1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

15
1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008 Distributed MPC Strategies With Application to Power System Automatic Generation Control Aswin N. Venkat, Ian A. Hiskens, Fellow, IEEE, James B. Rawlings, and Stephen J. Wright Abstract—A distributed model predictive control (MPC) frame- work, suitable for controlling large-scale networked systems such as power systems, is presented. The overall system is de- composed into subsystems, each with its own MPC controller. These subsystem-based MPCs work iteratively and cooperatively towards satisfying systemwide control objectives. If available computational time allows convergence, the proposed distributed MPC framework achieves performance equivalent to centralized MPC. Furthermore, the distributed MPC algorithm is feasible and closed-loop stable under intermediate termination. Automatic generation control (AGC) provides a practical example for illus- trating the efficacy of the proposed distributed MPC framework. Index Terms—Automatic generation control, distributed model predictive control, power system control. I. INTRODUCTION M ODEL predictive control (MPC) is widely recognized as a high performance, yet practical, control technology. This model-based control strategy uses a prediction of system response to establish an appropriate control response. An at- tractive attribute of MPC technology is its ability to systemati- cally account for process constraints. The effectiveness of MPC is dependent on a model of acceptable accuracy and the avail- ability of sufficiently fast computational resources. These re- quirements limit the application base for MPC. Even so, appli- cations abound in the process industries, and are becoming more widespread [7], [28]. Traditionally, control of large, networked systems is achieved by designing local, subsystem-based controllers that ignore the interactions between the different subsystems. A survey of de- centralized control methods for large-scale systems is available Manuscript received August 30, 2006; revised June 6, 2007. Manuscript received in final form October 10, 2008. First published June 10, 2008; cur- rent version published October 22, 2008. Recommended by Associate Editor D. Rivera. This work was supported in part by the Texas-Wisconsin Mod- eling and Control Consortium and the National Science Foundation (NSF) under Grant CTS-0456694. This paper was presented in part at the IFAC Sym- posium on Power Plants and Power Systems Control, Kananaskis, Canada, June 25-28, 2006. A. N. Venkat was with the Department of Chemical and Biological En- gineering, University of Wisconsin, Madison, WI-53706 USA. He is now with the Process Modeling, Control, and Optimization Group, Shell Global Solutions, Westhollow Technology Center, Houston, TX 77210 USA (e-mail: [email protected]). I. A. Hiskens is with Faculty of Electrical and Computer Engineering, Univer- sity of Wisconsin, Madison, WI 53706 USA (e-mail: [email protected]). J. B. Rawlings is with Faculty of Chemical and Biological Engineering, Uni- versity of Wisconsin, Madison, WI 53706 USA (e-mail: [email protected]. edu). S. J. Wright is with Faculty of Computer Sciences, University of Wisconsin, Madison, WI 53706 USA (e-mail: [email protected]). Digital Object Identifier 10.1109/TCST.2008.919414 in [29]. It is well known that a decentralized control philosophy may result in poor systemwide control performance if the sub- systems interact significantly. Centralized MPC, on the other hand, is impractical for control of large-scale, geographically expansive systems, such as power systems. A distributed MPC framework is appealing in this context; the distributed MPC controllers must, however, account for the interactions between the subsystems. These and other issues critical to the success of distributed MPC are examined in this paper. Each MPC, in addition to determining the optimal current response, also generates a prediction of future subsystem behavior. By suitably leveraging this prediction of future sub- system behavior, the various subsystem-based MPCs can be integrated and the overall system performance improved. A discussion on economic and performance benefits attainable by integrating subsystem-based MPCs is available in [17] and [24]. One of the goals of this paper, however, is to illustrate that a simple exchange of predicted subsystem trajectories (communication) does not necessarily improve overall system control performance. A few distributed MPC formulations are available in the literature. A distributed MPC framework was proposed in [13], for the class of systems that have independent subsystem dynamics but are linked through their cost functions. More recently in [12], an extension of the method described in [13] that handles systems with weakly interacting subsystem dynamics was proposed. Stability is proved through the use of a conservative, consistency constraint that forces the predicted and assumed input trajectories to be close to each other. Also, as pointed out by the author, the performance of the distributed MPC framework in [12] is, in most cases, different from that of centralized MPC. A distributed MPC algorithm for uncon- strained, linear time-invariant (LTI) systems was proposed in [8] and [20]. For the models considered in [8] and [20], the evolution of the states of each subsystem is assumed to be in- fluenced only by the states of interacting subsystems and local subsystem inputs. This choice of modeling framework can be restrictive. In many cases, such as the two area power network with FACTS device (see Section V-G3) and most chemical plants, the evolution of the subsystem states is also influenced by the inputs of interconnected subsystems. More crucially for the distributed MPC framework proposed in [8] and [20], the subsystem-based MPCs have no knowledge of each other’s cost/utility functions. It is known from noncooperative game theory that if such pure communication-based strategies (in which competing agents have no knowledge of each others cost functions) converge, they converge to the Nash equilibrium (NE) [2], [3]. In most cases involving a finite number of agents, 1063-6536/$25.00 © 2008 IEEE Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Transcript of 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

Page 1: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008

Distributed MPC Strategies With Application toPower System Automatic Generation Control

Aswin N. Venkat, Ian A. Hiskens, Fellow, IEEE, James B. Rawlings, and Stephen J. Wright

Abstract—A distributed model predictive control (MPC) frame-work, suitable for controlling large-scale networked systemssuch as power systems, is presented. The overall system is de-composed into subsystems, each with its own MPC controller.These subsystem-based MPCs work iteratively and cooperativelytowards satisfying systemwide control objectives. If availablecomputational time allows convergence, the proposed distributedMPC framework achieves performance equivalent to centralizedMPC. Furthermore, the distributed MPC algorithm is feasibleand closed-loop stable under intermediate termination. Automaticgeneration control (AGC) provides a practical example for illus-trating the efficacy of the proposed distributed MPC framework.

Index Terms—Automatic generation control, distributed modelpredictive control, power system control.

I. INTRODUCTION

M ODEL predictive control (MPC) is widely recognizedas a high performance, yet practical, control technology.

This model-based control strategy uses a prediction of systemresponse to establish an appropriate control response. An at-tractive attribute of MPC technology is its ability to systemati-cally account for process constraints. The effectiveness of MPCis dependent on a model of acceptable accuracy and the avail-ability of sufficiently fast computational resources. These re-quirements limit the application base for MPC. Even so, appli-cations abound in the process industries, and are becoming morewidespread [7], [28].

Traditionally, control of large, networked systems is achievedby designing local, subsystem-based controllers that ignore theinteractions between the different subsystems. A survey of de-centralized control methods for large-scale systems is available

Manuscript received August 30, 2006; revised June 6, 2007. Manuscriptreceived in final form October 10, 2008. First published June 10, 2008; cur-rent version published October 22, 2008. Recommended by Associate EditorD. Rivera. This work was supported in part by the Texas-Wisconsin Mod-eling and Control Consortium and the National Science Foundation (NSF)under Grant CTS-0456694. This paper was presented in part at the IFAC Sym-posium on Power Plants and Power Systems Control, Kananaskis, Canada,June 25-28, 2006.

A. N. Venkat was with the Department of Chemical and Biological En-gineering, University of Wisconsin, Madison, WI-53706 USA. He is nowwith the Process Modeling, Control, and Optimization Group, Shell GlobalSolutions, Westhollow Technology Center, Houston, TX 77210 USA (e-mail:[email protected]).

