Skip to main content

Advertisement

Log in

Unified value-based feedback, optimization and risk management in complex electric energy systems

  • Research Article
  • Published:
Optimization and Engineering Aims and scope Submit manuscript

Abstract

The ideas in this paper are motivated by an increased need for systematic data-enabled resource management of large-scale electric energy systems. The basic control objective is to manage uncertain disturbances, power imbalances in particular, by optimizing available power resources. To that end, we start with a centralized optimal control problem formulation of system-level performance objective subject to complex interconnection constraints and constraints representing highly heterogeneous internal dynamics of system components. To manage spatial complexity, an inherent multi-layered structure is utilized by modeling interconnection constraints in terms of unified power variables and their dynamics. Similarly, the internal dynamics of components and sub-systems (modules), including their primary automated feedback control, is modeled so that their input–output characterization is also expressed in terms of power variables. This representation is shown to be key to managing the multi-spatial complexity of the problem. In this unifying energy/power state space, the system constraints are all fundamentally convex, resulting in the convex dynamic optimization problem, for typically utilized quadratic cost functions. Based on this, an interactive multi-layered modeling and control method is introduced. While the approach is fundamentally based on the primal–dual decomposition of the centralized problem, this is formulated for the first time for the couple real-reactive power problem. It is also is proposed for the first time to utilize sensitivity functions of distributed agents for solving the primal distributed problem. Iterative communication complexity typically required for convergence of point-wise information exchange is replaced by the embedded distributed optimization by the modules when creating these functions. A theoretical proof of the convergence claim is given. Notably, the inherent multi-temporal complexity is managed by performing model predictive control (MPC)-based decision making when solving distributed primal problems. The formulation enables distributed decision-makers to value uncertainties and related risks according to their preferences. Ultimately, the distributed decision making results in creating a bid function to be used at the coordinating market-clearing level. The optimization approach in this paper provides a theoretical foundation for next-generation Supervisory Control and Data Acquisition (SCADA) in support of a Dynamic Monitoring and Decision Systems (DyMonDS) for a multi-layered interactive market implementation in which the grid users follow their sub-objectives and the higher layers coordinate interconnected sub-systems and the high-level system objectives. This forms a theoretically sound basis for designing IT-enabled protocols for secure operations, planning, and markets.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Notes

  1. \(|V|\angle \theta \) is a short hand for complex phasor domain representation of the signal \(|V| cos\left( \omega t + \theta \right) + j |V| sin\left( \omega t + \theta \right) \) where V is the amplitude of the signal, \(\omega \) is the carrier frequency, \(\theta \) is the voltage phase angle and \(j = \sqrt{-1}\)

References

  • Almassalkhi M, Frolik J, Hines P(2017) Packetized energy management: asynchronous and anonymous coordination of thermostatically controlled loads. In: 2017 American control conference (ACC). IEEE, pp 1431–1437

  • Anand S, Fernandes BG, Guerrero J (2012) Distributed control to ensure proportional load sharing and improve voltage regulation in low-voltage dc microgrids. IEEE Trans Power Electron 28(4):1900–1913

    Google Scholar 

  • Arroyo JM, Conejo AJ (2002) Multiperiod auction for a pool-based electricity market. IEEE Trans Power Syst 17(4):1225–1231

    Google Scholar 

  • Bachovchin KD, Ilić MD (2015) Automated modeling of power system dynamics using the lagrangian formulation. Int Trans Electric Energy Syst 25(10):2087–2108

    Google Scholar 

  • Beck A, Teboulle M (2009) A fast iterative shrinkage-thresholding algorithm for linear inverse problems. SIAM J Imaging Sci 2(1):183–202

    MathSciNet  MATH  Google Scholar 

  • Beck A, Nedić A, Ozdaglar A, Teboulle M (2014) An \( o (1/k) \) gradient method for network resource allocation problems. IEEE Trans Control Netw Syst 1(1):64–73

    MathSciNet  MATH  Google Scholar 

  • Bertsimas D, Litvinov E, Sun XA, Zhao J, Zheng T (2012) Adaptive robust optimization for the security constrained unit commitment problem. IEEE Trans Power Syst 28(1):52–63

    Google Scholar 

  • Chapman JW, Ilic MD, King CA, Eng L, Kaufman H (1993) Stabilizing a multimachine power system via decentralized feedback linearizing excitation control. IEEE Trans Power Syst 8(3):830–839

    Google Scholar 

  • Chen Z, Lei W, Yong F (2012) Real-time price-based demand response management for residential appliances via stochastic optimization and robust optimization. IEEE Trans Smart Grid 3(4):1822–1831

    Google Scholar 

  • Chen T, Li N, Giannakis GB (2018) Aggregating flexibility of heterogeneous energy resources in distribution networks. In: 2018 Annual American control conference (ACC). IEEE, pp 4604–4609

  • Chinchilla M, Arnaltes S, Burgos JC (2006) Control of permanent-magnet generators applied to variable-speed wind-energy systems connected to the grid. IEEE Trans Energy Convers 21(1):130–135

    Google Scholar 

  • Chow JH, Peponides G, Kokotovic PV, Avramovic B, Winkelman JR (1982) Time-scale modeling of dynamic networks with applications to power systems, vol 46. Springer, New York

    Google Scholar 

  • Cobos NG, Arroyo JM, Alguacil N, Wang J (2018) Robust energy and reserve scheduling considering bulk energy storage units and wind uncertainty. IEEE Trans Power Syst 33(5):5206–5216

    Google Scholar 

  • Coffrin C, Van Hentenryck P (2014) A linear-programming approximation of ac power flows. INFORMS J Comput 26(4):718–734

    MATH  Google Scholar 

  • Cvetković M, Ilić M (2015) Interaction variables for distributed numerical integration of nonlinear power system dynamics. In: 2015 53rd Annual Allerton conference on communication, control, and computing (Allerton). IEEE, pp 560–566

  • Dali M, Belhadj J, Roboam X (2010) Hybrid solar-wind system with battery storage operating in grid-connected and standalone mode: control and energy management-experimental investigation. Energy 35(6):2587–2595

    Google Scholar 

  • Dall’Anese E, Zhu H, Giannakis GB (2013) Distributed optimal power flow for smart microgrids. IEEE Trans Smart Grid 4(3):1464–1475

    Google Scholar 

  • De Persis C, Weitenberg ERA, Dörfler F (2018) A power consensus algorithm for dc microgrids. Automatica 89:364–375

    MathSciNet  MATH  Google Scholar 

  • Denholm P, O’Connell M, Brinkman G, Jorgenson J (2015) Overgeneration from solar energy in California. A field guide to the duck chart. Technical report. National Renewable Energy Lab. (NREL), Golden

  • Desoer CA, Katzenelson J (1965) Nonlinear rlc networks. Bell Syst Tech J 44(1):161–198

    MATH  Google Scholar 

  • Diao R, Lu S, Elizondo M, Mayhorn E, Zhang Y, Samaan N (2012) Electric water heater modeling and control strategies for demand response. In: 2012 IEEE power and energy society general meeting. IEEE, pp 1–8

