by the retailer, ρðeÞ would be $61/MWh.
1 0 0.234 61.0 23.4 15.6
The increase in the profit can be interpreted as an increase in the
2 0.01 0.225 60.6 23.6 15.7 risk. The percentage of purchase from the pool will increase while
3 0.02 0.216 60.5 23.9 15.7 J w is changed by the other options. In fact, choosing a low-cost
4 0.03 0.206 60.1 24.1 15.8 strategy means accepting a higher risk, and so a higher portion
5 0.04 0.196 60.2 24.4 15.5 of load procurements from the sources with uncertain prices. Also,
6 0.05 0.186 60.0 24.6 15.4 to attract more customers, the selling price to the customers de-
7 0.06 0.175 59.7 24.9 15.4 creases (Fig. 9). Fig. 9 shows that the selling price to the customer
8 0.08 0.152 59.1 25.4 15.4 decreases with the risk-taking strategy, because the percentage of the
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

9 0.1 0.129 63.2 21.8 15.0 retailer’s customers is high and consequently it is more profitable.
10 0.12 0.107 69.1 16.2 14.8
In Case 5, the purpose of the retailer is to select an opportunity
11 0.14 0.084 70.4 15.3 14.3
12 0.16 0.06 70.4 15.6 13.9 approach against the demand uncertainty (with α ¼ 0). Solving
13 0.18 0.034 70.4 16.0 13.6 Eq. (23) for different values of τ (changed from 0 to 45%), the
14 0.2 0.008 74.9 12.3 12.9 optimum decisions in opportunistic strategy are achieved for each
Jw . The optimum opportunity function values β̂ versus profit level

0.16 Opportunity value, (d,Jw)

0.14

0.12

0.10

0.08

0.06

0.04

0.02

0.00
2,528,400 2,578,968 2,629,536 2,680,104 2,730,672 2,755,956 2,781,240 2,831,808 2,907,660 3,034,080
Profit level, Jw($)

Fig. 8. Opportunity function (Case 4)

74 ($/MWh)
72
70
68
66
64
62
60
58
56
54
2,528,400 2,578,968 2,629,536 2,680,104 2,730,672 2,755,956 2,781,240 2,831,808 2,907,660 3,034,080
Profit level, Jw($)

Fig. 9. Selling price to customers (Case 4)

© ASCE 04017030-9 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


0.5 Opportunity value, (d,Jw)
0.45

0.4

0.35

0.3

0.25

0.2
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

0.15

0.1

0.05

0
2,528,400 2,654,820 2,781,240 2,907,660 3,034,080 3,160,500 3,286,920 3,413,340 3,539,760 3,666,180
Profit level, Jw($)

Fig. 10. Opportunity function (Case 5)

(J w ) are found and presented in Fig. 10. The value of β̂ begins from region. For a risk-seeking retailer, the proposed opportunistic
0 in τ ¼ 0 and reaches its maximum value, β̂ ¼ 0.464 in τ ¼ 0.45. formulation enables the retailer to benefit from unpredictable
Optimal procurement from different sources is presented in Table 5. price falls (Case 4) and demand spikes (Case 5).
In Case 6, the purpose of the retailer is to select an opportunity • The results demonstrate that in the robust strategy (related to a
strategy against the simultaneous price and demand uncertainty lower profit), purchases from sources with an uncertain price
for a target profit (J w ). First, if the retailer select the τ ¼ 0.15 (pool market) are decreased. Furthermore, the selling price
(J w ¼ $2,907,660), then Eq. (23) can be solved for the given value is decreased to attract more clients, because the percentage
of τ and it will try to find the minimum value of α. For this case, of a retailer’s customers is high and consequently it is more
α ¼ 0.118 (with ζ ¼ 0), so the problem is solved as a multiobjec- profitable. Unlike the robust strategy, in the risk-seeking strat-
tive problem. The Pareto-optimal front is shown in Fig. 11. Also, egy, purchases from sources with uncertain prices is increased,
Fig. 12 shows the selling prices to the customers (τ ¼ 0.15). which provides opportunities for windfall benefits. Further-
more, the selling price is increased.
• The results show that in Case 2 (robust strategy with demand
Discussion fall), with more deviation in the demand, purchases from the
IGDT methods have practical advantages over other methods (such pool market are decreased (i.e., a 45% decrease in load results
as scenario-based methods and fuzzy logic–based models). In the in a 15.8% decrease in pool purchases with 46.4% robustness).
case of risk-based planning of the retailer, an IGDT-based model Also, in the opportunistic strategy related to Case 5, with in-
determines optimal schedules in order to achieve a target profit, creasing demand, the purchase from the pool market is in-
whereas scenario-based methods find optimal schedules based creased (i.e., a 45% increase in load results in a 4.7%
on a limited number of possible price (and load) scenarios. In increase in pool purchases with 46.4% opportunity).
addition, unlike scenario-based models, IGDT-based methods
guarantee a predefined level of profit. The different scenario gen-
eration methods can result in different solutions, especially in the 0.14 values
stochastic methods. Also, in scenario generation methods, the un- 0.12
certain parameter should have a probability density function or a
specific membership function. In contrast, in the proposed method, 0.1
there is no assumption regarding the above issues. Therefore, the
results of these two methods are not comparable. Accordingly, 0.08
based on the obtained results, it can be discussed as follows:
0.06
• In the proposed strategy, the decision remains immune to a mini-
mum critical profit attributable to future price spikes (Case 1) 0.04
and demand falls (Case 2) within a maximized robustness
0.02

