Two-Stage Load Control for Severe Under-Frequency...

12
1 AbstractA two-stage load control scheme is presented to address severe under-frequency conditions. The first stage is event-based under-frequency load shedding that guarantees fast response to the detected high consequence disturbances, for initial system protection. The control signals are obtained by solving mixed integer programming problems. After the first stage, model predictive control is triggered in the second stage for step-by-step on-line closed-loop control of interruptible loads, the amount of which is identified via solving linear programming problems. Trajectory sensitivities are used in the optimization problems in both stages to facilitate the analysis. The proposed scheme has been tested on a 9-bus system and on the New England 39-bus system. Index TermsLoad control, under-frequency load shedding, trajectory sensitivities, interruptible loads, model predictive control, mixed integer programing. NOMENCLATURE A. Basic Variables and Indices x Vector of power system state variables y Vector of power system algebraic variables u Vector of power system control variables t Continuous time index k Time series index of discretized trajectory l Load index n Generator index; and 0 is for the system d Frequency limit index [h] Subscription indicating model predictive control step B. Time-Constant Symbols SL Load shedding amount CP Load importance based on cost f n,st Steady-state frequency of generator n f n,base_st Base-case steady-state frequency of generator n f n,d Frequency limit d of generator n This work was in part supported by the Power System Engineering Research Center (PSERC) through the project “ Next generation on-line dynamic security assessment”. L. Tang and J. McCalley are with the Department of Electrical and Computer Engineering, Iowa State University, Ames, IA 50011 USA (emails: [email protected]; [email protected]) . M Large positive value B Large positive value f n,ref Reference of generator n steady-state frequency f n,err Error tolerance of generator n steady-state frequency SL l,max Maximum shedding amount of load l T n,d,max Limits of generator n frequency violating limit d H n Inertia of generator n T C Control horizon of model predictive control T P Predicted time horizon of model predictive control δ Sampling interval of model predictive control C. Time-Variant Symbols x b /y b Base-case trajectory of x/ y x u /y u Trajectory sensitivity of x/y to u f n frequency trajectory of generator n f n,base Base-case trajectory of generator n u n,d u n,d (k ) = 0 if f n (k ) > f n,d ; and u n,d (k ) = 1 if f n (k ) < f n,d T n,d Duration of generator n frequency violating limit d r n,d r n,d (k ) = u n,d (k )∙T n,d (k ) to linearize the optimization t Time-domain simulation integration time step II. INT RODUCT ION Power systems experience under-frequency when load exceeds generation. In severe conditions, load control can contribute to balancing load and generation. Reference [1] summarized the historical severe under-frequency conditions in which load shedding is involved, including the 17,644 MW load shedding in the U.S. Northeast blackout of 2003. This paper designs a two-stage load control scheme for severe under-frequency conditions. The designed scheme consists of event-based under-frequency load shedding (UFLS) in the first stage and model predictive control (MPC) using interruptible loads in the second stage. UFLS is a remedial action for high consequence events that involve significant frequency decline. UFLS is the last automated measure to arrest frequency decline while preventing generator trip and subsequent blackouts [2]. NERC requires that each planning coordinator develop an UFLS program implemented by transmission owners and distribution providers [3]. In a severe under-frequency condition followed by UFLS, the system frequency recovers Two-Stage Load Control for Severe Under-Frequency Conditions Lei Tang, Member, IEEE, and James McCalley, Fellow, IEEE

Transcript of Two-Stage Load Control for Severe Under-Frequency...

Page 1: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

1

Abstract—A two-stage load control scheme is presented to

address severe under-frequency conditions. The first stage is

event-based under-frequency load shedding that guarantees

fast response to the detected high consequence disturbances,

for initial system protection . The control signals are obtained

by solving mixed integer programming problems. After the

first stage, model predictive control is triggered in the second

stage for step-by-step on-line closed-loop control of

interruptible loads, the amount of which is identified via

solving linear programming problems. Trajectory

sensitivities are used in the optimization problems in both

stages to facilitate the analysis. The proposed scheme has

been tested on a 9-bus system and on the New England 39-bus

system.

Index Terms—Load control , under-frequency load shedding,

trajectory sensitivities, interruptible loads, model predictive

control , mixed integer programing.

NOMENCLATURE

A. Basic Variables and Indices

x Vector of power system state variables

y Vector of power system algebraic variables

u Vector of power system control variables

t Continuous time index

k Time series index of discretized trajectory

l Load index

n Generator index; and 0 is for the system

d Frequency limit index

[h] Subscription indicating model predictive control

step

B. Time-Constant Symbols

SL Load shedding amount

CP Load importance based on cost

fn,st Steady-state frequency of generator n

fn,base_st Base-case steady-state frequency of generator n

fn,d Frequency limit d of generator n

This work was in part supported by the Power System Engineering Research

Center (PSERC) through the project “ Next generation on-line dynamic security

assessment”.

L. Tang and J. McCalley are with the Department of Electrical and Computer

Engineering, Iowa State University, Ames, IA 50011 USA (emails:

[email protected]; [email protected]) .

M Large positive value

B Large positive value

fn,ref Reference of generator n steady-state frequency

∆fn,err Error tolerance of generator n steady-state

frequency

SLl,max Maximum shedding amount of load l

Tn,d,max Limits of generator n frequency violating limit d

Hn Inertia of generator n

TC Control horizon of model predictive control

TP Predicted time horizon of model predictive control

δ Sampling interval of model predictive control

C. Time-Variant Symbols

xb/y

b Base-case trajectory of x/y

xu/yu Trajectory sensitivity of x/y to u

fn frequency trajectory of generator n

fn,base Base-case trajectory of generator n

un,d un,d(k) = 0 if fn(k) > fn,d; and un,d(k) = 1 if fn(k) < fn,d

Tn,d Duration of generator n frequency violating limit d

rn,d rn,d(k) = un,d(k)∙Tn,d(k) to linearize the optimization

∆t Time-domain simulation integration time step

II. INTRODUCTION

Power systems experience under-frequency when load

exceeds generation. In severe conditions, load control can

contribute to balancing load and generation. Reference [1]

summarized the historical severe under-frequency

conditions in which load shedding is involved, including the

17,644 MW load shedding in the U.S. Northeast blackout of

2003. This paper designs a two-stage load control scheme

for severe under-frequency conditions. The designed

scheme consists of event-based under-frequency load

shedding (UFLS) in the first stage and model predictive

control (MPC) using interruptible loads in the second stage.

UFLS is a remedial action for high consequence events

that involve significant frequency decline. UFLS is the last

automated measure to arrest frequency decline while

preventing generator trip and subsequent blackouts [2].

