An introduction to optimal power flow: Theory, formulation, and … · 2018. 12. 6. · the OPF...

27
Full Terms & Conditions of access and use can be found at http://www.tandfonline.com/action/journalInformation?journalCode=uiie20 IIE Transactions ISSN: 0740-817X (Print) 1545-8830 (Online) Journal homepage: http://www.tandfonline.com/loi/uiie20 An introduction to optimal power flow: Theory, formulation, and examples Stephen Frank & Steffen Rebennack To cite this article: Stephen Frank & Steffen Rebennack (2016) An introduction to optimal power flow: Theory, formulation, and examples, IIE Transactions, 48:12, 1172-1197, DOI: 10.1080/0740817X.2016.1189626 To link to this article: https://doi.org/10.1080/0740817X.2016.1189626 View supplementary material Accepted author version posted online: 21 May 2016. Published online: 11 Aug 2016. Submit your article to this journal Article views: 1911 View Crossmark data Citing articles: 10 View citing articles

Transcript of An introduction to optimal power flow: Theory, formulation, and … · 2018. 12. 6. · the OPF...

  • Full Terms & Conditions of access and use can be found athttp://www.tandfonline.com/action/journalInformation?journalCode=uiie20

    IIE Transactions

    ISSN: 0740-817X (Print) 1545-8830 (Online) Journal homepage: http://www.tandfonline.com/loi/uiie20

    An introduction to optimal power flow: Theory,formulation, and examples

    Stephen Frank & Steffen Rebennack

    To cite this article: Stephen Frank & Steffen Rebennack (2016) An introduction to optimalpower flow: Theory, formulation, and examples, IIE Transactions, 48:12, 1172-1197, DOI:10.1080/0740817X.2016.1189626

    To link to this article: https://doi.org/10.1080/0740817X.2016.1189626

    View supplementary material

    Accepted author version posted online: 21May 2016.Published online: 11 Aug 2016.

    Submit your article to this journal

    Article views: 1911

    View Crossmark data

    Citing articles: 10 View citing articles

    http://www.tandfonline.com/action/journalInformation?journalCode=uiie20http://www.tandfonline.com/loi/uiie20http://www.tandfonline.com/action/showCitFormats?doi=10.1080/0740817X.2016.1189626https://doi.org/10.1080/0740817X.2016.1189626http://www.tandfonline.com/doi/suppl/10.1080/0740817X.2016.1189626http://www.tandfonline.com/doi/suppl/10.1080/0740817X.2016.1189626http://www.tandfonline.com/action/authorSubmission?journalCode=uiie20&show=instructionshttp://www.tandfonline.com/action/authorSubmission?journalCode=uiie20&show=instructionshttp://crossmark.crossref.org/dialog/?doi=10.1080/0740817X.2016.1189626&domain=pdf&date_stamp=2016-05-21http://crossmark.crossref.org/dialog/?doi=10.1080/0740817X.2016.1189626&domain=pdf&date_stamp=2016-05-21http://www.tandfonline.com/doi/citedby/10.1080/0740817X.2016.1189626#tabModulehttp://www.tandfonline.com/doi/citedby/10.1080/0740817X.2016.1189626#tabModule

  • IIE TRANSACTIONS, VOL. , NO. , –http://dx.doi.org/./X..

    An introduction to optimal power flow: Theory, formulation, and examples

    Stephen Franka and Steffen Rebennackb

    aNational Renewable Energy Laboratory, Golden, CO, USA; bColorado School of Mines, Golden, CO, USA

    ARTICLE HISTORYReceived August Accepted April

    KEYWORDSPower flow; optimal powerflow; electric power systemsanalysis; electricalengineering; nonlinearprogramming; optimization;operations research

    ABSTRACTThe set of optimization problems in electric power systems engineering known collectively as OptimalPower Flow (OPF) is one of themost practically important andwell-researched subfields of constrainednon-linear optimization. OPF has enjoyed a rich history of research, innovation, and publication since its debutfive decades ago. Nevertheless, entry into OPF research is a daunting task for the uninitiated—both due tothe sheer volumeof literature andbecauseOPF’s ubiquitywithin the electric power systems community hasled authors to assume a great deal of prior knowledge that readers unfamiliar with electric power systemsmay not possess. This article provides an introduction to OPF from an operations research perspective; itdescribes a complete and concise basis of knowledge for beginningOPF research. The discussion is tailoredfor the operations researcher who has experience with nonlinear optimization but little knowledge of elec-trical engineering. Topics covered include power systems modeling, the power flow equations, typical OPFformulations, and common OPF extensions.

    1. Introduction

    The set of optimization problems in electric power systems engi-neering known collectively as Optimal Power Flow (OPF) is oneof the most practically important and well-researched subfieldsof constrained nonlinear optimization. Carpentier (1962) intro-duced OPF as an extension to the problem of optimal EconomicDispatch (ED) of generation in electric power systems. Carpen-tier’s key contribution was the inclusion of the electric powerflow equations in the ED formulation. Today, the defining fea-ture of OPF remains the presence of the power flow equations inthe set of equality constraints.

    OPF includes any optimization problem that seeks to opti-mize the operation of an electric power system subject to thephysical constraints imposed by electrical laws and engineer-ing limits. This general framework encompasses dozens of opti-mization problems for power systems planning and operation(Zhu, 2009; Zhang, 2010; Frank et al., 2012a). As illustratedin Fig. 1, the optimization of power system operation typicallyoccurs via incremental planning: long-term planning proce-dures make high-level decisions based on coarse system mod-els, whereas short-term procedures refine earlier decisions usingdetailed models but more limited decision spaces. OPF may beapplied to decision-making at nearly any planning horizon—from long-term transmission network capacity planning tominute-by-minute adjustment of real and reactive power dis-patch (Wood and Wollenberg, 1996; Glover et al., 2008; Zhu,2009).

    To date, thousands of articles and hundreds of textbookentries have been written about OPF. In its maturation overthe past five decades, OPF has served as a practical prov-ing ground for many popular nonlinear optimization algo-rithms, including gradientmethods (Dommel andTinney, 1968;

    CONTACT Stephen Frank [email protected] data for this article can be accessed at www.tandfonline.com/uiie.

    Peschon et al., 1972; Alsac and Stott, 1974), Newton-type meth-ods (Sun et al., 1984), sequential linear programming (Stott andHobson, 1978; Alsac et al., 1990), sequential quadratic program-ming (Burchett et al., 1982), both linear and nonlinear interiorpoint methods (Vargas et al., 1993; Granville, 1994; Torres andQuintana, 1998), and semi-definite programming (Lavaei andLow, 2012; Low, 2013). These algorithmic approaches, amongothers, are reviewed in several surveys (Huneault and Galiana,1991; Momoh et al., 1999a, 1999b; Zhang et al., 2007; Zhang,2010; Bienstock, 2013), including one recently published by theauthors Frank et al. (2012a, 2012b).

    Although OPF spans operations research and electrical engi-neering, the accessibility of the OPF literature skews heavilytoward the electrical engineering community. OPF has becomesufficiently familiar within the electric power systems commu-nity that the recent literature, including survey papers, assumesa great deal of prior knowledge on the part of the reader.Few papers even include a full OPF formulation, much lessexplain the particulars of the objective function or constraints.Even introductory textbooks (Wood and Wollenberg, 1996;Rau, 2003b; Zhu, 2009) require a strong background in powersystems analysis, specifically regarding the form, construction,and solution of the electric power flow equations. Althoughmany electrical engineers have this prior knowledge, an oper-ations researcher likely will not. We believe that this accessibil-ity gap has been detrimental to the involvement of the opera-tions research community in OPF research; our impression isthat most OPF articles continue to be published in engineeringjournals by electrical engineers.

    What is missing in the literature—and what we providein this introductory article—is a detailed introduction tothe OPF problem from an operations research perspective.

    Copyright © “IIE”

    http://dx.doi.org/10.1080/0740817X.2016.1189626mailto:[email protected]://www.tandfonline.com/uiie

  • IIE TRANSACTIONS 1173

    Planning Horizon

    Decreasing Uncertainty

    Years Months Weeks Days Hours Minutes Real-Time

    Market Clearing /Unit Commitment

    Security-Constrained UnitCommitment

    AutomaticGeneration

    Control

    EconomicDispatch

    Classic OptimalPower Flow

    Increasing M

    odel Resoluti

    on

    Long-termGenerationScheduling

    CapacityExpansionPlanning

    Reactive PowerPlanning

    Security-Constrained UnitCommitment

    Optimal ReactivePower Flow

    VoltageControl

    Reactive Power Dispatch

    Real Power Dispatch

    Figure . Optimization and control procedures for incremental planning of power system operation. Bold text indicates procedures that incorporate variants of optimalpower flow.

    Existing review articles and surveys (Zhang and Tolbert, 2005;Zhang et al., 2007; Bienstock, 2013), including the authors’recent survey Frank et al. (2012a, 2012b), focus heavily on opti-mization theory and tailored OPF solution algorithms. In con-trast, this article emphasizes the electrical engineering theoryand mechanics of the OPF formulation. The goal of this arti-cle is to provide a bridge between OPF theory and practice: in itwe outline the tool set required to understand, formulate, ana-lyze, and ultimately solve a typical OPF problem. Therefore, weaddress many topics given little attention in other introductorymaterials, including the construction of the admittance matrixfor electrical power flow, treatment of advanced controls suchas phase-shifting and tap-changing transformers, a qualitativecomparison of the various forms of the electric power flow equa-tions, and various practical considerations, such as the use of theper unit system.

    As we have written this article for the operations researcher,we assume that the reader has significant experience with non-linear optimization and advanced mathematical concepts butlittle background in electrical engineering. Specifically, this arti-cle requires a foundational understanding of

    � linear algebra (Greenberg, 1998);� complex number theory (Greenberg, 1998; Freitag andBusam, 2009);

    � analysis of differential equations in the frequency domain(Greenberg, 1998); and

    � linear and nonlinear optimization theory and application(Rardin, 1997; Nocedal and Wright, 2006).

    We expect that readers may not possess a working knowl-edge of electrical circuit theory; we therefore provide a brief

    introduction in Appendix B and recommend O’Malley (2011)for further reading. Readers interested in the technical detailsof electric power flow should also consult a good power systemstext such asGlover et al. (2008) orWood andWollenberg (1996).

    We begin in Section 2 with a description of power systemsmodels, including the classic formulation of the OPF prob-lem. Section 3 surveys some common applications of OPF andincludes full formulations for several of the decision processesshown in Fig. 1. Section 4 describes the bus admittance matrix,which is the foundation of the power flow equations. Section5 reviews the various forms of the power flow equations, withan emphasis on describing their relative advantages and dis-advantages. Building on the power flow equations, Section 6introduces OPF solution methods and practical considerations.Section 7 then provides a worked example of a classic OPF for-mulation. Finally, Section 8 concludes the article.

    We also include four Appendices that provide supplemen-tal information regarding electric power systems modeling andanalysis. Appendix A documents the notation used throughoutthe article, Appendix B reviews fundamental electric power sys-tem concepts relevant to OPF, Appendix C summarizes the per-unit system, and Appendix D describes common formats for theexchange of power system data.