Table 5. Percent of Load Procurement Strategies (Case 5) 0


0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
τ Pool Bilateral contract Self-generation values
(%) β̂ (%) (%) (%)
0 0 86.4 3.9 9.7 Fig. 11. Pareto optimal front of opportunity with price and demand
45 0.464 90.7 2.6 6.7 uncertainty for τ ¼ 0.15 (Case 6)

© ASCE 04017030-10 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


($/MWh)
74 α=0.045 α=0.012

72 α=0.062

α =0.028 α=0
70

68

66

64 α=0.091
α=0.118
α=0.079
62
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

60

58

56
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16
values

Fig. 12. Selling price to customers (Case 6)

Conclusion and Future Work equality and inequality constraints, respectively. fðX; dÞ describes
the relations between the decision variables (d) and input uncertain
This paper proposed an IGDT-based strategy for the retailer con- parameters (X). In the case of the uncertain input parameters, X
sidering uncertain demand and future electricity market prices. values are equal to their predicted values (X ¼ X̄). Then solving
Considering a self-generating unit, bilateral contracts, and market Eq. (24) gives the predicted value of y ¼ ȳ. However, if the value
pool as the purchase sources, the proposed technique helps the of X is unknown, then the IGDT method tries to find a solution for
retailer to recognize its purchase and selling-price strategy while the problem that is robust against the error in predicting the value of
evaluating its robustness against the high pool price and low de- X. In the IGDT, the robustness is defined as the immunity of sat-
mand (low-profit situation). The effectiveness of the proposed isfaction of a predefined constraint. The robustness of a decision d
framework has been tested in realistic examples. based on the requirement lc , α̂ðd; lc Þ, is defined as the maximum
Also, the proposed model considers the uncertainty of pool pri- value of α at which the decision maker is sure that the required
ces and demand. However, examining the effects of a renewable constraints are always satisfied, as follows (Soroudi and Amraee
portfolio (Zhou and Liu 2014), the bilateral contract price and 2013a; Aien et al. 2016):
the capacity of the bilateral contract are also subject to uncertainty.
The proposed formulation can be enhanced if call options and in- max α̂ðd; lc Þ
terruptible contracts are available for the retailer. Also, in future
∀ X ∈ Uðα; X̄Þ
work, based on the IGDT concept, other pricing schemes (such
as time-of-use pricing) can be considered. The load modeling issue fðX; dÞ ≤ lc
is an important challenge that should be considered in the time-of- lc ¼ ð1 þ ζÞ × ȳ
use pricing. If the selling price is too high, the clients could maybe
shift their loads or choose another rival retailer, for which the HðX; dÞ ¼ 0
chosen strategy is completely different. GðX; dÞ ≥ 0 ð25Þ