NERC requires that each planning coordinator develop an

UFLS program implemented by transmission owners and

distribution providers [3]. In a severe under-frequency

condition followed by UFLS, the system frequency recovers

Two-Stage Load Control for Severe

Under-Frequency Conditions

Lei Tang, Member, IEEE, and James McCalley, Fellow, IEEE

Page 2: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

2

to a level that largely depends on the total load shedding

amount rather than the load shedding locations. However,

the steady-state system frequency recovery is not the only

criteria for UFLS design. The system dynamics after UFLS

action are different with loads interrupted in different

locations. In addition, the cost of interrupting loads of the

same amount but at different locations may not be the same.

A well-designed UFLS program should minimize the

economic loss while satisfying both the steady-state and

dynamic criteria. Existing UFLS programs are considered to

be “response-based” in that they respond to a measured

system condition, in this case, frequency. For example, if the

frequency at a location drops below 59.3 Hz for 0.3 second,

there will be 10% load shedding with 0.1 second time delay

[4]. UFLS programs monitor local system frequency

response after the disturbances and incorporate controls to

react to actual system conditions [5], [6]. The relay settings

have several scales of frequency thresholds and intentional

time delay, to shed certain amounts of load according to

local measurements of frequency levels (f) and possibly

frequency change rates (df/dt). The relay settings are

designed based on operational experience and are fixed for

all scenarios. Response-based strategies such as UFLS are

used when gradual increase of remedial action is acceptable,

but they may not be fast enough to prevent instability

following severe disturbances causing large and rapid

frequency excursions [7].

Modern power grids have witnessed two recent changes

that motivate interest in new forms of frequency control. One

change is the increased penetration of renewable energy,

most of which is not contributing to automatic generation

control (AGC). Reference [8] discussed the future scenarios

in California with increased penetration of renewable energy

and the corresponding decrease in conventional generation.

Less conventional generation leads to diminished capability

of providing timely generation control. In case of severe

under-frequency conditions and when conventional

generation is close to their capacity limits (e.g., during low

wind, high load evening conditions when solar is almost

zero), AGC may not have sufficient control capacity for

frequency recovery, leaving the system vulnerable to

subsequent disturbances.

The other change is that demand-side controllable

resources have become more prevalent. Loads are already

participating in the regulation market for real-time frequency

regulation [9]. In [10], it was demonstrated how to include

loads, especially air-conditioning loads, as spinning reserve

for load curtailment during large contingencies. Similar

demand-side resources suitable for frequency regulation are

refrigerators, water heaters and various other household

appliances that are generally available when needed. Such

loads exist in the form of small and independent units that

can be controlled in a reliable, robust and relatively

continuous way. A controller was developed in [11] to

attach to appliances for response to under-frequency in

heavily loaded conditions. Reference [12] also suggested a

hierarchical structure in which substation level controllers

coordinate the interruptible loads under them, making

possible continuous load control. In [13], a load-side primary

frequency control was designed to minimize the aggregate

cost of tracking an operating point.

In this paper, interruptible loads are controlled to

supplement AGC. This is attractive for three reasons. First,

in severe under-frequency conditions when the system

needs quick frequency recovery to withstand the follow-up

disturbances, load control can respond faster than

generation-based AGC, which requires mechanical

movement of a valve or a water gate and is designed with

constrained speed to avoid following fast varying

components of load [14]. Once a load control decision is

made and transmitted to the local controller, the local

controller can drop the loads instantaneously by opening

the breaker that connects the loads. Second, load control is

available at many locations throughout the grid, in contrast

to generation, which is usually concentrated locationally.

Third, some types of loads, particularly heating and cooling

loads, have service quality that is almost insensitive to their

use for control, in contrast to renewables which are often

unwilling to control away from points of maximum energy

extraction. For example, based on contracted agreement with

customers, Southern California Edison curtailed

air-conditioning loads 37 times with duration of 5-20

minutes, and received no complaint [10].

The proposed load control scheme coordinates wide-area

loads for optimal control. It has two stages. The event-based

UFLS (EB-UFLS) in the first stage is driven by high

consequence events, such as large power plant tripping or

controlled islanding. It is event-based in that it reacts

immediately to pre-identified events, rather than waiting and

counting the duration of the frequency below a threshold to

shed the load. The UFLS actions are determined accurately

through solving optimization that coordinates wide-area

load resources. Event-based strategies have been widely

used for remedial action design to obtain very fast action in

order to maintain system integrity with minimum time delay

[15-20]. Anticipated high consequence events that can

cause severe under-frequency are analyzed before they

occur. The corresponding remedial actions, i.e., feeder

disconnection decisions, are obtained by solving mixed

integer programming (MIP) problems we call load shedding

Page 3: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

3

optimization (LSO). System dynamics in LSO are

approximated with linear constraints based on trajectory

sensitivity analysis. The designed remedial actions form a

lookup table for online decision-making. In real time, the

EB-UFLS program directly takes the remedial action

according to the lookup table once an analyzed event is

detected, to maintain system stability and protect

turbine-generators.

The second stage provides an alternative to AGC

considering future severe under-frequency scenarios when

high penetration of renewables reduces the AGC capability

to provide timely generation control. In the second stage,

MPC uses optimization to design controls at every control

step. MPC is closed-loop. It repetitively optimizes the

control outputs to regulate system frequency with real-time

measurements and an internal system dynamic model. In our

design, MPC determines control signals at each control step

by solving a linear programming (LP) problem we call model

predictive optimization (MPO) that coordinates wide-area

interruptible load resources . In this way, MPC recovers the

system frequency to the rated frequency (60 Hz),

coordinating with existing AGC in its control process. MPC

has been widely explored in process control with slow

system dynamics. MPO uses trajectory sensitivity to

approximate system frequency into LP formulation

efficiently solved in this paper. MPC has also been

implemented in power system real-time control and proven

to be an attractive control strategy, not only for load control

[12], but also for AGC [21] and reactive power control [22],

[23].

Compared with other optimization-based load controls for

frequency recovery, such as [6] and [13], which formulate

system dynamics integration into optimization constraints

that consider only swing equations , LSO and MPO interpret

system dynamics into linear constraints with trajectory

sensitivities that are calculated in off-line analysis. Thus, the

control actions can evaluate system dynamics with detailed

modeling without losing online computational efficiency.

The paper compares the two methods of system dynamics

modeling to show the advantages of us ing trajectory

sensitivities.

References [24], [25] introduce a frequency primary

response market that coordinates various frequency

controlling resources, including their pricing, as ancillary

services for under-frequency conditions. The proposed

two-stage load control scheme in this paper fits into the

frequency primary response market design that incorporates

demand response as a unique frequency controlling

resource in an individual module based on dynamic

simulation.

The paper is organized as follows. Section II introduces

trajectory sensitivity analysis and formulates LSO for

EB-UFLS. Section III introduces MPC and formulates MPO

for MPC-based load control. Section IV provides the overall

design of the proposed load control scheme. Section V

provides the results of case studies on two test systems.

Concluding remarks are provided in section VI.

III. LSO FOR EB-UFLS

This section provides the LSO formulation for EB-UFLS

using trajectory sensitivities to approximate system

dynamics when parameters change.

