Power System Observability and Dynamic State Estimation for...

10
Power System Observability and Dynamic State Estimation for Stability Monitoring Using Synchrophasor Measurements Kai Sun Junjian Qi ∗∗ Wei Kang ∗∗∗ Department of EECS, University of Tennessee, Knoxville, TN 37996 USA (e-mail: [email protected]) ∗∗ Energy Systems Division, Argonne National Laboratory, Argonne, IL 60439 USA (e-mail: [email protected]) ∗∗∗ Department of Applied Mathematics, Naval Postgraduate School, Monterey, CA 93943 USA (e-mail: [email protected]) Abstract: Growing penetration of intermittent resources such as renewable generations increases the risk of instability in a power grid. This paper introduces the concept of observability and its computational algorithms for a power grid monitored by the wide-area measurement system (WAMS) based on synchrophasors, e.g. phasor measurement units (PMUs). The goal is to estimate real-time states of generators, especially for potentially unstable trajectories, the information that is critical for the detection of rotor angle instability of the grid. The paper studies the number and siting of synchrophasors in a power grid so that the state of the system can be accurately estimated in the presence of instability. An unscented Kalman filter (UKF) is adopted as a tool to estimate the dynamic states that are not directly measured by synchrophasors. The theory and its computational algorithms are illustrated in detail by using a 9-bus 3-generator power system model and then tested on a 140-bus 48-generator Northeast Power Coordinating Council power grid model. Case studies on those two systems demonstrate the performance of the proposed approach using a limited number of synchrophasors for dynamic state estimation for stability assessment and its robustness against moderate inaccuracies in model parameters. Keywords: Observability; Dynamic State Estimation; Synchrophasor; PMUs; Rotor Angle Stability; Unscented Kalman Filter. 1. INTRODUCTION The penetration of a large number of intermittent power generations using renewable energy brings increasing un- certainties to daily operations of interconnected power grids. Early indicating wide-area stability problems is cru- cially important for control centers to prevent cascading power outages and blackouts (Dobson [2001], Qi [2013], Qi, Sun, Mei [2015]). In the US and many other countries, synchrophasors, e.g. Phasor Measurement Units (PMUs), are being installed and networked onto transmission sys- tems to provide high-resolution wide-area measurements synchronized using real-time GPS signals. Electricity util- ities are building PMU networks to collect data from dispersed PMUs and then send the data to online stability applications at control centers. As of today there are still large technical gaps in applying PMU data effectively for real-time stability analysis and for the control of power grids. Most existing PMU applications are based on direct visualization of PMU data for operators to monitor the grid, which, however, cannot provide in-depth informa- tion about the dynamic behavior of the whole grid under disturbances, e.g. contingencies, load changes and unpre- dictable variations with renewable generations. In fact, the data in time series collected from networked PMUs makes it possible to reliably predict the dynamic behavior of the entire grid. Thus, the observability based on data from networked PMUs and methods of real-time This work was supported in part by the University of Tennessee, Knoxville, the CURENT Engineering Research Center, and AFOSR. state estimation of the dynamic grid need to be stud- ied (Kang [2013]). The goal of this paper is to develop and validate the fundamental theory and the methodol- ogy for designing efficient PMU networks by maximizing the observability of potential instabilities and unmeasured states. In addition, a unscented Kalman filter (UKF) based data fusion algorithm is also developed. Some fundamental questions on designing a PMU network are such as how to evaluate networked PMUs in terms of the observability to the presence of instability, and how to develop computa- tionally efficient algorithms for estimating state variables not directly measured by PMUs. While the theory and algorithms for the filtering and optimal sensor design have been developed for many years, some recent results in Kang-Xu [2009a,b] can be used to quantitatively measure observability for large-scale nonlin- ear systems. Methodologies of optimal sensor placement based on observability Gramian have been developed for weather prediction (Kang-Xu [2012]) and chemical engi- neering (Singh [2005, 2006]). In power systems, most of research on PMU placement is for static state estimation and, hence, is mainly based on the topological observ- ability criterion and focuses on the binary connectivity graph (Baldwin [1993]). Under this framework, many op- timization approaches have been proposed, e.g. mixed in- teger programming (Xu [2004], Gou [2008]), binary search (Chakrabarti [2008]) and genetic algorithms (Milosevic [2003], Aminifar [2009]). This article has been accepted by the IFAC journal of Control Engineering Practice

Transcript of Power System Observability and Dynamic State Estimation for...

Page 1: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

Power System Observability and DynamicState Estimation for Stability MonitoringUsing Synchrophasor Measurements �

Kai Sun ∗ Junjian Qi ∗∗ Wei Kang ∗∗∗

∗ Department of EECS, University of Tennessee, Knoxville, TN 37996USA (e-mail: [email protected])

∗∗ Energy Systems Division, Argonne National Laboratory, Argonne,IL 60439 USA (e-mail: [email protected])

∗∗∗ Department of Applied Mathematics, Naval Postgraduate School,Monterey, CA 93943 USA (e-mail: [email protected])

Abstract: Growing penetration of intermittent resources such as renewable generationsincreases the risk of instability in a power grid. This paper introduces the concept of observabilityand its computational algorithms for a power grid monitored by the wide-area measurementsystem (WAMS) based on synchrophasors, e.g. phasor measurement units (PMUs). The goalis to estimate real-time states of generators, especially for potentially unstable trajectories,the information that is critical for the detection of rotor angle instability of the grid. Thepaper studies the number and siting of synchrophasors in a power grid so that the state of thesystem can be accurately estimated in the presence of instability. An unscented Kalman filter(UKF) is adopted as a tool to estimate the dynamic states that are not directly measured bysynchrophasors. The theory and its computational algorithms are illustrated in detail by usinga 9-bus 3-generator power system model and then tested on a 140-bus 48-generator NortheastPower Coordinating Council power grid model. Case studies on those two systems demonstratethe performance of the proposed approach using a limited number of synchrophasors for dynamicstate estimation for stability assessment and its robustness against moderate inaccuracies inmodel parameters.

Keywords: Observability; Dynamic State Estimation; Synchrophasor; PMUs; Rotor AngleStability; Unscented Kalman Filter.