I. A. Hiskens is with Faculty of Electrical and Computer Engineering, Univer-sity of Wisconsin, Madison, WI 53706 USA (e-mail: [email protected]).

J. B. Rawlings is with Faculty of Chemical and Biological Engineering, Uni-versity of Wisconsin, Madison, WI 53706 USA (e-mail: [email protected]).

S. J. Wright is with Faculty of Computer Sciences, University of Wisconsin,Madison, WI 53706 USA (e-mail: [email protected]).

Digital Object Identifier 10.1109/TCST.2008.919414

in [29]. It is well known that a decentralized control philosophymay result in poor systemwide control performance if the sub-systems interact significantly. Centralized MPC, on the otherhand, is impractical for control of large-scale, geographicallyexpansive systems, such as power systems. A distributed MPCframework is appealing in this context; the distributed MPCcontrollers must, however, account for the interactions betweenthe subsystems. These and other issues critical to the success ofdistributed MPC are examined in this paper.

Each MPC, in addition to determining the optimal currentresponse, also generates a prediction of future subsystembehavior. By suitably leveraging this prediction of future sub-system behavior, the various subsystem-based MPCs can beintegrated and the overall system performance improved. Adiscussion on economic and performance benefits attainableby integrating subsystem-based MPCs is available in [17] and[24]. One of the goals of this paper, however, is to illustratethat a simple exchange of predicted subsystem trajectories(communication) does not necessarily improve overall systemcontrol performance.

A few distributed MPC formulations are available in theliterature. A distributed MPC framework was proposed in[13], for the class of systems that have independent subsystemdynamics but are linked through their cost functions. Morerecently in [12], an extension of the method described in[13] that handles systems with weakly interacting subsystemdynamics was proposed. Stability is proved through the use ofa conservative, consistency constraint that forces the predictedand assumed input trajectories to be close to each other. Also,as pointed out by the author, the performance of the distributedMPC framework in [12] is, in most cases, different from thatof centralized MPC. A distributed MPC algorithm for uncon-strained, linear time-invariant (LTI) systems was proposed in[8] and [20]. For the models considered in [8] and [20], theevolution of the states of each subsystem is assumed to be in-fluenced only by the states of interacting subsystems and localsubsystem inputs. This choice of modeling framework can berestrictive. In many cases, such as the two area power networkwith FACTS device (see Section V-G3) and most chemicalplants, the evolution of the subsystem states is also influencedby the inputs of interconnected subsystems. More crucially forthe distributed MPC framework proposed in [8] and [20], thesubsystem-based MPCs have no knowledge of each other’scost/utility functions. It is known from noncooperative gametheory that if such pure communication-based strategies (inwhich competing agents have no knowledge of each others costfunctions) converge, they converge to the Nash equilibrium(NE) [2], [3]. In most cases involving a finite number of agents,

1063-6536/$25.00 © 2008 IEEE

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 2: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

VENKAT et al.: DISTRIBUTED MPC STRATEGIES WITH APPLICATION TO POWER SYSTEM AUTOMATIC GENERATION CONTROL 1193

the NE is different from the Pareto optimal (PO) solution[10], [11], [26]. In fact, nonconvergence or suboptimality ofpure communication-based strategies may result in unstableclosed-loop behavior in some cases. A four area power networkexample is used here (see Section V-G2) to illustrate instabilitydue to communication-based MPC. Such examples are not un-common. A distributed MPC framework in which the effect ofinterconnected subsystems are treated as bounded uncertaintieswas proposed in [21]. Stability and optimality properties havenot been established however.

Most interconnected power systems rely on automatic gen-eration control (AGC) for regulating system frequency and tie-line interchange [37]. These objectives are achieved by control-ling the real power output of generators throughout the system,taking into account restrictions on the amount and rate of gen-erator power deviations. To cope with the expansive nature ofpower systems, a distributed control structure has been adoptedfor AGC. The current form of AGC may not, however, be wellsuited to future power systems [1], with various trends set to im-pact its effectiveness.

Future power systems will see greater use of flexible ac trans-mission system (FACTS) devices [18]. These devices allow con-trol of power flows over selected paths through a transmissionnetwork, offering economic benefits [23] and improved secu-rity [14]. However, FACTS controllers must be coordinated withother power system controls, including AGC. On the other hand,greater utilization of intermittent renewable resources, such aswind generation, brings with it power flow fluctuations that aredifficult to regulate [27].

These changes provide an opportunity to rethink AGC. Dis-tributed MPC offers an effective means of achieving the desiredcontroller coordination and performance improvements, whilstalleviating the organizational and computational burden asso-ciated with centralized control. AGC therefore provides a veryrelevant example for illustrating the performance of distributedMPC in a power system setting.

This paper is organized as follows. In Section II, a briefdescription of the different modeling frameworks is presented.Notation used in this paper is introduced in Section III. InSection IV, a description of the different MPC-based sys-temwide control frameworks is provided. An implementablealgorithm for terminal penalty distributed MPC is described inSection V. Properties of this distributed MPC algorithm andclosed-loop properties of the resulting distributed controllerare established subsequently. Three examples are presentedto highlight the performance benefits of terminal penalty dis-tributed MPC. A framework for terminal control distributedMPC is introduced in Section VI. In Section VII, the main con-tributions of this study are summarized, and various extensionsare reported.

II. MODELS

Distributed MPC relies on decomposing the overall systemmodel into appropriate subsystem models. A system comprisedof interconnected subsystems will be used to establish theseconcepts.

Centralized Model: The overall system model is representedas a discrete, linear time-invariant (LTI) model of the form

in which denotes discrete time and

......

. . ....

......

. . ....

......

. . ....

......

. . ....

......

. . ....

For each subsystem , the triplet rep-resents the subsystem input, state, and output vector, respec-tively. The centralized model pair is assumed to be sta-bilizable and is detectable.1

Decentralized Model: In the decentralized modeling frame-work, it is assumed that the interaction between the subsystemsis negligible. Subsequently, the effect of the external subsystemson the local subsystem is ignored in this modeling framework.The decentralized model for subsystem is

Partitioned Model (PM): The PM for subsystem combinesthe effect of the local subsystem variables and the effect of thestates and inputs of the interconnected subsystems. The PM forsubsystem is obtained by considering the relevant partition ofthe centralized model and can be explicitly written as

(1a)

(1b)

III. NOTATION

For any matrix , and denote the max-imum and minimum (absolute) eigenvalue of , respectively.For any subsystem , let the predicted state and

1In the applications considered here, local measurements are typically asubset of subsystem states. The structure selected for the � matrix reflects thisobservation. A general � matrix may be used, but impacts possible choices fordistributed estimation techniques [35].

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 3: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

1194 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008

input at time instant , based on data at time be de-noted by and , re-spectively, where is the set of admissible controls for subsys-tems . We have the following definitions for the infinite horizonpredicted state and input trajectory vectors in the different MPCframeworks:

Centralized state trajectory:

Centralized input trajectory:

State trajectory

Input trajectory

Let denote the control horizon. Define. The following notation is used to represent the

finite horizon predicted state and input trajectory vectors in thedifferent MPC frameworks

Centralized state trajectory:

Centralized input trajectory:

State trajectory

Input trajectory

IV. MPC FRAMEWORKS FOR SYSTEMWIDE CONTROL

Let each set of admissible controls be a nonempty, com-pact, convex set with . The set of admissible controlsfor the whole plant is defined to be the Cartesian product ofthe admissible control sets .

