You are on page 1of 11

This article has been accepted for inclusion in a future issue of this journal.

Content is final as presented, with the exception of pagination.


IEEE TRANSACTIONS ON SMART GRID

Robust Transient Stability-Constrained Optimal


Power Flow With Uncertain Dynamic Loads
Yan Xu, Member, IEEE, Jin Ma, Member, IEEE, Zhao Yang Dong, Senior Member, IEEE,
and David J. Hill, Life Fellow, IEEE

AbstractLoad dynamics has a substantial impact on transient


stability of a power system, but has not been properly treated in
transient stability-constrained operational planning. This paper
proposes a robust transient stability-constrained optimal power
flow (OPF) model considering dynamic load models and uncertain variations in the model parameters (e.g., load composition).
The uncertainty is modeled by selecting a small yet representative
number of deterministic scenarios that approximate the whole
uncertainty space. A decomposition-based solution approach is
then developed, which iterates between a master problem corresponding to the ordinary OPF, and a set of subproblems
corresponding to transient stability assessment and stabilization constraint generation. The proposed model and solution
approach is validated on the New England 39-bus system. The
obtained solutions can well immunize against random realizations
of uncertain load model parameters while maintaining transient
stability.
Index TermsTransient stability, optimal power flow, dynamic
load model, model uncertainty.

I. I NTRODUCTION
RANSEINT
stability-constrained
optimal
power
flow (TSCOPF) is an effective tool for balancing
the economic concerns and the stability requirements in
power system operations. Given an unstable contingency, the
TSCOPF aims to optimize the system operating states while
retaining the ability to withstand the contingency should it
really occur [1].

Manuscript received August 19, 2015; revised October 28, 2015 and
November 30, 2015; accepted December 13, 2015. This work was supported in part by the Faculty of Engineering and Information Technologies,
University of Sydney through the Faculty Research Cluster Program and the
Early Career Researcher Development Scheme, and in part by China Southern
Power Grid Company under Project WYKJ00000027. The work of Y. Xu was
supported by the University of Sydney Postdoctoral Research Fellowship.
Paper no. TSG-00969-2015.
Y. Xu and Z. Y. Dong are with the School of Electrical and Information
Engineering, University of Sydney, Sydney, NSW 2006, Australia, and also
with the China Southern Power Grid Research Institute, Guangzhou 510000,
China (e-mail: eeyanxu@gmail.com; zydong@ieee.org).
J. Ma is with the School of Electrical and Information Engineering,
University of Sydney, Sydney, NSW 2006, Australia (e-mail:
j.ma@sydney.edu.au).
D. J. Hill is with the Department of Electrical and Electronic Engineering,
University of Hong Kong, Hong Kong, and also with the School of Electrical
and Information Engineering, University of Sydney, Sydney, NSW 2006,
Australia (e-mail: dhill@eee.hku.hk).
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TSG.2015.2510447

From a mathematics perspective, the TSCOPF is a largescale non-linear and non-convex programming problem containing differential-algebraic equations (DAEs), which is difficult to be directly solved. A variety of solution methodologies
have been proposed with varying pros and cons. In general, the methodologies can be classified into the following
categories: 1) numerical discretizing method [1][4], which
discretizes the differential equations into numerically equivalent algebraic ones, making the model tractable by classic
programming algorithms. However, the major drawback of
this approach lies in the heavy computational burden which
is proportional to the product of the number of generators
and the simulation steps; 2) direct method [5][7], which
analytically derives the required generation shifting amounts
for stabilizing the system and incorporates it into the ordinary OPF model. This approach is advantageous for its high
solution efficiency and transparency. However, it is usually
argued that it can only obtain a near-optimal solution; 3) model
reduction method [8][10], which approximates the original,
multi-machine model with a reduced, much smaller one; therefore the solution efficiency can be increased drastically but the
accuracy may be sacrificed; 4) evolutionary algorithm (EA)based method [11], [12], which drives an EA to stochastically
search the optima of the problem. Its advantage lies in higher
chance to obtain high-quality solutions, however it usually suffers from slow convergence and inconsistence issues. In this
regard, Xu et al. [13] developed a hybrid approach combining
an EA and classic programming algorithm to overcome these
drawbacks; 5) statistical method [14], [15], which acquires the
stable/unstable operating regions or rules through statistical
learning from a transient stability database and incorporates
the regions or rules as explicit constraints into the ordinary
OPF model. The stable/unstable operating region or rules are
transparent and interpretable which can also be used for stability monitoring. However, this method is sometimes sensitive
to the change of stability database. A systematical review,
discussion, and comparison on these methods can be found
in [16].
While the generator dynamics are well represented in
the TSCOPF models, the load dynamics has not been
properly treated yet. In the literature, almost all of the
reported methods assume the loads to be static (e.g., constant impedance [1][15]). But in real world, an actual load
bus is a mix of various static and dynamic components
such as induction motors. Such dynamic components have
a substantial impact on power systems short-term stability

c 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
1949-3053 
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
2