A. Introduction to Trajectory Sensitivity Analysis

Consider power system dynamics described by

differential algebraic equations (DAE) [26]

( , , )x F x y u (1)

0 ( , , )G x y u (2)

where x are state variables, such as rotor angles and speeds;

y are algebraic variables, such as voltages and power; and u

are control variables, such as loads. The discrete events are

modeled by switching between different sets of DAE. To

obtain the trajectory sensitivities of system variables x and y

with respect to parameter u, we take derivatives of (1) and (2)

with respect to u, i.e.,

u x u y u ux F x F y F (3)

0 x u y u uG x G y G (4)

Equations (3)-(4) are called augmentations of the original

DAE. They generate two new sets of unknown variables:

  u

xx

u

and

u

yy

u

. (5)

The efficient calculation of the original DAE along with

their augmentations is described in [26]. This paper uses a

simple procedure to approximate the trajectory sensitivities

as [27]

0

( ) ( ) ( ) ( )limuu

x u u x u x u u x ux

u u

(6)

which involves simulation of the system model for parameter

u and u+∆u where ∆u is infinitesimal increment of u. yu can be

obtained using the same method. Trajectory sensitivity

linearizes the system along a nominal trajectory, rather than

around an equilibrium point. Reference [28] summarized the

wide applications of trajectory sensitivity analysis in the

power system area.. One useful function of trajectory

sensitivity is to approximate the trajectory deviations after a

slight change of parameters. Given base-case time-domain

simulation trajectories xb(t), y

b(t) and parameter change Δu,

using Taylor series expansion, the system variable

Page 4: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

4

trajectories are approximated by ignoring higher order terms,

according to

x(t) = xb(t) + Dx(t) » xb(t) + xu(t)Du (7)

y(t) = yb(t) + Dy(t) » yb(t) + yu(t)Du (8)

Equations (7)-(8) are widely used in power system dynamic

security assessment to improve computational efficiency

[29-32]. This paper uses (7)-(8) to design the two-stage load

control scheme.

B. LSO Formulation

LSO determines optimal load shedding solutions for each

contingency to satisfy system performance criteria.

Performance requirements for frequency dynamics are

imposed within the LSO constraints. The formulation of LSO

that determines the optimal load shedding (SL) amount is as

follows:

minimize CPl×SL

l( )l=1

L

å (9)

subject to:

Frequency dynamics approximation:

,

1

( ) ( ) ( ) /L

n n base n l l

l

f k f k f k SL SL

(10)

,st , _ ,

1

/L

n n base st n st l l

l

f f f SL SL

(11)

Under-frequency time limits:

, ,( ) ( ) /n d n d nU k f f k M (12)

, , ,( ) ( ) ( ) 1n d n d n dU k u k U k (13)

0 £ rn,d

(k -1) £ B ×un,d

(k -1) (14)

0 £Tn,d

(k -1) - rn,d

(k -1) £ B ×[1-un,d

(k -1)] (15)

, , ,( ) ( 1) ( 1) ( 1)n d n d n dT k r k u k t k (16)

, , ,max( )n d n dT k T (17)

Reference frequency constraints:

, ,err ,st , ,errn ref n n n ref nf f f f f (18)

Load shedding constraints:

0 £ SLl£ SL

l ,max (19)

where k ∈ {1,…,Kn} is the time series index of discretized

trajectory n; n ∈ {0,…,N} is the generator index where 0 is

used to indicate the system frequency; and d ∈ {0,…,D} is

the time limit threshold index.

The decision variable SLl in the objective function (9) is

load shedding amount on load l. The coefficients CPl

indicate the load importance based on cost, which is

determined according to the demand bids in the day-ahead

market [33].

UFLS involves dropping blocks of loads by

disconnecting feeders. This discreteness creates modeling

difficulties since feeder flows cannot be known exactly until

real-time. Modeling load shedding amount as continuous

variables SLl in LSO requires on-line decisions of selecting

feeders to disconnect such that the total load shedding on a

bus is closest to but more than the calculated optimal values

that meet the minimum performance requirements. This

over-shedding is considered by MPC-based load control in

stage 2, using interruptible loads.

In (10)-(11), which approximate frequency dynamics , the

frequency response of generator n is approximated using

trajectory sensitivities according to (7)-(8). Variable fn,base(k)

is the base-case frequency along which the trajectory

sensitivities are calculated. The system frequency f0, which

is used to evaluate the system response, is the

center-of-inertia frequency, expressed in (20)

1

0

1

( )

( )

N

n n

n

N

n

n

H f k

f k

H

(20)

where Hn is the inertia of generator n. The frequencies

experience multiple swings of oscillation with decreased

magnitudes before reaching their steady states. Constraints

(10)-(17) only consider the early swings with the most severe

oscillations. This reduces LSO’s computational burden

through reducing Kn, which is the total number of time steps,

and thus reduces the number of constraints in LSO. Then the

steady-state frequencies (11) are constrained in (18).

The frequency dynamics are constrained in (10)-(17). The

post-contingency frequency response cannot exceed the

threshold fn,d for more than a certain time duration Δtd,max.

Different frequency thresholds have different time limits, as

indicated in Table I. Table I lists two sets of time limits in

different columns. One set is “time limits for generators” that

avoid undesirable generator trip, as considered in [2]. The

other set is “time limits for system” that guarantee

acceptable performance of system frequency f0. The system

performance criteria used by Northeast Power Coordinating

Council (NPCC) are used in our case studies [34].

TABLE I

UNDER-FREQUENCY T IME LIMITS

Index

(d)

Frequency

Limits (Hz)

Time Limits

for Generator

Time Limits

for System

Below 60.5 Safe Safe

1 Below 59.5 3 min 30s

2 Below 58.5 30s

10s

3 Below 58.0 0

4 Below 57.8 7.5s

Page 5: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

5

5 Below 57.5 45 cycles

6 Below 56.8 7.2 cycles

7 Below 56.5 0

To evaluate the duration, a set of binary variables ud,n (k)

indicates if the frequency is below the threshold fn,d or not, as

in (13). The variable fn,d is frequency threshold d of generator

n; and M is a large positive value (we have used M = 100)

such that

, ( )1

n d nf f k

M

. (21)

Thus, un,d (k) = 0 if fn (k) > fn,d and un,d (k) = 1 if fn (k) < fn,d .

Then Tn,d (k), duration below fn,d, is computed as

, , , ,( ) ( 1) ( 1) ( 1) ( 1)n d n d n d n dT k u k T k u k t k (22)

where Δt is integration time step. Once the frequency is

higher than fn,d, Tn,d (k) is reset to zero. The time-domain

simulation uses an implicit trapezoidal integration method

[35] with variable time steps. At each time step of the

integration, the step size is automatically determined

considering both the simulation accuracy and the

computational demand [36].

In order to avoid the nonlinearities in (22), the term un,d

(k-1)∙Tn,d (k-1), is re-defined as rn,d (k-1) to satisfy (14)-(15)

where B is chosen large enough (we have used B = 100) such

that

, ,maxn dB T . (23)

where Tn,d,max ,which can be found in Table I, is the time limit

of threshold d for generator n . This strategy was also

deployed in [6] for setting UFLS relays.