The stage cost at stage along the prediction horizon isdefined as

(2)

in which are symmetric weighting matrices andis detectable. The cost function for subsystem

is defined over an infinite horizon and is written as

(3)

with . For any system, the constrained stabi-lizable set (also termed Null controllable domain) is the setof all initial states that can be steered to the origin by

applying a sequence of admissible controls (see [32, Def. 2]). Itis assumed throughout that the initial system state vector

, in which denotes the constrained stabilizable set for theoverall system. A feasible solution to the corresponding opti-mization problem, therefore, exists. For notational simplicity,we drop the time dependence of the state and input trajecto-ries in each MPC framework. For instance, in the centralizedMPC framework, we write and . In thedistributed MPC framework, we use and

.Four MPC-based systemwide control frameworks are de-

scribed in the following. In each MPC framework, the controlleris defined by implementing the first input of the solution to thecorresponding optimization problem.

Centralized MPC: In the centralized MPC framework, theMPC for the overall system solves the following optimizationproblem:

subject to

where .For any system, centralized MPC achieves the best attainable

performance (Pareto optimal) as the effect of interconnectionsamong subsystems are accounted for exactly. Furthermore, anyconflicts among controller objectives are resolved optimally.

Decentralized MPC: In the decentralized MPC framework,each subsystem-based MPC solves the following optimizationproblem:

subject to

Each decentralized MPC solves an optimization problem tominimize its (local) cost function. The effects of the inter-connected subsystems are assumed to be negligible and areignored. In many situations, however, the previous assumptionis not valid and leads to reduced control performance.

Distributed MPC: The partitioned model for each subsystemis assumed to be available. Two formulations

for distributed MPC, namely communication-based MPCand cooperation-based MPC, are considered. Communica-tion-based strategies form the basis for the distributed MPCformulations in [8] and [20]. In the sequel, the suitability ofpure communication-based MPC, as a candidate systemwidecontrol formulation, is assessed. For both communication andcooperation-based MPC, several subsystem optimizations andexchanges of variables between subsystems are performedduring a sample time. An optimization and exchange of vari-ables is termed an iterate. We may choose not to iterate toconvergence. The iteration number is denoted by .

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 4: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

VENKAT et al.: DISTRIBUTED MPC STRATEGIES WITH APPLICATION TO POWER SYSTEM AUTOMATIC GENERATION CONTROL 1195

Communication-Based MPC: For communication-basedMPC,2 the optimal state-input trajectory for subsystem

at iterate is obtained as the solution to theoptimization problem

subject to

Each communication-based MPC utilizes the objectivefunction for that subsystem only. For each subsystem atiteration , only that subsystem input sequence is opti-mized and updated. The other subsystems’ inputs remain at

. If the communication-basediterates converge, then at convergence, a Nash equilibrium(NE) is achieved. In this work, the term communication-basedMPC alludes to the previous framework at convergence of theexchanged trajectories. Examples are presented in Section V-Gfor which communication-based MPC leads to either unaccept-able closed-loop performance or closed-loop instability.

Feasible Cooperation-Based MPC (FC-MPC): To arrive at areliable distributed MPC framework, we need to ensure that thesubsystems’ MPCs cooperate, rather than compete, with eachother in achieving systemwide objectives. The local controllerobjective is replaced by an objective that measures thesystemwide impact of local control actions. The simplest choicefor such an objective is a strict convex combination of the con-troller objectives, i.e.,

.In large-scale implementations, the system sampling interval

may be insufficient to allow convergence of an iterative, cooper-ation-based algorithm. In such cases, the cooperation-based al-gorithm has to be terminated prior to convergence of exchangedtrajectories. The final calculated input trajectories are used todefine a suitable distributed MPC control law. To enable inter-mediate termination, it is necessary that all iterates generatedby the cooperation-based algorithm are strictly systemwide fea-sible (i.e., satisfy all model and inequality constraints) and theresulting nominal distributed control law is closed-loop stable.Such a distributed MPC algorithm is presented in Section V.

For notational convenience, we drop the dependence of. It is shown in [34] that each

can be expressed as

(6)

We consider the more practical case of open-loop stable sys-tems first. A distributed MPC methodology capable of handlinglarge, open-loop unstable systems is described in Section VI.

For open-loop stable systems, the FC-MPC optimizationproblem for subsystem , denoted , is defined as

(7a)

2Similar strategies have been proposed by [8] and [20].

subject to (7b)

(7c)

The infinite horizon input trajectory is obtained by aug-menting with the input sequence .The infinite horizon state trajectory is derived from bypropagating the terminal state using (1) and

. The cost functionis obtained by eliminating the state trajectory from

(3) using (6) and the input, state parameterization describedbefore. The solution to the optimization problem is denotedby . By definition

V. TERMINAL PENALTY FC-MPC

A. Optimization

For the quadratic form of given by (3), the FC-MPCoptimization problem (7), for each subsystem ,can be written as

(8a)

subject to (8b)

in which

and

......

. . .. . .

...(9)

is a suitable terminal penalty matrix. Restricting attention (fornow) to open-loop stable systems simplifies the choice of . Foreach , let

. The terminal penalty can be obtained as the solution tothe centralized Lyapunov equation

(10)

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 5: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

1196 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008

in which . The central-ized Lyapunov equation (10) is solved offline. The solutionto (10) has to be recomputed if the subsystems’ models and/orcost functions are altered.

B. Algorithm and Properties

At time , let represent the maximum number of per-missible iterates for the sampling interval. The following algo-rithm is employed for cooperation-based distributed MPC.

Algorithm 1 (Terminal penalty FC-MPC)

Given

and

while for some and

do

, (see (8))

end (do)

for each

Transmit to each interconnected subsystem.

end (for)

end (while)

The state trajectory for subsystem generated by the input tra-jectories and initial state is represented as

. At each iterate in Algorithm 1, the statetrajectory for subsystem can be calculated as

. At each represents a de-sign limit on the number of iterates; the user may choose to ter-minate Algorithm 1 prior to this limit.

The infinite horizon input and state trajectories canbe obtained following the discussion in Section IV. Denote thecooperation-based cost function after iterates by

The following properties can be established for the FC-MPCformulation (8) employing Algorithm 1.

Lemma 1: Given the distributed MPC formulation definedin (7) and (8), , the sequence of cost functions

generated by Algorithm 1 is non-increasing with iteration number .

A proof is given in Appendix A.

Using Lemma 1 and the fact that is bounded below as-sures convergence of the sequence of cost functions with itera-tion number.

Consider the centralized MPC optimization problem obtainedby eliminating the subsystem states using the PM equations (1),

(11a)

subject to

(11b)

(11c)

From the definition of given by (3), we have .Hence, in (8). It follows thatis strictly convex. Using convexity of

and strict convexity of , the solutionto the centralized MPC optimization problem (11) exists and isunique. By definition, .

Lemma 2: Consider positive definite quadratic andlet , be convex, compact. Assume thesolution to Algorithm 1 after iterates is withan associated cost function value , inwhich . Denote the unique solution to (11)by , in which , and let

represent the optimal cost functionvalue. The solution obtained at convergence of Algorithm 1satisfies

A proof is given in Appendix A.