IEEE TRANSACTIONS ON SMART GRID

(both rotor angle and voltage stability, which are strongly


coupled). Following a short-circuit fault, the induction motors
may decelerate dramatically or even stall if the electrical
torque is insufficient for driving the mechanical load. This
in turn draws very high reactive current from the network that
strongly decreases the voltage magnitudes and impedes the
voltage recovery. The delayed voltage recovery and/or progressive voltage drop prevent power export of synchronous
machines, which may increase the rotor angle separation and
lead to the loss of synchronism [37]. There has been a long
recognized need for incorporating detailed load models for
stability analysis [17], [37]. As illustrated by post-event validation studies, e.g., the 1996 Western Electricity Coordinating
Council (WECC) blackout, the standard load models used in
the simulation failed to replicate the disturbance recordings but
lead to a converse conclusion [19]. Hence, without appropriate representation of load dynamics, the TSCOPF may fail
to maintain the transient stability if the contingency really
occurs.
On the other hand, due to the stochastic nature of the
load behavior, the load composition (or load model parameters) is uncertain and time-varying. It has been widely shown
that varying load composition can result in diverse dynamic
trajectories under the same disturbance [20]. Thus, the load
model uncertainties should be well accounted for in the
TSCOPF, which, would otherwise lead to conservative or
risky dispatching results. With increased installations of smart
meter in the Smart Grid era, it is possible to capture more
detailed load models to support system security operation and
control [34].
This paper proposes a robust TSCOPF model considering
dynamic load models and uncertain model variations. The
load is modeled by a composite of both static and dynamic
components. Rather than specifying the load model to have
a deterministic set of parameters, the load uncertainty is modeled by a set of strategically selected scenarios that represent
the whole uncertainty space. A decomposition-based solution
approach is then developed to convert the original large-scale
nonlinear programming problem into a set of deterministic
OPF and transient stability assessment (TSA) problems. The
proposed model and solution approach is demonstrated on
the New England 39-bus system. The robust degree of the
obtained solutions and the computation efficiency are also
analysed.
II. L OAD M ODELING
A. Composite Load Model
During the transient period, one key driving force for voltage instability is the tendency of dynamic loads to restore their
consumed power within a very short time frame. As already
described, short-term voltage instability and transient instability are strongly coupled and sometimes undistinguished.
According to the simulation studies in [18], given the same
fault, different load models (static or dynamic) can result in
contrary transient stability statuses. Consequently, it is necessary to include dynamic load models into the TSCOPF for
a reliable dispatching result.

Fig. 1.

Structure of the composite load model CLOD.

For a straightforward illustration of the problem, this paper


uses the composite load model CLOD [21]. This load model
aggregates all constant MVA, current, and admittance load into
a composite load consisting of induction motors, lighting, and
other typical equipments see Fig. 1. The model allows the
user to specify a minimum amount of data stating the general
character of the composite load. It uses this data internally to
establish the relative capacity of motor models for dynamic
simulation and to establish typical values for the detailed
parameter lists required in the detailed modeling [21].
Specifically, this composite load model is defined by
8 parameters [pL , pS , pD , pT , pC , Kp , Re , Xe ] corresponding respectively to the percentage value of large motors (LM),
small motors (SM), discharge lighting (DL), transformer exciting current (TX), constant power load (CP), the exponent of
voltage-dependent real power load, and branch resistance and
reactance (p.u. on load MW base). The model has been implemented by commercial software such as PSS/E as a standard
composite load model [21].
B. Uncertainty Modeling
During real-time operation, load composition is stochastic
and time-varying, and the variation can exert a substantial
influence on the system dynamic response [20], implying that
a single set of load model parameters is inadequate to capture
the full range of expected dynamic behavior of the system.
Therefore, for a robust generation dispatch in the presence of
uncertain load compositions, the load model parameters should
be specified in terms of their statistical properties.
In general, optimization under uncertainty can be classified
as stochastic and robust approaches [22]. The former relies on
a presumed probability density function (PDF) of the uncertain
parameters, which is difficult to accurately identify. Besides, it
usually suffers from a heavy computational burden due to the
need for a huge number of sampling scenarios. By contrast,
robust approach only requires knowledge of the variation range
of the uncertain parameters, e.g., the upper and lower bounds
representing the uncertainty spectrum, and is therefore more
computationally efficient. However, it usually results in the
worst-case scenario and conservative solutions.
Rather than adapting the uncertainty bounds into the
TSCOPF model which will yield tremendous mathematical
complexity, this paper employs an efficient scenario-based
scheme for uncertainty modeling. Specifically, the Taguchis
orthogonal array testing (TOAT) [23] is employed to select
a small number of testing scenarios with good statistical information in the uncertainty space. TOAT is originally used for
robust design, as proven in [24] it is able to select a small yet

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
XU et al.: ROBUST TSCOPF WITH UNCERTAIN DYNAMIC LOADS

TABLE I
T ESTING S CENARIOS D ETERMINED OF OA L8 (27 )

representative set of testing scenarios to approximate all the


possible combinations of uncertain variables in additive and
quadratic models. Here, additive or quadratic models mean that
the uncertain variables have an additive (linear) or quadratic
impact on the model output. Compared with stochastic optimization, TOAT uses only a few deterministic scenarios to
approximate the whole uncertainty space and is therefore much
more computationally efficient. Compared with robust optimization, TOAT is scenario-based and hence free from a strict
mathematical formulation and less prone to conservatism. In
the literature, TOAT has been applied in transmission network expansion planning [25] and conventional optimal power
flow (OPF) problems [26] for handling uncertainties arising
from stochastic renewable energy generation and load demand
variation.

3) In any two columns, combinations of two variable levels


occurs the same times, i.e., (1) (1), (1) (2), (2)
(1), and (2) (2) appear once.
4) The combinations determined by OA are uniformly
distributed over the uncertainty space.
5) If any two columns of an OA are exchanged or some
columns are ignored, the resulting array still satisfies the above
features.
As studied in [23] and [24], the selected testing scenarios are able to represent the whole variable combinations
with good statistical information. OAs can be constructed
by mathematical methods [23] or directly indexed from
OA libraries [27]. It is important to choose an appropriate
OA number B which indicates the representative levels of the
random variables. In general, if the uncertain variable has a linear effect on the system, m should have two testing levels,
and if m is symmetrically distributed, (m ) (m ) and
(m ) + (m ) should be used, where (m ) and (m ) are
the mean and standard deviation of m , respectively; if m has
three levels,
a quadratic effect on the system, m should have
distributed,
(

3/2 (m ),
and if m is symmetrically
m

(m ), and (m ) + 3/2 (m ) should be used. When using


the OA libraries, if there is not an OA with desired variable
number (M), according to feature 5), an OA with the next
larger M can be used by ignoring redundant columns.