  • Didsayabutra P, Lee W-J, Eua-Arporn B (2002) Defining the must-run and must-take units in a deregulated market. IEEE Trans Ind Appl 38(2):596–601

    Google Scholar 

  • Donadee J, Ilić MD (2014) Stochastic optimization of grid to vehicle frequency regulation capacity bids. IEEE Trans Smart Grid 5(2):1061–1069

    Google Scholar 

  • Dörfler F, Bullo F (2013) Novel insights into lossless ac and dc power flow. In: 2013 IEEE power and energy society general meeting. IEEE, pp 1–5

  • Dufour C, Abourida S, Belanger J(2005) Hardware-in-the-loop simulation of power drives with rt-lab. In: 2005 International conference on power electronics and drives systems, vol 2. IEEE, pp 1646–1651

  • Elizondo J, Zhang RY, White JK, Kirtley JL (2015) Robust small signal stability for microgrids under uncertainty. In 2015 IEEE 6th international symposium on power electronics for distributed generation systems (PEDG). IEEE, pp 1–8

  • Espinosa LAD, Almassalkhi M, Hines P, Heydari S, Frolik J (2017) Towards a macromodel for packetized energy management of resistive water heaters. In: 2017 51st Annual conference on information sciences and systems (CISS). IEEE, pp 1–6

  • Faruque MDO, Strasser T, Lauss G, Jalili-Marandi V, Forsyth P, Dufour C, Dinavahi V, Monti A, Kotsampopoulos P, Martinez JA et al (2015) Real-time simulation technologies for power systems design, testing, and analysis. IEEE Power Energy Technol Syst J 2(2):63–73

    Google Scholar 

  • Feng Y (2015) Solving for the low-voltage/large-angle power-flow solutions. PhD thesis, Arizona State University

  • Grant LL, Crow ML, Cheng MX (2015) Computationally efficient solvers for power system applications. In: 2015 IEEE power and energy conference at Illinois (PECI). IEEE, pp 1–5

  • Gribik PR, Hogan WW, Pope SL, et al (2007) Market-clearing electricity prices and energy uplift. Harvard Electricity Policy Group working paper, Cambridge, MA

  • Guedes RBL, Alberto LFC, Bretas NG (2005) Power system low-voltage solutions using an auxiliary gradient system for voltage collapse purposes. IEEE Trans Power Syst 20(3):1528–1537

    Google Scholar 

  • Guerrero JM, Chandorkar M, Lee T-L, Loh PC (2012) Advanced control architectures for intelligent microgrids—part I: decentralized and hierarchical control. IEEE Trans Ind Electron 60(4):1254–1262

    Google Scholar 

  • Heffner GC, Goldman CA, Moezzi MM (2005) Innovative approaches to verifying demand response of water heater load control. IEEE Trans Power Deliv 21(1):388–397

    Google Scholar 

  • Hestenes MR (1969) Multiplier and gradient methods. J Optim Theory Appl 4(5):303–320

    MathSciNet  MATH  Google Scholar 

  • Hug-Glanzmann G, Andersson G (2009) Decentralized optimal power flow control for overlapping areas in power systems. IEEE Trans Power Syst 24(1):327–336

    Google Scholar 

  • Ilic M (1992) Network theoretic conditions for existence and uniqueness of steady state solutions to electric power circuits. In: Proceedings of 1992 IEEE international symposium on circuits and systems, vol 6. IEEE, pp 2821–2828

  • Ilic M (2016) Ferc reliability technical conference panel I: 2016 state of reliability. Carnegie Mellon University, Pittsburgh

    Google Scholar 

  • Ilic MD (2007) From hierarchical to open access electric power systems. Proc IEEE 95(5):1060–1084

    Google Scholar 

  • Ilic MD, Zaborszky J (2000) Dynamics and control of large electric power systems. Wiley, New York

    Google Scholar 

  • Ilić MD, Liu Q (2012) Toward sensing, communications and control architectures for frequency regulation in systems with highly variable resources. In: Chakrabortty A, Ilić M (eds) Control and optimization methods for electric smart grids. Springer, New York, NY, pp 3–33

    Google Scholar 

  • Ilic MD, Jaddivada R (2018a) Fundamental modeling and conditions for realizable and efficient energy systems. In: 2018 IEEE conference on decision and control (CDC). IEEE, pp 5694–5701

  • Ilić MD, Jaddivada R (2018b) Multi-layered interactive energy space modeling for near-optimal electrification of terrestrial, shipboard and aircraft systems. Annu Rev Control 45:52–75

    MathSciNet  Google Scholar 

  • Ilic MD, Jaddivada R (2019) Exergy/energy dynamics-based integrative modeling and control method for difficult electric aircraft missions, September 12 2019. Provisional Patent Application No. 62/730203

  • Ilic MD, Joo JY, Andrews BW, Raghunathan B, Benitez D, Maus F (2012) Adaptive load management: a system for incorporating customer electrical demand information for demand and supply side energy management, April 5 2012. US Patent App. 12/895,780

  • Ilic MD et al (2016) Toward a unified modeling and control for sustainable and resilient electric energy systems. Found Trends Electric Energy Syst 1(1–2):1–141

    MathSciNet  Google Scholar 

  • Ilić M, Jaddivada R, Miao X (2017) Modeling and analysis methods for assessing stability of microgrids. IFAC-PapersOnLine 50(1):5448–5455

    Google Scholar 

  • Ilic M, Xia M, Jaddivada R (2018) Plug-and-play reconfigurable electric power microgrids, November 1 2018. US Patent App. 15/965,823

  • Ilic M, Jaddivada R, Miao X, Popli N (2019) Toward multi-layered MPC for complex electric energy systems. In: Handbook of model predictive control. Springer, pp 625–663

  • Ilic MD, Jaddivada R (2019) Toward technically feasible and economically efficient integration of distributed energy resources. In: 57th Annual Allerton conference on communication, control, and computing 2019. University of Illinois at Urbana-Champaign, Coordinated Science Laboratory (to appear)

  • Jiang R, Wang J, Guan Y (2011) Robust unit commitment with wind power and pumped storage hydro. IEEE Trans Power Syst 27(2):800–810

    Google Scholar 

  • Kim BH, Baldick R (2000) A comparison of distributed optimal power flow algorithms. IEEE Trans Power Syst 15(2):599–604

    Google Scholar 

  • Kocuk B, Dey SS, Sun XA (2016) Strong socp relaxations for the optimal power flow problem. Oper Res 64(6):1177–1196

    MathSciNet  MATH  Google Scholar 

  • Kosaraju KC, Pasumarthy R, Singh NM, Fradkov AL (2017) Control using new passivity property with differentiation at both ports. In: 2017 Indian control conference (ICC). IEEE, pp 7–11

  • Kraning M, Chu E, Lavaei J, Boyd S et al (2014) Dynamic network energy management via proximal message passing. Found Trends Optim 1(2):73–126

    Google Scholar 

  • Kundur P, Balu NJ, Lauby MG (1994) Power system stability and control, vol 7. McGraw-hill, New York

    Google Scholar 

  • Lam AYS, Zhang B, David NT (2012b) Distributed algorithms for optimal power flow problem. In: 2012 IEEE 51st IEEE conference on decision and control (CDC). IEEE, pp 430–437