C. Distributed MPC Control Law

At time , let the FC-MPC algorithm (Algorithm 1) be termi-nated after iterates, with

representing the solution to Algorithm 1 after coopera-tion-based iterates. The distributed MPC control law is obtainedthrough a receding horizon implementation of optimal controlwhereby the input applied to subsystem is

(12)

D. Feasibility of FC-MPC Optimizations

Since , there exists a set of feasible, open-loopinput trajectories , such that ,

and sufficiently large. Convexity of ,and Algorithm 1 guarantee that given a

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 6: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

VENKAT et al.: DISTRIBUTED MPC STRATEGIES WITH APPLICATION TO POWER SYSTEM AUTOMATIC GENERATION CONTROL 1197

feasible input sequence at time , a feasible input se-quence exists for all future times. One trivial choice for afeasible input sequence at is ,

. This choice follows from our assumptionthat each is nonempty and . Existence of afeasible input sequence for each subsystem at ensuresthat the FC-MPC optimization problem (7), (8) has a solutionfor each and all .

E. Initialization

At discrete time , define

(13)

It follows that consti-tute feasible subsystem input trajectories with an associated costfunction .

F. Nominal Closed-Loop Stability

Given the set of initial subsystem states ,. Define to be the value of the coopera-

tion-based cost function with the set of zero input trajectories, , . At time ,

let represent the value of the cooperation-basedcost function with the input trajectory initialization describedin (13). For notational convenience we drop the functiondependence of the generated state trajectories and write

, . The valueof the cooperation-based cost function after iterates isdenoted by . Thus

(14a)

(14b)

At , we have, using Lemma 1, that. It follows from (13) and Lemma 1 that

(15)

Using the previous relationship recursively from time to time0 gives

(16)

From (14), we have . Using(16), gives

. From the previous two cost relationships,

we obtain , whichshows that the closed-loop system is Lyapunov stable [36, p.265]. In fact, using the cost convergence relationship (15) theclosed-loop system is also attractive, which proves asymptoticstability under the distributed MPC control law.

Lemmas 1 and 2 can be used to establish the following(stronger) exponential closed-loop stability result.

Theorem 1: Given Algorithm 1 using the distributed MPCoptimization problem (8) with . In Algorithm 1, let

, . If is stable, is obtained from(10), and

then the origin is an exponentially stable equilibrium for theclosed-loop system

in which

for all and any .A proof is given in Appendix A.Remark 1: If is detectable, then the weaker re-

quirement , is sufficient toensure exponential stability of the closed-loop system under thedistributed MPC control law.

G. Examples

Power System Terminology and Control Area Model: For thepurposes of AGC, power systems are decomposed into controlareas, with tie-lines providing interconnections between areas[37]. Each area typically consists of numerous generators andloads. It is common, though, for all generators in an area to belumped as a single equivalent generator, and likewise for loads.Furthermore, because AGC operation is limited to relativelysmall system disturbances, use of linearized models is standard[37]. Those modeling simplifications are adopted in all subse-quent examples. Some basic power systems terminology is pro-vided in Table I. The notation is used to indicate a deviationfrom steady state. For example, represents a deviation in theangular frequency from its nominal operating value (60 Hz).

Consider any control area , interconnected tocontrol area through a tie line. A simplified model forsuch a control area is given byArea

(17a)

(17b)

(17c)

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 7: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

1198 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008

Fig. 1. Performance of different control frameworks rejecting a load disturbance in area 2. Change in frequency�� , tie-line power flow�� , and load refer-ence setpoints �� ��� .

TABLE IBASIC POWER SYSTEMS TERMINOLOGY

Tie-line power flow between areas and

(17d)

(17e)

Performance comparison. The cumulative stage cost is usedas an index for comparing the performance of different MPCframeworks. Define

(18)

where is the simulation horizon. For each example presentedin this paper, the model and controller parameters are omittedfor brevity; they are available in [33].

1) Two-Area Power System Network: An example with twocontrol areas interconnected through a tie line is considered ini-tially. A control horizon is used for each MPC. Thecontrolled variable (CV) for area 1 is the frequency deviation

and the CV for area 2 is the deviation in the tie-line powerflow between the two control areas . From the control areamodel (17), if and then .

For a 25% load increase in area 2, the load disturbancerejection performance of the FC-MPC formulation is evaluatedand compared against the performance of centralized MPC(cent-MPC), communication-based MPC (comm-MPC), andstandard AGC with anti-reset windup. The load referencesetpoint in each area is constrained between 0.3. In practice,a large load change, such as the one considered above, wouldresult in curtailment of AGC and initiation of emergencycontrol measures such as load shedding. The purpose of thisexaggerated load disturbance is to illustrate the influence ofinput constraints on the different control frameworks.

The relative performance of standard AGC, cent-MPC, andFC-MPC (terminated after one iterate) rejecting the load dis-turbance in area 2 is depicted in Fig. 1. The closed-loop trajec-tory of the FC-MPC controller, obtained by terminating Algo-rithm 1 after one iterate, is almost indistinguishable from theclosed-loop trajectory of cent-MPC. Standard AGC performsnearly as well as cent-MPC and FC-MPC in driving the localfrequency changes to zero. Under standard AGC, however, thesystem takes in excess of 400 s to drive the deviational tie-linepower flow to zero. With the cent-MPC or the FC-MPC frame-work, the tie-line power flow disturbance is rejected in about100 s. A closed-loop performance comparison of the different

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 8: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

VENKAT et al.: DISTRIBUTED MPC STRATEGIES WITH APPLICATION TO POWER SYSTEM AUTOMATIC GENERATION CONTROL 1199

TABLE IIPERFORMANCE OF DIFFERENT CONTROL FORMULATIONS W.R.T. CENT-MPC,

��% � �� � � ���� � � ���

Fig. 2. Four-area power system.

control frameworks is given in Table II. The comm-MPC frame-work stabilizes the system but incurs a control cost that is nearly18% greater than that incurred by FC-MPC (one iterate). If fiveiterates per sampling interval are allowed, the performance ofFC-MPC is almost identical to that of cent-MPC.

Notice from Fig. 1 that the initial response of AGC is to in-crease generation in both areas. This causes a large deviation inthe tie-line power flow. On the other hand, under comm-MPCand FC-MPC, initially reduces area 1 generation and

orders a large increase in area 2 generation (the areawhere the load disturbance occurred). This strategy enables amuch more rapid restoration of tie-line power flow.

2) Four-Area Power System Network: Consider the four-area power system shown in Fig. 2. The model for each controlarea follows from (17). In each control area, a change in localpower demand (load) alters the nominal operating frequency.The MPC in each control area manipulates the load referencesetpoint to drive the frequency deviations and tie-linepower flow deviations to zero. Power flow through the tielines gives rise to interactions among the control areas. Hence, aload change in area 1, for instance, causes a transient frequencychange in all control areas.

The relative performance of cent-MPC, comm-MPC, andFC-MPC is analyzed for a 25% load increase in area 2 and asimultaneous 25% load drop in area 3. This load disturbanceoccurs at 5 s. For each MPC, we choose a control horizon of

. In the comm-MPC and FC-MPC formulations, theload reference setpoint in each area is manipulatedto reject the load disturbance and drive the change in localfrequencies and tie-line power flows to zero. Inthe cent-MPC framework, a single MPC manipulates all four

. The load reference setpoint for each area is constrainedbetween 0.5.