III. M ATHEMATICAL M ODEL


The following TSCOPF model is proposed aiming to adjust
control variables u to make the system robust stable to the
random variations of load model , i.e., the system remains
transient stable regardless the variation of the load model
parameters.
min
u

(1)

s.t. g(x, u) = 0
h(x, u) 0
TSI(x, u, )

C. Taguchis Orthogonal Array Testing


Given an uncertainty space formed by M stochastic variables
= [1 , . . . M ], if B value levels are considered, e.g., two
levels can be the minimum and the maximum value of the
variable, the whole uncertainty space has BM combinations,
which is computationally unaffordable when M is large. With
TOAT, scenarios are determined by orthogonal arrays (OAs),
where an OA is a matrix LH (BM ) which uses H combinations
of levels (H<< BM ) to represent the whole uncertainty space
(H and M are respectively the number of rows and columns
of the matrix, and B is the number of the matrix element
levels) [25].
As an example, for a system with seven random variables
and each being represented by two value levels, there will be
a total of 27 combinations. With TOAT, an OA denoted as
L8 (27 ) can be formed to represent the all the combinations
with only eight scenarios as shown in Table I.
Commonly, an OA has the following features:
1) Each testing scenario is a row in the matrix, and there
are H (here, H=8) scenarios in total.
2) In each OA column, every value level of a variable occurs
H/B times, i.e., (1) and (2) both occur four times.

f (x, u)

(2)
(3)
(4)

where Eqs. (1) to (3) correspond to an ordinary OPF model,


and Eq. (4) implicitly represents the transient stability constraint through a transient stability index (TSI); x and u denote
the vectors of state variables and control variables, respectively; denotes the vector of uncertain dynamic load model
parameters, and is a threshold for stability requirement.
A. Objective Function
Eq. (1) is the objective function for total generation costs:
f =

NG 


ai P2Gi + bi PGi + ci


(5)

i=1

where ai , bi , ci are the generation cost coefficients of the i-th


generator, PGi is the active output of the i-th generator, and
NG is the total number of dispatchable generators.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
4

IEEE TRANSACTIONS ON SMART GRID

B. Equality Constraints
Eq. (2) corresponds to the power flow equations:

NB


V
Vj Gij cos ij + Bij sin ij = 0

Gi
Di
i

j=1

NB

Vj Gij sin ij Bij cos ij = 0


Q QDi Vi

Gi

(6)

j=1

where Vi is the voltage magnitude of the i-th bus, QGi is the


reactive output of the i-th reactive power source, PDi and QDi
are the active and reactive load of the i-th load bus, respectively; Gij and Bij are conductance and susceptance between
the i-th and the j-th bus, respectively; and NB is the total
number of network bus bars (including generator, load, and
connecting buses).

equivalent (SIME) [30], is a hybrid TSA technique which


integrates the full time-domain simulation (TDS) and the
equal-area criterion (EAC). Specifically, it drives the TDS to
obtain the multi-machine trajectories (complex system model
is therefore not a limitation), and then separates them into two
exclusive clusters: one composed of critical machines (CMs)
which are responsible for the loss of synchronism, and the
other composed of non-critical machines (NMs). The two
clusters of CMs and NMs are represented as two equivalent
machine trajectories:


Mi i (t); C (t) = (MC )1
Mi i (t)
C (t) = (MC )1
iC

N (t) = (MN )

iC

Mj j (t); N (t) = (MN )

jN

(8)
Mj j (t)

jN

(9)

C. Inequality Constraints
Eq. (3) corresponds to the steady-state lower and upper
limits of the operational variables:
min
PGi PGi Pmax
i = 1, 2, . . . , NG

Gi ,

Qmin Q Qmax , i = 1, 2, . . . , N
Gi
G
Gi
Gi
(7)
min V V max ,

V
i
=
1,
2,
.
.
.
,
N
i
B

i
i

min
Li Li Limax ,
i = 1, 2, . . . , NL
where Li is the apparent power across the i-th branch.
D. Transient Stability Constraints
Eq. (4) is the transient stability constraint which corresponds
to a large set of DAEs characterizing the systems dynamic
behavior. In most of the literature, the transient stability criterion is expressed as the maximum post-disturbance rotor angle
deviation bounded by a pre-defined threshold, e.g., 180 [1][4].
However, this threshold is usually system and/or operating
state dependent, which is not accurate in reflecting the transient stability: if this type of thresholds are set smaller, the
resulting system tends to be conservative, which hence leads
to higher operating costs; if too relaxed, the stability may not
be retained even if the threshold is not exceeded [9].
In this paper, the TSI is implicitly expressed in terms of
the transient stability margin defined by the extended equalarea criterion (EEAC) [28]. The motivation is twofold: first,
it provides a way for quantitative and fast assessment of transient stability, which can measure the degree of stability while
greatly increasing the computation efficiency; second, it can
derive the required generation shifting to stabilize the system, based on which equivalent linearized stability constraints
can be formulated. In this way, it is possible to decompose
the original large-scale problem into a series of smaller and
tractable problems, diminishing significantly the complexity
of the problem and enabling parallel computing to improve
solution efficiency.
IV. E XTENDED E QUAL -A REA C RITERION
A. Principle
The EEAC was proposed by Xue et al. in [28]. Its state-ofthe-art version IEEAC [29], also known as single machine

where subscripts C and N denote the CMs and NMs, respectively; MC and MN are respectively the inertia coefficient of
CMs and NMs, calculated as:


Mi ; MN =
Mj
(10)
MC =
jN

iC

The one-machine-infinite-bus (OMIB) equivalence of the


multi-machine trajectories is then constructed as follows:
(11)
(t) = C (t) N (t); (t) = C (t) N (t)



Pmi (t) (MN )1
Pmj (t)
Pm (t) = M (MC )1
jN

iC

Pe (t) = M (MC )1

Pei (t) (MN )1

(12)

Pej (t)

jN

iC

(13)
M = (MC MN ) (MC + MN )