    Different readers may have different goals in reviewing thisarticle. We recommend that all readers start with the Intro-duction and Section 2. Readers interested only in a brief OPFoverview can subsequently read Sections 3 and 6.2 and safelyskip the detailed derivations in other sections. Conversely, read-ers attempting to implement or test an OPF solution algo-rithm should read Sections 4 to 7 in detail. Readers interested

  • 1174 S. FRANK AND S. REBENNACK

    primarily in understanding linear approximations for the OPFproblem should review Section 5 and in particular Section 5.3,which discusses DC power flow.

    2. Modeling of power systems

    In this section, we introduce the classic network model for anelectric power system and describe conventional power flowand OPF. The development requires some basic knowledge ofelectrical circuit theory, the frequency domain (phasor) repre-sentation of electrical quantities, and the concept of complexelectric power. Readers unfamiliar with these topics should con-sult Appendix B first for a brief overview.

    2.1. Notation

    Throughout this article, italic roman font (A) indicates a vari-able or parameter, bold roman font (A) indicates a set, and a tildeover a symbol (ã) indicates a phasor quantity (complex number).Letter case does not differentiate variables from parameters; agiven quantity may be a variable in some cases and a parameterin others. Symbolic superscripts are used as qualifiers to differ-entiate similar variables, whereas numeric superscripts indicatemathematical operations. For example, the superscript L differ-entiates load power PL from net power P, but P2 indicates (net)power squared. Where applicable, electrical units are specifiedusing regular roman font. The unit for a quantity follows thenumeric quantity and is separated by a space; for example, 120 Vindicates 120 Volts.

    We use the following general notation for optimizationformulations:

    u vector of control variables (independent decisionvariables);

    x vector or state variables (dependent decision variables);f (u, x) objective function (scalar);g(u, x) vector function of equality constraints;h(u, x) vector function of inequality constraints.

    In order to remain consistent with the existing body of OPFliterature, this article uses notation that follows electrical engi-neering conventions rather than those of the operations researchcommunity. In particular, the symbols e and j represent mathe-matical constants:

    e Euler’s number (the base of the natural logarithm),e ≈ 2.71828; and

    j the imaginary unit or 90° operator, j = √−1.This differs from the use of e as the unit vector and j as an indexas is common in operations research literature. Appendix Aincludes a full listing of the notation used in this article, includ-ing relevant commentary on other notational differences anda listing of electrical engineering units used in power systemsanalysis.

    2.2. Network representation

    Electric power systems may be modeled as a network of elec-trical buses (nodes) interconnected by branches (arcs or edges)that represent transmission lines, cables, transformers, and sim-ilar power systems equipment. Buses represent physical pointsof interconnection among power systems equipment, whereas

    branches represent paths for the flow of electrical current. Thepurpose of an electric power system is to transfer electricalenergy from generation (supply) buses to load (demand) buseselsewhere in the network.

    Buses are referenced by node with index i ∈ N, whereasbranches are referenced as arcs between nodes (i, k) ∈ L, wherei, k ∈ N. The undirected graph (N, L) therefore describes theconnectivity of the electrical network. The number of buses andbranches are N = |N| and L = |L|, respectively.

    Each system bus i has an associated voltage Ṽi, which ismeasured with respect to the system reference (typically Earthground). When connected via the branch network, or “grid,”these voltages induce current in each branch in proportion to thebranch admittance. (Admittance, which is ameasure of how eas-ily a conductor permits the flowof electrical current, is describedin Appendix B.) The most common and concise mathematicaldescription of this phenomenon is the matrix equation

    Ĩ = ỸṼ , (1)in which Ṽ = (Ṽ1, . . . , ṼN ) is an N-dimensional vector ofphasor voltages at each system bus, Ĩ = (̃I1, . . . , ĨN ) is an N-dimensional vector of phasor currents injected into the networkat each system bus, and

    Ỹ =

    ⎛⎜⎜⎝Ỹ11 . . . Ỹ1N...

    . . ....

    ỸN1 . . . ỸNN

    ⎞⎟⎟⎠is theN × N complex bus admittancematrix, which is describedin detail in Section 4. At each bus i, injection current Ĩi repre-sents the net current supplied to the network: generation (sup-ply) minus load (demand). In this framework, voltages Ṽ arestate variables that fully characterize system power flow for agiven matrix Ỹ .

    It is more convenient to work with power flows than cur-rents because (i) injected powers are independent of systemvoltage angle whereas injected currents are not and (ii) work-ing directly with power allows straightforward computation ofrequired electrical energy via integration with respect to time.Therefore, power systems engineers transform Equation (1) byapplying the definition of complex power S = Ṽ Ĩ∗ to each sideof the matrix equation, as described in Appendix B.4. (Here andelsewhere in this article, the symbol ∗ denotes complex conjuga-tion rather than an optimal value; this use is typical in electricalengineering.) The result is the complex power flow equation

    S = Ṽ ◦ (ỸṼ )∗ , (2)in which S = P + jQ is a vector of complex power injections ateach bus and ◦ denotes element-wise vector multiplication. Ateach bus i, the total injected power is the difference between thegeneration SGi and the load S

    Li

    Si = SGi − SLi ⇔ Pi + jQi =(PGi − PLi

    )+ j (QGi − QLi ) .For numerical analysis or optimization, Equation (2) may bedecomposed into a set of equivalent, real-valued, nonlinearpower flow equations by separating its real and imaginary com-ponents (see Section 5).

  • IIE TRANSACTIONS 1175

    2.3. Conventional power flow

    The conventional Power Flow (PF) problem seeks a determin-istic solution to network Equation (2) using numerical analy-sis techniques. Conventional PF is a feasibility problem: thereis no objective function. Rather, the goal is to compute all sys-tem bus voltages and power injections. Here, we summarize thePF problem with polar voltage coordinates, which is the classicrepresentation.

    If each bus voltage is represented in polar form with magni-tude V and phase angle δ, then Equation (2) decomposes intothe set of power flow equations

    Pi (V, δ) = PGi − PLi ∀ i ∈ N, (3)Qi (V, δ) = QGi − QLi ∀ i ∈ N, (4)

    in which net real and reactive power injections Pi and Qi aretrigonometric functions of the system voltages. (See Section 5for the fully expanded equations.) Each systembus has four vari-ables (net real power injection Pi, net reactive power injectionQi, voltage magnitudeVi, and voltage angle δi) and is governedby two equations. Thus, a deterministic solution to the conven-tional PF problem requires fixing the values of two out of fourvariables at each bus.

    In conventional PF, all system buses are assigned to one ofthree bus types.

    Slack Bus: At the slack bus, or swing bus, the voltage magnitudeand angle are fixed and the power injections are free. The pur-pose of the slack bus is twofold. First, it provides a voltagereference (typically V = 1.0 p.u. and δ = 0◦) such that theremaining bus voltages are uniquely determined; we explainthe per unit system “p.u.” in Appendix C. Second, as it is theonly bus at which real power is free to vary, the slack bus isrequired to ensure that the power flow equations have a fea-sible solution. There is only one slack bus in a power systemmodel.

    Load Bus: At a load bus, or “PQ” bus, the power injections arefixed while the voltage magnitude and angle are free. Thereare a fixed number of PQ buses in the system; the symbolMdenotes this number.

    Voltage-Controlled Bus: At a voltage-controlled bus, or “PV”bus, the real power injection and voltage magnitude are fixedwhile the reactive power injection and the voltage angle arefree. (This corresponds to allowing a local source of reactivepower to regulate the voltage to a desired setpoint.) There areN − M − 1 PV buses in the system.

    Assigning buses in this way establishes an equal number ofequations and unknowns. Table 1 summarizes the known andunknown quantities for each of the bus types.

    Once all voltage magnitudes and angles in the system havebeen computed, the remaining power injections are trivial to

    Table . Power system bus types and characteristics for conventional power flow.

    Bus type Slack PQ PV

    Number of buses in system M N − M − 1Known quantities δ,V P,Q P,VUnknown quantities P,Q δ,V δ,QNumber of equations in conventional PF

    evaluate via Equations (3) and (4). Solving the PF thereforerequires determining N − 1 voltage angles (corresponding tothe PQ and PV buses) andM voltage magnitudes (correspond-ing to the PQ buses only). This is done by solving N + M − 1simultaneous nonlinear equations with known right-hand sidevalues. This equation set consists of the real power injectionEquation (3) at each PQ and PV bus and the reactive powerinjectionEquation (4) at eachPQbus. Section 6.1 discusses solu-tion methods for these equations.

    Even though the power flow equations are nonlinear, thereexists only one physically meaningful solution for most powersystems models given an equal number of equations andunknowns. Although other mathematically valid solutionssometimes exist, they have no realistic physical interpretation.(An example would be any solution that returns a negative volt-age magnitude, as magnitudes are by definition non-negative.)Hence, in practice, conventional PF is an exactly determinedproblem.

    2.4. OPF

    OPF combines an objective function with the PF Equations (3)and (4) to form an optimization problem. The presence of thePF equations is the feature that distinguishes OPF from otherclasses of power systems problems, such as classic ED, UnitCommitment (UC), and market-clearing problems.

    Most OPF variants build upon the classic formulation ofCarpentier (1962) and Dommel and Tinney (1968). The clas-sic formulation is an extension of classic ED: its objective is tominimize the total cost of electricity generation while maintain-ing the electric power system within safe operating limits. Thepower system is modeled as a set of buses N connected by a setof branches L, with controllable generators located at a subsetG ⊆ N of the system buses. The operating cost of each gener-ator is a (typically quadratic) function of its real output power:Ci(PGi ). The objective is tominimize the total cost of generation.

    The classic form of the formulation is

    min∑i∈G

    Ci(PGi), (5)

    s.t. Pi (V, δ) = PGi − PLi ∀ i ∈ N, (6)Qi (V, δ) = QGi − QLi ∀ i ∈ N, (7)PG,mini ≤ PGi ≤ PG,maxi ∀ i ∈ G, (8)QG,mini ≤ QGi ≤ QG,maxi ∀ i ∈ G, (9)Vmini ≤ Vi ≤ Vmaxi ∀ i ∈ N, (10)δmini ≤ δi ≤ δmaxi ∀ i ∈ N. (11)

    Constraints (6) and (7) are the PF equations in polar form. Theremaining constraints represent bounds on the system voltagesand powers. Typically, load real and reactive power are fixedwhile generator real and reactive power are control variablessubject to minimum and maximum limits. The voltage magni-tude and angle at the system slack bus (by convention, bus 1) arealso fixed, usually to Ṽ1 = 1.0∠0.

  • 1176 S. FRANK AND S. REBENNACK

    Early methods partition the decision variables into a set ofcontrol variables u (typically the controllable bus power injec-tions) and a set of state variables x (the voltage magnitudes andangles; Dommel and Tinney, 1968; Burchett et al, 1982). Underthis framework, the vector of control variables (independentdecision variables) for the classic formulation is

    u = (PGi:i∈G,QGi:i∈G)and the vector of state variables (dependent decision variables)is

    x = (δ2, . . . , δN,V2, . . . ,VN ) .