  • Lam A, Dominguez-Garcia A, Zhang B, Tse D (2012a) Optimal distributed voltage regulation in power distribution networks. Technical report

  • LaWhite N, Ilic MD (1997) Vector space decomposition of reactive power for periodic nonsinusoidal signals. IEEE Trans Circuits Syst I Fund Theory Appl 44(4):338–346

    Google Scholar 

  • Li N, Qu G, Dahleh M (2014) Real-time decentralized voltage control in distribution networks. In: 2014 52nd annual Allerton conference on communication, control, and computing (Allerton). IEEE, pp 582–588

  • Liu J, Zhang W, Rizzoni G (2017) Robust stability analysis of dc microgrids with constant power loads. IEEE Trans Power Syst 33(1):851–860

    Google Scholar 

  • Luh PB, Blankson WE, Chen Y, Yan JH, Stern GA, Chang S-C, Zhao F (2006) Payment cost minimization auction for deregulated electricity markets using surrogate optimization. IEEE Trans Power Syst 21(2):568–578

    Google Scholar 

  • Ma W, Thorp JS (1993) An efficient algorithm to locate all the load flow solutions. IEEE Trans Power Syst 8(3):1077–1083

    Google Scholar 

  • Martin A, Müller JC, Pokutta S (2014) Strict linear prices in non-convex european day-ahead electricity markets. Optimn Methods Softw 29(1):189–221

    MathSciNet  MATH  Google Scholar 

  • Mehta D, Molzahn DK, Turitsyn K (2016) Recent advances in computational methods for the power flow equations. In: 2016 American control conference (ACC). IEEE, pp 1753–1765

  • Melamed M, Ben-Tal A, Golany B (2018) A multi-period unit commitment problem under a new hybrid uncertainty set for a renewable energy source. Renew Energy 118:909–917

    Google Scholar 

  • Miao X, Ilić MD (2017) Distributed model predictive control of synchronous machines for stabilizing microgrids. In: 2017 North American power symposium (NAPS). IEEE, pp 1–6

  • Miao X, Ilić M D (2015) Model predictive excitation control for constrained frequency and voltage stabilization. In: 2015 IEEE power and energy society general meeting. IEEE, pp 1–5

  • Miao X, Liu Q, Ilic M (2019) Enhanced automatic generation control (e-agc) for electric power systems with large intermittent renewable energy sources. arXiv preprint arXiv:1902.07644

  • Millner AR, Smith CL, Jaddivada R, Ilić MD (2018) Component standards for stable microgrids. IEEE Trans Power Syst 34(2):852–863

    Google Scholar 

  • Molina-Garcia A, Bouffard F, Kirschen DS (2010) Decentralized demand-side contribution to primary frequency control. IEEE Trans Power Syst 26(1):411–419

    Google Scholar 

  • Molzahn DK, Hiskens IA et al (2019) A survey of relaxations and approximations of the power flow equations. Found Trends Electric Energy Syst 4(1–2):1–221

    Google Scholar 

  • Molzahn DK, Dörfler F, Sandberg H, Low SH, Chakrabarti S, Baldick R, Lavaei J (2017) A survey of distributed optimization and control algorithms for electric power systems. IEEE Trans Smart Grid 8(6):2941–2962

    Google Scholar 

  • Mukherjee JC, Gupta A (2014) A review of charge scheduling of electric vehicles in smart grid. IEEE Syst J 9(4):1541–1553

    Google Scholar 

  • Navidi T, El Gamal A, Rajagopal R(2018) A two-layer decentralized control architecture for der coordination. In: 2018 IEEE conference on decision and control (CDC). IEEE, pp 6019–6024

  • Paliwal P, Patidar NP, Nema RK (2014) Planning of grid integrated distributed generators: a review of technology, objectives and techniques. Renew Sustain Energy Rev 40:557–570

    Google Scholar 

  • Pauphilet J, Kiner D, Faille D, El Ghaoui L (2016) A tractable numerical strategy for robust MILP and application to energy management. In: 2016 IEEE 55th conference on decision and control (CDC). IEEE, pp 1490–1495

  • Pena R, Clare JC, Asher GM (1996) Doubly fed induction generator using back-to-back pwm converters and its application to variable-speed wind-energy generation. IEE Proc Electric Power Appl 143(3):231–241

    Google Scholar 

  • Penfield P, Spence R, Duinker S (1970) A generalized form of Tellegen’s theorem. IEEE Trans Circuit Theory 17(3):302–305

    MathSciNet  Google Scholar 

  • Peng Q, Low SH (2016) Distributed optimal power flow algorithm for radial networks, I: balanced single phase case. IEEE Trans Smart Grid 9(1):111–121

    Google Scholar 

  • Popli N, Ilić MD (2015) Enabling convex energy bids for flexible ramp product via smart local automation. IFAC-PapersOnLine 48(30):245–250

    Google Scholar 

  • Popli N, Jaddivada R, O’Sullivan F, Ilic M (2019) Accommodating wind and solar in low-flexibility systems: unbundling flexibilities for sequential dispatch. IEEE Trans Power Syst (Submitted)

  • Rotting TA, Gjelsvik A (1992) Stochastic dual dynamic programming for seasonal scheduling in the norwegian power system. IEEE Trans Power Syst 7(1):273–279

    Google Scholar 

  • Ruiz C, Conejo AJ, Gabriel SA (2012) Pricing non-convexities in an electricity pool. IEEE Trans Power Syst 27(3):1334–1342

    Google Scholar 

  • Salcedo RO, Nowocin JK, Smith CL, Rekha RP, Corbett EG, Limpaecher ER, LaPenta JM (2016) Development of a real-time hardware-in-the-loop power systems simulation platform to evaluate commercial microgrid controllers. Technical report. Massachusetts Inst of Tech Lexington Lincoln lab

  • Salcedo R, Corbett E, Smith C, Limpaecher E, Rekha R, Nowocin J, Lauss G, Fonkwe E, Almeida M, Gartner P et al (2019) Banshee distribution network benchmark and prototyping platform for hardware-in-the-loop integration of microgrid and device controllers. J Eng 2019(8):5365–5373

    Google Scholar 

  • Sanandaji BM, Hao H, Poolla K (2014) Fast regulation service provision via aggregation of thermostatically controlled loads. In: 2014 47th Hawaii international conference on system sciences. IEEE, pp 2388–2397

  • Sang-Jin O, Yoo C-H, Chung I-Y, Won D-J (2013) Hardware-in-the-loop simulation of distributed intelligent energy management system for microgrids. Energies 6(7):3263–3283

    Google Scholar 

  • Siano P (2014) Demand response and smart grids—a survey. Renew Sustain Energy Rev 30:461–478

    Google Scholar 

  • Sun B, Luh PB, Jia Q-S, Yan B (2015) Event-based optimization within the lagrangian relaxation framework for energy savings in HVAC systems. IEEE Trans Autom Sci Eng 12(4):1396–1406

    Google Scholar 

  • Tashtoush B, Molhim M, Al-Rousan M (2005) Dynamic model of an HVAC system for control analysis. Energy 30(10):1729–1745