The performances of cent-MPC, comm-MPC, and FC-MPC(one iterate) are shown in Fig. 3. Only and are

TABLE IIIPERFORMANCE OF DIFFERENT MPC FRAMEWORKS RELATIVE TO CENT-MPC,

��% � �� � � ���� � � ���

shown as the frequency and tie-line power flow deviations inthe other areas display similar qualitative behavior. Likewise,only and are shown as other load referencesetpoints behave similarly. The control costs are given inTable III. Under comm-MPC, the load reference setpoints forareas 2 and 3 switch repeatedly between their upper and lowersaturation limits. Consequently, the power system network isunstable under comm-MPC. The closed-loop performance ofthe FC-MPC formulation, terminated after just one iterate,is within 26% of cent-MPC performance. If the FC-MPCalgorithm is terminated after five iterates, the performanceof FC-MPC is within 4% of cent-MPC performance. By al-lowing the cooperation-based iterative process to converge, theclosed-loop performance of FC-MPC can be driven to withinany prespecified tolerance of cent-MPC performance.

3) Two-Area Power System With FACTS Device: In thisexample, we revisit the two area network considered inSection V-G1. In this case though, a FACTS device is employedby area 1 to manipulate the effective impedance of the tieline and control power flow between the two interconnectedcontrol areas. The control area models follow from (17). Inorder to incorporate the FACTS device, though, (17a) in area1 is replaced by

and in area 2 by

where is the impedence deviation induced by the FACTSdevice. The tie-line power flow deviation becomes

Notice that if , the model reverts to (17). The MPCfor area 1 manipulates and to drive and therelative phase difference to zero. The MPCfor area 2 manipulates to drive to zero.

The relative performance of cent-MPC, comm-MPC, andFC-MPC rejecting a simultaneous 25% increase in the loadof areas 1 and 2 is investigated. The closed-loop performanceof the different MPC frameworks is shown in Fig. 4. The

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 9: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

1200 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008

Fig. 3. Performance of different control frameworks rejecting a load disturbance in areas 2 and 3. Change in frequency�� , tie-line power flow�� , and loadreference setpoints �� ��� .

Fig. 4. Performance of different control frameworks rejecting a load disturbance in area 2. Change in relative phase difference �� , frequency �� , tie-lineimpedence �� due to the FACTS device and load reference setpoint �� .

associated control costs are given in Table IV. The perfor-mance of FC-MPC (one iterate) is within 28% of cent-MPCperformance. The performance of comm-MPC, on the otherhand, is highly oscillatory and significantly worse than that ofFC-MPC (one iterate). While comm-MPC is stabilizing, thesystem takes nearly 400 s to reject the load disturbance. WithFC-MPC (one iterate), the load disturbance is rejected in lessthan 80 s. If five iterates per sampling interval are possible, the

FC-MPC framework achieves performance that is within 2.5%of cent-MPC performance.

VI. TERMINAL CONTROL FC-MPC

The terminal penalty-based FC-MPC framework consideredearlier utilizes a suboptimal parameterization of the postulatedinput trajectories. Accordingly, performance is infinite horizonoptimal only in the limit as . Otherwise, convergence

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 10: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

VENKAT et al.: DISTRIBUTED MPC STRATEGIES WITH APPLICATION TO POWER SYSTEM AUTOMATIC GENERATION CONTROL 1201

TABLE IVPERFORMANCE OF DIFFERENT MPC FRAMEWORKS RELATIVE TO CENT-MPC,

��% � �� � � ���� � � ���

achieves performance that is within a prespecified tolerance of amodified infinite horizon optimal control problem (11). The mo-tivation behind terminal control-based FC-MPC is to achieve in-finite horizon optimal (centralized, constrained, LQR [30]) per-formance at convergence using finite values of .

For terminal control FC-MPC, the unconstrained centralizedfeedback law is employed as the terminal feedback law. Theidea is to force the collection of subsystem-based MPCs to drivethe system state to a neighborhood of the origin in which theunconstrained centralized feedback law is feasible. From [15],we know that such a neighborhood of the origin is well definedand can be computed offline. Following the description in [15],we use to denote the maximal output admissible setfor the overall system . Since ,and are convex, we have from [15, Th. 2.1] that is convex.We assume that each is a polytope, i.e.,

. The determination of , in this case, involves thesolution to a set of linear programs. Because is detectableonly (and not observable), is a cylinder with infiniteextent along directions in the unobservable subspace.

Let denote the optimal, centralized linear quadratic reg-ulator (LQR) gain and let denote the solution to the corre-sponding centralized discrete steady-state Riccati equation, i.e.,

(19a)

(19b)

in which and. Conditions for existence of a

solution to (19) are well known [4], [9]. Using a subsystem-wisepartitioning for and gives

.... . .

. . ....

.... . .

. . ....

The terminal control law for subsystem at timeis, therefore,

. To arrive at the terminal control FC-MPC optimizationproblem, we use existing definitions in (8) and redefine

The terminal control FC-MPC optimization problem is thengiven by (8), with these modifications. Algorithm 1 is againutilized for terminal control FC-MPC.

Initialization: To initialize Algorithm 1 for terminal controlFC-MPC, it is necessary to calculate a set of subsystem input tra-jectories that steers the terminal system state (i.e., the predictedstate at the end of the control horizon of each subsystem-basedMPC) inside . For the initial system state

, such a set of subsystem input trajectories can becomputed by solving a simple quadratic program (QP). One for-mulation for this initialization QP is described as follows:

(20a)

subject to (20b)

(20c)

in which

.... . .

. . ....

...

......

...

, with defined in (6), anddefined in (8). The definition of is such that

. The QP (20) is a centralized cal-culation; distributed versions for this initialization QP can bederived using techniques similar to those presented here, butare not pursued in this paper.

Define the steerable set

such that

The set denotes the set of all for which the initializationQP (20) is feasible for a given . We have . Constrainedstabilizability, therefore, follows.

At each iterate of the terminal control FC-MPC algorithm,the validity of the terminal set constraint

must be verified. Two approaches are availablefor ensuring the validity of the postulated terminal controllaw without explicitly enforcing a terminal set constraint. Inthe first approach, the value of is altered online to ensurevalidity of the terminal set constraint. At each iterate, a sub-system-based procedure is used to verify the validity of thepostulated terminal control law. If the selected control horizonis not sufficient to ensure feasibility of the terminal control law,

is increased and the subsystems’ terminal control FC-MPCoptimizations are resolved using the new value of . Strategiesfor increasing online to enable efficient implementation havebeen investigated in [30] for single MPCs.

Rather than increase online, a second approach may beadopted. The idea in this case is to restrict the set of permissibleinitial states to a positively invariant set in which the terminalset constraint is feasible for each subsystem .This positively invariant set depends on the choice of . Fora given , we first construct the steerable set . Next, wedetermine the set of all possible combinations of system states

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 11: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

1202 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008

Fig. 5. Performance of FC-MPC (tc) and CLQR, rejecting a load disturbance in areas 2 and 3. Change in local frequency �� , tie-line power flow �� , andload reference setpoint �� .

and assumed subsystem input trajectories for which the solutionto the terminal control FC-MPC optimization problem for eachsubsystem satisfies the terminal set constraint. Finally, the do-main of the controller, which is the largest positively invariantset for which the terminal control FC-MPC control law is sta-bilizing, is constructed. To construct this invariant set, one mayemploy standard techniques available in the literature for back-ward construction of polytopic invariant sets under constraints[6], [16], [22]. Space restrictions preclude further developmentof either approach in this paper; details of both are available in[33].