    Although not considered in the earliest papers, more recentOPF formulations may include branch current limits and, ifapplicable, box constraints related to the operational limits of PFcontrol devices. These and other side constraints are discussedin Section 5.5.

    2.5. Challenges

    Power systems have evolved significantly since the early days ofOPF, in particular through the addition of advanced controls.Modern grids include control devices that are difficult to incor-porate into OPF formulations: on-load tap changers (Acha et al.,2000), phase shifters (Momoh et al., 2001), series and shuntcapacitors (Momoh et al., 1997), and flexible AC transmissionsystems devices (Xiao et al., 2002). As such controls exert alarge influence on system power flows, they cannot be neglectedin practical OPF formulations. Unfortunately, many academicpapers neglect the modeling of advanced controls and treat onlythe classic formulation, limiting their application in a practicalsetting.

    Nevertheless, some researchers have dedicated considerableeffort to developing accurate but efficient models for advancedcontrols (Acha et al., 2000; Lehmköster, 2002; Xiao et al., 2002).Most such models employ auxiliary power injections at certainsystembuses coupledwith side constraints to enforce power bal-ance. However, even a modest number of such constraints cansignificantly increase OPF problem complexity (Azevedo et al.,2010). Many devices also have discrete control settings, which ifaccurately modeled create a large and intractable Mixed-IntegerNonlinear Programming (MINLP) problem (Soler et al., 2012).A typical approach is to model the control space as contin-uous and round the optimal solution to the nearest discretevalue (Adibi et al., 2003), but this heuristic can yield suboptimalor infeasible solutions (Capitanescu and Wehenkel, 2010; Soleret al., 2012).

    Even without variable phase angles, variable tap ratios, orbranch current limits, the classic OPF formulation is diffi-cult to solve. The power flow constraints (6) and (7) are bothnonlinear and non-convex, and the presence of trigonomet-ric functions complicates the construction of approximations.Moreover, becauseOPF is tightly constrained, local optima oftenprovide only incremental improvements with respect to the sys-tem starting point. Thus, the optimum may only improve a fewpercent upon the base case, as illustrated by numerical results in

    Alsac and Stott (1974), Sun et al. (1984), Granville (1994), andmany other papers.

    For these reasons, OPF problems have historically beensolved using tailored algorithms rather than general-purposesolvers. Although effective for research problems, few such algo-rithms are considered sufficiently reliable for industrial deploy-ment. In particular, solution methods for both conventionalPF and OPF become significantly less reliable when the powersystem is under stress (Bienstock, 2013). Algorithm robust-ness is also a key concern: small changes in the power systemstate may lead to the loss of local feasibility (a critical issue forlocal nonlinear solvers) or large changes in the optimal solution(Almeida and Galiana, 1996). In addition, real power systemmodels include thousands of buses; problems of this size presentcomputational challenges for state-of-the-art nonlinear solvers.

    In actual practice, almost all practical OPF problems aresolved using a linear (DC) power flow approximation (Section5.3; Rau, 2003a; Stott et al., 2009). The results of the DC-OPF arethen used either for contingency analysis with conventional AC-PF or, less commonly, to initialize, or “warm start,” a nonlinearOPF (AC-OPF). Unfortunately, the DC formulation can exhibitsignificant inaccuracy, particularly for heavily loaded transmis-sion lines (Momoh et al., 1997; Stott et al., 2009). Nor is the DCmodel applicable to reactive power dispatch. The search for reli-able AC-OPF solutionmethods therefore remains extremely rel-evant to the power systems community.

    3. Applications of OPF

    In addition to the classic ED formulation, several other OPFvariants are common in both industry and research. Theseinclude Security-Constrained Economic Dispatch (SCED),Security-Constrained Unit Commitment (SCUC), OptimalReactive Power Flow (ORPF), and Reactive Power Planning(RPP).

    3.1. Security-Constrained Economic Dispatch

    SCED, sometimes referred to as security-constrained optimalpower flow, is an OPF formulation that includes power systemcontingency constraints (Alsac and Stott, 1974). A contingencyis defined as an event that removes one or more generators ortransmission lines from the power system, increasing the stresson the remaining network. SCED seeks an optimal solution thatremains feasible under any of a pre-specified set of likely con-tingency events. SCED is a restriction of the classic OPF for-mulation: for the same objective function, the optimal solutionto SCED will be no better than the optimal solution withoutconsidering contingencies. The justification for the restrictionis that SCED mitigates the risk of a system failure (blackout)should one of the contingencies occur.

    SCED formulations typically have the same objective func-tion and decision variables u as the classic formulation, exceptthat the slack bus real and reactive power are considered statevariables, as they must be allowed to change in order for the sys-tem to remain feasible during each contingency.However, SCEDintroduces NC additional sets of state variables x and accompa-nying sets of power flow constraints, where NC is the number of

  • IIE TRANSACTIONS 1177

    contingencies. SCED can be expressed in a general way as

    min f (u, x0),s.t. g0(u, x0) = 0,

    h0(u, x0) ≤ 0,gc(u, xc) = 0 ∀ c ∈ C,hc(u, xc) ≤ 0 ∀ c ∈ C, (12)

    where C = {1, . . . ,NC} is the set of contingencies to consider.Each contingency has a distinct admittance matrix Ỹc, typicallywith less connectivity than the original system. Apart from thecontingency index, f , g, and h are defined as the objective func-tion, equality constraints, and inequality constraints in the clas-sic OPF formulation (5) to (11), respectively.

    For each contingency c ∈ C, the post-contingency powerflowmust remain feasible for the original decision variables u:

    (i) the power flow equations must have a solution;(ii) the contingency state variables xc must remain within

    limits; and(iii) any inequality constraints, such as branch flow limits,

    must be satisfied.One very common contingency set considers independently theloss of each generator and each non-radial transmission line.The resulting SCED solution is said to be N − 1 reliable, sincefor every possible single element contingency c the network hasa feasible solution xc given the optimal dispatch decisions u.

    Typically, the limits on the contingency-dependent state vari-ables xc and other functional inequality constraints are relaxedfor the contingency cases compared to the base case. Forexample, system voltages are allowed to dip further during anemergency than under normal operating conditions. The relax-ation of system limits is justified because operation under acontingency is temporary: when a contingency occurs, opera-tors immediately begin re-configuring the system to return allbranches and buses to normal operating limits. For the same rea-son, the SCED objective function considers only the base case:contingency cases are improbable and transient and do not con-tribute to the expected cost function.

    SCED has interesting connections to other areas of optimiza-tion. The motivation for SCED is theoretically similar to that ofRobustOptimization (RO) (Bertsimas et al., 2011), althoughROtypically addresses continuous uncertain parameters rather thandiscrete scenarios. Additionally, because the constraints are sep-arable for a fixed u, SCED lends itself well to parallelization anddecomposition algorithms (Qiu et al., 2005).

    3.2. Security-Constrained Unit Commitment

    In electric power systems operation, unit commitment (UC)refers to the scheduling of generating units such that total oper-ating cost is minimized. UC differs from ED in that it operatesacross multiple time periods and schedules the on–off status ofeach generator in addition to its power output. UCmust addressgenerator startup and shutdown time and costs, limits on gen-erator cycling, ramp rate limits, reserve margin requirements,and other scheduling constraints. UC is a large-scale, multi-period MINLP. Many UC formulations relax certain aspects of

    the problem in order to obtain a mixed-integer linear programinstead—for instance, by using linearized cost functions.

    If the power flow equations are added to the UC problem, theformulation becomes SCUC. In SCUC, a power flow is appliedat each time period to ensure that the scheduled generation sat-isfies not only the scheduling constraints but also system volt-age and branch flow limits. In other words, SCUC ensures thatthe UC algorithm produces a generation schedule that can bephysically realized in the power system. Due to its complexity,research on SCUC has accelerated only with the advent of fastercomputing capabilities and algorithmic advancements.

    In SCUC, we introduce a time index t ∈ T and a set of binarycontrol variables wit to the OPF formulation. Each wit indicateswhether or not generator i is committed for time period t . Themodified formulation becomes

    min∑t∈T

    ∑i∈G

    (witCi

    (PGit)+CSUi wit (1 − wi,t−1)

    +CSDi (1 − wit )wi,t−1), (13)

    s.t. Pit (Vt , δt , ϕt ,Tt ) = PGit − PLit ∀ i ∈ N,∀ t ∈ T,(14)

    Qit (Vt , δt , ϕt ,Tt ) = QGit − QLit ∀ i ∈ N,∀ t ∈ T,(15)

    witPG,mini ≤ PGit ≤ witPG,maxi ∀ i ∈ G,∀ t ∈ T,(16)

    witQG,mini ≤ QGit ≤ witQG,maxi ∀ i ∈ G,∀ t ∈ T,(17)

    Vmini ≤ Vit ≤ Vmaxi ∀ i ∈ N,∀ t ∈ T, (18)δmini ≤ δit ≤ δmaxi ∀ i ∈ N,∀ t ∈ T, (19)ϕminik ≤ ϕikt ≤ ϕmaxik ∀ ik ∈ H,∀ t ∈ T, (20)Tminik ≤ Tikt ≤ Tmaxik ∀ ik ∈ K,∀ t ∈ T, (21)Iikt (Vt , δt ) ≤ Imaxik ∀ ik ∈ L,∀ t ∈ T, (22)PDowni ≤ PGit − PGi,t−1 ≤ PUpi ∀ i ∈ G,∀ t ∈ T,

    (23)∑i∈G

    witPG,maxi −∑i∈G

    PGit ≥ PReserve ∀ t ∈ T. (24)

    The objective function (13) includes terms for unit startup costsCSU and shutdown costsCSD in addition to the generation costsin each time period. The PF Equations (14) and (15) are gener-alized to include the effects of phase-shifting and tap-changingtransformers (Section 5.5. We abuse notation slightly by using(Vt , δt , ϕt ,Tt ) to represent the vector of all voltage magnitudes,voltage angles, and engineering control settings for time periodt .) The generation limits (16) and (17) are modified such thatuncommitted units must have zero real and reactive power gen-eration. Box constraints (20) and (21) enforce engineering con-trol limits and (22) limits the current on each transmission line;these constraints and associated sets H and K are described indetail in Section 5.5. Constraint (23) specifies positive and nega-tive generator ramp limits PUp and PDown, respectively; these are

  • 1178 S. FRANK AND S. REBENNACK

    physical limitations of the generators. Constraint (24) requires aspinning reserve margin of at least PReserve; sometimes this con-straint is written such that PReserve is a fraction of the total loadin each time period.

    The SCUC formulation (13)–(24) is one ofmany possible for-mulations. Some formulations include more precise ramp limitsand startup and shutdown characteristics; others include con-straints governing generator minimum uptime and downtime.Due to of the scale and presence of binary decision variables,SCUC is one of the most difficult power systems optimizationproblems. Bai andWei (2009) and Zhu (2009) providemore dis-cussion of SCUC, including detailed formulations.