    Google Scholar 

  • Thatte AA, Xie L, Viassolo DE, Singh S (2013) Risk measure based robust bidding strategy for arbitrage using a wind farm and energy storage. IEEE Trans Smart Grid 4(4):2191–2199

    Google Scholar 

  • Toledo OM, Filho DO, Diniz ASAC (2010) Distributed photovoltaic generation and energy storage systems: a review. Renew Sustain Energy Rev 14(1):506–511

    Google Scholar 

  • US Federal Energy Regulatory Commission et al (2016) Electric storage participation in markets operated by regional transmission organization and independent system operators. Washington, DC, p 139

  • van der Meer AA, Gibescu M, van der Meijden MAMM, Kling WL, Ferreira JA (2014) Advanced hybrid transient stability and emt simulation for VSC-HVDC systems. IEEE Trans Power Deliv 30(3):1057–1066

    Google Scholar 

  • van der Schaft AJ, Van Der Schaft AJ (2000) L2-gain and passivity techniques in nonlinear control, vol 2. Springer, New York

    MATH  Google Scholar 

  • Wang Q, Zhang C, Ding Y, Xydis G, Wang J, Østergaard J (2015) Review of real-time electricity markets for integrating distributed energy resources and demand response. Appl Energy 138:695–706

    Google Scholar 

  • Wang Y, Lei W, Wang S (2016) A fully-decentralized consensus-based admm approach for dc-opf with demand response. IEEE Trans Smart Grid 8(6):2637–2647

    Google Scholar 

  • Wei J, Kundur D (2012) Two-tier hierarchical cyber-physical security analysis framework for smart grid. In: 2012 IEEE power and energy society general meeting. IEEE, pp 1–5

  • Wei E, Ozdaglar A (2012) Distributed alternating direction method of multipliers. In: 2012 IEEE 51st IEEE conference on decision and control (CDC). IEEE, pp 5445–5450

  • Werbos PJ (2018) Ai intelligence for the grid 16 years later: progress, challenges and lessons for other sectors. In: 2018 International joint conference on neural networks (IJCNN). IEEE, pp 1–8

  • Wood AJ, Wollenberg BF, Sheblé GB (2013) Power generation, operation, and control. Wiley, New York

    Google Scholar 

  • Wyatt JL, Ilic M (1990) Time-domain reactive power concepts for nonlinear, nonsinusoidal or nonperiodic networks. In: IEEE international symposium on circuits and systems. IEEE, pp 387–390

  • Xie L, Yingzhong G, Zhu X, Genton MG (2013) Short-term spatio-temporal wind power forecast in robust look-ahead power system dispatch. IEEE Trans Smart Grid 5(1):511–520

    Google Scholar 

  • Xu B, Korpas M, Botterud A, O’Sullivan F (2019) A Lagrangian policy for optimal energy storage control. arXiv preprint arXiv:1901.09507

  • Yuan H, Li F, Wei Y, Zhu J (2016) Novel linearized power flow and linearized opf models for active distribution networks with application in distribution lmp. IEEE Trans Smart Grid 9(1):438–448

    Google Scholar 

  • Zavala VM (2016) New architectures for hierarchical predictive control. IFAC-PapersOnLine 49(7):43–48

    MathSciNet  Google Scholar 

  • Zhang W, Kalsi K, Fuller J, Elizondo M, Chassin D (2012a) Aggregate model for heterogeneous thermostatically controlled loads with demand response. In: 2012 IEEE power and energy society general meeting. IEEE, pp 1–8

  • Zhang W, Kalsi K, Fuller J, Elizondo M, Chassin D (2012b) Aggregate model for heterogeneous thermostatically controlled loads with demand response. In: 2012 IEEE Power and energy society general meeting. IEEE, pp 1–8

  • Zhang X, Shi W, Yan B, Malkawi A, Li N (2017) Decentralized and distributed temperature control via HVAC systems in energy efficient buildings. arXiv preprint arXiv:1702.03308

  • Zhao C, Wang J, Watson J-P, Guan Y (2013) Multi-stage robust unit commitment considering wind and demand response uncertainties. IEEE Trans Power Syst 28(3):2708–2717

    Google Scholar 

  • Zhao C, Topcu U, Li N, Low S (2014) Design and stability of load-side primary frequency control in power systems. IEEE Trans Autom Control 59(5):1177–1189

    MathSciNet  MATH  Google Scholar 

  • Zhao C, Dall’Anese E, Low SH (2017) Convex relaxation of OPF in multiphase radial networks with delta connection. In: Proceedings of the 10th bulk power systems dynamics and control symposium

  • Zhao L, Zeng B (2012) Robust unit commitment problem with demand response and wind energy. In: 2012 IEEE power and energy society general meeting. IEEE, pp 1–8

  • Zhiwei X, Callaway DS, Zechun H, Song Y (2016) Hierarchical coordination of heterogeneous flexible loads. IEEE Trans Power Syst 31(6):4206–4216

    Google Scholar 

  • Zhou X, Dall’Anese E, Chen L, Simonetto A (2017) An incentive-based online optimization framework for distribution grids. IEEE Trans Autom Control 63(7):2019–2031

    MathSciNet  MATH  Google Scholar 

  • Zonetti D, Ortega R, Schiffer J (2017) A tool for stability and power-sharing analysis of a generalized class of droop controllers for high-voltage direct-current transmission systems. IEEE Trans Control Netw Syst 5(3):1110–1119

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This material is based upon work supported by the Department of Energy under Air Force Contract No. FA8721-05-C-0002 and/or FA8702-15-D-0001. Any opinions, findings, conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the Department of Energy.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rupamathi Jaddivada.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix

1.1 A Theoretical basis for the provability of DyMonDS-based bids

Let us first pose the problem for allocation of resources in general form. From the benchmark problem described in Eq. (19), one can see that the coupling constraints are all linear in interaction space. Let us use the notation \(z_i[k]\) to represent interactions including \(E_i, p_i, P_i, \dot{Q}_i\) for all time samples \(kT_t \in [t, t+H]\). The internal variables of including \(x_i[k], u_i[k], m_i[k]\) for the same time samples are all abstracted in the vector \(x_i\) (with slight abuse of notation). Let xz denote the vectors \(x_i, z_i\) for all components stacked up.

The centralized problem can then be posed in a general form as shown below:

$$\begin{aligned} \begin{array}{l} {(C):}\,\,{\mathop {\min }\limits _{{z_i}}} \sum \limits _{i \in \mathcal {N}} {f_i(z_i)} \\ s.t.\\ \quad\quad g(z) \le 0 \quad \left( \lambda \right) \\ \qquad h_i(x_i,z_i) \le 0 \quad \left( \mu _i\right) \\ \end{array} \end{aligned}$$
(36)

The first constraint function g(z) is a coupling constraint while the second one \(h_i(x_i,z_i)\) is specific to each agent i. Solving the problem in Eq. (36) turns out to be complicated since the number of components given by \(|\mathcal {N}|\) is extremely large. While the coupling constraints given by g(z) are linear, the number of internal constraints modeled through \(h_i\) for each agent i can be high, leading to a large-scale problem to be solved at once.