(14)

where and here denote the rotor angle and angular speed
of the OMIB, respectively; Pm and Pe denote the mechanical
and electrical power of the OMIB, respectively.
Finally, the EAC is applied to the OMIB P plane
for quantifying the transient stability degree and extracting
stability information of the original multi-machine system.
B. Transient Stability Assessment
IEEAC or SIME provides the following TSA outputs [30]:
1) CMs and NMs, which determines an instability mode.
2) The time to instability Tu indicating the time that system loss synchronism. At this time, the curve of Pe
crosses Pm :
Pa (Tu ) = Pm (Tu ) Pe (Tu ) = 0, P a (Tu ) > 0 (15)
3) The time to first-swing stability Tr indicating the time
that the system can be declared as first-swing stable. At
this time, the curve Pe stops its excursion and return
back before crossing Pm :
Pa (Tr ) = Pm (Tr ) Pe (Tr ) < 0, (Tr ) = 0

(16)

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
XU et al.: ROBUST TSCOPF WITH UNCERTAIN DYNAMIC LOADS

Note that Tu and Tr can be used to terminate the


TDS much earlier, thereby saving computation time.
Specifically, during the TDS process, the machine angles
are continuously checked against (15) or (16), if either
condition is satisfied, the TDS can be stopped immediately without progressing ahead.
4) The transient stability margin quantifying the degree of
stability, determined by the decelerating area Adec minus
the accelerating area Aacc of the OMIB P plane:
= Adec Aacc

if (15) meets
M((Tu ))2 /2,
=
(17)
|Pa (Tr )|((Tu ) (Tr ))/2, otherwise
where Tu is obtained from a less unstable case or estimated through triangle extrapolation [30]. < 0 means
the system is unstable; otherwise stable.
For correct identification of CMs and NMs, during the
TDS, several decomposition patterns of the two machine
angle clusters are constructed based on the rotor angle deviations between adjacent machines rotor trajectories [30] (e.g.,
top 3 largest angle deviations). Each candidate clustering
of machines is replaced by the OMIB, and the candidate
OMIB that has lowest stability margin is declared as the right
clustering.
Based on IEEAC, TSA can be achieved quantitatively
and much more efficiently while maintaining necessary engineering accuracy. Its effectiveness and accuracy have been
verified by several industry-grade software packages such as
FASTEST [31] and DSATools [38], which have been practically implemented in many countries.

Eq. (18)-(21) reveal that by shifting real power output of CMs to NMs, the transient stability can be
restored [6], [28][30].
Numerous simulations have shown an approximate linear
relationship between changes of stability margin and OMIB
mechanical power at pre-contingency state [6], [28][30], i.e.,
= Pm (t0 )

(22)

where is the approximate linear sensitivity of the stability


margin with respect to generation change.
Simply, the sensitivity value around the operating point n
can be numerically estimated via two successive runs:
n =

(n2) (n1)
Pm (t0 )(n2) Pm (t0 )(n1)

(23)

With n , the required generation shifting for stabilizing can


be analytically determined. Specifically, to stabilize an unstable state with a stability margin us (us < 0), if the desired
stability margin is ( > 0), the required increment in stability
margin is us + . Based on (21)-(23), the required
generation shifting between CMs and NMs can be calculated
as follows:
 


M
us +
M 1

PC
+
(24)
n
MC
MN
Note that the power loss is temporarily neglected in (20)
since it is unknown before the generation shifting. The power
loss will be implicitly compensated by NMs through (24)
which will be added as a constraint to the OPF model.

C. Transient Stability Control


IEEAC can provide significant information for transient
stability control. In the literature, it has been adopted as
a direct [6], [7] or a model reduction method [8][10] for
TSCOPF.
Specifically, stabilizing an unstable system consists of
modifying the pre-contingency condition until stability margin becomes positive. This can be achieved by increasing
the decelerating area Adec and/or decreasing the accelerating
area Aacc of the OMIB P plane. In practice, this can be
realized by reducing the OMIB mechanical power Pm (t0 ) as
follows:
M
M
Pm (t0 ) =
PC
PN
(18)
MC
MN
where t0 denotes the pre-contingency state, PC and PN are
respectively the changes in the total power of CMs and NMs:


PC =
Pmi (t0 ); PN =
Pmj (t0 )
(19)
jN

iC

To maintain the power balance, the following condition


should be satisfied while neglecting power loss:
PN = PC

(20)

Substituting (20) into (18):






M
M
M
M
PC =
PN
+
+
Pm (t0 ) =
MC
MN
MC
MN
(21)

V. C RITICAL U NCERTAIN PARAMETERS I DENTIFICATION


While a dynamic load model contains several parameters
(for the composite load model used in this paper, it has a total
of 8 parameters), it may be the case that not all of them
has a remarkable impact on the system stability [32]. In this
regard, it is necessary to identify the critical model parameters and include only them into the TSCOPF model. This
could reduce the number of uncertainty parameters M and
therefore the number of testing scenarios. Such critical parameters are also known as well-conditioned parameters which
means they exert a large influence on the system trajectory,
and the influence can be quantified in terms of trajectory
sensitivities [20], [32].
Trajectory sensitivity is the time-varying sensitivity
of (small) changes in a parameter with respect to the
system dynamic behavior [20]. It can either be calculated
analytically [20] or numerically approximated through two
successive TDS runs:
(x0 , t, + ) (x0 , t, )
(x0 , t, )
=
x

(25)

where (x0 , t)/x denotes the trajectory sensitivity along


time t given the initial system condition x0 , denotes the
dynamic trajectory of the system, is the perturbation.
Normally, it can be empirically set to a small value, e.g., 5%
for the problem studied in this paper.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
6

IEEE TRANSACTIONS ON SMART GRID

At the initial iteration, (3) only comprises the steady-state


operational constraints (7), while the transient stability constraints are disregarded. After the sub-problems are solved,
additional constraints may be generated, and if any, are
attached to (3) in subsequent iterations to eliminate stability
violations.
B. Sub-Problem

Fig. 2.