In this way, rn,d (k-1) is equivalent to un,d (k-1)∙Tn,d (k-1)

because

rn,d

(k -1) =Tn,d

(k -1) un,d

(k -1) =1

0 un,d

(k -1) = 0

ìíï

îï

(24)

After the linearization, (22) becomes (16).

The steady-state frequencies should reach the reference

frequency, within a tolerance, as in (18) where fn,ref is the

reference, and fn,err is the tolerance.

IV. MPC-BASED LOAD CONTROL

A. Introduction to MPC

In MPC, a sequence of controls optimizes the predicted

behavior of the monitored system variables [37]. This

closed-loop control process is shown in Fig. 1. The central

feature of MPC is the incorporation of an explicit process

model into the control optimization, which enables the

controllers, in principle, to predict and regulate the system

process dynamics, such as frequencies in our case.

Fig. 1 Closed-loop control process of MPC

MPC is conventionally implemented using linear system

models. In our case, the system model is nonlinear, so we

convert the system frequency behaviors into linear

constraints in MPO using trajectory sensitivities. MPC

based on trajectory sensitivity has been implemented in the

power systems area before [38].

An introduction to MPC for nonlinear systems and the

standard optimization formulation can be found in [39], [40].

In the MPC optimization, TP is the prediction horizon, and TC

is the control horizon. At time tk, MPC solves the optimal

control problem over a finite predicted time horizon [tk,

tk+TP], to design control output u for [tk, tk+TC] (where TC ≤

TP), according to the following:

minimize ( ( ), ( ), ( ))k P

k

t T

tF x y u d

(25)

subject to:

( ( ), ( ), ( )), ( ) ( )k kx f x y u x t x t (26)

0 ( ( ), ( ), ( )), ( ) ( )k kg x y u y t y t (27)

( ) , [ , ]Cu t t T U (28)

( ) ( ), [ , ]k C C Pu u t T t T t T (29)

( ) , ( ) , [ , ]k k Px y t t T X Y (30)

The function F(∙) evaluates the control performance

which in our case is the cost of load control. The tildes (~) in

the formulation indicate real-time measurements as opposed

to internal model variables, where the difference between the

two is a result of the inaccuracies in the model and/or the

model parameters. The optimization is solved for every

sampling interval δ (δ ≤ TC) when real-time measurements are

obtained and used in the formulation. The optimized control

u are applied only for [tk, tk+δ], since updated control will be

obtained from solving the next optimization problem at time

tk+1 = tk+δ, based on real-time measurements at that time, with

tk+TC and tk+TP shifted forward to tk+1+TC and tk+1+TP.

The on-line numerical solution of the nonlinear

optimization of (25)-(30) can be computationally expensive

for large systems, since DAE are involved. Approximation

Page 6: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

6

using trajectory sensitivities reduces the computational

burden by replacing (26)-(27) with (7)-(8). This

approximation has also been deployed in MPC-based power

system load control [41-45] and reactive power

compensation for voltage stability [23], [44].

B. MPO Formulation for MPC-based Load Control

After the frequencies have recovered from the large

excursions in the first stage, MPC in the second stage

neglects binary variables that count durations in LSO and

formulates MPO into LP problems for online computing.

Since the second stage aims to recover the system

frequency level, MPO regulates only steady-state system

frequency based on trajectory sensitivity analysis ,

neglecting the dynamics of generator frequencies.

The MPO formulation for MPC is

minimize CPl

[h] ×SLl

[h]( )l=1

L

å (31)

subject to:

[ ] [ ] [ ] [ ] [ ]

0, 0, _ 0,st

1

/L

h h h h h

ref base st l

l

f f f SL SL

(32)

[ ] [ ] [ ]

,min ,max

h h h

l l lSL SL SL (33)

where superscript [h] is the index of control steps. The

cost-oriented coefficients [ ]h

lCP are determined by real-time

pricing, or incentives paid to participating customers [46].

For example, Midcontinent Independent System Operator

(MISO) introduces in [47] how to control the demand

resources in its demand response program. The

implementation details including pricing can be found in the

MISO demand response Business Practices Manual [48]. In

our case we used identical coefficients. The reference

frequency [ ]

0,

h

reff does not stay the same throughout the

control process. We use a changing reference strategy

whereby the reference frequency [ ]

0,

h

reff starts from the

steady-state frequency before the initialization of MPC and

increases by [ ]

0,

h

incf at each MPC step until the reference

frequency reaches 60 Hz. This changing-reference strategy,

which is also employed in [38] for controlling shunt

capacitors for voltage recovery, is implemented to avoid

large load interruption at each control step. The base-case

system steady-state frequency [ ]

0, _

h

base stf is real-time system

frequency at the time of MPO formulation. Setting [ ]

,min 0h

lSL allows MPO to recover interruptible loads in

over-frequency conditions. The sensitivity of system

steady-state system frequency to load [ ] [ ]

0,st /h hf SL is not

associated with any particular load l, because the load

changes of the same amount but at different locations

equally contribute to the steady-state system frequency

recovery.

MPO (31)-(33) can be solved heuristically using an

iterative method. The total amount of load control is

determined by reformulating (32) into

[ ] [ ]

0, 0, _[ ]

[ ] [ ]1 0,st /

h hLref base sth

l h hl

f fSL

f SL

. (34)

Since the interruptible loads of the same amount but at

different locations equally contribute to the system

steady-state frequency recovery, at each iteration MPO

chooses the most economical action, i.e., fully curtailing the

available load l with the smallest coefficient [ ]h

lCP while

satisfying load l limit (33). The iteration ends with partial

curtailment to satisfy(34).

Fig. 2 illustrates the intended post-contingency

frequency response under MPC. The controls are

determined by solving (31)-(33). After the application of the

designed control at tk, the real-time measurements differ from

the predicted system response due to inaccuracies in the

model and/or the model parameters. In the next control step

at tk+1 = tk+δ, real-time measurements are used to initialize (32),

and this process is repeated until the control objective is

achieved. The parameter δ is chosen such that the frequency

reaches a new steady state before occurrence of the next

load curtailment. Using the changing-reference strategy,

each MPO predicts system frequency for the next sampling

interval δ, and the control is also designed for the next

sampling interval δ, thus, TP = TC = δ holds in our case for

frequency regulation.

Fig. 2 MPC-based load control for frequency regulation.

V. IMPLEMENTATION OF TWO-STAGE LOAD CONTROL

Fig. 3 shows the time frames of all controls for frequency

regulation.

Page 7: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

7

Fig. 3 Frequency control time frames.