1. INTRODUCTION

The penetration of a large number of intermittent powergenerations using renewable energy brings increasing un-certainties to daily operations of interconnected powergrids. Early indicating wide-area stability problems is cru-cially important for control centers to prevent cascadingpower outages and blackouts (Dobson [2001], Qi [2013],Qi, Sun, Mei [2015]). In the US and many other countries,synchrophasors, e.g. Phasor Measurement Units (PMUs),are being installed and networked onto transmission sys-tems to provide high-resolution wide-area measurementssynchronized using real-time GPS signals. Electricity util-ities are building PMU networks to collect data fromdispersed PMUs and then send the data to online stabilityapplications at control centers. As of today there are stilllarge technical gaps in applying PMU data effectively forreal-time stability analysis and for the control of powergrids. Most existing PMU applications are based on directvisualization of PMU data for operators to monitor thegrid, which, however, cannot provide in-depth informa-tion about the dynamic behavior of the whole grid underdisturbances, e.g. contingencies, load changes and unpre-dictable variations with renewable generations.

In fact, the data in time series collected from networkedPMUs makes it possible to reliably predict the dynamicbehavior of the entire grid. Thus, the observability basedon data from networked PMUs and methods of real-time� This work was supported in part by the University of Tennessee,Knoxville, the CURENT Engineering Research Center, and AFOSR.

state estimation of the dynamic grid need to be stud-ied (Kang [2013]). The goal of this paper is to developand validate the fundamental theory and the methodol-ogy for designing efficient PMU networks by maximizingthe observability of potential instabilities and unmeasuredstates. In addition, a unscented Kalman filter (UKF) baseddata fusion algorithm is also developed. Some fundamentalquestions on designing a PMU network are such as how toevaluate networked PMUs in terms of the observability tothe presence of instability, and how to develop computa-tionally efficient algorithms for estimating state variablesnot directly measured by PMUs.

While the theory and algorithms for the filtering andoptimal sensor design have been developed for many years,some recent results in Kang-Xu [2009a,b] can be used toquantitatively measure observability for large-scale nonlin-ear systems. Methodologies of optimal sensor placementbased on observability Gramian have been developed forweather prediction (Kang-Xu [2012]) and chemical engi-neering (Singh [2005, 2006]). In power systems, most ofresearch on PMU placement is for static state estimationand, hence, is mainly based on the topological observ-ability criterion and focuses on the binary connectivitygraph (Baldwin [1993]). Under this framework, many op-timization approaches have been proposed, e.g. mixed in-teger programming (Xu [2004], Gou [2008]), binary search(Chakrabarti [2008]) and genetic algorithms (Milosevic[2003], Aminifar [2009]).

This article has been accepted by the IFAC journal of Control Engineering Practice

Page 2: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

In this paper, we extend existing results on observabil-ity and dynamic state estimation for nonlinear dynamicsystems to explore and analyze the observability of powergrids and to address some of the aforementioned issues onPMU network design. In Section 2, we first introduce theconcept of observability and its computational algorithm.This concept is fundamental to the evaluation of PMUnetworks. It is utilized to determine the number and sitingof PMUs so that the state of the system can be accuratelyestimated in the presence of instability. Then, an UKFis introduced as a tool to estimate the states that arenot directly measured by PMUs. In Sections 3 and 4, weillustrate the methodology developed in detail using a 9-bus 3-generator power system model. In Section 5, we testits performance on a 140-bus 48-generator power system,i.e. a reduced model of the Northeast Power CoordinatingCouncil (NPCC) power grid in North America.

2. OBSERVABILITY AND UKF

This section introduces the concept of unobservabilityindex and an UKF based state estimation method.

2.1 Observability

In Kang-Xu [2009a,b], a quantitative measure of partialobservability is defined for general dynamical systems. Forpower systems, we adopt a simplified version of that defini-tion. Consider any system defined by ordinary differentialequations (1), where x ∈ IRn is the state variable andy(t) ∈ IRm is the output given by sensor measurement.

dx(t)

dt= f(t, x(t)),

y(t) = h(x(t)).

(1)

In this section, we define the observability with initialcondition x(0), which uniquely determines the trajectoryof a system. The definition can be easily modified to definethe observability for x(t0) with any given time t = t0.Suppose x and y lie in normed spaces with norms denotedby ||x|| and ||y||Y .

Definition 1. Given ρ > 0 and a nominal trajectory x(t)of (1). Let

εN = inf ||h(x(t))− h(x(t))||Y (2)

where x(t) satisfies

dx

dt= f(t, x) (3)

Let ||x(0)−x(0)|| = ρ, and define ρ/ε as the unobservabil-ity index (UI) of x(0).

Remark. As a quantitative measure of observability, theratio ρ/ε can be interpreted as follows: if the maximumerror of the measured output, or sensor error, is ε, thenthe worst estimation error of x(0) is ρ. Therefore, a smallvalue of ρ/ε implies strong observability of x(0). Forlinear systems with a L2-norm, it can be proved that thereciprocal of the unobservability index is the square rootof the smallest eigenvalue of the observability gramian.

Definition 1 can be numerically implemented for nonlinearsystems. In this paper, two algorithms are used to computethe UI, namely the empirical gramian method and themethod of pseudospectral dynamical optimization. Empir-ical gramian is a method of first order approximation forthe UI. The idea is to approximate the UI using the small-est eigenvalue of a gramian matrix (Krener [2009], Kang-Xu [2009a,b]). Due to its simplicity, this method is used

for most cases in this paper. However, for the problem ofrobust observability, that method is not accurate enough.Alternatively, we use a more sophisticated algorithm basedon pseudospectral method (Gong [2006]).

2.2 Unscented Kalman Filter (UKF)

For systems with strong or reasonable observability, filterscan be used as virtual sensors to estimate the variablesthat are not directly measured. Kalman filter was devel-oped originally for linear systems. Modified Kalman filtersare widely used for nonlinear systems. Extended Kalmanfilter has been utilized in power system state estimation forstability control purposes (e.g. Baechle [2014]). However,it requires the linearization of system models, which maynot be easily available during real-time operations forlarge-scale systems like power grids. This paper adoptsUKF, which does not require online computation of systemlinearization. UKF is used as virtual sensors to estimateboth state variables and uncertain parameters. It is alsoused as a tool to verify the observability computed usingdifferent methods. Consider a nonlinear system

xn = f(xn−1, wn−1)yn = h(xn−1, vn−1)

(4)

where x, y, w, and v represent the state, measurement, pro-cess noise, and measurement noise, respectively. The UKFis “founded on the intuition that it is easier to approxi-mate a probability distribution than it is to approximatean arbitrary nonlinear function or transformation” Julier-Uhlmann [2004]. The algorithm of an UKF is outlined asfollows. Details can be found in Julier-Uhlmann [2004].