Proposed decomposition-based solution approach.

Based on the concept of the trajectory sensitivity, a sensitivity index (SI) can be calculated to rank the relative influence
of the load model parameters on system transient stability:
SI =

TSI(x, u, + ) TSI(x, u, )
TSI
=

(26)

Depending on practical needs, the top few parameters


can be selected as critical model parameters and included
in the optimization model. Specifically, if some parameters
have remarkably lower SI values, they should be obviously
excluded. While if the parameters SI values are very close,
it is usually needed to define a definite number for critical
parameters.
VI. S OLUTION A PPROACH
The proposed robust TSCOPF is a large-scale non-linear
multi-state programming problem with DAEs and uncertain
parameters, which is hardly to be solved. This paper develops
a decomposition-based solution approach. As shown in Fig. 2,
the principle is to divide the original problem into a master
problem and a set of tractable smaller sub-problems, which
interact among each other iteratively.
A. Master Problem
The master problem is to solve the ordinary OPF

model (1)-(3) and the output is an operating state (x, u).


It should be noted that the OPF problem is a non-convex
non-linear programming model; therefore it is impractical to
obtain global optima. Although some EA-based methods claim
that global optimality can be acquired, the robustness and convergence cannot be always guaranteed in finite time. More
importantly, it is difficult to rigorously prove the global optimality of these solutions in real applications. Nowadays, the
most mature and practical OPF algorithm is interior-point (IP)
method. Although it can only ensure local optima and convergence, its reliability, solution quality and speed have been
proven and well examined in both academia and industry.
Consequently, IP method is used to solve the master problem
in this paper. Therefore, the solution of the whole TSCOPF
model is also locally convergent.

The sub-problems evaluate the stability of the master problem solution under each contingency and scenario and generate
stabilization constraints if unstable.
For each testing scenario determined by the TOAT, the
EEAC-based TSA is performed and if (4) is not satisfied, the
stabilization constraint, i.e., generation shifting between CMs
and NMs, is generated. To simultaneously stabilize multiple
contingencies/scenarios with least number of additional constraints, the contingencies/scenarios can be grouped according
to their resulting instability modes. Those having a common
instability mode are divided into one group. Each group is then
represented by the severest mode which is the one with the
smallest stability margin in that group. For each representative
instability mode, the stabilization constraint is:
 




M
M 1
us +
Pmi (t0 )

Pmi (t0 )
+
n
MC
MN
iC

iC

(27)
where P mi (t0 ) denotes the active power output of generator i
obtained from the master problem.
The stabilization constraint (27) conveys information about
how the generation dispatch should be modified to retain the
transient stability. It is important to note that (27) is a linear
constraint. Hence, it can be directly attached to (3) in the
master problem. In such a way, the original problem size can
be reduced to one similar to an ordinary OPF. Note that the
power loss neglected in (20) will be implicitly compensated
after adding (27) to the master problem.
C. Computation Steps
1) Prepare the computation data, including the system
data (network data and dynamic models), contingency set,
and the mean value and the standard deviation of
the load model parameters. The load model parameters can
be obtained from load composition statistics, load modeling practices [33], and/or advanced smart grid functions (e.g.,
smart meter data-based home appliance load modeling [34]).
2) For each load model parameter, estimate their impact
(around their mean value) on the transient stability of the system. This corresponds to calculating the SI based on Eq. (26).
Only those with high SIs are included in the uncertainty
parameter set whose dimension is denoted as M and the
remaining load model parameters are fixed at their mean
values.
3) For the uncertainty parameters, select an appropriate representative level (i.e., OA level), denoted as B. It will be shown
later that B=2 can be sufficient as the load model parameters
exhibit a quasi-linear effect on the stability margin.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
XU et al.: ROBUST TSCOPF WITH UNCERTAIN DYNAMIC LOADS

TABLE II
C ONTINGENCY S ET

TABLE III
BASE C ASE -I NITIAL G ENERATION D ISPATCH (MW) [13]

4) Given the values of M and B, form an OA LH (BM ) which


has H rows, each representing a testing scenario.
5) Solve the master problem [Eq. (1)-(3)] using the IP

algorithm, and obtain the current operating point (x, u).


6) Given the master solution, for each contingency, perform
the TSA using EEAC for each testing scenario of uncertain
load model parameters. Given K contingencies, there will be
KH EEAC simulation runs in total. The earlier-termination
criteria (15) and (16) can be applied to save the total CPU time.
In the meantime, the decomposition nature of the approach
also allows the parallel computing to speed-up the solution
procedure [35].
7) If all the contingencies under all the testing scenarios are
transient stable, stop; otherwise, go the next step.
8) For each representative instability mode, compute the
stabilization constraint (27) according to Section IV-C.
9) Add the generated stabilization constraints to the inequality constraint set (3), and go back to Step 5).
It should be indicated that the stabilization constraint
derived from EEAC is an approximation of the original
multi-machine angle trajectories, given the nonlinearity of the
TSCOPF problem, the complete contingency stabilization may
not be achieved immediately after being added to the master
problem (similar issues are also reported in [7] and [8]), so
the whole computation is processed iteratively until all the
contingencies/scenarios are stabilized.
VII. N UMERICAL R ESULTS

TABLE IV
M EAN VALUE () OF THE L OAD M ODEL PARAMETERS

Fig. 3.

System trajectories for base case with dynamic load model.