    More broadly, SCUC belongs to a class of problems knownas multi-period or dynamic OPF. Dynamic OPF problemsspan multiple time periods. Mathematically, each time periodrequires a single copy of the underlying PF equations; thedimensionality therefore scales linearly with the time horizon.Coupling constraints then link the different time periods. Inthe case of the SCUC, the ramping constraints (23) providethis coupling. Grid-scale storage systems, such as batteries, alsorequire coupling constraints that create dynamic OPF problems(Snyder, 2013). Other examples include system design problemssuch as discussed in Frank and Rebennack (2015), where multi-ple OPF problems are coupled by binary design variables.

    3.3. Optimal Reactive Power Flow

    ORPF, also known as reactive power dispatch or VAR control,seeks to optimize the system reactive power generation in orderto minimize the total system losses. In ORPF, the system realpower generation is determined a priori, from the outcome of,for example, a DC-OPF algorithm, UC, or another form of ED.A basic ORPF formulation is

    min P1, (25)

    s.t. Pi (V, δ, ϕ,T ) = PGi − PLi ∀ i ∈ N, (26)Qi (V, δ, ϕ,T ) = QGi − QLi ∀ i ∈ N, (27)QG,mini ≤ QGi ≤ QG,maxi ∀ i ∈ G, (28)Vmini ≤ Vi ≤ Vmaxi ∀ i ∈ N, (29)δmini ≤ δi ≤ δmaxi ∀ i ∈ N, (30)ϕminik ≤ ϕik ≤ ϕmaxik ∀ ik ∈ H, (31)Tminik ≤ Tik ≤ Tmaxik ∀ ik ∈ K, (32)Iik (V, δ) ≤ Imaxik ∀ ik ∈ L. (33)

    The power flow Equations (26) and (27) are again generalizedto include the effects of phase-shifting and tap-changing trans-formers (Section 5.5). The vector of control variables is

    u = (P1,QGi:i∈G, ϕik:ik∈H,Tik:ik∈K) ,while the vector of state variables x = (δ,V ) is identical to theclassic formulation. In ORPF, all real power load and generationis fixed except for the real power at the slack bus, P1. MinimizingP1 is therefore equivalent to minimizing total system loss.

    One motivation for using ORPF is the reduction of the vari-able space compared with fully coupled OPF (Contaxis et al.,

    1986); another is the ability to reschedule reactive power to opti-mally respond to changes in the system load without changingpreviously established real power setpoints. Many interior pointalgorithms for OPF have focused specifically on ORPF (Franket al., 2012a). Zhu (2009, Ch. 10) discusses several approximateORPF formulations and their solution methods.

    3.4. Reactive Power Planning

    RPP extends the ORPF problem to the optimal allocation ofnew reactive power sources—such as capacitor banks—withina power system in order to minimize either system losses ortotal costs. RPP modifies ORPF to include a set of possible newreactive power sources; the presence or absence of each newsource is modeled with a binary variable. The combinatorialnature of installing new reactive power sources has inspiredmany papers that apply heuristic methods to RPP (Frank et al.,2012b).

    A basic RPP formulation that minimizes total costs is

    min C1 (P1) +∑i∈Q

    wiCInstalli , (34)

    s.t. Pi (V, δ, ϕ,T ) = PGi − PLi ∀ i ∈ N, (35)Qi (V, δ, ϕ,T )=QGi +QNewi −QLi ∀ i ∈ N, (36)QG,mini ≤ QGi ≤ QG,maxi ∀ i ∈ G, (37)wiQNew,mini ≤ QNewi ≤ wiQNew,maxi ∀ i ∈ Q, (38)Vmini ≤ Vi ≤ Vmaxi ∀ i ∈ N, (39)δmini ≤ δi ≤ δmaxi ∀ i ∈ N, (40)ϕminik ≤ ϕik ≤ ϕmaxik ∀ ik ∈ H, (41)Tminik ≤ Tik ≤ Tmaxik ∀ ik ∈ K, (42)Iik (V, δ) ≤ Imaxik ∀ ik ∈ L, (43)

    where CInstalli represents the capital cost of each new reactivepower source i ∈ Q; QNewi is the amount of reactive powerprovided by each new reactive power source, subject to limitsQNew,mini andQ

    New,maxi ; andwi is a binary variable governing the

    decision to install each new reactive power source. Themodifiedvector of control variables is

    u = (P1,QGi:i∈G,wi:i∈Q,QNewi:i∈Q, ϕik:ik∈H,Tik:ik∈K) .Some variants of RPP also include real power dispatch in thedecision variables or include multiple load scenarios.

    By necessity, RPP optimizes with respect to uncertain futureconditions—typically reactive power requirements for worst-case scenarios. This uncertainty, together with the problemcomplexity, make RPP a very challenging optimization problem(Zhang et al., 2007). Zhang and Tolbert (2005) review both for-mulations and solution techniques for RPP.

    4. The admittancematrix

    The PF equations are the defining constraints in OPF, and thebus admittance matrix Ỹ in turn forms the core of the PF equa-tions. OPF data sources do not typically provide Ỹ directly,

  • IIE TRANSACTIONS 1179

    Bus i Bus k

    Vi Vk

    Iik Ikiyik

    yShik2

    yShik2

    Figure . � branch model.

    and we therefore summarize the theory and mechanics of itsconstruction here. Readers uninterested in the electrical engi-neering details may wish to skip this section; such readers mayreference Equations (53) to (56) as needed for the mathematicaldefinition of the matrix elements.

    The elements of Ỹ derive from the application of Ohm’s andKirchoff ’s laws (Appendix B) to a steady-state AC electricalnetwork. Recall from Section 2.2 that the set of buses N andset of branches L form undirected graph (N, L) that describesthe electrical network. Each branch (i, k) ∈ L has an associatedseries admittance ỹik that governs the voltage–current relation-ship between buses i and k. According to Kirchoff ’s Voltage Law(KVL) and Ohm’s law,

    Ĩik =(Ṽi − Ṽk

    )ỹik, (44)

    in which Ĩik is the current flowing through branch ik from busi to bus k. Branches may also have an associated shunt admit-tance ỹShik , which represents leakage of current from within thebranch to the reference node. In physical terms, this leakageoccurs all along the branch, but in the model the shunt admit-tance is applied in equal parts to the buses at each end of thebranch as illustrated in Fig. 2. This representation is called the� branch model.

    By Kirchoff ’s Current Law (KCL), the current Ĩi injected intobus i must exactly equal the sum of all currents flowing out ofbus i via the various series and shunt admittances, including anyshunt admittance ỹSi associated with the bus itself. Neglectingoff-nominal branch turns ratios (described in Section 4.1), thebus injection current is

    Ĩi = Ṽi⎛⎝ỹSi + 12 ∑

    k:(i,k)∈LỹShik +

    12

    ∑k:(k,i)∈L

    ỹShki

    ⎞⎠+

    ∑k:(i,k)∈L

    (Ṽi − Ṽk

    )ỹik +

    ∑k:(k,i)∈L

    (Ṽi − Ṽk

    )ỹki. (45)

    Matrix Equation (1) is equivalent to Equation (45) if the ele-ments of Ỹ are defined as

    Ỹii = sum of all admittances connected to bus i−Ỹik = sum of all admittances connected between bus i and

    bus k (46)

    Since only a single branch (i, k) connects bus i to bus k, theoff-diagonal elements become Ỹik = Ỹki = −ỹik. If there is noconnection between buses i and k, Ỹik = 0. Typically, each busconnects to only a few branches, such that L ∈ O(N). Thus, Ỹ

    is sparse, having dimension N × N but only N + 2L non-zeroentries.

    4.1. Branchmodels

    The � model of Fig. 2 is sufficient for modeling the majorityof power systems branch elements, including transmission lines,cables, and transformers.Most power systems transformers havenominal turns ratios; that is, the voltage ratio across the trans-former exactly equals change in system voltage base across thetransformer (a 1:1 voltage ratio in per unit, with no phase shift).As the per unit system automatically accounts for nominal turnsratios (seeAppendixC), no corrections to the admittancematrixare necessary.

    Some transformers, however, do not have exactly a 1:1voltage ratio in per unit. Such transformers are labeled “off-nominal”; this category includes fixed-tap transformers withother than unity turns ratios, tap-changing transformers, andphase-shifting transformers. In practical power system models,improperly neglecting off-nominal turns ratiosmay severely dis-tort computed PF, such that any resulting OPF solution may beunusable. Therefore, off-nominal transformers requiremodifiedentries in Ỹ to account for the additional voltage magnitude orphase angle change relative to the nominal case.

    The generalized � branch model in Fig. 3 extends the �model to accommodate both nominal and off-nominal turnsratios. The model includes series admittance ỹik, shunt admit-tance ỹShik , and an ideal transformer. Bus i is the tap bus and busk is the impedance bus, or Z bus. The transformer turns ratio inper unit is a : 1, where a is a complex exponential consisting ofmagnitude T and phase shift ϕ

    a = Te jϕ,

    such that in the figure Ṽi = aṼ ′i and Ĩ = Ĩ′ik/a∗. (To simplify themathematical presentation, we abuse notation throughout thissection by omitting the subscript ik on a, T , and ϕ.) SelectingT = 1 and ϕ = 0 yields the nominal turns ratio.

    In order to model the effects of the off-nominal turns ratio,we require a partial admittance matrix Ỹ ′ such that(

    ĨikĨki

    )=(Ỹ ′ii Ỹ ′ikỸ ′ki Ỹ

    ′kk

    )(ṼiṼk

    ).

    The elements of Ỹ ′ are derived from the application of KVL,KCL, and Ohm’s law. From KCL at node i and the definitionsof Ṽ ′i and Ĩ′i ,

    Ĩik = 1aa∗ Ṽi(ỹShik2

    + ỹik)

    − 1a∗Ṽkỹik. (47)

    Similarly, at bus k,

    Ĩki = −1aṼiỹik + Ṽk(ỹShik2

    + ỹik)

    . (48)

  • 1180 S. FRANK AND S. REBENNACK

    Bus i Bus k

    Vi Vi Vk

    Iik Iik Ikiyik

    yShik2

    yShik2

    a : 1

    Figure . Generalized� branch model, including off-nominal turns ratio.

    Rearranging Equations (47) and (48) to form a matrix equationyields Ỹ ′,

    (ĨikĨki

    )=

    ⎛⎜⎜⎜⎝1aa∗

    (ỹShik2

    + ỹik)

    − 1a∗

    ỹik

    −1aỹik

    ỹShik2

    + ỹik

    ⎞⎟⎟⎟⎠(ṼiṼk

    )= Ỹ ′

    (ṼiṼk

    ).

    (49)When constructing the full admittance matrix Ỹ , the cor-

    rected relationships of Equation (49) must be preserved. Thisis done via appropriate substitutions in Equation (45); Section4.2 gives formulas for the resulting admittance matrix entries. Abranchwith off-nominalmagnitude only (real valued a) leaves Ỹa symmetric matrix, but a phase-shifting transformer (complexa) does not.