• Based on the previous-step estimation of the state,xn−1, and the covariance matrix, P xx

n−1, calculate aset of sigma points as

σi = xn−1 ±√

NxP xxn−1, i = 1, 2, . . . , Nx. (5)

• Propagate all the sigma points through the nonlineardynamic and the output equations,

zi = f(σi, 0), gi = h(σi, 0), i = 1, 2, . . . , Nx. (6)

• Calculate predicted means of the state and output,

xn =1

2Nx

2Nx∑i=1

zi, yn =1

2Nx

2Nx∑i=1

gi. (7)

• Predictions of the covariance matrices are given by

P xxn =

1

2Nx

2Nx∑i=1

(zi − xn)(zi − xn)

T

P yyn =

1

2Nx

2Nx∑i=1

(gi − yn)(gi − yn)

T

P xyn =

1

2Nx

2Nx∑i=1

(zi − xn)(gi − yn)

T .

(8)

Once the predictions of xn, Pxxn , P yy

n , and P xyn are avail-

able, the update is given by

xn = xn +K(yn − yn) (9)

where K = P xyn [P yy

n ]−1, P xxn = P xx

n −KP xyn KT .

3. OBSERVABILITY ANALYSIS ON THEBOUNDARY OF THE DOMAIN OF ATTRACTION

To provide adequate information for stability monitoring,it is important to verify that the system is reasonablyobservable both before and at the time of loss of instability.

Page 3: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

Given a set of PMUs, the UI can be computed to deter-mine if the system is adequately observable around theboundary of the domain of attraction (DOA) about systemstates. Theoretically speaking, the UI should be evaluatedfor sampling points of enough density on the boundary ofthe DOA since the system may pass any of those pointsto lose stability. However, the DOA usually exists in thehigh dimensional state space. Discretizations of geometricsurfaces in high dimensions suffer the curse of dimension-ality, i.e. the number of sampling points increases at anexponential rate so that it quickly exceeds a computer’scapability. Therefore, reducing the dimension of the sys-tem model may significantly simplify the problem. Forthe 9-bus system, we directly reduce dimensions of thestate equations for visualization protected onto a 3D statespace. Later, in Section 5 for the 140-bus NPCC system,we investigate those sampling points on the boundary thatare related to only credible “N-1” contingencies (i.e. lossof a single component) based on standard power systemprotection settings.

3.1 Identification of the boundary of the DOA

First, consider the following n-bus power system whosegenerators are represented by the 2nd-order classic model,

δi = ωi − ωR,

ωi =ωR

2Hi

[(Tmi − Tei −Di

ωi − ωR

ωR

] (10)

where i is the generator serial number, rotor angle δiand rotor speed ωi are state variables, Tmi is mechanicaltorque, Tei is electric air-gap torque, ωR is the rated valueof angular frequency, Hi is the inertia constant, and Di isthe damping factor, and

Tei = E2i Gii +

n∑j=1,j �=i

EiEjYij cos(θij − δi + δj). (11)

Proposition 1 Sun [2014]. Suppose (δi(t), ωi(t)), i =1, 2, · · · , n, is a solution of (10). Let δ0 and ω0 be constantnumbers.(a) (δi(t), ωi(t)), i = 1, 2, · · · , n, is a solution of (10), where

δi(t) = δi(t) + δ0, ωi(t) = ωi(t). (12)

(b) If Di = 0, then (δi(t), ωi(t)), i = 1, 2, · · · , n, is asolution of (10), where

δi(t) = δi(t) + ω0t, ωi(t) = ωi(t) + ω0. (13)

The 9-bus system model from Anderson-Fouad [1994], asshown in Fig 1, is used to illustrate the evaluation of theproposed UI. Parameters are: ωR = 2π ·60 Hz, D1 = D2 =D3 = 0, H1=47.28 s, H2=12.8 s, H3=6.02 s, E1=1.0566p.u., E2=1.0502 p.u., E3=1.0170 p.u., Power Base = 100MVA, Tm1=0.716 p.u., Tm2=1.63 p.u., Tm3=0.85 p.u..

The reduced Y matrix is[0.8455− j2.9883 0.2871 + j1.5129 0.2096 + j1.22560.2871 + j1.5129 0.4200− j2.7239 0.2133 + j1.08790.2096 + j1.2256 0.2133 + j1.0879 0.2770− j2.3681

]. (14)

The system has an equilibrium

[ δ1 δ2 δ3 ]T= [ 2.2717◦ 19.7315◦ 13.1752◦ ]T

ωi = ωR, i = 1, 2, 3.(15)

Around any equilibrium of nonlinear systems there isa DOA. The trajectory starting from any point in itconverges to the equilibrium. However, there is no generalway of computing the DOA. Although Zubov’s equationdefines its boundary, this equation is extremely difficult tosolve, numerically or analytically, if not impossible. On theother hand, practical criteria based on experimentationscan be used to approximate a DOA. It is desired that

Fig. 1. A 9-bus 3-generator system and parameters

the data collected by PMUs make the system stronglyobservable at the boundary of the DOA, which impliesthat all state information can be reliably estimated foranalysis and control before a trajectory becomes unstable.

In the following, we find a layer outside the boundary of theDOA, starting from which all trajectories lost its stability.Adequate simulations on the 9-bus system indicate thatstability is lost quickly after at least one angle relative toa preselected reference angle becomes > 650◦. Therefore,this practical criterion is used to compute an envelopoutside the DOA: all initial angles so that at least onerelative angle is between 650◦ and 750◦ at t = 5s.

Using Proposition 1, we can reduce the dimension in thedescription of the DOA by 2. For example, the state spaceof a 9-bus model has six state variables, (δi, ωi), i = 1, 2, 3.The boundary of the DOA is a 5D surface, which cannotbe visualized using 3D graphs. However, we can focus ontrajectories (δi(t), ωi(t)) in which δ3(0) and ω3(0) havethese fixed values, i.e. from the equilibrium (15),

δ3(0) = 13.1752◦, ω3(0) = ωR. (16)

More specifically, we find the envelope of the DOA in the4D subspace [ δ1(0) δ2(0) ω1(0) ω2(0) ] with fixed initialvalues for δ3 and ω3. Any initial values in the envelopeof the DOA in the original 6D space can be transferred,using (12)-(13), into a trajectory with initial state satis-fying (16). Therefore, every trajectory in the 6D space isequivalent to a trajectory with an initial state in a reduced4D subspace. For the purpose of finding the DOA, thisimplies that we only compute an envelope in the δ1δ2ω1ω2-subspace.