The dotted lines indicate the effect of the scheme on

original primary (governor) and secondary (AGC) control. In

real-time operation, the designed EB-UFLS is applied

immediately following the contingency detection. Governor

response begins at the same time. MPC begins when the

system reaches the steady state after EB-UFLS and

governor response. From Fig. 3, EB-UFLS reduces the

amount of governor action required, and MPC reduces the

amount of AGC action required. Since both MPC and AGC

are closed-loop control, they can coordinate with each other

to reach the reference frequency. Several factors determine

the duration of MPC, including frequency reference and

length of sampling interval δ [10], [12].

The flowchart to implement the proposed two-stage load

control scheme is shown in Fig. 4. Off-line analysis solves

LSO for identified contingencies under forecasted operating

conditions if any time limit in Table I is violated. The

designed controls form a lookup table that is stored in the

dynamic database for online decision-making. MPC is

implemented to correct the power imbalance existing after

the governor control. The trajectory sensitivities for

MPO-based MPC are calculated off-lines to improve the

computational efficiency of online MPC. The two stages

coordinate with two different time-scales for frequency

recovery in severe under-frequency conditions. EB-UFLS

guarantees fast response with open-loop control. Due to the

discreteness of feeder disconnection and open-loop nature

in EB-UFLS, EB-UFLS under-sheds and is followed by MPC.

MPC provides refined closed-loop control and coordinates

with existing AGC.

According to

Table I

Determine available

loads, day ahead schedule

and credible contingencies

Limit violated?

Calculate trajectory

sensitivities for EB-UFLS

Time-domain simulation

Form & Solve LSO

Time-domain simulation

with EB-UFLS

Next contingency?

Calculate trajectory

sensitivities for MPO

End

Stage 1: EB-UFLS

Stage 2: MPC

Severe disturbance capture

Off-line analysis Real-time operation

EB-UFLS designed?

Real-time monitoring

EB-UFLS solutions & trajectory

sensitivities for MPO

YesNo

No

YesNo

Yes

Information

storage

Lookup table

+

trajectory

sensitivities

Fig. 4 Flowchart of the two-stage load control scheme implementation.

VI. CASE STUDIES

The goal of the case studies is to illustrate the

performance of the proposed load control scheme in terms of

frequency recovery, governor response and AGC response.

Two systems are used to test the proposed scheme: A 9-bus

system and the New England 39-bus system. A

MATLAB-based program PSAT [49] performs time-domain

simulation for the test systems. The mathematical expression

of the simulation models can be found in [50]. The machines

use a 4th order model; the turbine governors use Type II

model; the prime movers are constants; and the loads use

ZIP (polynomial) models with frequency sensitivities . Since

the time-domain simulation and trajectory sensitivity

calculation are needed only in off-line analysis, the

complexity of dynamic models does not affect the online

computing in real-time operation.

A. 9-Bus System

The one-line diagram of a 9-bus system, which has 6

generators, is shown in Fig. 5. Two contingencies are

analyzed; and each is to trip one generator, as described in

Table II.

Page 8: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

8

5 6

2

3

4

5 6

7 8 9

1

32

1

Cont.1Cont.2

4

Fig. 5 One-line diagram of 9-bus system.

TABLE II

CONTINGENCY LIST FOR 9-BUS SYSTEM

Contingency # Description Gen. % of System Gen.

Contingency 1 Trip generator 5 16 MW 2.78%

Contingency 2 Trip generator 6 22 MW 3.83%

After tripping one generator at 0.2 second, due to the

imbalance between the load and the generation, the system

experiences sudden decline of generator speeds. The

governors react to the contingency based on their droop

characteristic [51], such that after the contingency, the

frequencies recover to a level lower than the rated 60 Hz. Fig.

6 shows the system frequencies following the two

contingencies. For contingency 1, all frequencies meet the

performance requirements in Table I, thus only MPC-based

load control is needed. For contingency 2, the steady-state

frequencies are below 59.5 Hz, which does not meet the

performance requirements in Table I, thus both stages are

needed in load control.

0 5 10 1559

59.5

60

(a)

Time [Second]

Fre

qu

en

cy [

Hz]

0 5 10 1559

59.5

60

(b)

Time [Second]

System freqency

w/o EB-UFLS

System freqency

w/ EB-UFLS

Fig. 6 Frequency responses after tripping generators. Plot (a) corresponds to

generator frequencies after the two contingencies. Plot (b) corresponds to system

frequencies in contingency 2 with and without EB-UFLS.

EB-UFLS is designed for contingency 2. MPC is triggered

for both contingencies 1 and 2 after 15 seconds, at the end of

the transient period. The MPC control the interruptible loads

step by step to recovery the system steady-state frequency.

Using the changing reference strategy, the reference

frequency [ ]

,

h

n reff starts from the steady-state frequency

before the initialization of MPC and increases by [ ]

,

h

n incf = 0.1

Hz at each MPC step. When the frequencies exceed 59.9 Hz,

increasing [ ]

,

h

n reff by 0.1 Hz will cause over adjustment,

thus [ ]

,

h

n incf is reduced such that [ ]

,

h

n reff is 60 Hz. MPO recovers

interruptible loads in case of over-frequency due to

over-adjustment or operating condition change. The

sampling interval δ (which is 20 seconds in this case) is

chosen to be large enough for transients to cease after each

control action. MPC remains in effect until the system

frequency is recovered to 60 Hz.

AGC signals are applied every 1s. The load control

scheme is tested under two conditions. The first condition

has sufficient reserve for AGC-based frequency regulation.

The second condition assumes high penetration of

renewable energy, when conventional generation is close to

their capacity due to lack of renewable generation, as

discussed in section I, thus AGC reaches the limit before the

frequency recovery. The AGC ramp rates are determined in

accordance with NERC operating guide [52], which requires

that the frequency is back to the rated value within 10

minutes in contingency conditions. Historical severe events

have shown that frequency recovery could take more than

10 minutes, e.g., the WECC event on June 14, 2004 took 18

minutes to recover [53].

In the simulation, to reflect the discreteness of load blocks

in reality, as discussed in section III-B, the EB-UFLS

over-sheds by a random percentage within [0%, 5%] at each

bus. To account for the time needed to detect, communicate,

process, and decide, the action of EB-UFLS has 0.2-second

delay after the contingency. This delay is captured in the

off-line analysis by the time-domain simulation and

trajectory sensitivity calculation as a switching event in (1)

-(4). The simulation stops when the system steady-state

frequency is within the range of [59.99, 60.01] Hz. There is no

over-frequency condition that needs interruptible load

recovery in this case. For contingency 1, the load

curtailments in MPC, with sufficient AGC and limited AGC,

are shown in Table III. For contingency 2, the EB-UFLS

(including the over-shedding) and load curtailments in MPC,

with sufficient AGC and limited AGC, are shown in Table IV.

The curtailment percentages of the total system load are also

listed in both tables.