    ... Transmission lines and cablesThe line characteristics for transmission lines and cables aremost often specified as a series impedance Rik + jXik and abranch shunt admittance jbShik , which is sometimes given as “linecharging” reactive power. The� branch series admittance ỹik forinclusion in Ỹ is then

    ỹik = 1Rik + jXik =Rik

    R2ik + X2ik− j Xik

    R2ik + X2ik. (50)

    As is typical for � branch models, shunt susceptance bShik isdivided into two equal parts that are applied to the buses at eachend of the branch. For short lines, branch shunt susceptance isnegligible and therefore is usually omitted. Transmission linesand cables have nominal turns ratios: T = 1.0 and ϕ = 0.

    ... TransformersLike cables, power systems transformers have a seriesimpedance Rik + jXik that models the electrical character-istics of the transformer windings. In PF analysis, however,the transformer series resistance is often neglected, yieldingỹik = − j/Xik. Although transformer models may also include ashunt admittance ỹShik that represents the losses and magnetizingcharacteristics of the transformer core, this shunt admittance isnearly always neglected as well.

    Off-nominal transformers have either T = 1 and/or ϕ = 0.In OPF, both T and ϕ may be control variables: controllable Tmodels an on-load tap changer, while controllable ϕ models aphase-shifting transformer.

    4.2. Construction equations for admittancematrix

    Any branch model may be distilled into a generic series admit-tance ỹik, shunt admittance ỹShik , and complex turns ratio (nom-inal or off-nominal) aik = Tike jϕik . Using Equation (46) andincorporating the corrected off-nominal voltage–current rela-tionships introduced in Section 4.1, the entries of Ỹ become

    Ỹii = ỹSi +∑

    k:(i,k)∈L

    1|aik|2

    (ỹik + 12 ỹ

    Shik

    )+

    ∑k:(k,i)∈L

    (ỹki + 12 ỹ

    Shki

    ),

    (51)

    Ỹik = −∑

    k:(i,k)∈L

    1a∗ik

    ỹik −∑

    k:(k,i)∈L

    1aik

    ỹki, i = k, (52)

    where aik = 1 for any branch with a nominal turns ratio. Equa-tions (51) and (52) may be separated into real and imaginaryparts using the definition Ỹ = G + jB and the identity aik =Tik(cosϕik + j sinϕik

    ),

    Gii = gSi +∑

    k:(i,k)∈L

    1T 2ik

    (gik + 12g

    Shik

    )+

    ∑k:(k,i)∈L

    (gki + 12g

    Shki

    ), (53)

    Gik = −∑

    k:(i,k)∈L

    1Tik

    (gik cosϕik − bik sinϕik

    )−

    ∑k:(k,i)∈L

    1Tki

    (gki cosϕki + bki sinϕki

    ), i = k (54)

    Bii = bSi +∑

    k:(i,k)∈L

    1T 2ik

    (bik + 12b

    Shik

    )+

    ∑k:(k,i)∈L

    (bki + 12b

    Shki

    ),

    (55)

    Bik = −∑

    k:(i,k)∈L

    1Tik

    (gik sinϕik + bik cosϕik

    )−

    ∑k:(k,i)∈L

    1Tki

    (−gki sinϕki + bki cosϕki) , i = k. (56)AppendixDdescribes how to obtain the required parameters forconstructing the admittance matrix from publicly available datasources.

    5. The PF equations

    The AC power flow equations transform the complex-domainmatrix function (2) into a set of real-valued simultaneous equa-tions suitable for inclusion in mathematical programming for-mulations. The most prevalent and compact form of the PFequations is the bus injection model. However, the less compact

  • IIE TRANSACTIONS 1181

    branch flow model also describes Equation (2) and has severaladvantages with respect to convex relaxation of OPF problems.

    5.1. Bus injectionmodel

    The bus injection model is usually synonymous with the term“AC power flow equations” and provides a compact represen-tation of power system behavior in terms of the real and reac-tive power injection at each system bus. Use of the bus injectionmodel is ubiquitous in conventional PF and in the early OPFliterature.

    For a given Ỹ , Equation (2) may be decomposed into a set ofequations for the real and reactive power injections by evaluat-ing the real and imaginary parts of S, respectively. The result-ing pair of equations may be written in several equivalent formsdepending on whether the voltages and admittance matrix ele-ments are expressed in polar or rectangular coordinates. In theliterature, the most common forms of the AC PF equations are,in order, as follows.

    1. Selection of polar coordinates for voltage, Ṽi = Vi∠δi,and rectangular coordinates for admittance, Ỹik = Gik +jBik:

    Pi (V, δ) = ViN∑k=1

    Vk(Gik cos (δi − δk)

    + Bik sin (δi − δk)) ∀ i ∈ N, (57)

    Qi (V, δ) = ViN∑k=1

    Vk(Gik sin (δi − δk)

    − Bik cos (δi − δk)) ∀ i ∈ N. (58)

    2. Selection of polar coordinates for voltage, Ṽi = Vi∠δi,and polar coordinates for admittance, Ỹik = Yik∠θik:

    Pi (V, δ) = ViN∑k=1

    VkYik cos (δi − δk − θik) ∀ i ∈ N,

    (59)

    Qi (V, δ) = ViN∑k=1

    VkYik sin (δi − δk − θik) ∀ i ∈ N.

    (60)

    3. Selection of rectangular coordinates for voltage, Ṽi =Ei + jFi, and rectangular coordinates for admittance,Ỹik = Gik + jBik:

    Pi (E, F ) =N∑k=1

    Gik (EiEk + FiFk)

    + Bik (FiEk − EiFk) ∀ i ∈ N, (61)

    Qi (E, F ) =N∑k=1

    Gik (FiEk − EiFk)

    − Bik (EiEk + FiFk) ∀ i ∈ N. (62)Power systems texts (Wood and Wollenberg, 1996; Gloveret al., 2008; Zhu, 2009) provide exact derivations of these threeforms of the AC PF equations. (The fourth form—selection ofrectangular coordinates for voltage and polar coordinates for

    admittance—is theoretically possible but has no advantages forpractical use.) Each form of the equations involves real-valuedquantities only. However, all forms are equivalent and give theexact solution to the PF under the analysis assumptions outlinedin Appendix B.3.

    The bus injectionmodel of theACPF equations yields exactly2N simultaneous equations in 4N variables. Of these, only N +M − 1 equations in are needed for conventional PF. Despite theequations’ nonlinearity and complexity, efficient solution meth-ods exist for the conventional PF problem (see Section 6.1).

    5.2. Branch flowmodel

    The branch flowmodel is an alternative formof the PF equationsthat expresses the real and reactive PF in each system branch, asderived fromEquation (44). Baran andWu (1989a, 1989b) origi-nated thismodel as part of an optimal capacitor placement prob-lem in radial distribution systems, although earlier researchershad used similar, linear network flow models derived from DCPF (Section 5.3; Azevedo et al. (2010)). The branch flow modelhas received considerable recent interest because it offers advan-tages for convex relaxation of OPF problems (Farivar and Low,2013a, 2013b).

    Low (2013) summarizes the branch flow model using threesets of complex equations

    Ṽi − Ṽk = Z̃ikĨik ∀ (i, k) ∈ L, (63)Sik = ṼĩI∗ik ∀ (i, k) ∈ L, (64)Si=

    ∑k:(i,k)∈L

    Sik−∑

    k:(i,k)∈L

    (Ski−Z̃ki

    ∣∣̃Iki∣∣2)+ (̃ySi )∗ ∣∣Ṽi∣∣2 ∀ i ∈ N(65)

    in which Z̃ik = 1/̃yik is the complex series impedance of eachsystem branch. (Note that Equations (63) to (65) neglect branchshunt admittances and off-nominal turns ratios.) The systemvariables are 2L real-valued components of directed branch cur-rents i → k, 2L directed branch real and reactive power flowsi → k, N bus voltage magnitudes, N bus voltage angles, and 2Nbus real and reactive power injections, for a total of 4N + 4Lvariables. Similarly, there are 2L directed branch current defi-nitions i → k, 2L directed branch real and reactive power defi-nitions i → k, and 2N bus power balance equations, for a totalof 2N + 4L simultaneous equations. As with the bus injectionmodel, the system is underdetermined by exactly 2N variables.However, the dimensionality of the system of equations is con-siderably larger.

    Frank and Rebennack (2015) present an alternative form ofthe branch flow model that (i) uses rectangular coordinates forthe bus voltages and (ii) substitutes the real and reactive PF at thereceiving of each branch as variables in place of the branch cur-rents. The branch power definitions follow directly from Equa-tion (49) and the definition of complex power,

    Pik = 1T 2ik

    (gik +

    gSik2

    ) (E2i + F2i

    )+ 1Tik

    (−gik cosϕik + bik sinϕik) (EiEk + FiFk)

  • 1182 S. FRANK AND S. REBENNACK

    + 1Tik

    (gik sinϕik + bik cosϕik

    )(EiFk − FiEk)

    ∀ (i, k) ∈ L, (66)

    Pki =(gik +

    gSik2

    ) (E2k + F2k

    )+ 1Tik

    (−gik cosϕik − bik sinϕik) (EiEk + FiFk)+ 1Tik

    (gik sinϕik − bik cosϕik

    )(EiFk − FiEk)

    ∀ (i, k) ∈ L, (67)

    Qik = − 1T 2ik

    (bik +

    bSik2

    ) (E2i + F2i

    )+ 1Tik

    (gik sinϕik + bik cosϕik

    )(EiEk + FiFk)

    + 1Tik

    (gik cosϕik − bik sinϕik

    )(EiFk − FiEk)

    ∀ (i, k) ∈ L, (68)

    Qki = −(bik +

    bSik2

    ) (E2k + F2k

    )+ 1Tik

    (−gik sinϕik + bik cosϕik) (EiEk + FiFk)+ 1Tik

    (−gik cosϕik − bik sinϕik) (EiFk − FiEk)∀ (i, k) ∈ L. (69)

    Bus power balance equations (65) are still required but insteadtake the form

    Pi =∑

    k:(i,k)∈LPik +

    ∑k:(i,k)∈L

    Pik∀ i ∈ N, (70)

    Qi =∑

    k:(i,k)∈LQik +

    ∑k:(i,k)∈L

    Qik∀ i ∈ N. (71)

    Unlike Equations (63) to (65), Equations (66) to (71) do includethe effects of branch shunt admittances and off-nominal turnsratios. The system of equations still contains 4N + 4L variables(2N bus voltage components, 2N bus real and reactive powerinjections, and 4L directed branch real and reactive power flows)and 2N + 4L simultaneous equations.

    5.3. DC PF

    The AC PF equations are nonlinear. For conventional PF, thisnonlinearity requires the use of an iterative numerical method;for OPF it implies both a nonlinear formulation and non-convexity in the feasible region. In order to simplify the systemrepresentation, power systems engineers have developed a linearapproximation to the PF equations. This approximation is called

    DC PF—so named because the equations resemble the PF in adirect current (DC) network.However, theDCPF equations stillmodel an AC power system.