In this general formulation and the proofs to follow, a few assumptions are made:

Assumption 1

The cost function \(f_i: \mathbb {R}^{n_{i}} \rightarrow \mathbb {R}\) is a smooth convex function which is continuously differentiable with Lipschitz continuous gradients \(L(f_i) > 0\) given as

$$\begin{aligned} \left\| \nabla {f_i}({x_i}) - \nabla {f_i}({y_i})\right\| \le L(f_i)\left\| {x_i} - {y_i}\right\| \quad \forall x_i,y_i\in \mathbb {R}^{n_i} \end{aligned}$$
(37)

where \(\Vert .\Vert \) denotes the standard Euclidean norm.

Assumption 2

The constraint set given in (36) is a closed convex set.

Exploiting the linearity of the coupling constraints, we have proposed a distributed approach towards solving the problem (C). As explained in Sect. 6, we propose to have the system operator solve the master problem comprising the coupling constraints alone which are of the same order as the number of nodes in the network. The increased cardinality of the problem created by local constraints and their preferences are rather abstracted through bid functions as computed by solving the agent-level problem in Eq. (38)

$$\begin{aligned} \begin{array}{l} {(Agent-i)}:{\mathop {\min }\limits _{{x_i},z_i}} {C_i}({z_i}) + {\lambda }_j^T g_j(z_i)\\ \qquad \qquad s.t.\\ \qquad \qquad h_i(x_i, z_i) \le 0 \\ \end{array} \end{aligned}$$
(38)

Here, \(\lambda _j\) corresponds to the vector of all the elements of vector \(\lambda \) for which there exists dependence of the constraint on the elements of \(z_i\) i.e. \(j \in \mathcal {J}\) to denote constraint indexes to which \(z_i\) contributes.

By solving the problem \(\left( Agent-i\right) \) for perturbations in each of the elements in \(\lambda _j\), one can obtain the bid functions \(B^j_i(z_i)\) for use by the operator to solve the system level problem. The willingness of the agent to participate depends on its internal constraints but also on the price incentive. The bid is such that it makes marginally the same amount of cost as it gets paid through the respective \(\lambda _j\). As a result,

$$\begin{aligned} \frac{\partial B_i^j}{\partial z_i} = -\lambda _j = a^j_iz_i + b^j_i \end{aligned}$$
(39)

The coefficients \(a^j_i\) and \(b^j_i\) are computed empirically by obtaining the points \(z_i\) by solving \(\left( Agent-i\right) \) for slight perturbations in \(\lambda \). This function is integrated with respect to \(z_i\) to obtain

$$\begin{aligned} B_i(z_i) = \sum _{j \in \mathcal {J}} B_i^j(z_i) = \frac{1}{2}z_i^T a_i z_i + b_i z_i \end{aligned}$$
(40)

Here, \(a_i = \sum _{j \in \mathcal {J}}a_i^j\) and \(b_i = \sum _{j \in \mathcal {J}}b_i^j\). Along with the bid functions, the feasibility region of the agent i is projected onto the \(z_i\) plane represented by the space \(\varPi _{i}\). This projected space varies with time as operating conditions change. This plane is found by utilizing the constraints Eq. (19f). These bid functions along with the time varying limits of the feasible region of interaction variables are all collected by the system operator to solve the following problem:

$$\begin{aligned} \begin{array}{*{20}{l}} (S): {\mathop {\min }\limits _{z_i}}\sum \limits _{i \in \mathcal {N} } {{B_i}(z_i)}\\ {s.t.}\\ \qquad \qquad g(z) \le 0 \qquad \left( \lambda \right) \\ \qquad \qquad z_i \in \varPi _{g,i} \end{array} \end{aligned}$$
(41)

Assumption 3

The problem \(\left( C\right) \) is solvable and there exists a unique minimizer \(x^*\).

Notice that the projected space \(\varPi _{i}\) is also convex since the projection map is a linear operator, the constraint set of problem (S) is convex. Since the bid functions are by construction convex, there would be a unique minimizer of \(\left( S\right) \). We next prove that the optimal point obtained by the proposed distributed scheme is the same as the one obtained by the centralized approach.

1.2 A.1 Proof for approaching the centralized solution optimum operating point through proposed primal–dual decomposition strategy

Theorem 1

Under assumptions 12, the solution obtained by solving\(\left( S\right) \)results in the same solution as that of the\(\left( C\right) \)when the bids\(B_i(z_i)\)are created by solving problem in\(\left( Agent-i\right) \)in response to small perturbations around\(\lambda = \lambda ^*\)where\(\lambda ^*\)is the dual solution of the problem\(\left( C\right) \).

Proof

Writing the KKT conditions for the problem \(\left( C\right) \), we obtain:

Stationarity:

$$\begin{aligned}&\nabla _{z_i} C_i(z_i) + \lambda ^T \nabla _{z_i} g(z) + \mu _i^T \nabla _{z_i} h_{i} (z_i,x_i) = 0 \qquad \forall i \in \mathcal {N} \end{aligned}$$
(42a)
$$\begin{aligned}&\mu _i^T \nabla _{x_i} h_{i} (z_i,x_i) = 0 \end{aligned}$$
(42b)

Primal feasibility:

$$\begin{aligned} \qquad g(z)\le & {} 0 \end{aligned}$$
(42c)
$$\begin{aligned} h_{i}(x_i,z_i)\le & {} 0 \end{aligned}$$
(42d)

Dual feasibility:

$$\begin{aligned} \lambda\ge & {} 0 \qquad \,\,\, \end{aligned}$$
(42e)
$$\begin{aligned} \mu _{i}\ge & {} 0 \,\,\,\forall i\in \mathcal {N} \end{aligned}$$
(42f)

Complementarity slackness:

$$\begin{aligned} \lambda ^T g(z)= & {} 0 \end{aligned}$$
(42g)
$$\begin{aligned} \mu _i^T h_{i} (x_i,z_i)= & {} 0 \,\,\,\forall i \in \mathcal {N} \end{aligned}$$
(42h)

Let the solution of above set of equations be \(x_i^*, z_i^*, \lambda ^*, \mu _i^*\).

Now, solving the problem posed in \((Agent-i)\) for a given value of \(\lambda \), we obtain the solution of the agent-level problem. As explained previously, the bid function is constructed by integrating the empirically constructed sensitivity given by

$$\begin{aligned} \lambda (z_i) = -\nabla _{z_i} B_i(z_i) = a_i z_i + b_i \end{aligned}$$
(43)

This results in the bid function \(B_i(z_i) = \frac{1}{2}a_iz_i^2 + b_iz_i\).