Contingency 1

Contingency 2

Contingency 2

Page 9: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

9

TABLE III

LOAD CURTAILMENTS ON 9-BUS SYSTEM FOR CONTINGENCY 1

Steps MPC-1 MPC-2 MPC-3 MPC-4 MPC-5

Time (s) 15 35 55 75 95

Curt. MW* 2.770 2.752 2.741 2.740 2.251

% of Load* 0.506% 0.504% 0.502% 0.502% 0.412%

Curt. MW** 2.769 2.749 2.743 2.752 2.319

% of Load** 0.505% 0.503% 0.502% 0.504% 0.425%

* Sufficient AGC; ** limited AGC

TABLE IV

LOAD CURTAILMENTS ON 9-BUS SYSTEM FOR CONTINGENCY 2

Steps EB-UFLS MPC-1 MPC-2

Time (s) 0.4 15 35

Curt. MW* 2.047 3.795 3.775

% of Load* 0.375% 0.695% 0.691%

Curt. MW** 2.046 3.801 3.783

% of Load** 0.374% 0.696% 0.693%

Steps MPC-3 MPC-4 MPC-5

Time (s) 55 75 95

Curt. MW* 3.725 3.725 1.351

% of Load* 0.682% 0.682% 0.248%

Curt. MW** 3.721 3.721 1.401

% of Load** 0.681% 0.681% 0.256%

* Sufficient AGC; ** limited AGC

The system frequencies for the two contingencies are

shown in Fig. 7 and Fig. 8. For comparison, Fig. 7 and Fig. 8

also show the effect of AGC without MPC. It is observed

that, MPC recovers the system frequency to 60 Hz much

faster than the situation with only AGC. In case of

insufficient reserve, as in Fig. 7(b) and Fig. 8(b), AGC loses

the capability of frequency regulation when it reaches the

limit.

0 20 40 60 80 100 120

59.5

60

(a)

Fre

qu

en

cy [

Hz]

Contingency 1: AGC only

Contingency 1: MPC + AGC

0 20 40 60 80 100 120

59.5

60

(b)

Fre

qu

en

cy [

Hz]

Contingency 1: limited AGC only

Contingency 1: MPC + limited AGC

Fig. 7 System frequencies in contingency 1.

0 20 40 60 80 100 120

59.5

60

(a)

Fre

qu

en

cy [

Hz]

Contingency 2: AGC only

Contingency 2: MPC + AGC

0 20 40 60 80 100 120

59.5

60

(b)

Time [Second]

Fre

qu

en

cy [

Hz]

Contingency 2: limited AGC only

Contingency 2: MPC + limited AGC

Fig. 8 System frequencies in contingency 2.

To show the advantage of using trajectory sensitivities in

the proposed load control scheme, we compare the

dynamics constraints (10)-(11) with the discretized swing

equations that neglect controls such as voltage regulator

and only consider loads of constant power. In the

comparison, the discretized swing equations, as introduced

in [6], replace constraints (10)-(11) in LSO. The re-formulated

optimization (which we call opt-UFLS) with discretized

swing equations determines control action that results in

unacceptable frequency response, as shown in Fig. 9.

Though opt-UFLS considers the time limits in Table I, the

system dynamic behavior expressed with discretized swing

equations is different from the simulation result obtained

from the method based on trajectory sensitivity analysis,

which is more flexible with system modeling complexity.

0 5 10 15 20 25 3059

59.5

60

Time [Second]

Fre

qu

en

cy [

Hz]

System freqency w/o UFLS

System freqency w/ EB-UFLS

System freqency w/ opt-UFLS

Fig. 9 Frequency response comparison after EB-UFLS and opt-UFLS.

B. New England 39-Bus System

The one-line diagram of the New England 39-bus system

is shown in Fig. 10. The simulated contingency is to trip the

Page 10: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

10

generator on bus 31 at 0.2 second. Before the contingency,

the tripped generator was operating at 368.9 MW, 5.10% of

the total generation. The frequencies of all generators

without the proposed load control are shown in Fig. 11(a),

and the corresponding system frequency is shown in Fig.

11(b).

1

8

2

1

39

9

37

25

3

4

5

8

18

26

2717

16

15

14

1213

6

711

10

30

2

1031

32

28

24

19

20

5

4

33

21

29

6

6

35

22

23

7

36

3

38

34

Cont.

Fig. 10 One-line diagram of the New England 39-bus system.

From Fig. 11, without load control the system frequency

cannot reach 59.5 Hz in 30 seconds, as required in Table I. To

mitigate the impact after the contingency, off-line analysis

solves LSO to determine the optimal amount of load to shed.

Fig. 11(b) compares the system frequencies with and without

EB-UFLS.

After the application of the EB-UFLS, the system

frequency comes to a secure level as shown in Fig. 11(b).

MPC is triggered after 30s when the frequencies almost

reach their steady states. MPC curtails the load step by step

to recovery the system steady-state frequency. It is found

that a 20-second sampling interval δ also works well for this

case. As in the 9-bus case, the reference

frequency [ ]

,

h

n reff increases by [ ]

,

h

n incf = 0.1Hz at each MPC

step. When the frequencies reach over 59.9 Hz, [ ]

,

h

n incf is

reduced such as [ ]

,

h

n reff is 60 Hz. The simulation stops when

the system steady-state frequency is within the range of

[59.99, 60.01] Hz. There is no over-frequency condition that

needs interruptible load recovery in this case. The EB-UFLS

(including the over-shedding) and load curtailments in MPC,

with sufficient AGC and limited AGC, are shown in Table V.

The load curtailment percentages of the total system load are

also listed in the table. Fig. 12 shows the system frequency

simulation results.

0 10 20 30 40

58.5

59

59.5

60

Time [Second]

Fre

qu

en

cy [

Hz]

(a)

0 10 20 30 40

58.5

59

59.5

60

Time [Second]

(b)

System frequency

w/o UFLS

System frequency

w/ UFLS

Fig. 11 Frequency response comparison. Plot (a) corresponds to generator

frequencies without load control. Plot (b) corresponds to comparison of system

frequency with and without EB-UFLS.

TABLE V

LOAD CURTAILMENTS ON 39-BUS SYSTEM

Steps EB-UFLS MPC-1 MPC-2

Time (s) 0.4 30 50

Curt. MW* 99.41 48.29 48.21

% of Load* 1.417% 0.688% 0.687%

Curt. MW** 99.36 48.32 48.19

% of Load** 1.416% 0.689% 0.687%

Steps MPC -3 MPC -4 MPC -5

Time (s) 70 90 110

Curt. MW* 48.00 47.68 15.94

% of Load* 0.684% 0.680% 0.227%

Curt. MW** 48.23 48.14 18.56

% of Load** 0.688% 0.686% 0.265%

* Sufficient AGC; ** limited AGC

0 20 40 60 80 100 120 14058.5

59

59.5

60

Fre

qu

en

cy [

Hz]

(a)

AGC only

MPC + AGC

0 20 40 60 80 100 120 14058.5

59

59.5

60

Time [Second]

Fre

qu

en

cy [

Hz]

(b)

Limited AGC only

MPC + limited AGC

Fig. 12 System frequencies.