    The conventional development of the DC PF equationsrequires several assumptions regarding the power system (Rau,2003b; Zhu, 2009):

    1. All system branch resistances are approximately zero;that is, the transmission system is assumed to be lossless.As a result, all θik = ±90◦ and all Gik = 0.

    2. The differences between adjacent bus voltage anglesare small, such that sin(δi − δk) ≈ δi − δk and cos(δi −δk) ≈ 1.

    3. The system bus voltages are approximately equal to 1.0.This assumption requires that there is sufficient reactivepower generation in the system to maintain a level volt-age profile.

    4. Reactive PF is neglected.Applying these assumptions to Equation (57) produces the DCpower flow equation

    Pi (δ) ≈N∑k=1

    Bik (δi − δk) . (72)

    Under normal operating conditions, DC PF models realpower transfer quite accurately. It has been successfully used inmany OPF applications that require rapid and robust solutions,including in commercial software. However, the assumptionsrequired for DC PF can lead to significant errors for stressedsystems.

    Neglecting off-nominal turns ratios, the exact equation forbranch power transfer is

    Pik = gikV 2i − gikViVk cos (δi − δk) − bikViVk sin (δi − δk),(73)

    cf. Equation (57), and the DC PF approximation is

    Pik ≈ −bik (δi − δk) . (74)The bik term dominates the exact expression because V 2i ≈ViVk cos (δi − δk) and therefore the first two terms in Equation(73) largely cancel. We observe that Equation (74) overestimatesthe magnitude of the branch power transfer (73) if

    (i) The bus voltages at either end of the branch are depressedrelative to the assumed value of 1.0 p.u.; or

    (ii) The angle difference between the buses is too large.Observation (ii) follows from the relationship |sin (δi − δk)| ≤|δi − δk|. Depressed voltages and larger than normal angle dif-ferences are common in stressed power systems. In particular,large differences in voltage in different areas of the system canlead to significant error (Stott et al., 2009). Therefore, the DCPF equations should not be used for OPF under stressed systemconditions unless they have been carefully evaluated for accu-racy in the system under test.

    As an example, consider an arbitrary transmission line frombus i to bus k with series admittance 0.05 − j2.0. Let Ṽi =0.95∠0◦ and Ṽk = 0.90∠−20◦. (Although these numbers do notrepresent normal operation, they are plausible for a stressedpower system. Operating voltages as low as 0.9 p.u. are allowablein emergency conditions, and angle differences of up to ±30°can occur on long, heavily loaded transmission lines.) The exact

  • IIE TRANSACTIONS 1183

    power transfer for this line isPik = 0.05 × 0.952 − 0.05 × 0.95 × 0.90 cos (0◦ + 20◦)

    −2.0 × 0.95 × 0.90 sin (0◦ + 20◦) = 0.590 p.u.The DC approximate power transfer is

    Pik ≈ −2.0 sin (0◦ + 20◦) ≈ 0.684 p.u.The error in the approximate power transfer is 16%; most of thiserror is attributable to the voltage difference. In systems withseverely depressed voltages, large absolute errors in line powerare typical, although the relative error in the proportion of totalpower flowing in each line may be much smaller (Stott et al.,2009).

    Even under normal operation, the approximation of a loss-less transmission network can also lead to significant errors ingenerator scheduling, branch PF estimates, and marginal fuelcost estimates. Power transfer errors for certain critically loadedbranches can be much higher than the average branch error.Therefore, in practical DC PF models an estimate of the lossesmust be reintroduced using approximate methods, especially ifthe network is large (Stott et al., 2009). For further discussionregarding the advantages and disadvantages of DC PF, includ-ing loss approximation methods, we refer the interested readerto Rau (2003a) and Stott et al. (2009). Throughout the rest of thisarticle, we use the AC PF equations.

    5.4. The PF equations as constraints

    In OPF, the PF equations form the core of the set of equalityconstraints. A large majority of OPF formulations use the businjection model because (i) it is more compact and (ii) it moreclosely resembles conventional PF, which facilitates the devel-opment of tailored algorithms. The advantages of the branchflow model are the explicit modeling of branch power flows,which facilitates inclusion of transmission capacity constraints,and that the form of the equations enables certain convexrelaxations.

    Traditionally, the branch impedances and elements of Ỹ areconsidered constant: most algorithmic development and analy-sis in the engineering literature assumes constant Ỹ . In newerOPF formulations Ỹ may also contain control (decision) vari-ables due to phase-shifting or tap-changing transformers, whichsignificantly complicates the problem by introducing bilinear,trilinear, or other nonlinear terms.

    ... Considerations for the bus injectionmodelThe key consideration when using the bus injection model isthe choice of polar or rectangular coordinates for voltage. Theadvantage of voltage polar coordinates is that constraints on thevoltage magnitude can be enforced directly,

    Vi ≥ Vmini ,Vi ≤ Vmaxi .

    In voltage rectangular coordinates, on the other hand, voltagemagnitude limits require the functional inequality constraints√

    E2i + F2i ≥ Vmini ,√E2i + F2i ≤ Vmaxi .

    Similarly, if the voltage magnitude is fixed (for instance at aPV bus; see Section 2.3), then in polar coordinates Vi can bereplaced with a constant value. In rectangular coordinates, how-ever, a fixed voltage magnitude requires the equality constraint√

    E2i + F2i = Vi.

    Thus, for a fixed voltage magnitude, use of voltage polar coor-dinates leads to a reduction of variables whereas the use of volt-age rectangular coordinates leads to an increase in (nonlinear,non-convex) equality constraints. For this reason, polar coor-dinates are preferred both for conventional PF and most OPFformulations.

    There is, however, one compelling reason to use voltage rect-angular coordinates: expressing voltage in rectangular coordi-nates eliminates trigonometric functions from the PF equations.The resulting PF Equations (61) and (62) are quadratic, whichpresents several advantages (Torres and Quintana, 1998):

    1. The elimination of trigonometric functions speeds eval-uation of the equations.

    2. The second order Taylor series expansion of a quadraticfunction is exact; this yields an efficiency advantage inhigher-order interior-point algorithms for OPF.

    3. The Hessian matrix for a quadratic function is constantand must be evaluated only once. This simplifies theapplication of Newton’s method to the Karush–Kuhn–Tucker (KKT) conditions of the OPF formulation.

    4. If the side constraints are also quadratic, then the entireOPF problembecomes aQuadratically Constrained Pro-gram, although it remains non-convex. This enables theuse of specialized solution algorithms.

    In some cases, these computational advantages outweigh thedisadvantages associated with increased dimensionality andenforcing voltage magnitude constraints. Table 2 summarizesthe differences between the two voltage coordinate choices.

    For the admittance matrix elements, rectangular coordinates(56) and (57) are typically favored because they facilitate theuse of certain approximations in fast-decoupled solution meth-ods for conventional PF (Zhu, 2009). These approximations arealso useful in the development of the DC PF equations (Section5.3). Finally, rectangular coordinates facilitate the inclusion oftransformer voltage ratios and phase angles as decision vari-ables. However, none of these advantages strongly affects the ACPF equations as implemented in many OPF formulations, andsome models use the polar form of the Equations (59) and (60).

    Table . Comparison of the selection of voltage polar coordinates versus voltagerectangular coordinates for the power flow equations. Bold entries indicate themore favorable characteristic.

    Polar coords. Rectangular coords.

    Number of variables inconventional PF

    N + M − 1 2N − 2Voltage magnitude limits Simple bounds Nonlinear functional

    inequality constraintFixed voltage magnitude Variable elimination Nonlinear functional

    by substitution equality constraintNature of PF equations Trigonometric QuadraticFirst derivative of PF

    equationsTrigonometric Linear

    Second derivative of PFequations

    Trigonometric Constant

  • 1184 S. FRANK AND S. REBENNACK

    ... Considerations for the branch flowmodelAs solving OPF problems is NP-hard, convex relaxations areof particular interest, especially when conditions are knownthat cause these relaxations to be exact. There are several dif-ferent approaches to convexify the OPF problem, dependenton the formulation used (Low, 2013). Examples of such relax-ations include semidefinite programming, second-order coneprogramming, and convex Lagrangian dual. Quite surprisingly,the exactness of some of these relaxations has been proven forradial networks (under mild conditions). However, for meshnetworks under the classic bus injection model, examples withlarge duality gap have been published and, as a consequence, therelaxations are no longer exact.

    In contrast, the branch flow model allows for exact relax-ations and convexifications for both radial and mesh networks(under mild conditions, for instance, no upper bounds on loads;Farivar and Low (2013a, 2013b)). Moreover, the interpretationof the decision variables in a branch flow model (for instance,branch power and current flows) is more intuitive comparedwith the semidefinite matrix and second-order cone program-ming relaxations and tend to be numerically more stable com-pared with second-order cone programming relaxations for thebus injection models (Low, 2014).

    If the rectangular coordinate branch flow model (66)–(69)is used, then a linear relaxation may be obtained by introduc-ing auxiliary variables for the quadratic terms and subsequentlyapplying the McCormick inequalities (Frank and Rebennack,2015). Although the relaxation is not exact, in practice it canbe made relatively tight by introducing tailored cuts that con-strain the branch power flows. The linear relaxation allows rapidgeneration of lower bounds on the objective function value andfacilitates mixed integer-linear relaxations for more complexOPF variants.

    5.5. Side constraints

    In addition to the PF equations, a typical OPF problem includesa number of side constraints. The simplest of these are box con-straints (8) to (11) that define the upper and lower limits for thecontrollable generation, the bus voltagemagnitudes, and the busvoltage angles.

    If the system contains controllable phase-shifting or tap-changing transformers, then the corresponding phase angles ϕand tap ratios T are introduced into the set of control variables.The control variable vector u becomes

    u = (PGi:i∈G,QGi:i∈G, ϕik:ik∈H,Tik:ik∈K) ,whereH and K are the sets of branches with controllable-phaseshifting transformers and tap-changing transformers, respec-tively. Since ϕ and T alter the elements of admittance matrix Ỹ ,the left-hand sides of Equations (6) and (7) become functionsof ϕ and T : Pi (V, δ, ϕ,T ) and Qi (V, δ, ϕ,T ), respectively. Theformulation is also augmented with bound constraints on thephase angles

    ϕminik ≤ ϕik ≤ ϕmaxik ∀ ik ∈ H (75)and the tap ratios

    Tminik ≤ Tik ≤ Tmaxik ∀ ik ∈ K . (76)

    Adding controllable phase-shifting or tap-changing transform-ers significantly complicates the OPF problem, rendering manycommon approximations and relaxations invalid. Some OPFresearch has therefore explored alternative models for con-trollable transformers and other advanced PF control devices,such as the inclusion of auxiliary buses or idealization of suchtransformers as controlled power injections (Acha et al., 2000;Lehmköster, 2002; Xiao et al., 2002).