Analytically, one can write the KKT conditions of the problem \(\left( Agent-i\right) \) to notice that the gradient of the bid function satisfies the following relation:

$$\begin{aligned} -\nabla _{z_i} B_i(z_i) \nabla _{z_i} g(z) + \nabla _{z_i} C_{i} (z_i) + \lambda ^T \nabla _{z_i}h(x_i,z_i) = 0 \end{aligned}$$
(44a)

while also satisfying the device-specific KKT conditions given by the set \(S_i\) i.e.

$$\begin{aligned} S_i = \left\{ \left( x_i, z_i, \lambda _i, \mu _i\right) | \hbox {(42b)}, \hbox {(42d)}; \hbox {(42f)}; \hbox {(42h)} \right\} \end{aligned}$$
(44b)

The set above which when projected onto the space of the \(z_i\) plane results in \(z_i \in \varPi _{g,i}\), representing the space of projection of the half space represented by \(g_i(x_i,z_i)\) onto the \(z_i\) plane in order to commit operating conditions dependent limits of \(z_i\) as a part of the bid function. Clearly, \(z_i \in \varPi _{i} \implies \left( x_i, z_i, \lambda _i, \mu _i \right) \in S_i\) for some values of \(x_i, \lambda _i, \mu _i\).

Now considering these bid functions and the projected constraint on \(z_i\) in the problem \(\left( S\right) \). The KKT conditions of \(\left( S\right) \)can be written as:

Stationarity:

$$\begin{aligned} \nabla B_i(z_i) + \lambda ^T \nabla _{z_i} g(z_i) \quad \forall i \end{aligned}$$
(45a)

Primal, dual feasibility and complementary slackness:

Equations (42c), (42e), (42g)

An added constraint of the bid function that \(z_i\) belongs to the projections of the set \(S_i\) ensures the other internal KKT constraints in Eq. (36) remain feasible. Upon substituting the relations in Eq. (44) into the stationarity condition in Eq. (45a) for each agent i and summing them up, one could obtain the KKT conditions in Eq. (42). Thus proved. \(\square \)

Remark 14

When the internal constraints \(g_i, h_i\) are non-convex, multiple agent-level solutions may exist and thus centralized and distributed solutions may not be the same.

B Proof for the rate of convergence of the DyMonDS-based sensitivity bids

1.1 B.1 Agent-level bid function

Following the definitions and proof sketch in Beck and Teboulle (2009), we now analyse the convergence rates for the proposed approach. First, we define the indicator function for the constraint set of the problem \(\left( Agent-i\right) \) as \(H_i(x_i,z_i)\). With this, we obtain a function to be minimized at the agent-level as

$$\begin{aligned} F_i(x_i, z_i, \lambda _j) = f_i(x_i, z_i) + H_i(x_i,z_i) + \lambda _j^T g_j(z_i) \end{aligned}$$
(46)

Here, it is assumed that each agent’s cost function is strongly convex, smooth and differentiable with a Lipschitz constant \(L_i(f_i)\).

Definition 1

The quadratic approximation of \(F_i(x_i,z_i,\lambda )\) around the point \(y_i\) for \(L_i \ge L_i(f_i)\) is given as

$$\begin{aligned} Q_{L,i}(x_i,z_i, \lambda _j)= & {} f_i(x_i,y_i,\lambda _j) + \left\langle z_i-y_i, \nabla _{z_i} f_i(x_i,y_i,\lambda )\right\rangle \nonumber \\&+ \frac{L_i}{2} \Vert z_i -y_i\Vert _2^2 + H_i(x_i,z_i) + \lambda _j^Tg_j(z_i) \end{aligned}$$
(47)

It admits a unique minimizer

$$\begin{aligned} {p_{L,i}}(x_i,{y_i},\lambda _j ) = \mathop {\arg \min }\limits _{z_i} {Q_{L,i}}({x_i},{z_i},\lambda _j) \end{aligned}$$
(48)

By denoting \(z_i\) as the minimizer and by using the notation \(\gamma _i\left( x_i,y_i,\right) \in \frac{\partial H_i}{\partial z_i}(x_i,z_i)\), the optimality condition can be obtained from Eq. (48) by taking the partial derivative w.r.t \(z_i\) and equating to zero.

$$\begin{aligned} \nabla _{z_i} f_i(x_i,y_i) + L_i\left( z_i-y_i\right) + \gamma _i\left( x_i,y_i\right) + \lambda _j^T\frac{\partial g_j}{\partial z_i} = 0 \end{aligned}$$
(49)

Since the coupling constraints are linear, the partial derivative \(\frac{\partial g_j}{\partial z_i}\) is a constant matrix. As explained previously, the marginal bid function at the agent level is equal to the expression for \(\lambda _j\), for which the analytical expression can be decomposed into a slope \(a_i\) and intercept value \(b_i\)

$$\begin{aligned} \frac{\partial B^j_i}{\partial z_i} = -\lambda _j = \underbrace{\left( \frac{\partial g_j}{\partial z_i}\right) ^{\dagger }L_i}_{a^j_{i}} z_i + \underbrace{\left( \frac{\partial g_j}{\partial z_i}\right) ^{\dagger }\left( \nabla _z f_i(x_i,y_i) + \gamma _i\left( x_i,y_i\right) - Ly_i\right) }_{b^j_{i}} \end{aligned}$$
(50)

Here, \((.)^{\dagger }\) represents the pseudo inverse operation. The bid function then can be re-expressed as

$$\begin{aligned} B_i(z_i) = \sum _{j \in \mathcal {J}} B_i^j(z_i) = \frac{1}{2}z_i^T a_i(y_i) z_i + b_i(y_i) z_i \end{aligned}$$
(51)

By collecting these bid functions from all the agents, the system cost function is constructed as

$$\begin{aligned} f^s(z) = \sum _{i \in \mathcal {N}} = \frac{1}{2} z_i^Ta_i(y_i)z_i + b_i(y_i)^Tz_i \end{aligned}$$
(52)

Remark 15

Notice that the system-lvel cost function is convex by construction. At each of the agents, the respective quadratic functions constructed, approximate the variation of their cost and the internal constraints around y.

1.2 B.2 System-level optimality conditions

Similarly, defining the function \(F^s\left( z\right) \) for the problem to be solved by the system operator as

$$\begin{aligned} F^s(z^k) = f^s(z^k,z^{k-1}) + G(z^k) \end{aligned}$$
(53)

where \(G(z^k)\) represents the indicator function of the coupling constraints \(g(z^k) \le 0\). We have \(Q^s\left( z^k,y\right) \) to denote the quadratic approximation \(F^s(z^k)\) around a point y for some \(L^s>0\) Around an arbitrary point y, the cost function coefficients are computed using the initial guess \(z^0\).

$$\begin{aligned} Q^s(z^k,y^k) = f^s(z^k,z^{k-1}) + \left\langle z^k-y, \nabla _z f^s\left( y,z^0\right) \right\rangle + \frac{L^s}{2} \Vert z^k-y^k\Vert ^2 + G(z^k) \end{aligned}$$
(54)

The unique minimizer of the quadratic approximation of \(F^s(z^k)\) around a point y is denoted as \(p_L^s\left( y\right) \). The optimality conditions can explicitly be written as

$$\begin{aligned} \nabla _z f^s(y,z^{0})+ L^s\left( z^k-y\right) + \gamma ^s\left( y\right) = 0 \end{aligned}$$
(55)

Let us now define the sequence \(\left\{ z^k\right\} \) given by \(z^k = p^s_L(z^{k-1})\)

Remark 16

The series \(F^s\left( z^k\right) \) is contracting.