For the nominal case, the set of shifted input trajectories (13),obtained using the solution to Algorithm 1 for terminal con-trol FC-MPC at time , is a feasible set of input trajectories attime . For this case, therefore, the initialization QP (20)has to be solved only once at . Lemmas 1 and 2 estab-lished for terminal penalty FC-MPC (see Section V) are alsovalid for terminal control FC-MPC. At convergence of the ex-changed input trajectories, the performance of terminal controlFC-MPC is within a prespecified tolerance of the centralized,constrained LQR [30], [32] performance. If is stabiliz-able, and are detectable, and

, the terminal control FC-MPC control lawis nominally asymptotically stable for all values of the iterationnumber .

Unstable Four-Area Power Network: Consider the four-areapower network described in Section V-G2. In this case though,the value of was increased to force the system to be open-loop unstable. At time 10 s, the load in area 2 increases by15% and simultaneously, the load in area 3 decreases by 15%.The load disturbance rejection performance of terminal con-trol FC-MPC [FC-MPC (tc)] is investigated and compared tothe performance of the benchmark centralized constrained LQR(CLQR) [30].

TABLE VPERFORMANCE OF TERMINAL CONTROL FC-MPC RELATIVE TO CENTRALIZED

CONSTRAINED LQR (CLQR) FOR CONTROL OF UNSTABLE FOUR AREA

NETWORK. ��% � �� � � ���� �� ���

Fig. 5 depicts the stabilizing and disturbance rejection perfor-mance of FC-MPC (tc) and CLQR. Only quantities relating toarea 2 are shown as variables in other areas displayed similarqualitative behavior. The associated control costs are given inTable V. For terminal control FC-MPC terminated after one it-erate, the load disturbance rejection performance is within 13%of CLQR performance. If five iterates per sampling interval arepossible, the incurred performance loss drops to 1.5%.

VII. DISCUSSION AND CONCLUSION

Centralized MPC is not well suited for control of large-scale,geographically expansive systems such as power systems. How-ever, performance benefits obtained with centralized MPC canbe realized through distributed MPC strategies. For distributedMPC, the overall system is decomposed into interconnectedsubsystems. Iterative optimization and exchange of informa-tion among the subsystems is performed. An MPC optimiza-tion problem is solved within each subsystem, using local mea-surements and the latest available external information (from theprevious iterate).

Various forms of distributed MPC have been considered.It is shown that communication-based MPC is an unreli-able strategy for systemwide control and may even resultin closed-loop instability. Feasible cooperation-based MPC(FC-MPC), on the other hand, precludes the possibility of

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 12: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

VENKAT et al.: DISTRIBUTED MPC STRATEGIES WITH APPLICATION TO POWER SYSTEM AUTOMATIC GENERATION CONTROL 1203

parochial controller behavior by forcing the MPCs to cooperatetowards achieving systemwide control objectives. A terminalpenalty version of FC-MPC was initially established. Thesolution obtained at convergence of the FC-MPC algorithmis identical to the centralized MPC solution (and therefore,Pareto optimal). In addition, the FC-MPC algorithm can beterminated prior to convergence without compromising fea-sibility or closed-loop stability of the resulting distributedcontroller. This feature allows the practitioner to terminatethe algorithm at the end of the sampling interval, even ifconvergence is not achieved. The FC-MPC framework allowssmooth transitioning from completely decentralized control tocompletely centralized control. For each subsystem , by setting

in the FC-MPC optimizationproblem, we revert to decentralized MPC. On the other hand,by iterating the FC-MPC algorithm to convergence, centralizedMPC performance is realized. Intermediate termination of theFC-MPC algorithm results in performance between decentral-ized MPC and centralized MPC control limits.

Several extensions for the terminal penalty distributed MPCframework are possible. The proposed distributed MPC frame-work can be extended to penalize and constrain the rate ofchange of inputs. The state for subsystem is augmented withthe input from the previous time step (see [25]). Incorporationof the rate of change of input penalty results in additional termsin the FC-MPC cost function and additional input constraints.All established properties apply however. Details can be foundin [33, Ch. 10]. To ensure closed-loop stability while dealingwith open-loop unstable systems, a terminal state constraintthat forces the unstable modes to be at the origin at the endof the control horizon is necessary. The control horizon mustsatisfy , in which is the number of unstable modes forthe system. The FC-MPC optimization problem of (8) is solvedwith an additional coupled input constraint which forces theunstable modes to the origin at the end of the control horizon.The details for the terminal penalty-based FC-MPC optimiza-tion problem for open-loop unstable systems are available in[33, Ch. 10]. It follows that all iterates generated by Algo-rithm 1 (solving the modified FC-MPC optimization problemwith coupled input constraints) are systemwide feasible, thecooperation-based cost function isa non-increasing function of the iteration number , and thesequence of iterates converges. An important distinction, whicharises due to the presence of the coupled input constraint,is that the limit points of Algorithm 1 need not be optimal.The distributed MPC control law based on any intermediateiterate is feasible and closed-loop stable, but may not achievecentralized MPC performance at convergence of the iterates.

Because terminal penalty FC-MPC is reliant on a suboptimalparametrization of postulated control trajectories, it cannotachieve infinite horizon optimal performance for finite valuesof . In Section VI, a terminal control FC-MPC framework,which achieves infinite horizon optimal performance at conver-gence with finite values of , was described. Unlike terminalpenalty FC-MPC, the proposed terminal control FC-MPC for-mulation also allows the handling of unstable systems withoutthe need for a coupled input constraint. Consequently forunstable systems, optimality at convergence can be guaranteedwith terminal control FC-MPC. For small values of , the

performance of terminal control FC-MPC is observed to besuperior to that of terminal penalty FC-MPC. An alternatestrategy for terminal control FC-MPC is to explicitly enforce aterminal constraint that forces each subsystem-based estimateof the state vector to be in . For small , this strategytypically leads to excessively aggressive controller response,which is undesirable. Enforcing the terminal set constraintexplicitly also introduces a coupled input constraint. For thisformulation, feasibility and stability of the resulting controllaw can be shown. Optimality at convergence, however, is notnecessarily obtained. Further details are available in [33].

Examples were presented to illustrate the applicability andeffectiveness of the proposed distributed MPC framework forAGC. First, a two-area network was considered. Both commu-nication-based MPC and cooperation-based MPC outperformedAGC due to their ability to handle process constraints. Thecontroller defined by terminating Algorithm 1 after five iteratesachieved performance that was almost identical to centralizedMPC. Next, the performance of the different MPC frameworkswas evaluated for a four-area network. For this case, communi-cation-based MPC led to closed-loop instability. FC-MPC (oneiterate) stabilized the system and achieved performance thatwas within 26% of centralized MPC performance. The two-areanetwork considered earlier, with an additional FACTS deviceto control tie-line impedance, was examined subsequently.Communication-based MPC stabilized the system but gaveunacceptable closed-loop performance. The FC-MPC frame-work was shown to allow coordination of FACTS controls withAGC. The controller defined by terminating Algorithm 1 afterjust one iterate gave an improvement in performance of around190% compared to communication-based MPC. For this case,therefore, the cooperative aspect of FC-MPC was very impor-tant for achieving acceptable response. Finally, terminal controlFC-MPC was employed for control of an open-loop unstablefour area network. Terminal control FC-MPC, terminated afterfive iterates gave performance that was within 1.5% of theinfinite horizon optimal control performance. At convergence,the performance of terminal control FC-MPC is always withina prespecified tolerance of the infinite horizon optimal controlperformance.