Appendix I. Comparison among IGDT, RO, where ζ = degree that the decision maker tolerates the deterioration
and Min-Max Regret Optimization of the objective function because of the forecasting error of the
input parameter X.
There are similarities and differences among the IGDT technique,
robust optimization, and min-max regret optimization in determining
Robust Optimization Formulation
the decision maker’s strategy. The general formulations of these three
models have been presented for comparison purposes as follows: Consider a function like z ¼ fðX; yÞ, which is linear in X and non-
linear in y. The values of X are subject to the uncertainty while the
IGDT Formulation values of y are known. The uncertainty of X is modeled with an
uncertainty set X ∈ UðXÞ, where UðXÞ is a set that parameter X
Consider a typical optimization function as follows: can take value from it. The maximization of z ¼ fðX; yÞ can be
formulated via the following formulation:
y ¼ minfðX; dÞ
HðX; dÞ ¼ 0 max z ¼ fðX; yÞ
GðX; dÞ ≥ 0 ð24Þ X ∈ UðXÞ ð26Þ

where X = vector of input parameters (which are subject to severe Since the value of z is linear with respect to X, it can be refor-
uncertainty) and d = vector of decision variables. H and G are the mulated as follows:

© ASCE 04017030-11 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


max z T X
X E X
I
profit ¼ ρðe; iÞD̄ðe; i; tÞ − λðtÞPP ðtÞ
~ yÞ
z ≤ fðX; t¼1 e¼1 i¼1
~ yÞ ¼ AðyÞ × X~ þ gðyÞ X
B X
G 
fðX;

− γðb; tÞPB ðb; tÞkðtÞ − πðgÞEG ðg; tÞ ð30Þ
X~ ∈ UðXÞ ¼ fXjjX − X̄j ≤ X g ð27Þ b¼1 g¼1

⌢ For the sake of simplicity, the above equation can be stated as


~ X̄; X = uncertain value, predicted value, and maximum
where X;
possible deviation of variable X from X̄, respectively. The robust P½λðtÞ; Dðe; i; tÞ ¼ max fI½ρðe; iÞ; Dðe; i; tÞ − C½λðtÞ; Xg
optimization seeks a solution that not only maximizes the objective ρðe;iÞ;X

function z but also assures the decision maker that if there exist ð31Þ
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

some prediction errors about the values of X, the z remains opti-


mum with high probability. The robust counterpart is defined as In which P, I, and C = profit, income, and cost, respectively.
follows: Also, X is the vector of decision variables.
For a given decision the regret, Reg½ρðe; iÞ; X of a retailer is
max z defined as
~ yÞ
z ≤ fðX; Reg½ρðe; iÞ; X∶ ¼ P½λðtÞ; Dðe; i; tÞ
X ⌢
fðX; yÞ ¼ AðyÞ × X̄ þ gðyÞ − max ai ðyÞ × X i × wi − fI½ρðe; iÞ; Dðe; i; tÞ − C½λðtÞ; Xg ð32Þ
i
X
wi ≤ Γ In the above definition, the maximum total profit is evaluated
i by Eq. (31) under perfect information (i.e., the profit obtained
0 ≤ wi ≤ 1 ð28Þ if the retailer had known the electricity price and demand before
making any decision). On the other hand, I½ðe; iÞ; Dðe; I; tÞ −
C½λðtÞ; X represents the total profit corresponding to the decision
Γ = degree of conservativeness and y = set of decision variables. ½ρðe; iÞ; X under the realized electricity price and demand.
Based on Eq. (28), two nested optimization problems are to be Reg½ρðe; iÞ; X measures the loss of profit attributable to imperfect
solved. Inserting a dual form of Eq. (28) information of the uncertain parameters. Under the min-max regret
criterion, the objective function is to minimize the largest regret
value Reg½ρðe; iÞ; X over all possible scenarios as follows:
max z
~ yÞ
z ≤ fðX; min max Reg½ρðe; iÞ; X ð33Þ
½ρðe;iÞ;X λ;D
X
fðX; yÞ ¼ AðyÞ × X̄ þ gðyÞ − Γβ − ξi
i