Following Proposition 1, we consider a “slice” or sectionof the DOA in the subspace in which δ3 and ω3 are fixed.The DOA in the full state space can be achieved bythe transformation (12)-(13). This method of dimensionreduction is especially efficient for 5D or 6D problemsbecause the domain with a reduced dimension can bevisualized by using 3D graphics. Using the 9-bus modelas an example, an envelope is a 5D surface in the 6Dstate space. The envelope with a reduced dimension isa 3D surface in a 4D space, i.e. the δ1δ2ω1ω2-subspace.We cannot visualize its geometry in a 4D space, but canvisualize its sections in a sequence of 3D subspaces. As anexample, we assign a sequence of values to ω2,

ω2 = ωR +ΔωΔω = −1,−0.5, 0, 0.5, 1 Hz.

(17)

For each fixed value of ω2, compute an envelope for theDOA in δ1δ2ω1-space. This section of envelope is a 2Dsurface. The following algorithm can search for sampling

Page 4: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

points in the envelope. The grid points for ω1 is a sequencein the interval [ωR − 2, ωR + 2] (Hz):

ωk1 = ωR − 2 + 0.1k Hz, k = 0, 1, 2, · · · , 21. (18)

For each value of ω1, a 360◦ search around the equilibrium(15) is carried out numerically for a sequence of directions,

vj = [ cos(θj) sin(θj) ] , θj = 3◦ × j, j = 1, 2, · · · , 120.For each ωk

1 in (18), one point in the envelope of the DOAis found in the direction of vj ,

[ 2.2717◦ 19.7315◦ ] + r × vj (19)

in which the radius r is determined as follows.

(1) Find rmin and rmax so that the system is stable andunstable respectively with r = rmin and rmax in initialcondition (19).

(2) Let r = (rmax + rmin)/2. Solve ODE (1) using (19).(3) Check Δ = max{|δ2(T ) − δ1(T )|, |δ3(T ) − δ1(T )|},

where T = 5. If Δ > 750◦, then rmax = r and goto step (2); if Δ < 650◦, then rmin = r and go to step(2); if 650◦ ≤ Δ ≤ 750◦, stop, and the point (19) isin the envelope of the DOA.

For each value of ω2 and ω1, this algorithm is appliedto 120 different directions, vj . In each 3D section of theenvelope with a fixed ω2, a total of 22 × 120 = 2, 640sampling points are found in the envelope. We computedfive such sections using values from (17) so that the overallenvelope of the DOA is approximated by 13, 200 samplingpoints. 2 of the 5 sections of the envelope of the DOA areillustrated in Fig. 2. From the figures, the boundary is notsmooth. This is in fact possible for nonlinear dynamicalsystems in which the boundary consists of trajectoriesas well as equilibrium points. However, details about thesystem behavior around the nonsmooth boundary pointsrequire further investigation that is out the scope of thispaper. Trajectories starting from this surface diverge afterabout t = 5s. A typical trajectory is shown in Fig. 3.

Fig. 2. Envelopes of the DOA in δ1δ2ω1-space with Δω2 =−1 and 0 Hz

0 1 2 3 4 5−100

0

100

200

300

400

500

600

700

time

Rela

tive

Angl

es

Fig. 3. Relative angles of an unstable trajectory. Solid line:δ2 − δ2; dashed line: δ3 − δ1

3.2 Observability and nonlinear estimation

For the purpose of detecting instability, the PMUs shouldbe installed so that trajectories near the DOA should beobservable. In the following, the quantitative observabilityin Definition 1 is applied to all points in the envelope ofthe DOA. Assume that a PMU is installed at Generator 1and δ1 and ω1 can be directly measured. Then the output

function is y = [ δ1 ω1 ]T. We assume that PMU collects

data at its typical data rate 30 Hz. The norms in Definition1 are defined as

||y(t)||2Y =1

30

30∑j=1

[ δ1 ω1 ]W21

[δ1ω1

](20)

where W1 is the weight matrix

W1 =[1/Rδ 00 1/Rω

],

Rδ = 0.01 · π

180Rω = 2π · 5 · 10−3.

(21)

This weight matrix is chosen based on the assumption thatthe error of δ is bounded by 0.01◦ and the error for ω isbounded by 5× 10−3 Hz. The metric for state variables is

||x||2 = xTW 22 x (22)

where x = [ δ1 δ2 δ3 ω1 ω2 ω3 ]T

and the weight matrix,W2, is defined as follows,

W2 =

⎡⎣

50

RδI3 0

050

RωI3

⎤⎦ , I3 is an identity matrix. (23)

This weight matrix implies that 50 × 0.01 = 0.5◦ and50 × 5 × 10−3 = 0.25 Hz are considered good accuracyin estimation. The UI is a number describing the smallestinput-to-output gain from the initial state to the variablesmeasured by sensors. It means that the worst estimationerror of the state variable is UI times the sensor error, intheir corresponding metrics. For the purpose of instabilityanalysis, the goal is to have reasonable estimate that cantell the trends of the state variables. For this purpose, tra-jectories with 0≤UI<1 are strongly observable; 1≤UI≤30are reasonably observable; UI>30 are weakly observable,sometimes unobservable.

The algorithm of computing the first order approximationof UI using empirical gramian is outlined as follows. It isinspired by the work of Krener [2009]. We use weightednorms based on the noise covariance. Let ρ > 0 be a smallnumber that represents the error tolerance in estimation.Let x ∈ IRn represent the state variable consisting of δiand ωi. Let {X1, X2, · · · , Xn} be a basis of IRn. In thecase of the 9-bus model, n = 6. Given any initial conditionx0 on the envelope of DOA, we solve the ODE model ofthe system to compute

Δi(t) =1

2ρ(y(t;x0 + ρXi)− y(t;x0 − ρXi))

Gij =< Δi,Δj >W1, Sij =< Xi, Xj >W2

.(24)

For the 9-bus model,

< Δi,Δj >W1=

1

30

30∑k=1

Δi(tk)TW1Δj(tk)

< Xi, Xj >W2= XTi W2Xj .

(25)

Let σmin be the smallest eigenvalue of G relative to S, i.e.

Gξ = σminSξ. (26)

For some ξ �= 0, the UI is approximated by1√σmin

.