Page 11: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

11

From the case studies on the two test systems, the load

control scheme meets the control performance as required in

Table I, regardless of the limit on AGC. MPC curtails the load

gradually in closed loop and reacts faster than AGC.

The online computing time for MPO is important. Unlike

LSO that requires integer variables to model the system

dynamic details, MPO is formulated as an LP optimization

regulating system steady-state frequency recovery. The

number of variables in MPO is the number of interruptible

loads; and the constraints include an equation to

approximate the system steady-state frequency and the

limits on load control amount. This LP problem is solved

heuristically using an iterative method. After all the

interruptible loads are sorted only once based on [ ]h

lCP in

(31) , each iteration has the same amount of calculation

regardless of system size and complexity; and the sorting

can be done before MPO with known [ ]h

lCP . The MPO

computing time is proportional to the number of interruptible

loads that are changed in the MPO solution. In the worse

situation, the MPO computing time is proportional to the

number of the total interruptible loads. In MATLAB for

Windows environment with Intel Core 2 Duo CPU T6500 and

4GB RAM, the average computing time of MPO is 0.0008

second for each iteration. It is suggested to have MPO

computing time within 10% of the sampling interval δ, which

is 20 seconds in our case studies for both systems. The MPO

computing time for the two test systems is much less than

the suggested 2-second limit. MPO can meet the 2-second

limit with a maximum of 2500 interruptible loads in the worst

situation.

If the MPO computing time is limited by existing

computing resources for very large systems with more than

2500 interruptible loads, the sampling interval δ can be

enlarged. In addition, a large system can be scaled into

multiple areas to implement decentralized MPC that further

improves the computational efficiency [54]. Reference [55]

implements a decentralized MPC for load frequency control

in an interconnected power system. Each area in the

interconnected power system has its own local model

predictive controller, which is designed independently and

can communicate with other areas . The decentralized MPC

has also been implemented in AGC to improve computation

efficiency for large systems [56].

VII. CONCLUSIONS

This paper has presented a two-stage load control

scheme. The load control solutions are obtained by solving

LSO and MPO, respectively, coordinating wide-area load

resources. Interruptible loads participate in the second stage

of the load control scheme. Trajectory sensitivities

interpolate the frequency trajectory variations with the

change of loads for better computational efficiency. The

EB-UFLS guarantees fast response to severe

under-frequency conditions to maintain system stability

with minimum time delay. With the weakened AGC-based

frequency regulation due to high penetration of renewables,

MPC using interruptible loads in the second stage can

supplement AGC for quick frequency recovery.

VIII. REFERENCES

[1] NERC. History of UFLS Events from 2001 through 2011. Available:

www.nerc.com/pa/RAPA/ri/Pages/UnderFrequencyLoadShedding.aspx

[2] IEEE PES Power System Relaying Committee, "IEEE guide for abnormal

frequency protection for power generating plants," in IEEE Std

C37.106-2003 (Revision of ANSI/IEEE C37.106-1987), ed, 2004, pp.

0_1-34.

[3] NERC, "Automatic underfrequency load shedding " 2010.

[4] General Electric, "Load shedding, load restoration and generator protection

using solid-state and electromechanical underfrequency relays."

[5] A. Denis Lee Hau, "A general-order system frequency response model

incorporating load shedding: analytic modeling and applications," IEEE

Transactions on Power Systems, vol. 21, pp. 709-717, 2006.

[6] F. Ceja-Gomez, S. Qadri, and F. Galiana, "Under-frequency load shedding

via integer programming," IEEE Transactions on Power Systems, vol. 27,

pp. 1387-1394, 2012.

[7] WECC, "WECC guide: remedial action schedule desing guide," 2006.

[8] G. Hoste, M. Dvorak, and M. Z. Jacobson, "Matching hourly and peak

demand by combining different renewable energy sources," ed: VPUE

Final Report, Stanford University, Palo Alto, California, 2009.

[9] A. Malins, "Demand side developments," National Grid Company, 2003.

[10] J. Eto, J. Nelson-Hoffman, C. Torres, S. Hirth, B. Yinger, J. Kueck, B.

Kirby, C. Bernier, R. Wright, and A. Barat, "Demand response spinning

reserve demonstration," Lawrence Berkeley National Laboratory, 2007.

[11] PNNL. (2006). Grid Friendly Controller Helps Balance Energy Supply

and Demand [Online]. Available: http://readthis.pnl.gov/MarketSource/

ReadThis/B3099_not_print_quality.pdf

[12] I. Hiskens and B. Gong, "MPC-based load shedding for voltage stability

enhancement," in IEEE Conference on Decision and Control, 2005, pp.

4463-4468.

[13] Z. Changhong, U. Topcu, L. Na, and S. Low, "Design and stability of

load-side primary frequency control in power systems," IEEE

Transactions on Automatic Control, vol. 59, pp. 1177-1189, 2014.

[14] N. Jaleeli, L. VanSlyck, D. N. Ewart, L. Fink, and A. Hoffmann,

"Understanding automatic generation control," IEEE Transactions on

Power Systems, vol. 7, pp. 1106-1122, 1992.

[15] WECC. (2002, May) WECC Policy Regarding Extreme Contingencies

and Unplanned Events [Online]. Available: http://www.wecc.biz/library

/WECC%20Documents/Miscellaneous%20Operating%20and%20Planni

ng%20Policies%20and%20Procedures/Extreme%20Contingency.pdf.

[16] A. Heniche, I. Kamwa, and M. Dobrescu, "Hydro-Quebec's defense plan:

present and future," in IEEE Power and Energy Society General Meeting,

2013, pp. 1-5.

[17] "Defense plans against extreme contingencies," CIGRE Task Force,

Technical Brochure C2.02.24, 2007.

[18] N. Varghese, J. Licheng, S. Ghosh, G. Lin, and P. Bunthath, "The

CAISO experience of implementing automated remedial action schemes in

energy management systems," in IEEE Power & Energy Society General

Meeting, 2009, pp. 1-5.

[19] Y. Wang, I. Pordanjani, and W. Xu, "An event-driven demand response

scheme for power system security enhancement," IEEE Transactions on

Smart Grid, vol. 2, pp. 23-29, 2011.

[20] Y. Dai, Y. Xu, Z. Dong, K. Wong, and L. Zhuang, "Real-time prediction

of event-driven load shedding for frequency stability enhancement of power

systems," IET Generation, Transmission & Distribution, vol. 6, pp.

914-921, 2012.

Page 12: Two-Stage Load Control for Severe Under-Frequency …home.eng.iastate.edu/~Jdm/WebJournalPapers/2StageLoadCntrl.pdffrequency control was designed to minimize the aggregate cost of

12

[21] A. Venkat, I. Hiskens, J. Rawlings, and S. Wright, "Distributed MPC

strategies with application to power system automatic generation control,"

IEEE Transactions on Control Systems Technology, vol. 16, pp.

1192-1206, 2008.