A. Simulation Tools and Test System


The numerical simulation is conducted on a 64-bit PC with
3.10 GHz CPU and 4.0 GB RAM. TDS is performed using the
commercial software PSS/E [21], and the IEEAC algorithm
is realized in the MATLAB platform. OPF is solved using
MATPOWER with the in-built IP algorithm [36].
The method is demonstrated on the New England 10machine system which is a popular benchmark in TSCOPF
research [5], [7], [8], [12], [13]. Three commonly studied
contingencies in the literature are considered (see Table II).
The base case is set to the three-contingency constrained solution from [13] (see Table III), which is the best dispatch
result reported in the literature. We assume the mean value
of parameters in Table IV and the standard deviation
to be 5%.
B. Transient Stability With Dynamic Loads
Without modeling dynamic load components, the base operating state is stable for all of the three contingencies. While
with the dynamic load model, all of the three contingencies
cause the system to become unstable - see Fig. 3, where

it is worth noting that the time to instability Tu allows


earlier-termination of the TDS.
The simulations clearly indicate that dynamic loads exert
a substantial impact on the system stability and justify the
needs for incorporating dynamic load models in the TSCOPF.

C. Single Contingency Case


A single contingency C1 is considered first. Based on
Eq. (26), the SIs are calculated for each load component,
and the values are: -18.4 (LM), -15.9 (SM), -0.0004 (TX),
3.68 (DL), and -6.74 (CP). This implies that load components have varying impact of degree on system stability.
Consequently, LM, SM, CP, and DL are included into the
uncertainty vector, while TX is set to its mean value. To determine a proper representative levels of the OA level B, the
quantitative impact of the load model parameters on the transient stability margin is then investigated. To this end, EEAC
are performed under continuously varying percentage values
for each load component, and the results are shown in Fig. 4.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
8

IEEE TRANSACTIONS ON SMART GRID

TABLE VI
M ULTI -C ONTINGENCY S OLUTION R ESULTS (MW)

Fig. 4.

Transient stability margin with varying load component proportion.


TABLE V
S INGLE C ONTINGENCY (C1) S OLUTION R ESULTS (MW)

According to Fig. 4, except for TX, the change in proportion


of other load components all have an apparent approximatelinear impact on the transient stability margin: for LM, SM,
CP, increased percentage value reduces the stability margin;
for DL, increased percentage value increases the stability
margin.
According to the TOAT theory, if the uncertain parameter
has a linear effect on the system, two value levels are sufficient
to represent the uncertain parameter space. Therefore, corresponding to an OA, B is two and M is four, which respectively
denote the number of testing levels and uncertain variables.
While there is not a coincident OA in the libraries, a redundant
OA L8 (27 ) can be used (see Table I) with the last three columns
being discarded. Determined by L8 (27 ), there are a total of
eight testing scenarios, (1) and (2) correspond to the lower
level ( ) and upper level ( + ) of the uncertain variable
m , respectively.
The robust TSCOPF model is solved using the proposed
approach, and the results are given in Table V. The CMs for
this contingency is G4, G5, G6, and G7, and the total generation shifting amount from the CMs to NMs is 320MW. The
increased cost over the base case is 542.35 $/Hr or 0.89%,
which is very limited for stability enhancement purpose.

D. Multi-Contingency Case
All of the three contingencies are included simultaneously
in the TSCOPF model. The solution results are given in
Table VI and the system trajectories under each contingency
are shown in Fig. 5. The CMs are: G2, G3, G4, and G5 for C2;
and G9 for C3. The increased cost over the base case is
1815.7 $/Hr or 2.98%.
The obtained solution is stable for all of the three contingencies under all of the eight testing scenarios. To illustrate, the
system trajectories under the eighth testing scenario are shown
in Fig. 5. It is important to note that the time to first swing
stability Tr allows earlier-termination of the TDS, which can

Fig. 5. System trajectories for the multi-contingency TSCOPF solution under


the eighth testing scenario.

significantly save the CPU time. Namely, when the TDS process till Tr , it can be stopped immediately without progressing
ahead.
E. Result Validation
To validate the robustness of the solution to the load model
uncertainty, a series of random validation scenarios are generated by Monte Carlo sampling given and of the load
model parameters, and each scenario is assessed for the three
contingencies. If the system is stable, the validation scenario
is defined as feasible, otherwise defined as infeasible. The
robustness degree is defined as follows:
=

Fi
100%
F

(28)

where F is the total number of validation scenarios (set to


100) and Fi is the number of feasible (stable) scenarios for
i-th contingency.
The robustness of the solutions obtained in this paper
and [13] are compared in Table VII. Note that the solution
reported in [13] is the best (most economical) result available in the literature and the only one that stabilizes three
contingencies simultaneously.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
XU et al.: ROBUST TSCOPF WITH UNCERTAIN DYNAMIC LOADS

TABLE VII
ROBUSTNESS D EGREE

In average, the robustness degree of the dispatch result


from the conventional TSCOPF model without considering
dynamic load models is 16%. By marked contrast, the robustness degree for the proposed robust TSCOPF model is 98.67%,
validating that it is well immunized against the load model
uncertainty.
F. Computation Efficiency Analysis
Compared with traditional TSCOPF methods, the proposed
one in this paper makes full use of EEAC and decomposition strategy to convert the whole large-scale problem into
smaller tractable problems. In each iteration, the computation
involves solving an OPF model (master problem) and performing TSA for each contingency under each load composition
scenario, and for each unstable scenario, transient stability
margin sensitivity is calculated for deriving stabilization constraint. So, the total computation time can be estimated as
follows:
Ttotal

NIt

= (TOPF + TTSA K H) NIt +
(TTSA Nui )
i=1

(29)
where TOPF , TTSA respectively denote the CPU time for OPF
solution, and TSA for a contingency; K and H respectively
denote the number of contingencies and the testing scenarios
determined by TOAT; NIt denotes the total iteration number,
and Nui denotes the number of unstable scenarios in the i-th
iteration. The second term in (29) represents that an unstable
scenario requires an additional TDS to calculate the stability
margin sensitivity for deriving the stabilization constraint.
For the test in this paper, the total computation time is 34.3s
and 110.4s for single- and multi-contingency cases, respectively, which are much faster than conventional methods such
as discretizing and EA-based methods. Further discussions are
given below.
For contingency number K, it is 3 in this paper. In practice,
although a larger system has more contingencies, only unstable ones are to be included in the TSCOPF. In fact, since
the transient stability is considered at the planning stage, for
a well-planned system, the number of unstable contingencies
can be limited. Besides, since different contingencies may have
different probabilities of occurrence, it is sensible to include
only those more likely (or higher impact) ones in TSCOPF.
For testing scenario H, it is 8 in this paper. According to
OA tables [27], H can be very limited even for a large model.
For iteration number NIt , given the effectiveness of EEAC
stabilization constraint, this number is generally small. It is
2 in this paper which is consistent with other literature [6][9].