Once again, the empirical gramian algorithm is used in thecomputation. We focus on the observability in the time

Page 5: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

0 5 10 15 200

200

400

600

800

1000

1200

1400

0 5 10 15 20 25 300

200

400

600

800

1000

1200

(a) Δω2 = −1 and −0.5 Hz

0 5 10 15 20 250

200

400

600

800

1000

1200

0 5 10 15 20 25 30 35 400

500

1000

1500

(b)Δω2 = 0 and 0.5 Hz

0 10 20 30 40 50 600

500

1000

1500

2000

2500

(c)Δω2 = 1 Hz

Fig. 4. The histograms of the unobservability index

interval [4, 5] (s), i.e. the last second before the trajectoryloses its stability. Each value of Δω2 = −1,−0.5, 0, 0.5, 1Hz determines a 3D section of the envelope as shown inFig. 2. For each section, a histogram about the UI is drawnas shown in Fig. 4. For all trajectories initiated from thesesections, we compute their UIs. The histograms show thevalue of the UI vs. the number of initial states in theenvelope. Except for a few special cases, all trajectoriesare either strongly or reasonably observable (UI<30). Theworst observability is over 40 when Δω2 = 0.5 and 1 Hz.

UKF is applied to observable trajectories. As expected,for almost all initial states in the envelope the estimatesfor both the angles and the angular velocities track thetrue value closely when the system loses stability. However,for trajectories with high UIs the estimator is not alwaysreliable. In some cases, UKF works fine, but in some othercases it fails to track the true trajectory. For instance, weapplied UKF to the trajectory with the worst observability,UI=63.95. The result is fine as shown in Figs. 5 and 6.Although the errors are not small, the estimates are ableto follow the trend of the true trajectory, an importantfeature used in power system stability analysis.

4 4.2 4.4 4.6 4.8 5100

200

300

400

500

600

700

800

t

δ a

nd δ

est

4 4.2 4.4 4.6 4.8 52.3

2.35

2.4

2.45

2.5

2.55x 10

4

t

ω a

nd ω

est

Fig. 5. UKF estimates (star) with UI= 63.95

If PMU is installed at Generator 2 (the least observabletrajectory has UI=40.22), UKF cannot track the trends ofthe trajectory when it goes unstable (Figs. 7 and 8).

For the same trajectories, the observability is changed ifthe PMU is installed at a different generator. We computed

4 4.2 4.4 4.6 4.8 5−2

−1

0

1

2

t

erro

r of

δes

t

4 4.2 4.4 4.6 4.8 5−20

−10

0

10

20

t

erro

r of

ωes

t

Fig. 6. UKF estimation errors with UI = 63.9457

4 4.2 4.4 4.6 4.8 50

100

200

300

400

500

600

700

800

900

t

δ a

nd δ

est

4 4.2 4.4 4.6 4.8 52.3

2.35

2.4

2.45

2.5

2.55x 10

4

t

ω a

nd ω

est

Fig. 7. UKF estimates (star) with UI= 40.22

4 4.2 4.4 4.6 4.8 5−200

0

200

400

600

t

erro

r of

δes

t

4 4.2 4.4 4.6 4.8 5−1000

0

1000

2000

t

erro

r of

ωes

t

Fig. 8. UKF estimation errors with UI= 40.22

the UI for all three generators. The results are similarto what is shown above, i.e. almost all trajectories arereasonably observable. Among all 13, 200 trajectories fromthe envelope of the DOA with a PMU located in any ofthe three generators, there are only three trajectories withweak observability, UI> 30. The case studies on the 9-bussystem show that, even if the system is mostly observable,it cannot be concluded that the overall system is alwaysobservable. There could be very small neighborhoods inwhich observability drops significantly.

4. OBSERVABILITY AND ESTIMATION IN THEPRESENCE OF SYSTEM UNCERTAINTIES

System instabilities can be triggered by two different rea-sons, an initial state outside the DOA or a change ofsystem parameters. The former is studied in the previoussection. In this section, we first study the robustness ofthe observability in the presence of unknown parameterchanges, and then introduce an adaptive nonlinear es-timation method to provide accurate state estimate. Inaddition, some unknown parameters can also be estimatedusing sensor information.

4.1 Robustness of observability

Let δei and ωei , i = 1, 2, 3, be the equilibrium point (15)

of the system in (1) with a reduced Y matrix (14). At

Page 6: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

time t = 0, we assume that the system’s admittancematrix is changed due to, e.g., a contingency, i.e. Y =Y + ΔY , which is not known to the operator. If theestimation is still based on the original Y matrix, howrobustness is the observability? To quantitatively measurethe robustness of observability, we use the remaindersof trajectories, an approach from Kang [2011]. Supposethe PMU measures δ1(t) and ω1(t). Let δi(t) and ωi(t),i = 1, 2, 3, be a trajectory with the new Y matrix startingfrom the equilibrium (15). Suppose δ∗i (t) and ω∗

i (t) be thebest estimate of δi(t) and ωi(t) using the original Y matrix,

min || [ δ1(t)− δ1(t) ω1(t)− ω1(t)] ||W1

δi(t) and ωi(t) satisfy (10)-(14)(27)

where W1 is defined in (21). Let δri (t) = δi(t) − δ∗i (t) andωri (t) = ωi(t)− ω∗

i (t) be the remainder, then

δi(t) = δ∗i (t) + δri (t), ωi(t) = ω∗i (t) + ωr

i (t). (28)

According to (27), δ∗i (t) and ω∗i (t) are the best estimates

of δi(t) and ωi(t) using the matrix Y and the sensorinformation. The estimation error is the remainder, δri (t)and ωr

i (t). This error is not directly caused by the outputnoise, i.e. this error cannot be reduced no matter howaccurate the output is measured. Therefore, the remainderis a measure of the robustness of observability. If theremainder is small, it implies that a nonlinear estimatoris able to accurately estimate the state variables in thepresence of small unknown parameter change.