[22] J. Wen, Q. Wu, D. Turner, S. Cheng, and J. Fitch, "Optimal coordinated

voltage control for power system voltage stability," IEEE Transactions on

Power Systems, vol. 19, pp. 1115-1122, 2004.

[23] J. Licheng, R. Kumar, and N. Elia, "Model predictive control-based

real-time power system protection schemes," IEEE Transactions on

Power Systems, vol. 25, pp. 988-998, 2010.

[24] E. Ela, V. Gevorgian, A. Tuohy, B. Kirby, M. Milligan, and M.

O'Malley, "Market designs for the primary frequency response ancillary

service - part I: motivation and design," IEEE Transactions on Power

Systems, vol. 29, pp. 421-431, 2014.

[25] E. Ela, V. Gevorgian, A. Tuohy, B. Kirby, M. Milligan, and M.

O'Malley, "Market designs for the primary frequency response ancillary

service - part II: case studies," IEEE Transactions on Power Systems, vol.

29, pp. 432-440, 2014.

[26] I. Hiskens and M. Pai, "Trajectory sensitivity analysis of hybrid systems,"

IEEE Transactions on Circuits and Systems I: Fundamental Theory and

Applications, vol. 47, pp. 204-220, 2000.

[27] I. Hiskens and M. Akke, "Analysis of the Nordel power grid disturbance of

January 1, 1997 using trajectory sensitivities," IEEE Transactions on

Power Systems, vol. 14, pp. 987-994, 1999.

[28] L. Tang and J. McCalley, "Trajectory sensitivities: applications in power

systems and estimation accuracy refinement," in IEEE Power and Energy

Society General Meeting, 2013, pp. 1-5.

[29] L. Tang, "Dynamic security assessment processing system," Ph.D.,

Department of Electrical and Computer Engineering, Iowa State

University, Ames, IA, 2014.

[30] L. Tang and J. McCalley, "Quantitative transient voltage dip assessment

of contingencies using trajectory sensitivities," International Journal of

Electrical Power & Energy Systems, vol. 61, pp. 298-304, 2014.

[31] L. Tang and J. McCalley, "An efficient transient stability constrained

optimal power flow using trajectory sensitivity," in North American

Power Symposium (NAPS), Champaign, IL, 2012, pp. 1-6.

[32] L. Tang and J. McCalley, "Transient stability constrained optimal power

flow for cascading outages," in IEEE PES General Meeting, Washington

DC, 2014.

[33] X. Ma, H. Song, M. Hong, J. Wan, Y. Chen, and E. Zak, "The

security-constrained commitment and dispatch for Midwest ISO day-ahead

co-optimized energy and ancillary service market," in IEEE PES General

Meeting, 2009.

[34] Northeast Power Coordinating Council, "NPCC regional reliability

reference directory # 12 under-frequency load shedding program

requirements," 2009.

[35] P. J. Davis and P. Rabinowitz, Methods of numerical integration: Courier

Dover Publications, 2007.

[36] T. Hull, W. Enright, B. Fellen, and A. Sedgwick, "Comparing numerical

methods for ordinary differential equations," SIAM Journal on Numerical

Analysis, vol. 9, pp. 603-637, 1972.

[37] E. Camacho and C. Bordons, Model Predictive Control: Springer Berlin,

1999.

[38] M. Zima and G. Andersson, "Model predictive control employing

trajectory sensitivities for power systems applications," in IEEE

Conference on Decision and Control, 2005, pp. 4452-4456.

[39] F. Allgower, R. Findeisen, and Z. K. Nagy, "Nonlinear model predictive

control: from theory to application," Journal - Chinese Institute of

Chemical Engineering, vol. 35, pp. 299-316, 2004.

[40] R. Findeisen and F. Allgöwer, "An introduction to nonlinear model

predictive control," in 21st Benelux Meeting on Systems and Control,

Veldhoven, Netherlands, 2002, pp. 1-23.

[41] M. Zima and G. Andersson, "Stability assessment and emergency control

method using trajectory sensitivities," in IEEE Bologna Power Tech

Conference Proceedings, 2003, pp. 1-7.

[42] G. Bo and I. Hiskens, "Two-stage model predictive control for voltage

collapse prevention," in North American Power Symposium, 2008, pp.

1-7.

[43] I. Hiskens and B. Gong, "Voltage stability enhancement via model

predictive control of load," Intelligent Automation and Soft Computing,

vol. 12, pp. 117-124, 2006.

[44] A. Beccuti, T. Geyer, and M. Morari, "A hybrid system approach to power

systems voltage control," in IEEE Conference on Decision and Control,

2005, pp. 6774-6779.

[45] B. Gong, "Voltage stability enhancement via model predictive control,"

Ph.D., University of Wisconsin-Madison, 2008.

[46] U.S. Department of Energy. (2006. Feb.). Benefits of Demand Response in

Electricity Markets and Recommendations for Achieving Them [Online].

Available: http://energy.gov/sites/prod/files/oeprod/Document

sandMedia/DOE_Benefits_of_Demand_Response_in_Electricity_Markets

_and_Recommendations_for_Achieving_Them_Report_to_Congress.pdf

[47] MISO. Level 100 - Demand Response as Resource [Online]. Available:

https://www.misoenergy.org/Training/MarketParticipantTraining/Pages/

MarketParticipantTraining.aspx

[48] MISO. Business Practices Manual - Demand Response [Online].

Available: https://www.misoenergy.org/Library/BusinessPracticesMa

nuals/Pages/BusinessPracticesManuals.aspx

[49] F. Milano, "An open source power system analysis toolbox," IEEE

Transactions on Power Systems, vol. 20, pp. 1199-1206, 2005.

[50] F. Milano, Power System Modelling and Scripting: Springer, 2010.

[51] A. Wood and B. Wollenberg, Power Geneartion Opeartion & Control:

John Wiley & Sons, 2006.

[52] NERC. (2012, Mar.) NERC Operating Manual [Online]. Available:

http://www.nerc.com/files/opman_3_2012.pdf

[53] Z. Zhong, C. Xu, B. Billian, L. Zhang, S. Tsai, R. Conners, V. Centeno,

A. Phadke, and Y. Liu, "Power system frequency monitoring network

(FNET) implementation," IEEE Transactions on Power Systems, vol. 20,

pp. 1914-1921, 2005.

[54] P. Christofides, R. Scattolini, D. Muñoz de la Peña, and J. Liu,

"Distributed model predictive control: a tutorial review and future research

directions," Computers & Chemical Engineering, vol. 51, pp. 21-41,

2013.

[55] T. Mohamed, H. Bevrani, A. Hassan, and T. Hiyama, "Decentralized

model predictive based load frequency control in an interconnected power

system," Energy Conversion and Management, vol. 52, pp. 1208-1214,

2011.

[56] A. Venkat, I. Hiskens, J. Rawlings, and S. Wright, "Distributed MPC

strategies with application to power system automatic generation control,"

IEEE Transactions on Automatic Control, vol. 16, pp. 1192-1206, 2008.