APPENDIX ATERMINAL PENALTY FC-MPC

Lemma 3 (Minimum Principle for Constrained, Convex Op-timization): Let be a convex set and let be a convex func-tion over . A necessary and sufficient condition for to be aglobal minimum of over is

A proof is given in [5, p. 194].Proof of Lemma 1: From Algorithm 1, we know that

(21)

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 13: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

1204 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008

Therefore, from the definition of (Algorithm 1), we have

By convexity of

(22)

in which equality is obtained if.

Proof of Lemma 2: Since the level set

is closed and bounded (hence compact), a limit point for Al-gorithm 1 exists. We know that is the unique so-lution for the centralized MPC optimization problem (11). Let

. Define .Assume that the sequence , generated by Al-gorithm 1, converges to a feasible subset of the non-optimallevel set

Since is strictly convex and by assumption of non-opti-mality . Let be generated byAlgorithm 1 for large. To establish convergence of Algorithm1 to a point rather than a limit set, we assume the contrary andshow a contradiction. Suppose that Algorithm 1 does not con-verge to a point. Our assumption here implies that there exists

generated by the next iterate of Algorithm1 with . Consider the set of op-timization problems

(23a)

(23b)

(23c)

We have in which. By assumption, there exists at

least one for which . WLOG let . Bydefinition, . Itfollows that . Since

. Using convexity of , wehave

in which the strict inequality follows from for at leastone . Hence, a contradiction. Suppose now that

.From uniqueness of the optimizer,

. Since , generatedusing Algorithm 1, converges to , we have

(24a)

(24b)

(24c)

From Lemma 3

Define and, . We have, from our

assumption , thatfor at least one index .

A second-order Taylor series expansion aroundgives

......

(25)

Using (25) and optimality of gives

(26)

in which is a positive definite function (from(25)). We have from (26) that , which im-plies . It follows, therefore, that

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 14: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

VENKAT et al.: DISTRIBUTED MPC STRATEGIES WITH APPLICATION TO POWER SYSTEM AUTOMATIC GENERATION CONTROL 1205

. Using the previous relationgives . Hence,

.Lemma 4: Let the input constraints in (8) be specified in terms

of a collection of linear inequalities. Consider the closed ball, in which is chosen such that the input constraints

in each FC-MPC optimization problem (8) are inactive for each. The distributed MPC control law defined by the

FC-MPC formulation of Theorem 1 is a Lipschitz continuousfunction of , for all .

A proof is available in [33, Ch. 10].Proof of Theorem 1: Since and is stable,

[31]. The constrained stabilizable set for the system is. To prove exponential stability, we use the value function

as a candidate Lyapunov function. We need to show[36, p. 267] that there exists constants , such that

(27a)

(27b)

in which .Let be chosen such that the input constraints remain in-

active for . Such an exists because the origin is Lya-punov stable and . Since is compact

, there exists such that . Forany satisfying .For , we have from Lemma 4 thatis a Lipschitz continuous function of . There exists, there-fore, a constant , such that

. Define , in whichand independent of . The previous definition gives

and all . For , define.

By definition, . We have

. Similarly,

define ,. By definition . Since is stable, there

exists , such that [19, Corollary 5.6.13, p.199], in which . Hence

since .Let and

. Then

in which.

Also, . Furthermore

(28)

which proves the theorem.

REFERENCES

[1] N. Atic, D. Rerkpreedapong, A. Hasanovic, and A. Feliachi, “NERCcompliant decentralized load frequency control design using modelpredictive control,” in Proc. IEEE PES General Meet., Jun. 2003, pp.554–559.

[2] T. Basar, “Asynchronous algorithms in non-cooperative games,” J.Econom. Dyn. Control, vol. 12, pp. 167–172, 1988.

[3] T. Basar and G. J. Olsder, Dynamic Noncooperative Game Theory.Philadelphia, PA: SIAM, 1999.

[4] D. P. Bertsekas, Dynamic Programming. Englewood Cliffs, NJ:Prentice-Hall, 1987.

[5] D. P. Bertsekas, Nonlinear Programming, 2nd ed. Belmont, MA:Athena Scientific, 1999.

[6] F. Blanchini, “Set invariance in control,” Automatica, vol. 35, pp.1747–1767, 1999.

[7] E. Camacho and C. Bordons, Model Predictive Control, 2nd ed.Berlin, Germany: Springer, 2004.

[8] E. Camponogara, D. Jia, B. H. Krogh, and S. Talukdar, “Distributedmodel predictive control,” IEEE Control Syst. Mag., vol. 9, no. 1, pp.44–52, Jan. 2002.

[9] S. W. Chan, G. C. Goodwin, and K. S. Sin, “Convergence propertiesof the Riccati difference equation in optimal filtering of nonstabilizablesystems,” IEEE Trans. Autom. Control, vol. 29, no. 2, pp. 110–118, Feb.1984.

[10] J. E. Cohen, “Cooperation and self interest: Pareto-inefficiency of Nashequilibria in finite random games,” in Proc. Nat. Acad. Sci., Aug. 1998,vol. 95, pp. 9724–9731.

[11] P. Dubey and J. Rogawski, “Inefficiency of smooth market mecha-nisms,” J. Math. Econom., vol. 19, pp. 285–304, 1990.

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.

Page 15: 1192 IEEE TRANSACTIONS ON CONTROL SYSTEMS …

1206 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 16, NO. 6, NOVEMBER 2008

[12] W. B. Dunbar, “Distributed receding horizon control of dynamicallycoupled nonlinear systems,” IEEE Trans. Autom. Control, vol. 52, no.7, pp. 1249–1263, Jul. 2007.

[13] W. B. Dunbar and R. M. Murray, “Distributed receding horizon controlfor multi-vehicle formation stabilization,” Automatica, vol. 42, no. 4,pp. 549–558, Apr. 2006.

[14] M. Ghandhari, G. Andersson, and I. A. Hiskens, “Control Lyapunovfunctions for controllable series devices,” IEEE Trans. Power Syst., vol.16, no. 4, pp. 689–694, Nov. 2001.

[15] E. G. Gilbert and K. T. Tan, “Linear systems with state and controlconstraints: The theory and application of maximal output admissiblesets,” IEEE Trans. Autom. Control, vol. 36, no. 9, pp. 1008–1020, Sep.1991.

[16] P.-O. Gutman and M. Cwikel, “An algorithm to find maximal stateconstraint sets for discrete-time linear dynamical systems with boundedcontrols and states,” IEEE Trans. Autom. Control, vol. 32, no. 3, pp.251–254, Mar. 1987.

[17] V. Havlena and J. Lu, “A distributed automation framework for plant-wide control, optimisation, scheduling and planning,” presented at the16th IFAC World Congr., Prague, Czech Republic, Jul. 2005.

[18] N. G. Hingorani and L. Gyugyi, Understanding FACTS. Piscataway,NJ: IEEE Press, 2000.