To compute the remainder and the best estimate, we mustsolve (27). It is a problem of nonlinear dynamic optimiza-tion. An analytic solution does not exist. Here, compu-tational dynamic optimization is applied. To solve (27),this paper adopts a pseudo-spectral method of computa-tional optimal control based on Legendre-Gauss-Lobattoquadrature nodes, which has been actively studied in theliterature (see, for instance, references Fahroo-Ross [1998],Gong [2006]). It consists of two parts, a discretizationmethod and a nonlinear programming (NLP) solver. Thediscretization is based on a set of nodes in the timeinterval, t0 < t1 < · · · < tN . The nodes used in apseudo-spectral method are determined based upon or-thogonal polynomials, for instance the Legendre-Gauss-Lobatto nodes that consists of {−1, 1} and all zeros of thederivative of a Legendre polynomial. A state trajectoryx(t) is approximated by N vectors {xNj}Nj=0 at the nodes,i.e.

xNj ≈ x(tj), uNj ≈ u(tj), j = 0, 1, · · · , N (29)

The differential equations in the power system model aretransformed into a set of nonlinear algebraic equations of{xNj}Nj=0. As a result, the optimization problem (27) isapproximated by a NLP, which is a finite dimensional opti-mization constrained by algebraic equations. Then an NLPnumerical solver using sequential quadratic programmingis applied to find an estimated solution of (27) .

As an example, we compute the remainder for a ΔY inwhich all entries are zero except ΔY12 = ΔY21 = γY12

and γ = −0.01, i.e. the entries Y12 and Y21 in Y arereduced by 1% in magnitude. For the time interval [0, 1],the remainder is shown in Fig. 9. The magnitude of theremainders depends on the length of the time intervaland are found to increase with the final time T . Usethe final-time error as a metric for the remainder, i.e.|| [ δr1(T ) δr2(T ) δr3(T ) ] || and || [ ωr

1(T ) ωr2(T ) ωr

3(T ) ] ||, theresults for T = 1, 2, 3, 4, 5 (s) is also shown in Fig. 9,where “∗” represents the norm of angles in “deg” and‘o” represents the norm of angular velocities in “deg/s”,which respectively increase linearly and at a higher order.The result implies that the UI is “unstable” in the sense

Table 1. UI with parameter variations

20% change UI 20% change UIY12 0.8572 Y12 and Y13 1.2585Y13 1.7593 Y12 and Y23 0.7026Y23 0.8055 Y13 and Y23 1.3826

that the remainder’s magnitude, i.e. the error of the bestestimate, increases with time. Therefore, the observabilitydecreases as time goes on, and hence is not robust to thevariation of system parameters.

0 0.2 0.4 0.6 0.8 1−4

−2

0

2

t

err

ors

of

ω

0 0.2 0.4 0.6 0.8 1−1.5

−1

−0.5

0

0.5

t

err

ors

of

δ

1 1.5 2 2.5 3 3.5 4 4.5 50

5

10

15

20

25

30

35

40

45

50

Length of time interval

Err

ors

at

fin

al tim

e

Fig. 9. The remainders for the first 1 s (left) and forT = 1, 2, 3, 4, 5 (right)

4.2 The observability of inaccurate parameters and adaptiveestimation

Given the robustness study, it is important to make astate estimator adaptive to an unknown parameter change.In fact, we can compute the observability of both thestate variables and the parameters in the system model.More specifically, suppose the inaccurate parameter is themagnitude of Y12, then Y12 = γY12. If γ = 1, the nominalvalue of Y is the same as the true value. If γ �= 1, theparameter in the system is varied. In this case, we wantto estimate the change and then adaptively adjust theestimates. For this purpose, we define the observabilityof both state variables ωi and δi and r using Definition 1and this norm

||x||2 = xTW2x+W3r2 (30)

where W2 is from (23) and W3 = 1/0.05. The nominalvalue, γ, in the simulations is 0.8, i.e. we assume a 20%parameter variation that is unknown. Observability ofthe parameter variation as well as the state variablesis computed and listed in Table 1. From the table wecan conclude that parameter variations are observable.For instance, if Y12 is varied by 20%, the unobservabilityindex is 0.9853, which implies strong observability. If twoparameters are unexpectedly changed by 20%, the systemis still observable. The result is also shown in Table1. Because the inaccurate parameters are observable, itmakes sense to apply an estimator that is adaptive to thesystem change. In this case the PMU measurement servestwo purposes: for the estimation of the state variables andfor the real-time update of parameter changes. Once again,UKF is applied to estimate the value of the state variableas well as the parameter variation.

We tested UKF by changing one parameter in Y matrixby 20%. For example, the true value of Y12 is 80% ofY12 given to the UKF. Started from the incorrect modelparameters, the UKF uses the PMU measurement withnoise to estimate ω1 and δi as well as Y12. The processis able to correct the parameter automatically so that theestimates gradually approach the true value. The result isshown in Fig. 10. Similar tests are carried out for variations

Page 7: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

of all other entries in Y . All estimates are convergent witha behavior similar to what shown in Fig. 10.

0 1 2 3 4 50.75

0.8

0.85

0.9

0.95

1

1.05

1.1

t

Pa

ram

ete

r va

ria

tio

n

0 1 2 3 4 5−5

0

5

10

t

err

or

of

δest

0 1 2 3 4 5−50

0

50

100

t

err

or

of

ωest

Fig. 10. Adaptive estimation: inaccurate Y12

5. TESTS ON THE 140-BUS NPCC SYSTEM MODEL

Fig. 11. 140-bus 48-generator NPCC system model

Here, we test the proposed UI and UKF-based stateestimation method on the 140-bus 48-generator NPCCsystem model from Chow [2008] as shown in Fig. 11. Thedynamic states of this system is estimated by the square-root unscented Kalman filter (Merwe [2001], Qi [2015]).

5.1 Realistic Generator and Measurement Models

Regarding generator models, 27 of 48 generators are repre-sented in a detailed, widely accepted 4th-order round-rotordifferential equation model in (31) in a d-q reference framerevolving with the rotor and the rest are represented usingthe 2nd-order classic model in (10). Thus the number ofstates is 150 = 4× 27 + 2× 21.

δi = ωi − ωR

ωi =ωR

2Hi

(Tmi − Tei −Di

ωi − ωR

ωR

)

e′qi =1

T ′d0i

(Efdi − e′qi − (xdi − x′

di)idi)

e′di =1

T ′q0i

(−e′di + (xqi − x′qi)iqi

)

(31)

where e′qi and e′di are the transient voltage along q and daxes, iqi and idi are stator currents at q and d axes, Efdi

is the internal field voltage, T ′q0i and T ′

d0i are the open-circuit time constants for q and d axes, xqi and xdi are thesynchronous reactances, and x′

qi and x′di are the transient

reactances at the q and d axes, respectively.