Notice that \(F^s(z^k) \le Q^s\left( z^k,z^{k-1}\right) \le Q^s\left( z^{k-1},z^{k-1}\right) = F^s(z^{k-1})\)

We will now derive its convergence rate.

1.3 B.3 Convergence proof

We begin by revising one of the lemmas in Beck and Teboulle (2009) for the distributed setting here, which will be utilized in the rest of our convergence proof:

Lemma 1

For some\(L^s>0\)and\(y\in \mathbb {R}^{n_z}\)if\(F^s\left( p_L^s\left( y\right) \right) \le Q_L^s\left( p_L^s\left( y\right) ,y\right) \). Then, \(\forall z \in \mathbb {R}^{n_z}\),

$$\begin{aligned} F^s(z^k) - F^s\left( p_L^s\left( y\right) \right) \ge \frac{L^s}{2}\Vert p^s_L(y) - y\Vert ^2 + L^s\left\langle y-z^k, p^s_{L}(y)- y\right\rangle \end{aligned}$$
(56)

Proof

Notice that the condition in Lemma holds true if \(L^s = L^s_y = diag(a_i(y_i))\) i.e. the matrix created by the quadratic coefficient of the bid function. Because of the convexity of \(f^s\) with respect to \(z^k\) and since G is an indicator function of the convex set of coupling constraints, we can establish the following relations:

$$\begin{aligned} f^s(z^k,z^{k-1})\ge & {} f^s(y, z^0) + \left\langle z^k - y, \nabla f^s(y, z^{0})\right\rangle \end{aligned}$$
(57)
$$\begin{aligned} g(z^k)\ge & {} g(p^s_L(y)) + \left\langle z^k - p^s_L(y), \gamma (y) \right\rangle \end{aligned}$$
(58)

We can now write the expression for the objective function as

$$\begin{aligned} F^s(z^k) \ge f^s(y, z^{0}) + \left\langle z^k - y, \nabla f^s(y, z^{0})\right\rangle + g(p_L(y)) + \left\langle z^k - p_L(y), \gamma (y) \right\rangle \end{aligned}$$
(59)

Furthermore, from the definition of Q, we have

$$\begin{aligned} Q(p_L(y),y) = f^s(y,y^{0}) + \left\langle p_L(y) - y, \nabla f^s(y, z^{0}) \right\rangle + \frac{L^s}{2} \Vert p_L(y) - y\Vert ^2+ g(p_L(y)) \end{aligned}$$
(60)

Subtracting the two equations, we obtain

$$\begin{aligned} F^s(z^k) - Q(p_L(y),y) \ge \left\langle z^k - p_L(y^k), \nabla f^s(y, z^{0}) + \gamma (y) \right\rangle - \frac{L^s}{2}\Vert p_L(y) - y\Vert ^2 \end{aligned}$$
(61)

Utilizing the assumption stated in the lemma and substituting the optimality condition in Eq. (55) in the right hand side, we obtain

$$\begin{aligned} F^s(z^k) - F^s(p_L(y))\ge & {} L \left\langle z^k - p_L(y^k), z^k - y \right\rangle - \frac{L^s}{2}\Vert p_L(y) - y\Vert ^2\nonumber \\= & {} \frac{L^s}{2}\Vert p_L(y) - y\Vert ^2 + L^s \left\langle y - z^k, p_L(y)-y \right\rangle \end{aligned}$$
(62)

\(\square \)

Theorem 2

Let\(\left\{ z^k\right\} \)denote the series produced by the update rule\(z^k = p_L^s\left( z^{k-1}\right) \). Then, the system-level objective function converges at the rate of\(\mathcal {O}\left( 1/k\right) \)as follows:

$$\begin{aligned} F^s\left( z^k\right) -F^s\left( z^*\right) \le \frac{{\Vert z^0-z^*\Vert }_{L0}}{2k} \end{aligned}$$
(63)

where\(z^*\)is the optimal point and\(L_0\)is the diagonal matrix consisting of the quadratic coefficients of the bid functions of each of the agents.

Proof

Invoking the Lemma 1 with \(z^k = z^*, y = z^n, L^s = L_n = diag(a_i(z^n_i))\), we have

$$\begin{aligned} {F^s}({z^*}) - {F^s}\left( {{z^{n + 1}}} \right)\ge & {} \frac{{{L_n}}}{2}\left( \Vert {{z^{n + 1}} - {z^n}\Vert ^2 + 2{L_n}\left\langle {{z^n} - {z^*},{z^{n + 1}} - {z^n}} \right\rangle } \right) \nonumber \\= & {} \frac{L_n}{2} \left( {\Vert {z^*} - {z^{n + 1}}\Vert }^2- {\Vert {z^*} - {z^{n}\Vert }}^2\right) \nonumber \\= & {} \frac{1}{2} \left( {\Vert {z^*} - {z^{n + 1}}\Vert ^2_{L_n}}- {\Vert {z^*} - {z^{n}\Vert ^2_{L_n}}}\right) \end{aligned}$$
(64)

Next, invoking the Lemma 1 with \(z^k = z^n, y = z^n, L^s = L_n = diag(a_i(z^n_i))\), we have

$$\begin{aligned} {F^s}({z^n}) - {F^s}\left( {{z^{n + 1}}} \right) \ge \frac{1}{2} {\Vert {z^{n + 1}} - {z^n}\Vert ^2}_{L_n} \end{aligned}$$
(65)

Summing the inequality in Eq. (64) over \(n = 0,1,\ldots k-1\), we have

$$\begin{aligned} k{F^s}({z^*}) - \sum \limits _{n = 0}^{k - 1} {{F^s}\left( {{z^{n + 1}}} \right) } \ge \Vert z^* - z^k\Vert ^2_{L_{k-1}} - \Vert z^* - z^0\Vert ^2_{L_{0}} \end{aligned}$$
(66)

Next, multiplying the inequality in Eq. (65) by n and summing the result over \(n = 0,1,, \ldots k-1\), we have

$$\begin{aligned} -k{F^s}({z^k}) + \sum \limits _{n = 0}^{k - 1} {{F^s}\left( {{z^{n + 1}}} \right) } \ge \sum \limits _{n = 0}^{k - 1} {n \Vert z^{n+1} - z^0\Vert _{L_n}^2} \end{aligned}$$
(67)

Now, by adding equations (66) and (67), we obtain

$$\begin{aligned} F^s\left( z^k\right) -F^s\left( z^*\right) \le \frac{{\Vert z^*-z^0\Vert }_{L0}}{2k} \end{aligned}$$
(68)

\(\square \)

Here, the cleared values at each iteration are communicated to the agents which are utilized to compute the new bid functions. As a result, at \(k^{th}\) iteration, the bids are computed using the solution obtain by system coordinator at \((k-1)^{th}\) iteration. The system level objective function convergence stated above indicates that these bid functions converge at a rate of \(\mathcal {O}(1/k)\).

Note also that error at each iteration is because of the quadratic approximation of the bid function by the agents. The error in the This result combined with the existence of unique solution as obtained by the centralized and distributed solution strategies as shown in Theorem 1 ensures that the DyMonDS approach leads to the centralized optimal solution.