[19] R. A. Horn and C. R. Johnson, Matrix Analysis. Cambridge, U.K.:Cambridge Univ. Press, 1985.

[20] D. Jia and B. H. Krogh, “Distributed model predictive control,” in Proc.Amer. Control Conf., Jun. 2001, pp. 2767–2772.

[21] D. Jia and B. H. Krogh, “Min-max feedback model predictive controlfor distributed control with communication,” in Proc. Amer. ControlConf., May 2002, pp. 4507–4512.

[22] S. S. Keerthi and E. G. Gilbert, “Computation of minimum-timefeedback control laws for discrete time-systems with state-controlconstraints,” IEEE Trans. Autom. Control, vol. 32, no. 5, pp. 432–435,May 1987.

[23] B. H. Krogh and P. V. Kokotovic, “Feedback control of overloadednetworks,” IEEE Trans. Autom. Control, vol. 29, no. 8, pp. 704–711,Aug. 1984.

[24] R. Kulhavý, J. Lu, and T. Samad, “Emerging technologies for enter-prise optimization in the process industries,” in Proc. Chem. ProcessControl—VI: 6th Int. Conf. Chem. Process Control, J. B. Rawlings, B.A. Ogunnaike, and J. W. Eaton, Eds., Jan. 2001, vol. 98, no. 326, pp.352–363, AIChE Symp. Series.

[25] K. R. Muske and J. B. Rawlings, “Model predictive control with linearmodels,” AIChE J., vol. 39, no. 2, pp. 262–287, 1993.

[26] R. Neck and E. Dockner, “Conflict and cooperation in a model of stabi-lization policies: A differential game approach,” J. Econom. Dyn. Con-trol, vol. 11, pp. 153–158, 1987.

[27] R. Piwko, D. Osborn, R. Gramlich, G. Jordan, D. Hawkins, and K.Porter, “Wind energy delivery issues,” IEEE Power Energy Mag., vol.3, no. 6, pp. 47–56, Nov./Dec. 2005.

[28] S. J. Qin and T. A. Badgwell, “A survey of industrial model predictivecontrol technology,” Control Eng. Practice, vol. 11, no. 7, pp. 733–764,2003.

[29] N. R. Sandell, Jr., P. Varaiya, M. Athans, and M. Safonov, “Survey ofdecentralized control methods for larger scale systems,” IEEE Trans.Autom. Control, vol. 23, no. 2, pp. 108–128, Feb. 1978.

[30] P. O. M. Scokaert and J. B. Rawlings, “Constrained linear quadraticregulation,” IEEE Trans. Autom. Control, vol. 43, no. 8, pp. 1163–1169,Aug. 1998.

[31] E. D. Sontag, Mathematical Control Theory, 2nd ed. New York:Springer-Verlag, 1998.

[32] M. Sznaier and M. J. Damborg, “Heuristically enhanced feedback con-trol of constrained discrete-time linear systems,” Automatica, vol. 26,no. 3, pp. 521–532, 1990.

[33] A. N. Venkat, “Distributed model predictive control: Theory and appli-cations” Ph.D. dissertation, Dept. Chem. Biol. Eng., Univ. Wisconsin-Madison, Madison, Oct. 2006 [Online]. Available: http://jbrwww.che.wisc.edu/theses/venkat.pdf

[34] A. N. Venkat, I. A. Hiskens, R. B. Rawlings, and S. J. Wright, “Dis-tributed MPC strategies for automatic generation control,” presentedat the IFAC Symp. Power Plants Power Syst. Control Kananaskis,Canada, Jun. 2006.

[35] A. N. Venkat, I. A. Hiskens, R. B. Rawlings, and S. J. Wright, “Dis-tributed output feedback MPC strategies for power system control,” inProc. IEEE Conf. Dec. Control, Dec. 2006, pp. 4038–4045.

[36] M. Vidyasagar, Nonlinear Systems Analysis, 2nd ed. EnglewoodCliffs, NJ: Prentice-Hall, 1993.

[37] A. J. Wood and B. F. Wollenberg, Power Generation Operation andControl. New York: Wiley, 1996.

Aswin N. Venkat received the B.Tech. degree inchemical engineering from the Indian Institute ofTechnology, Mumbai, India, in 2001, and the Ph.D.degree in chemical engineering from the Universityof Wisconsin, Madison, in 2006.

He is currently with the Process Modeling,Control, and Optimization Group, Shell Global So-lutions (US) Inc., Westhollow Technology Center,Houston, TX. His research interests include theareas of model predictive control, state estima-tion, plantwide control and monitoring, real time

optimization, and resource scheduling.

Ian A. Hiskens (S’77–M’80–SM’96–F’06) re-ceived the B.Eng. degree in electrical engineeringand the B.App.Sc. degree in mathematics fromthe Capricornia Institute of Advanced Education,Rockhampton, Australia, in 1980 and 1983, respec-tively, and the Ph.D. degree from the University ofNewcastle, Newcastle, Australia, in 1991.

He is currently a Professor with the Department ofElectrical and Computer Engineering, University ofWisconsin-Madison, Madison. He has held prior ap-pointments with the Queensland Electricity Supply

Industry, Australia, from 1980 to 1992, the University of Newcastle, from 1992to 1999, and the University of Illinois at Urbana-Champaign, from 1999 to 2002.His major research interests include the area of power system analysis, in par-ticular system dynamic performance, security, and numerical techniques. Otherresearch interests include nonlinear and hybrid dynamical systems.

Prof. Hiskens was an Associate Editor of the IEEE TRANSACTIONS ON

CIRCUITS AND SYSTEMS—I: REGULAR PAPERS from 2002 to 2005, and is cur-rently an Associate Editor of the IEEE TRANSACTIONS ON CONTROL SYSTEMS

TECHNOLOGY. He is also the Treasurer of the IEEE Systems Council.

James B. Rawlings received the B.S. degree from theUniversity of Texas, Austin, in 1979, and the Ph.D.degree from the University of Wisconsin, Madison,in 1985, both in chemical engineering.

He is currently the Paul A. Elfers Professorof Chemical and Biological Engineering, Uni-versity of Wisconsin, and the Codirector of theTexas-Wisconsin Modeling and Control Consortium(TWMCC). He has held prior appointments at theUniversity of Stuttgart, Stuttgart, Germany, as aNATO Postdoctoral Fellow and as a faculty member

at the University of Texas. His research interests include the areas of chemicalprocess modeling, molecular-scale chemical reaction engineering, monitoringand control, nonlinear model predictive control, and moving horizon stateestimation.

Stephen J. Wright received the B.Sc. (Hons.)and Ph.D. degrees from the University of Queens-land, Queensland, Australia, in 1981 and 1984,respectively.

After holding positions at North CarolinaState University, Argonne National Laboratory,and the University of Chicago, he joined theComputer Sciences Department, University of Wis-consin-Madison, Madison, as a Professor in 2001.His research interests include theory, algorithms,and applications of computational optimization.

Prof. Wright is Chair of the Mathematical Programming Society and hasserved on the editorial boards of Mathematical Programming, Series A and B,the SIAM Journal on Optimization, the SIAM Journal on Scientific Computing,and other journals. He also serves on the Board of Trustees of the Society forIndustrial and Applied Mathematics (SIAM).

Authorized licensed use limited to: University of Wisconsin. Downloaded on April 12,2010 at 20:19:35 UTC from IEEE Xplore. Restrictions apply.