Tmi and Efdi are considered as inputs. For generator iwhere a PMU is installed, the terminal voltage phaosrEti = eRi+jeIi and terminal current phasor Iti = iRi+jiIiare measured. See Qi [2015] for the detailed system modelabout those input, output, and state vectors. Gaussiannoise is added to the outputs. The variance for eRi andeIi is (0.05p.u.)

2 and that for iRi and iIi is (0.05p.u.)2.

5.2 Test Results with Accurate Model Parameters

Since the DOA for such a high-dimensional system is dif-ficult to visualize or analyze, we only study its projectiononto a subspace about specific directions to disturb thesystem, i.e. about all “N-1” contingencies each losing a sin-gle transmission line following a fault: adding a three-phaseshort-circuit fault to one end of each line and tripping theline on both ends to clear that fault after a fault clearingtime Tc. We do not consider faults that cause instabilitywith a Tc even as short as 0.05s, which is shorter thanthe response time of most realistic protective relays. Suchfaults are typically very close to generators and shouldbe avoided all the time. Finally, there are totally 202contingencies to be considered in the tests.

(a) log(UI) with small contingencies (Tc = CCT/5)

(b) log(UI) with large contingencies (Tc = CCT )

Fig. 12. Comparison of the unobservability indices

In order to find the boundary of the DOA following each ofthose contingencies, we first search for the critical clearingtime denoted by CCT , i.e. the maximum Tc withoutcausing instability. To place PMUs in a realistic powergrid, we want to demonstrate the necessity of evaluatingobservability from a nonlinear system perspective, so weevaluate the proposed UI for both small (very stable) andlarge (marginally stable or unstable) contingencies: wecalculate the UI for both Tc = CCT and Tc = CCT/5for each strategy of PMU placement. We consider the

Page 8: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

number of PMUs to be placed respectively from 1 to 48at generator buses. The placement is optimized using themethod in Qi [2015].

Fig. 12 visualizes log(UI) for each contingency and foreach optimized PMU placement strategy, whose value<1.477, i.e. UI=30 (colored green) corresponds to reason-able observability. From the figure, using 16 PMUs (i.e.1/3 of generators being measured directly) or more withoptimized PMU placement, most UIs are less than 30(corresponding to green and colder colors) for both smalland large contingencies. Compared with plot (b) for smallcontingencies, plot (a) for large contingencies has somecolder colors penetrating into the left warm color regions,and even with PMUs fewer than 16, some contingenciesstill have UIs < 30. An explanation is that when thesystem is under large disturbances, especially close toloss of stability, trajectories of state variables may exhibitsimpler patterns and might become easier to observe. Theestimation results with 16 PMUs for the contingency 26with UI=22.83 are illustrated in Fig. 13(a)-(b). Fig.14compares the state estimates with the true values for 5generators with larger errors, where generators 15, 23 and24 represented using the 4th order model each have statevariables δi, fi = ωi/2π, e

′qi, and e′di, and generators 34

and 47 represented using the 2nd model only have thefirst two state variables.

(a) Tc = CCT/5

(b) Tc = CCT

Fig. 13. Contingency 26: true angles of all generators vs.the estimates by 16 PMUs

5.3 Test Results with Inaccurate Model Parameters

We also test the case with inaccurate model parameters.To quantitatively compare the estimation results, we de-fine the following overall system state estimation errorindex

εx =

√√√√√g∑

i=1

Ts∑t=1

(xesti,t − xtrue

i,t )2

g Ts(32)

Fig. 14. Contingency 26: true states vs. estimates on 5generators (15, 23, 24, 34 and 47) with larger errors

Fig. 15. Contingency 26: State estimation error indexes forrandom line impedance changes (16 PMUs)

where state variable xi can be δi, fi = ωi/2π, e′qi, or e′di;

xesti,t is the estimated state and xtrue

i,t is the correspondingtrue value for generator i at time step t; Ts is the totalnumber of time steps in the simulation period.

Consider four types of model inaccuracies:

• Type-A: all line impedances have random relativeerrors uniformly from [−Δline,+Δline]

Page 9: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

Fig. 16. Contingency 26 (16 PMUs): state estimation errorindexes vs. relative load errors

(a) Relative line impedance errors up to 25%

(b) Relative bus load errors up to 10%

Fig. 17. Contingency 26: true angles of all generators vs.the estimates by 16 PMUs with inaccurate parameters

• Type-B: all bus loads have random relative errorsuniformly from [−Δload,+Δload]

• Type-C1: the actual system has random one line outof service, which is not told from the model; i.e. an”N-1” condition.

• Type-C2: the actual system has random two lines outof service, which is not told from the model; i.e. an”N-2” condition.

We still test the state estimation using 16 PMUs for thecontingency 26 with Tc = CCT . Respectively, we generate50 random cases for each of the four types.

For Type-A and Type-B, it is found that estimationerror indexes become significant when Δline = 0.25 andΔload = 0.1 (i.e. up to 25% and 10% errors of the true lineimpedances and bus loads), as shown by Fig. 15 and Fig. 16on the estimation error indexes with bar charts indicatingtheir standard deviations. Within those thresholds, theangle estimation error is less than 10 degree, which willnot significantly impact the judgment of system stability.Fig. 17 illustrates the estimates vs. the true states of twocases respectively from Type-A having Δline = 0.25 andfrom Type-B having Δload = 0.1. The estimates have moreoscillations in first 2 seconds but finally correctly trackall states. For both cases, 16 PMUs are able to robustlyjudge stability of the system even with slightly inaccurateparameters.

The histograms about εδ and εω of 50 random cases ofType-C1 and Type-C2 are shown in Fig. 18. 45 out of50 (90%) Type-C1 cases and 32 out of 50 (64%) Type-C2 cases have εδ ≤10 degree and εf ≤ 0.2 Hz. These testsshow that, for a majority of unknown outages of one or twolines, state estimates using an inaccurate system model notconsidering the outage have errors at a low level so as notto influence the judgment of system stability significantly.