β þ ξ i ≥ Aðyi Þ × X i ð29Þ Appendix II. ε-Constraint Method

A multiobjective problem simultaneously maximizes a number of


objective functions over a set of feasible solutions, Ω (Aghaei et al.
Min-Max Regret Optimization
2011)
Even though the robust optimization approach has been widely
used in power system operations and other fields, it is often con- maxfðxÞ ¼ ½f 1 ðxÞ; : : : ; fm ðxÞT ð34Þ
sidered to be overconservative, e.g., achieving robustness with a
significant profit reduction. In the min-max regret approach, the such that
uncertain parameter describes using the confidence intervals as that
described in the robust optimization approach, and aims to mini- x∈Ω
mize the maximum regret (Savage 1951; Aissi et al. 2009). By ap- Gk ðxÞ ¼ 0
plying the min-max regret approach, the objective is to obtain a
solution that minimizes the worst-case regret over all possible sce- k ¼ 1; : : : ; K
narios while ensuring system robustness. That is, for the same given Hz ðxÞ ≤ 0
risk level (e.g., for the same given price uncertainty set), by using
the min-max regret approach, a risk-averse market participant will z ¼ 1; : : : ; Z ð35Þ
get the minimum regret as compared to the robust and stochastic
optimization approaches under the worst-case scenario. Mean- where x = set of controllable quantities; fðxÞ = vector of the ob-
while, the min-max regret approach can provide computationally jective functions; and Gk ðxÞ and Hz ðxÞ = constraints.
less conservative objective value, i.e., larger expected profit, In this field, the optimality concept is replaced with the Pareto
than the worst-case robust optimization approach for the same optimality. With Pareto-optimal solutions (nondominated or nonin-
given uncertainty set (Aissi et al. 2009). Therefore, the min-max ferior) it is not possible to increase the performance of one objective
regret approach is less conservative than the robust optimiza- function without influencing negatively certain others. In this pa-
tion approach and more robust than the stochastic optimization per, noninferior solutions are calculated by using the ε-constraint
approach. method (Aghaei et al. 2011), in which one objective function (m) is
As mentioned in the text, the retailer’s profit can be abstracted as selected and the others become new constraints. Accordingly, the
follows: formulation becomes as follows:

© ASCE 04017030-12 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


maxfm ðxÞ ð36Þ α= horizon of the uncertain pool price;
such that γðb; tÞ = energy price of contract b at time t ($=MWh);
λðtÞ = pool price at time t ($=MWh);
fw ðxÞ ≥ εw ~ =
λðtÞ estimated pool price at time t ($=MWh);
w ¼ 1; : : : ; M μ= horizon of the uncertain variable;
and πðgÞ = cost associated with block g of the self-generating
facility ($=MWh);
w≠m ð37Þ ρðeÞ = selling price to client group e ($=MWh);
with the previous constraints (36)–(38). ρðe; iÞ = price of the i interval of the price-quota curve for client
Moreover group e ($=MWh);
ρ̄ðe; iÞ = upper limit of the price in block i of the price-quota
εw ¼ εw − Δεw
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