For Nui , its maximum initial value is KH, but it will be


gradually decreased during the iterations (i.e., the unstable
scenarios are stabilized after each iteration).
Solving OPF can be very fast, in this paper, TOPF is 0.1s
using MATPOWER. The major computation time is therefore spent on the TSA phase. For TTSA , it is 0.91.1s using
PSS/E package. Note that this time can be further reduced
if earlier-termination is used. In the meantime, if a more
specific and powerful TSA engine such as DSATools [38]
is used, the speed can be much faster. For the same task,
DSATools only cost 0.12s for one contingency. Consequently,
the total computation time will become 3.8s and 12s for singleand multi-contingency cases, which are fully compatible for
on-line use.
Moreover, it should be pointed out that the proposed method
can be parallelized given its decomposition structure. Hence,
if H CPU cores are available, the speed can be increased H
times by distributing the sub-problems to the H CPU cores.
A parallel computing platform for PSS/E based TSA has been
developed in our previous work [35]. It is worth indicating
that the parallel computing is widely available at a modern
power system control center, so the proposed approach can
make full use of such computing power.

VIII. C ONCLUSION
Load dynamics have a substantial impact on system stability. While conventional TSCOPF models neglect this concern,
this paper shows that the dispatch results from such models
may not be able to sustain the stability if the dynamic load is
considered. A robust TSCOPF model is proposed to consider
the dynamic load and its parameter uncertainty. In terms of the
ability to handle dynamic load model uncertainties, the proposed model reports a robustness degree of 98.7% compared
against 16% for the conventional model. In terms of the economic operation, for single contingency case, the increased
cost by the proposed model over the conventional model is
0.89%, and for the multi-contingency case, the cost increment
is 2.98%, which are quite limited in the context of stability
enhancement. The computational efficiency of the proposed
approach is very high due to the use of only a small number of testing scenarios and the EEAC-based decomposition
structure.
On the other hand, existing TSCOPF methods are deterministic, i.e., the system is either stable or unstable. Besides,
the probabilistic nature of the contingencies is not modelled
which may result in a costly solution. In the future, it would be
interesting to investigate this problem based on a probabilistic
basis or risk-based criterion.

R EFERENCES
[1] D. Gan, R. J. Thomas, and R. D. Zimmerman, Stability-constrained
optimal power flow, IEEE Trans. Power Syst., vol. 15, no. 2,
pp. 535540, May 2000.
[2] L. Chen, Y. Taka, H. Okamoto, R. Tanabe, and A. Ono, Optimal operation solutions of power systems with transient stability constraints,
IEEE Trans. Circuits Syst. I, Fundam. Theory Appl., vol. 48, no. 3,
pp. 327339, Mar. 2001.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
10