(a) εδ for Type-C1 (left) and Type-C2 (right)

(b) εf for Type-C1 (left) and Type-C2 (right)

Fig. 18. Contingency 26 (16 PMUs): Histograms on stateestimation error indexes

6. CONCLUSION

The concept of observability and its computational algo-rithms can be used as a tool to evaluate the effectiveness ofa PMU network. For a 9-bus model with three generators,a single PMU makes the entire system observable at theboundary of the DOA. Using the data from the PMU, allstate variables can be estimated using virtue sensors such

Page 10: Power System Observability and Dynamic State Estimation for …web.eecs.utk.edu/~kaisun/papers/2016_CEP.pdf · 2016-02-10 · Power System Observability and Dynamic State Estimation

as a UKF. The computation shows that observability is notrobust to model uncertainties. However, the unknown pa-rameter variations are observable. It implies that adaptiveestimation method should be used to provide reliable stateestimates. The proposed methodology has also been testedon a realistic 140-bus 48-generator NPCC system model.The results indicate that 1/3 of generators being measureddirectly by PMUs with an optimized placement can makethe system state reasonably observable for most of “N-1” contingencies especially when the system approachesinstability, i.e. Tc = CCT . Also, the proposed UKF-basedestimation method can track system states for stabilitymonitoring even with some inaccurate model parametersin line impedances, bus loads and the network topology.These features are important for real-time power systemstability monitoring and control (Sun [2011a,b, 2012]).

REFERENCES

W. Kang. L. Xu (2009). Computational Analysis ofControl Systems Using Dynamic Optimization. arXiv :0906.0215v2.

W. Kang. L. Xu (2009) A Quantitative Measure ofObservability and Controllability. The 48th IEEE CDC,Shanghai, China.

Arthur J. Krener, K. Ide (2009). Measures of unobserv-ability. The 48th IEEE CDC, Shanghai, China.

W. Kang, L. Xu (2012) Optimal placement of mobilesensors for data assimilations. Tellus A, 64, 17133.

A. K. Singh, J. Hahn (2005). Determining optimal sensorlocations for state and parameter estimation for sta-ble nonlinear systems. Ind. Eng. Chem. Res., 44(15),56455659.

A. K. Singh, J. Hahn (2006). Sensor location for stablenonlinear dynamic systems: Multiple sensor case. Ind.Eng. Chem. Res., 45(10), 36153623.

T. Baldwin, L. Mili, J. Boisen, M.B. Boisen, R. Adapa(1993) Power system observability with minimal pha-sor measurement placement. IEEE Trans. Power Syst.,8(2), 707–715.

B. Xu, A. Abur (2004). Observability analysis and mea-surement placement for systems with PMUs. in IEEEPower Syst. Conf. Expo., Vol. 2, 943–946.

B. Gou (2008). Optimal placement of PMUs by integerlinear programming. IEEE Trans. Power Syst., 23(3),1525–1526.

S. Chakrabarti, E. Kyriakides. (2008). Optimal placementof phasor measurement units for power system observ-ability. IEEE Trans. Power Syst., 23(3), 1433–1440.B. Milosevic, M. Begovic (2003) Nondominated sort-ing genetic algorithm for optimal phasor measurementplacement. IEEE Trans. Power Syst., 18(1), 69–75.

F. Aminifar, C. Lucas, A. Khodaei, M. Fotuhi-Firuzabad(2009). Optimal placement of phasor measurement unitsusing immunity genetic algorithm. IEEE Trans. PowerDel., 24(3), 1014–1020.

Q. Gong, W. Kang, I. M. Ross (2006). A Pseudospec-tral Method for the Optimal Control of ConstrainedFeedback Linearizable Systems. IEEE Trans. AutomaticControl, 51(7), 1115-1129.

S. J. Julier, J. K. Uhlmann (2004). Unscented filtering andnonlinear estimation. Proceedings of the IEEE, 92(3),401-422.

P. M. Anderson, A. A. Fouad (1994). Power SystemsControl and Stability, IEEE Press.

W. Kang, K. Sun (2013) Detection of Instability UsingSynchrophasors: A Theoretical Investigation on Observ-ability with Synchrophasor Networks. EPRI TechnicalReport, No. 3002002061.

W. Kang (2011) The consistency of partial observabilityfor PDEs. arXiv : 1111.5846v1.

F. Fahroo, I. M. Ross (1998). Costate Estimation by aLegendre Pseudospectral Method. The AIAA Guidance,Navigation and Control Conference, Boston, MA.

J. Qi, K. Sun, W. Kang (2015) Optimal PMU Placementfor Power System Dynamic State Estimation by UsingEmpirical Observability Gramian. IEEE Trans. PowerSystems, vol. 30, pp. 2041-2054, July 2015.

K. Sun, W. Kang (2014) Observability and EstimationMethods Using Synchrophasor. The 19th IFAC WorldCongress, Cape Town, South Africa.

K. Sun, K. Hur, P. Zhang (2011) A New Unified Schemefor Controlled Power System Separation Using Syn-chronized Phasor Measurements. IEEE Trans. PowerSystems 26(3), 1544-1554

K. Sun, S. T. Lee, P. Zhang (2011) An Adaptive PowerSystem Equivalent for Real-time Estimation of StabilityMargin using Phase-Plane Trajectories. IEEE Trans.Power Systems, 26(2), 915-923

K. Sun, X. Luo, J. Wong (2012) Early Warning of Wide-Area Angular Stability Problems Using Synchrophasors.IEEE Power & Energy Society General Meeting. SanDiego, CA, USA

J. Chow, G. Rogers (2008) User manual for power systemtoolbox, version 3.0, 1991–2008.

M. Baechle, V. Knazkins, M. Larsson, P. Korba, (2014)Improved Rotor Angular Speed Measurement: a Key forProper Power Grid Stabilization. IEEE Power & EnergySociety General Meeting. National Harbor, MD, USA

Dobson, I., Carreras, B.A., Lynch, V. E., Newman, D.E.(2001) An initial model for complex dynamics in electricpower system blackouts. 34th Hawaii Int. Conf. Syst.Sci., 710718.

Qi, J., Dobson, I., Mei, S. (2013) Towards estimating thestatistics of simulated cascades of outages with branch-ing processes. IEEE Transactions on Power Systems,28(3), 3410-3419.

Qi, J., Sun, K., Mei, S. (2015) An interaction model forsimulation and mitigation of cascading failures. IEEETransactions on Power Systems, 30(2), 804-819.

Van Der Merwe, R., Wan, E.A. (2001) The square-root unscented Kalman filter for state and parameter-estimation. IEEE Int. Conf. Acoustics, Speech, andSignal Processing (ICASSP), 6, 34613464.

Qi(2015) Qi, J., Sun, K., Wang, J., Liu, H. (2015) Dynamicstate estimation for multi-machine power system by un-scented Kalman filter with enhanced numerical stability.arXiv:1509.07394.