    Many recent OPF formulations also include line limits; thatis, upper bounds on the branch current magnitudes. Since linecurrent is voltage dependent, an exact bound requires a functioninequality constraint. By Ohm’s law, the current magnitude inbranch ik is

    Iik =∣∣Ṽi − Ṽk∣∣ yik,

    in which yik is the magnitude of the branch admittance. Thus,one version of the exact bound is∣∣Ṽi − Ṽk∣∣ yik ≤ Imaxik ,

    ⇔√

    (Vi cos δi −Vk cos δk)2+(Vi sin δi−Vk sin δk)2≤Imaxikyik

    ,

    ⇔ (Vi cos δi −Vk cos δk)2

    + (Vi sin δi −Vk sin δk)2 ≤(Imaxik

    )2y2ik

    ∀ ik ∈ L. (77)

    (For branches with off-nominal turns ratios, the tap bus voltageVi and angle δi in Equation (77) must be corrected for the off-nominal turns ratio. In this case, Vi is replaced by V ′i = Vi/Tikand δi is replaced by δ′i = δi − ϕik.)

    Rather than bounding the square of the current as is givenin Equation (77), most formulations instead bound the real andreactive PF in the line

    P2ik + Q2ik ≤(Smaxik

    )2.

    In almost all cases, this simplification is sufficiently accurate.

    6. Solutionmethods

    In this section, we provide a concise discussion of OPF solu-tionmethods and their relationship to conventional PF. AsmanyOPF algorithms rely on or incorporate aspects of conventionalPF algorithms, we summarize conventional PF solution meth-ods first.

    6.1. Solutionmethods for conventional PF

    Recall from Section 2.3 that conventional PF seeks a numericalsolution to the exactly determined system of complex nonlinearequations

    S = Ṽ ◦ (ỸṼ )∗ .Newton’s method is typically used to solve this system and isfully described in power systems texts (Wood and Wollenberg,1996; Glover et al., 2008; Zhu, 2009). Here, we summarize thesolution of the AC PF equations with voltage polar coordinates.The solution method for voltage rectangular coordinates is sim-ilar; Zhu (2009) provides a good summary.

  • IIE TRANSACTIONS 1185

    The first-order Taylor series approximation of Equations (57)and (58) about the current estimate ofV and δ yields

    (�P�Q

    )≈

    ⎛⎜⎜⎝∂P∂δ

    ∂P∂V

    ∂Q∂δ

    ∂Q∂V

    ⎞⎟⎟⎠(�δ�V)

    ,

    ≈ J(

    �δ

    �V

    ), (78)

    in which J is the Jacobian matrix of system (57)–(58). At eachiteration, the mismatches in the PF equations are

    �Pi =(PGi − PLi

    )− Pi (V, δ) , (79)�Qi =

    (QGi − QLi

    )− Qi (V, δ) . (80)Newton’s method consists of iteratively solving Equation (78)for the �δ and �V required to correct the mismatch in thePF equations computed from Equations (79) and (80). Newton’smethod is locally quadratically convergent. Therefore, given asufficiently good starting point, the method reliably finds thecorrect solution to the PF equations. However, convergence isnot guaranteed: the method can fail to converge if the startingvoltage estimate is poor, as can occur in a stressed power sys-tem (Bienstock, 2013). To address the convergence issue, a fewresearchers have applied globally convergent solution methods(Han, 1977) to conventional PF and OPF, including the applica-tion of homotopy (Cvijić et al., 2012), a backtracking line searchmethod (Pajic and Clements, 2005), and trust region methods(Pajic and Clements, 2005; Sousa et al., 2011). Suchmethods aregenerally robust with respect to poor quality starting points butincur a significant computational penalty.

    ... Decoupled PF algorithmsIn practical power systems, real power injections are stronglycoupled to voltage angles and reactive power injections arestrongly coupled to voltage magnitudes. Conversely, real powerinjections areweakly coupled to voltagemagnitudes and reactivepower injections are weakly coupled to voltage angles. This fea-ture has led to the development of decoupled solution methodsfor the PF equations (Glover et al., 2008; Zhu, 2009). The mostbasic decoupling method is to use a set of approximate Taylorseries expansions of the form

    �P ≈ ∂P∂δ

    �δ,

    �Q ≈ ∂Q∂V

    �V.

    This allows the use of separate Newton updates for δ andV withcorrespondingly smaller matrices—a significant computationaladvantage.

    Although decoupled PF uses an approximate updatemethod,it still computes exact real and reactive power mismatches �Pand�Q from Equations (79) and (80) and updates bothV and δat each iteration. Decoupled PF therefore is locally convergent tothe exact solution to the PF. However, due to the approximatedJacobian matrix, more iterations are required for convergence(Glover et al., 2008). Zhu (2009) discusses several decoupled PFvariants in detail.

    ... DC PF algorithmsIn DC PF, there is no distinction between PV and PQ busesbecause all voltagemagnitudes are considered to be 1.0 and reac-tive power is neglected. As with the AC PF, the slack bus angleis fixed. As the DC PF equations are linear, they may be solveddirectly for the voltage angles using

    δ = B−1P.

    (This is simply a solvedmatrix representation of Equation (72).)

    6.2. Solutionmethods for OPF

    The existing OPF literature describes an immense variety ofapproaches and solution methods. As mentioned in the Intro-duction, our intent in this article is to provide an operationsresearcher with sufficient information to formulate OPF prob-lems, understand the challenges and open problems in the field,and comprehend the electrical engineering-dominated litera-ture on OPF solutionmethods. Therefore, we refrain from sum-marizing the vast literature on tailored OPF solution methodsand instead refer the reader to the two surveys (Frank et al.,2012a, 2012b). We do, however, highlight several popular meth-ods and their relationship to conventional PF.

    Many classic nonlinear optimization techniques have beenapplied to the OPF problem, including gradient descent meth-ods, Newton’s method, Sequential Linear Programming (SLP),and Sequential Quadratic Programming (SQP). Recall fromSection 2.4 that many OPF algorithms partition the decisionvariables into a set of control variables u and a set of state vari-ables x. At each search step, the algorithm fixes u and derives xby solving a conventional PF. When this approach is used, theJacobian matrix J from conventional PF plays several importantroles.

    1. It provides the linearization of the PF equations requiredfor SLP and SQP iterations.

    2. It provides sensitivities in the PF injections with respectto the state variables.

    3. It provides a direct calculation of portions of the Hessianmatrix of the Lagrangian function for OPF algorithmsbased on Newton’s method (see Sun et al. (1984)).

    4. It is therefore often used to improve computationalefficiency in computing the KKT conditions for theLagrangian function.

    More recently, nonlinear Interior-Point Methods (IPMs) andvarious convex relaxation approaches such as semidefinite pro-gramming have become popular for OPF (Frank et al., 2012a).Many such methods rely heavily on specific characteristics ofthe PF equations in order to establish convergence propertiesor evaluate the tightness of the relaxation. Many meta-heuristicmethods and hybrid methods combining meta-heuristics withdeterministic approaches have also been applied to OPF (Franket al., 2012b). In the case ofmeta-heuristics, the typical approachis again to select a trial vector for u and compute the correspond-ing state vector x and objective function value using conven-tional PF.

  • 1186 S. FRANK AND S. REBENNACK

    6.3. Practical considerations

    In our experience, there are several practical and computationalaspects of OPF stemming from the power flow equations thatcan cause confusion or lead to solution errors. Two of these arethe use of the per unit system and formats for the exchange ofOPF data, discussed in Appendices C and D, respectively. Wenote several others here.

    ... Decoupled solutionmethods for OPFLike decoupled PF, decoupled OPF algorithms take advantageof the strong P–δ and Q–V relationships in the PF equationsby formulating separate real and reactive OPF subproblems.These subproblems and their optima are assumed to be indepen-dent. Unlike decoupled PF, however, decoupled OPF solves thesubproblems sequentially rather than simultaneously: the realsubproblem solves for the optimal values of P and δ while hold-ing Q and V constant, and the reactive subproblem solves forthe optimal values of Q and V while holding P and δ con-stant (Sun et al., 1984; Contaxis et al., 1986). To enable thisdecomposition, decoupled OPF neglects the weak P–V andQ–δrelationships in the constraints, introducing slight errors in thecomputed power flows. Decoupled OPF is therefore distinctlydifferent from decoupled PF in that the decoupled OPF solutionis inexact.

    The error in the decoupled OPF solution with respect tothe true optimum is a function of the accuracy of the decou-pling assumptions and can be significant if the system includesadvanced control devices (Zhang, 2010). These assumptionsshould therefore be evaluated for accuracy if a decoupled OPFapproach is considered. In the OPF literature, it is not alwaysimmediately apparent whether decoupled OPF is in use orwhether a decoupled PF procedure is used within the solutionalgorithm for a coupled OPF. Due to the implications for theOPF solution quality, the careful reader should try to discernwhich is the case.

    ... Degrees versus radiansPower systems engineers usually report angles in degrees,including in data files for OPF (see Appendix D). For com-putation, these angles must be converted to radians, for tworeasons:

    1. Nearly all optimization software and algebraic modelinglanguages—including AMPL and GAMS—implementtrigonometric functions in radians, not degrees.

    2. Even when using the DC PF equations (which requireno trigonometric function evaluations), radians must beused due to the derivation of the equations. If degreesare used, the powers computed from DC PF will have ascaling error of 180/π .

    PF software typically handles these conversions transparently,accepting input and giving output in degrees. Thus, it can be dif-ficult to remember that general-purpose optimization softwarerequires an explicit conversion.

    ... System initializationIn both conventional PF and OPF, the convergence of the PFequations depends strongly on the selection of a starting point.

    Bus 1

    Bus 2

    Bus 3 Bus 5

    Bus 4

    Branch

    (1,2)

    Branch (1,3)

    Branch (2,4)

    Branc

    h (3,4

    )

    Branch (3,5)

    Bra

    nch

    (4,5

    )

    Figure . Bus and branch indices in an example five-bus electrical network.

    Given a starting point far from the correct solution, the PF equa-tions may converge to a meaningless solution or may not con-verge at all. In the absence of a starting point, standard practiceis to initialize all voltage magnitudes to 1.0 p.u. and all voltageangles to zero; this is called a “cold start” or a “flat start.” Thealternative is a “warm start,” in which the voltages and anglesare initialized to the solution of a pre-solved power flow. Warmstarts are often used in online OPF to minimize computationtime and ensure that the search begins from the current systemoperating condition. Warm starts also aid the algorithm in con-verging to the nearest local minimum, which is often desirableduring real-time operation of power systems.

    If a good starting point is unavailable or the OPF algorithmencounters convergence problems, it is also possible to switchto a more robust solution method. Trust region methods andtheir variants coupled with IPMs are a popular and effectiveapproach (Pajic and Clements, 2005; Wang et al., 2007; Sousaet al., 2011). The primary disadvantage of such methods is a sig-nificant computational penalty compared with their less-robustcounterparts.

    7. Numerical example

    We now present a small numerical example to illustrate somekey elements of a typical OPF formulation. The network of Fig. 4provides the basis for the example OPF problem. It has N = 5buses and L = 6 branches, with corresponding sets

    N = {1, 2, 3, 4, 5}and

    L = {(1, 2), (1, 3), (2, 4), (3, 4), (3, 5), (4, 5)} .