ð38Þ curve pertaining to the client group e ($=MWh); and


where εw = lower limit of fw ðxÞ; εw = global noninferior value of ζ= horizon of the uncertain electricity demand.
f w ðxÞ; and Δεw = trade-off presence of the decision maker.
A payoff table can finally be introduced, representing the results
from the individual optimization of each objective function. References
Aghaei, J., et al. (2016). “Optimal robust unit commitment of CHP plants in
Notation electricity markets using information gap decision theory.” IEEE Trans.
Smart Grid, 1–9.
The following symbols are used in this paper: Aghaei, J., Shayanfar, H. A., and Amjady, N. (2011). “Multi-objective elec-
B = number of bilateral contracts; tricity market clearing considering dynamic security by lexicographic
optimization and augmented epsilon constraint method.” Appl. Soft
b = set of bilateral contracts;
Comput., 11(4), 3846–3858.
CB = bilateral contracts cost ($); Ahmadi, A., Charwand, M., and Aghaei, J. (2013). “Risk-constrained op-
CG = energy operation cost of the generating unit ($); timal strategy for retailer forward contract portfolio.” Int. J. Electr.
CP = net cost of trading in the pool ($); Power Energy Syst., 53(4), 704–713.
Dðe; tÞ = energy supplied by the retailer to client group e in Aien, M., Hajebrahimi, A., and Fotuhi-Firuzabad, M. (2016). “A compre-
period t (MWh); hensive review on uncertainty modeling techniques in power system
studies.” Renew. Sust. Energy Rev., 57(1), 1077–1089.
D̄ðe; i; tÞ = estimation of the energy associated with block i of the
Aissi, H., Bazgan, C., and Vanderpooten, D. (2009). “Min-max and min-
price-quota curve of client group e in period t (MWh);
max regret versions of combinatorial optimization problems: A survey.”
E = number of client groups; Eur. J. Oper. Res., 197(2), 427–438.
EG ðg; tÞ = energy pertaining to block g of the self-generating Ben-Haim, Y. (2006). Information gap decision theory, 2nd Ed., Academic,
facility at time t (MWh); San Diego.
Emax
G ðgÞ = size of production block g (MWh); Boroumand, R. H., and Zachmann, G. A. (2012). “Retailers’ risk manage-
e = set of client groups; ment and vertical arrangements in electricity markets.” Energy Policy,
40(1), 465–472.
G = number of production blocks of the generating facility;
Carrión, M., Arroyo, A. J., and Conejo, A. J. (2009). “A bilevel stochastic
g = set of the generating blocks of the self-generating programming approach for retailer futures market trading.” IEEE Trans.
facility; Power Syst., 24(3), 1446–1456.
I = number of blocks in the price-quota curves; Carrión, M., Conejo, A. J., and Arroyo, A. J. (2007). “Forward contracting
i = set of blocks in the price-quota curves; and selling price determination for a retailer.” IEEE Trans. Power Syst.,
J 0 = expected maximum profit based on the forecasted 22(4), 1050–1057.
prices ($); Charwand, M., Ahmadi, A., Heidari, A., and Esmaeel-Nezhad, A. (2014a).
J c = critical profit ($); “Benders decomposition and normal boundary intersection method for
multi-objective decision making framework for an electricity retailer in
J w = target profit ($); energy markets.” IEEE Syst. J., 9(4), 1475–1484.
Jðd; xÞ = retailer’s profit function ($); Charwand, M., Ahmadi, A., Sharaf, A. M., Gitizadeh, M., and Nezhad,
kðtÞ = duration of period t; A. E. (2016). “Robust hydrothermal scheduling under load uncertainty
PB ðb; tÞ = procured power from bilateral contract b at time t using information gap decision theory.” Int. Trans. Electr. Energy Syst.,
(MW); 26(3), 464–485.
PP ðtÞ = purchased energy from the pool at period t (MWh); Charwand, M., Ahmadi, A., Siano, P., Dargahi, V., and Sarno, D. (2015).
“Exploring the trade-off between competing objectives for electricity
B ðbÞ = maximum power pertaining to contract b (MW);
Pmax
energy retailers through a novel multi-objective framework.” Energy
B ðbÞ = minimum power pertaining to contract b (MW);
Pmin Converse. Manage., 91, 12–18.
Rðe; tÞ = revenue obtained from selling to client group e in Charwand, M., and Moshavash, Z. (2014b). “Midterm decision-making
period t ($); framework for an electricity retailer based on information gap decision
T = number of time periods; theory.” Int. J. Electr. Power Energy Syst., 63, 185–195.
t = set of hours; Feuerriegel, S., and Neumanno, D. (2014). “Measuring the financial
impact of demand response for electricity retailers.” Energy Policy,
x = uncertain parameter;
65(1), 359–368.
Yðe; iÞ = binary variable, which is equal to 1 if the selling price Gabriel, S., Sahakij, P., and Balakrishnan, S. (2004). “Optimal retailer load
offered by the retailer to client group e belongs to block estimates using stochastic dynamic programming.” J. Energy Eng.,
i of the price-quota curve, and 0 otherwise; 10.1061/(ASCE)0733-9402(2004)130:1(1), 1–17.
ZðbÞ = binary variable, which is equal to 1 if bilateral contract GAMS (General Algebraic Modeling System) [Computer software]. GAMS
b is selected, and 0 otherwise; Development Corp., Washington, DC.