Remark 17

The computational complexity through this approach of minimal coordination is of the order \(\mathcal {O}\left( n_z^3\right) + \sum _{i=1}^{|\mathcal {N}|} \mathcal {O}\left( n_{x_i}^3\right) \), where \(n_z, n_{x,i}\) respectively represent the number of interaction variables, and number of state variables in each of the components of the network. Typically \(n_{x_i}\) for each i is much smaller than \(n_{z}\) of the entire system.

The centralized solution complexity would be of the order \(\mathcal {O}\left( \left( n_z + \sum _{i=1}^{|\mathcal {N}|} n_{x_i}\right) ^3\right) \)!

In order to compare with iterative approaches such as ADMM, simple update equations are typically utilized in addition to solving the local agent problems with a computational cost of \(\sum _{i=1}^{|\mathcal {N}|}\mathcal {O}\left( n_{x_i}^3\right) \). However, the number of iterations needed is typically of the order \(1e^6\), for obtaining an accuracy of \(1e^{-3}\) in the solution vector.

1.4 B.4 Proof-of-concept numerical simulations

In order to show the effectiveness of the proposed DyMonDS-based interactive scheme to solving an optimization problem coupled across multiple agents, we consider a communication network shown in Fig. 8.

Fig. 8
figure 8

Test network: Source destination pair is displayed with same color; \(L_l: x_i, x_j \ldots \) denotes that the flow on link l is due to the flow from sources \(i, j, \ldots \) (Beck et al. 2014)

It has two source-load pairs and the objective is to maximize the network utility. Each link has dedicated source power injections as shown in the Fig. 8 with an upper limit on the capacity of the flow through the wires. Let each of the sources have the utility functions \(u_i(x_1) = C_ilog(x_1 + 0.1)\) where \(C_1 = 10\) and \(C_2 = 20\). Let the set of sources be denoted using the set \(\mathcal {S}\) and let the set of sources utilizing the link l be denoted using \(\mathcal {S}(l)\). Similarly, let the length of the path in use by source i be denoted as \(\mathcal {L}(i)\). Letting the capacity of all the links be equal to 1, the network utility maximization (NUM) problem can be posed as

$$\begin{aligned} \begin{array}{l} {\mathop {\max }\limits _{x}} \sum \limits _{i \in \mathcal{S}} {{C_i}\log ({x_1} + 0.1)} \\ \sum \limits _{i \in \mathcal{S}(l)} {{x_i} \le 1\quad \forall l \in \mathcal{L}} \\ {x_i} \ge 0\quad \forall i \in \mathcal{S} \end{array} \end{aligned}$$
(69)

In order to apply DyMonDS-based method, this formulation is a degenerate case of the one posed in Eq. (36) where there are no internal constraints and all the variables appear the intersections. Thus, \(x_i\) in this formulation is to be treated like \(z_i\) in the formulation in Then the agent-specific formulation is to just optimize its own utility function given the lagrange multipliers corresponding to its coupling constraints. This can be posed as follows:

$$\begin{aligned} \begin{array}{l} {\mathop {\max }\limits _{x_i}} {{C_i}\log ({x_i} + 0.1)} - \sum _{j\in \mathcal {L}(i)} \lambda _j x_i \\ {x_i} \ge 0 \end{array} \end{aligned}$$
(70)

This problem is solved to obtain bids using the relations in Eq. (50) which reduces to the following for this degenerate case of having no internal constraints

$$\begin{aligned} \frac{\partial B_i}{\partial x_i} = \underbrace{\nabla ^2_xf_i(y)}_{a^{k}_i} x_i + \underbrace{\left( \nabla _x f_i(y) - \nabla ^2_x f_i(y) y\right) }_{b^k_i} \end{aligned}$$
(71)

Here \(f_i(x_i) = C_i log(x_i + 0.1)\), y is the solution obtained by solving the agent level problem from \(\lambda ^{k-1}\) in the previous iterate utilized to obtain the bid function for the next iteration. The bid constructed as \(B^k_i(x_i) = \frac{1}{2}a_i^k x_i^2 + b_i^kx_i\), which are collected from all the sources and then optimized by the system coordinator by solving the following problem

$$\begin{aligned} \begin{array}{l} {\mathop {\max }\limits _{x}} \sum \limits _{i \in \mathcal{S}} {{B_i}({x_i})} \\ \sum \limits _{i \in \mathcal{S}(l)} {{x_i} \le 1\quad \forall l \in \mathcal{L}} \\ {x_i} \ge 0 \quad \forall i \in \mathcal{S} \end{array} \end{aligned}$$
(72)

Traditionally utilized first order gradient methods such as the distributed dual gradient and the fast dual gradient method are simulated for comparizon with the proposed DyMonDS-based approach. For the dual gradient algorithm and fast gradient algorithm, the step size for lagrange multiplier increments have been assumed to equal to \(\alpha =\frac{2 \sigma }{N_p}{N_s}\) where \(\sigma = \nabla ^2 f_i(1)\) is the strong convexity constant for the utility functions in used (Beck et al. 2014). \(N_p\) and \(N_s\) respectively are the longest path lengths among all sources and maximum number of sources sharing particular link respectively.

For our DyMonDS approach, we do not have to go through the hassle of selecting a step-size however. For all the methods, all of the following termination conditions have been utilized (Beck et al. 2014):

  • primal objective function values satisfy \(\left| \frac{f(x^{k+1})}{f(x^k)} - 1\right| \le 0.01\)

  • dual variables satisfy the worst case differences \(\Vert \lambda ^{k+1} - \lambda ^{k}\Vert _{\infty } \le 0.01\)

  • primal feasibility satisfies \(A x^k - c < 0.01\) for all the links

Fig. 9
figure 9

Comparison of the rate of convergence of the proposed DyMonDS approach with that of two other distributed methods for a system with 2 sources and 5 links

All the tested methods converge to a value equal to \(x_1 = 0.3\) and \(x_2 = 0.7\) for when \(C_1 = 10\) and \(C_2 = 20\) in the utility functions used in the formulations of NUM in Eq. (69). The convergence rates in terms of the objective function values for all three methods is shown in Fig. 9. Notice the effectiveness of the DyMonDS-based approach which only takes about 10 iterations for convergences in comparison to the other methods which can go upto thousands of iterations. Furthermore, just with a couple of iterations, the approach tends to converge to a near-optimal point which is permissible for large-scale systems.

Similar to the analysis in Beck et al. (2014), we have further produced random networks with a random number of sources in the range [1, 25] and a random number of links in the range of [1, 40] in order to test the scalability. For each of the 50 trials performed, the number of iterations for convergence is shown in the Fig. 10. As anticipated the DyMonDS approach requires very few number of iterations irrespective of the size of the system.

Fig. 10
figure 10

Comparison of iterations for random network

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ilic, M., Jaddivada, R. Unified value-based feedback, optimization and risk management in complex electric energy systems. Optim Eng 21, 427–483 (2020). https://doi.org/10.1007/s11081-020-09486-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11081-020-09486-y

Keywords

Navigation