    Using this network, we derive the admittance matrix (Section7.1), define the PF equations (Section 7.2), and develop the clas-sic OPF problem and its optimal solution (Section 7.3).

    7.1. Admittancematrix

    Table 3 provides a set of branch data for the example system.Note that branch (3, 4) is a phase-shifting transformer andbranch (4, 5) has an off-nominal voltage ratio. In addition to the

  • IIE TRANSACTIONS 1187

    Table . Branch impedance data for the network of Fig. . All quantities except phase angles are given in per unit. Dots indicate nominal voltage ratios and phase angles.

    From bus To bus Series resistance Series reactance Shunt susceptance Voltage ratio Phase anglei k Rik Xik b

    Shik Tik ϕik

    . . . · · . . . · · . . . · · . . . · −3.0◦ . . . . · . . . · ·

    branch data, bus 2 has a shunt susceptance of j0.30 pu and bus3 has a shunt conductance of 0.05 pu.

    To compute the admittance matrix for this system, we firstcompute the series admittance ỹik of each branch using Equation(50). For example, the series admittance of branch (1, 3) is

    ỹ13 = 0.0230.0232 + 0.1452 − j0.145

    0.0232 + 0.1452 ≈ 1.067 − j6.727.

    The remaining branches have series admittances

    ỹ12 ≈ 0.000 − j3.333,ỹ24 ≈ 5.660 − j30.189,ỹ34 ≈ 0.294 − j3.824,ỹ35 ≈ 0.000 − j3.125,

    andỹ45 ≈ 0.000 − j2.000.

    Next, we construct Ỹ using Equations (51) and (52). Forexample, diagonal element Ỹ33 consists of summing the admit-tances of branches (1, 3), (3, 4), and (3, 5), plus the con-tributions of the shunt conductance at bus 3 and the shuntsusceptance of branch (1, 3). Branch admittances ỹ34 and ỹ35have off-nominal turns ratios

    a34 = 1.0e− j3.0◦ ≈ 0.999 − j0.052and

    a35 = 0.98e− j0.0◦ = 0.980,respectively. (Note that a34a∗34 = 1.0 if rounding errors areneglected.) Therefore, matrix element Ỹ33 is

    Ỹ33 ≈(1.067 − j6.727)+ j 0.04

    2

    + 0.294 − j3.824(0.999 − j0.052) (0.999 + j0.052) − j3.1250.9802

    + 0.05 ≈ 1.41 − j13.78.An example off-diagonal element is Ỹ34, which from Equation(52) is

    Ỹ34 ≈ − 0.294 − j3.824(0.999 + j0.052) ≈ −0.09 + j3.83.The full admittance matrix is

    Ỹ ≈

    ⎛⎜⎜⎜⎜⎝1.07 − j10.04 0.00 + j3.33 −1.07 + j6.73 0 00.00 + j3.33 5.66 − j33.22 0 −5.66 + j30.19 0

    −1.07 + j6.73 0 1.41 − j13.78 −0.09 + j3.83 0.00 + j3.190 −5.66 + j30.19 −0.49 + j3.80 5.95 − j36.01 0.00 + j2.000 0 0.00 + j3.19 0.00 + j2.00 0.00 − j5.13

    ⎞⎟⎟⎟⎟⎠.

    7.2. PF equations

    Using the previously developed admittancematrix, we can writethe real and reactive PF equations for any bus in the five-bus example system. For example, from Equation (57), the realpower injection at bus 1 is

    P1 (V, δ) = V15∑

    k=1Vk(G1k cos (δ1 − δk) + B1k sin (δ1 − δk)

    ),

    ≈ 1.07V 21 cos (δ1 − δ1) − 1.07V1V3 cos (δ1 − δ3)− 10.04V 21 sin (δ1 − δ1)+ 3.33V1V2 sin (δ1 − δ2) + 6.73V1V3 sin (δ1 − δ3),

    ≈ 1.07V 21 − 1.07V1V3 cos (δ1 − δ3)+ 3.33V1V2 sin (δ1 − δ2) + 6.73V1V3 sin (δ1 − δ3).

    Similarly, fromEquation (58), the reactive power injection at bus1 is

    Q1 (V, δ) = V15∑

    k=1Vk(G1k sin (δ1 − δk) − B1k cos (δ1 − δk)

    ),

    ≈ 1.07V 21 sin (δ1 − δ1) − 1.07V1V3 sin (δ1 − δ3)+ 10.04V 21 cos (δ1 − δ1)− 3.33V1V2 cos (δ1 − δ2) − 6.73V1V3 cos (δ1 − δ3),

    ≈ −1.07V1V3 sin (δ1 − δ3) + 10.04V 21− 3.33V1V2 cos (δ1−δ2)−6.73V1V3 cos (δ1 − δ3).

    7.3. OPF formulation and solution

    We now develop the classic OPF formulation for the five-busexample system presented in Fig. 4, with the addition of twoadvanced control elements. The branch impedance data are asgiven in Table 3, except that we assign ϕ34 and T35 to be deci-sion variables representing a phase-shifting transformer and anon-load tap changer, respectively. ϕ34 and T35 have limits

    −30.0◦ ≤ ϕ34 ≤ 30.0◦and

    0.95 ≤ T35 ≤ 1.05.Tables 4 and 5 give the bus data (voltage limits, load, and gener-ation). The system power base is 100 MW.

  • 1188 S. FRANK AND S. REBENNACK

    Table . Bus data for the network of Fig. . All quantities are given in per unit. Dotsindicate zero values.

    Bus Load real power Load reactive power Min. bus voltage Max. bus voltagei PLi Q

    Li V

    mini V

    maxi

    · · 1.00 1.00 · · 0.95 1.05 · · 0.95 1.05 0.900 0.400 0.95 1.05 0.239 0.129 0.95 1.05

    Table . Generator data for the network of Fig. . All quantities are given in per unit.

    Min. generator Max. generator Min. generator Max. generatorBus real power real power reactive power reactive poweri PG,mini P

    G,maxi Q

    G,mini Q

    G,maxi

    −∞ ∞ −∞ ∞ 0.10 0.40 −0.20 0.30 0.05 0.40 −0.20 0.20

    Given this data, the sets defining the formulation are

    N = {1, 2, 3, 4, 5} ,G = {1, 3, 4} ,L = {(1, 2), (1, 3), (2, 4), (3, 4), (3, 5), (4, 5)} ,H = {(3, 4)} ,

    andK = {(3, 5)} .

    Let the three generator cost functions, in thousands of dollars,be

    C1(PG1) = 0.35PG1 ,

    C3(PG3) = 0.20PG3 + 0.40 (PG3 )2 ,

    C4(PG4) = 0.30PG4 + 0.50 (PG4 )2 ,

    in which the PGi are expressed in per-unit.To develop the full formulation, it is first necessary to re-write

    Ỹ as derived in Section 7.1 to explicitly include ϕ34 and T35. Let

    a34 = cosϕ34 + j sinϕ34and

    a35 = T35.(Note that a34a∗34 = 1.0, 1/a34 = cosϕ34 − j sinϕ34 = a∗34, and1/a∗34 = cosϕ34 + j sinϕ34 = a34.) Then, using Equations (51)and (52) and simplifying,

    Ỹ33 = 1.41 − j10.53 − j 1T 235× 3.13,

    Ỹ34 = −0.29 cos ϕ34 − 3.82 sinϕ34+ j (3.82 cosϕ34 − 0.29 sinϕ34) ,

    Ỹ43 = −0.29 cos ϕ34 + 3.82 sinϕ34+ j (3.82 cosϕ34 + 0.29 sinϕ34) ,

    Ỹ35 = j 1T35 × 3.13,and

    Ỹ53 = j 1T35 × 3.13.

    Ỹ44, Ỹ55, and the other remaining matrix elements areunchanged.

    Bus 1 is the system slack bus and therefore Ṽ1 is fixed to1.0∠0.0◦. To construct the formulation, we round all numeri-cal values to two decimal places. (This rounding does not affectmodel feasibility because sufficient degrees of freedom exist inthe state variables.) Following Equations (5) to (11), (75), and(76), the full formulation is

    min 0.35PG1 + 0.20PG3 + 0.40(PG3)2 + 0.30PG4 + 0.50 (PG4 )2 ,

    s.t. PG1 = 1.07 − 1.07V3 cos (−δ3) + 3.33V2 sin (−δ2)+ 6.73V3 sin (−δ3) ,

    0 = 5.66V 22 − 5.66V2V4 cos (δ2 − δ4)+ 3.33V2 sin (δ2) + 30.19V2V4 sin (δ2 − δ4) ,

    PG3 = 1.41V 23 − 1.07V3 cos (δ3)+ (−0.29 cosϕ34 − 3.82 sinϕ34)V3V4 cos (δ3 − δ4)+ 6.73V3 sin (δ3) + (3.82 cosϕ34 − 0.29 sinϕ34)×V3V4 sin (δ3 − δ4)+ 3.13

    T35V3V5 sin (δ3 − δ5) ,

    PG4 − 0.900 = 5.95V 24 − 5.66V4V2 cos (δ4 − δ2)+ (−0.29 cosϕ34 + 3.82 sinϕ34)V4V3 cos (δ4 − δ3)+ 30.19V4V2 sin (δ4 − δ2)+ (3.82 cosϕ34 + 0.29 sinϕ34)V4V3 sin (δ4 − δ3)+ 2.00V4V5 sin (δ4 − δ5) ,

    −0.239 = 3.13T35

    V5V3 sin (δ5 − δ3) + 2.00V5V4 sin (δ5 − δ4) ,

    QG1 = 10.04 − 1.07V3 sin (−δ3) − 3.33V2 cos (−δ2)− 6.73V3 cos (−δ3) ,

    0 = −5.66V2V4 sin (δ2 − δ4) + 33.22V 22− 3.33V2 cos (δ2) − 30.19V2V4 cos (δ2 − δ4) ,

    QG3 = −1.07V3 sin (δ3) + (−0.29 cosϕ34 − 3.82 sinϕ34)×V3V4 sin (δ3 − δ4)

    +(10.53 + 3.13

    T 235

    )V 23 − 6.73V3 cos (δ3)

    − (3.82 cosϕ34 − 0.29 sinϕ34)V3V4 cos (δ3 − δ4)− 3.13

    T35V3V5 cos (δ3 − δ5) ,

    QG4 − 0.940 = −5.66V4V2 sin (δ4 − δ2)+ (−0.29 cosϕ34 + 3.82 sinϕ34)V4V3 sin (δ4 − δ3)+ 36.01V 24 − 30.19V4V2 cos (δ4 − δ2)− (3.82 cosϕ34 + 0.29 sinϕ34)V4V3 cos (δ4 − δ3)− 2.00V4V5 cos (δ4 − δ5) ,

    −0.129 = 5.13V 25 −3.13T35

    V5V3 cos (δ5 − δ3)

    − 2.00V5V4 cos (δ5 − δ4) ,0.10 ≤ PG3 ≤ 0.40,0.05 ≤ PG4 ≤ 0.40,−0.20 ≤ QG3 ≤