© ASCE 04017030-13 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1


Garcia-Bertrand, R. (2013). “Sale prices setting tool for retailers.” IEEE Nojavan, S., Ivatloo, B. M., and Zare, K. (2015b). “Robust optimization
Trans. Power Syst., 4(4), 2028–2035. based price-taker retailer bidding strategy under pool market price un-
Hatami, A. R., Seifi, H., and Sheikh-El-Eslami, M. K. (2009). “Optimal certainty.” Int. J. Electr. Power Energy Syst., 73, 955–963.
selling price and energy procurement strategies for a retailer in an Savage, L. J. (1951). “The theory of statistical decision.” J. Amer. Statist.
electricity market.” Electr. Power Syst. Res., 79(1), 246–254. Assoc., 46(253), 55–67.
He, J., Cai, L., Cheng, P., and Fan, J. (2015). “Optimal investment for retail Shang, J., et al. (2014). “Lower partial moment and information entropy-
company in electricity market.” IEEE Trans. Ind. Informat., 11(5), based dynamic electricity purchasing strategy.” J. Energy Eng., 10.1061
1210–1219. /(ASCE)EY.1943-7897.0000208, 04014027.
Ivatloo, B. M., Zareipour, H., Amjady, N., and Ehsan, M. (2013). Soroudi, A., and Amraee, T. (2013a). “Decision making under uncertainty
“Application of information-gap decision theory to risk-constrained in energy systems: State of the art.” Renewable Sustainable Energy
self-scheduling of Gencos.” IEEE Trans. Power Syst., 28(2), Rev., 28, 376–384.
1093–1102. Soroudi, A., and Ehsan, M. (2013b). “IGDT based robust decision making
Jia, L., and Tong, L. (2016). “Dynamic pricing and distributed energy tool for DNOs in load procurement under severe uncertainty.” IEEE
Downloaded from ascelibrary.org by University of California, San Diego on 05/12/17. Copyright ASCE. For personal use only; all rights reserved.

management for demand response.” IEEE Trans. Smart Grid, 7(2), Trans. Smart Grid, 4(2), 886–895.
1128–1136. Wei, W., Liu, F., and Mei, S. (2015). “Energy pricing and dispatch for smart
grid retailers under demand response and market price uncertainty.”
Kazemi, M., Ivatloo, B. M., and Ehsan, M. (2015). “Risk-constrained stra-
IEEE Trans. Smart Grid, 6(3), 1364–1374.
tegic bidding of GenCos considering demand response.” IEEE Trans.
Wei, W., Liu, F., and Mei, S. (2015). “Offering non-dominated strategies
Power Syst., 30(1), 376–384.
under uncertain market prices.” IEEE Trans. Power Syst., 30(5),
Kettunen, J., Salo, A., and Bunn, D. W. (2010). “Optimization of electricity
2820–2821.
retailer’s contract portfolio subject to risk preferences.” IEEE Trans. Yu, M., and Hong, S. S. (2016). “A real-time demand-response algorithm
Power Syst., 25(1), 117–128. for smart grids: A Stackelberg game approach.” IEEE Trans. Smart
Kharrati, S., Kazemi, M.,, and Ehsan, M. (2016). “Medium-term retailer’s Grid, 7(2), 879–888.
planning and participation strategy considering electricity market uncer- Zare, K., Moghaddam, M. P., and Sheikh-El-Eslami, M. K. (2010).
tainties.” Int. Tran. Electr. Energy Syst., 26(5), 920–933. “Electricity procurement for large consumers based on information
Lim, A. E. B., Shanthikumar, J. G., and Vahn, G. Y. (2011). “Conditional gap decision theory.” Energy Policy, 38(1), 234–242.
value-at-risk in portfolio optimization: Coherent but fragile.” Opera- Zare, K., Moghaddam, M. P., and Sheikh-El-Eslami, M. K. (2011). “Risk-
tions Research Letters, 39(2), 163–171. based electricity procurement for large consumers.” IEEE Trans. Power
Mazer, A. (2007). Electric power planning for regulated and deregulated Syst., 26(4), 1826–1835.
market, Wiley, Hoboken, NJ. Zhou, Y., and Liu, T. (2014). “Impacts of the renewable portfolio standard
Nojavan, S., Ivatloo, B. M., and Zare, K. (2015a). “Optimal bidding on regional electricity markets.” J. Energy Eng., 10.1061/(ASCE)EY
strategy of electricity retailers using robust optimisation approach .1943-7897.0000202, B4014009.
considering time-of-use rate demand response programs under Zugno, M., Morales, J. M., Pinson, P., and Madsen, H. (2013). “A bilevel
market price uncertainties.” IET Gener. Transm. Distrib., 9(4), model for electricity retailers’ participation in a demand response mar-
328–338. ket environment.” Energy Econ., 36, 182–197.

© ASCE 04017030-14 J. Energy Eng.

J. Energy Eng., 2017, 143(5): -1--1

You might also like