[3] Y. Yuan, J. Kubokawa, and H. Sasaki, A solution of optimal power


flow with multicontingency transient stability constraints, IEEE Trans.
Power Syst., vol. 18, no. 3, pp. 10941102, Aug. 2003.
[4] Q. Jiang and Z. Huang, An enhanced numerical discretization method
for transient stability constrained optimal power flow, IEEE Trans.
Power Syst., vol. 25, no. 4, pp. 17901797, Nov. 2010.
[5] T. B. Nguyen and M. A. Pai, Dynamic security-constrained rescheduling of power systems using trajectory sensitivities, IEEE Trans. Power
Syst., vol. 18, no. 2, pp. 848854, May 2003.
[6] D. Ruiz-Vega and M. Pavella, A comprehensive approach to transient
stability control. I. Near optimal preventive control, IEEE Trans. Power
Syst., vol. 18, no. 4, pp. 14461453, Nov. 2003.
[7] A. Pizano-Martinez, C. R. Fuerte-Esquivel, and D. Ruiz-Vega, A new
practical approach to transient stability-constrained optimal power flow,
IEEE Trans. Power Syst., vol. 26, no. 3, pp. 16861696, Aug. 2011.
[8] A. Pizano-Martinez, C. R. Fuerte-Esquivel, and D. Ruiz-Vega, Global
transient stability-constrained optimal power flow using an OMIB reference trajectory, IEEE Trans. Power Syst., vol. 25, no. 1, pp. 392403,
Feb. 2010.
[9] R. Zrate-Miano, T. Van Cutsem, F. Milano, and A. J. Conejo,
Securing transient stability using time-domain simulations within
an optimal power flow, IEEE Trans. Power Syst., vol. 25, no. 1,
pp. 243253, Feb. 2010.
[10] X. Tu, L.-A. Dessaint, and I. Kamwa, Fast approach for transient stability constrained optimal power flow based on dynamic reduction method,
IET Gener. Transm. Distrib., vol. 8, no. 7, pp. 12931305, Jul. 2014.
[11] N. Mo, Z. Y. Zou, K. W. Chan, and T. Y. G. Pong, Transient stability constrained optimal power flow using particle swarm optimisation,
IET Gener. Transm. Distrib., vol. 1, no. 3, pp. 476483, May 2007.
[12] H. R. Cai, C. Y. Chung, and K. P. Wong, Application of differential evolution algorithm for transient stability constrained optimal power flow,
IEEE Trans. Power Syst., vol. 23, no. 2, pp. 719728, May 2008.
[13] Y. Xu, Z. Y. Dong, K. Meng, J. H. Zhao, and K. P. Wong, A hybrid
method for transient stability-constrained optimal power flow computation, IEEE Trans. Power Syst., vol. 27, no. 4, pp. 17691777,
Nov. 2012.
[14] I. Genc, R. Diao, V. Vittal, S. Kolluri, and S. Mandal, Decision treebased preventive and corrective control applications for dynamic security
enhancement in power systems, IEEE Trans. Power Syst., vol. 25, no. 3,
pp. 16111619, Aug. 2010.
[15] Y. Xu et al., Preventive dynamic security control of power systems
based on pattern discovery technique, IEEE Trans. Power Syst., vol. 27,
no. 3, pp. 12361244, Aug. 2012.
[16] Y. Xu, Z. Y. Dong, Z. Xu, R. Zhang, and K. P. Wong, Power system transient stability-constrained optimal power flow: A comprehensive
review, in Proc. IEEE PES Gen. Meeting, San Diego, CA, USA,
Jul. 2012, pp. 17.
[17] J. V. Milanovic, K. Yamashita, S. M. Villanueva, S. Z. Djokic, and
L. M. Korunovic, International industry practice on power system load
modeling, IEEE Trans. Power Syst., vol. 28, no. 3, pp. 30383046,
Aug. 2013.
[18] Y. Xu et al., Multi-objective dynamic VAR planning against short-term
voltage instability using a decomposition-based evolutionary algorithm,
IEEE Trans. Power Syst., vol. 29, no. 6, pp. 28132822, Nov. 2014.
[19] D. N. Kosterev, C. W. Taylor, and W. A. Mittelstadt, Model validation
for the August 10, 1996 WSCC system outage, IEEE Trans. Power
Syst., vol. 14, no. 3, pp. 967979, Aug. 1999.
[20] I. A. Hiskens and J. Alseddiqui, Sensitivity, approximation, and uncertainty in power system dynamic simulation, IEEE Trans. Power Syst.,
vol. 21, no. 4, pp. 18081820, Nov. 2006.
[21] PSSE 33.0 Program Application Guide: Volume-II, Siemens Power
Technol. Int., Schenectady, NY, USA, May 2011, pp. 2432.
[22] Q. P. Zhang, J. Wang, and A. L. Liu, Stochastic optimization for
unit commitmentA review, IEEE Trans. Power Syst., vol. 30, no. 4,
pp. 19131924, Jul. 2015.
[23] G. S. Peace, Taguchi Methods: A Hand on Approach. Reading, MA,
USA: Addison-Wesley, 1993.
[24] Q. Wu, On the optimality of orthogonal experimental design,
Acta Math. Applagatae Sinica, vol. 1, no. 4, pp. 283299, Nov. 1978.
[25] H. Yu, C. Y. Chung, and K. P. Wong, Robust transmission network
expansion planning method with Taguchis orthogonal array testing,
IEEE Trans. Power Syst., vol. 26, no. 3, pp. 15731580, Aug. 2011.
[26] H. Yu and W. D. Rosehart, An optimal power flow algorithm to
achieve robust operation considering load and renewable generation
uncertainties, IEEE Trans. Power Syst., vol. 27, no. 4, pp. 18081817,
Nov. 2012.

IEEE TRANSACTIONS ON SMART GRID

[27] (Aug. 2015). Orthogonal Arrays (Taguchi Designs). [Online]. Available:


http://www.york.ac.uk/depts/maths/tables/orthogonal.htm
[28] Y. Xue, T. Van Cutsem, and M. Ribbens-Pavella, A simple direct
method for fast transient stability assessment of large power systems,
IEEE Trans. Power Syst., vol. 3, no. 2, pp. 400412, May 1988.
[29] Y. Xue, Integrated extended equal area criterionTheory and application, in Proc. 5th Symp. Spec. Elect. Oper. Expansion Plan., Recife,
Brazil, 1996, pp. 16.
[30] M. Pavella, D. Ernst, and D. Ruiz-Vega, Transient Stability of Power
Systems: A Unified Approach to Assessment and Control. Norwell, MA,
USA: Kluwer, 2000.
[31] Y. Xue, Fast analysis of stability using EEAC and simulation technologies, in Proc. Int. Conf. Power Syst. Tech., Beijing, China, 1998,
pp. 1216.
[32] J. Ma, D. Han, R.-M. He, Z.-Y. Dong, and D. J. Hill, Reducing identified parameters of measurement-based composite load model, IEEE
Trans. Power Syst., vol. 23, no. 1, pp. 7683, Feb. 2008.
[33] R. He, J. Ma, and D. J. Hill, Composite load modeling via measurement approach, IEEE Trans. Power Syst., vol. 21, no. 2, pp. 663672,
May 2006.
[34] Z. Guo, Z. J. Wang, and A. Kashani, Home appliance load modeling
from aggregated smart meter data, IEEE Trans. Power Syst., vol. 30,
no. 1, pp. 254262, Jan. 2015.
[35] K. Meng, Z. Y. Dong, K. P. Wong, Y. Xu, and F. J. Luo, Speed-up
the computing efficiency of power system simulator for engineeringbased power system transient stability simulations, IET Gener. Transm.
Distrib., vol. 4, no. 5, pp. 652661, May 2010.
[36] R. D. Zimmerman, C. E. Murillo-Snchez, and R. J. Thomas,
MATPOWER: Steady-state operations, planning, and analysis tools
for power systems research and education, IEEE Trans. Power Syst.,
vol. 26, no. 1, pp. 1219, Feb. 2011.
[37] D. Han, J. Ma, R.-M. He, and Z.-Y. Dong, A real application of
measurement-based load modeling in large-scale power grids and its
validation, IEEE Trans. Power Syst., vol. 24, no. 4, pp. 17561764,
Nov. 2009.
[38] (Aug. 2015). DSATools Dynamic Security Assessment Software. [Online].
Available: http://www.dsatools.com/

Yan Xu (S10M13) received the B.E. and M.E. degrees from the South
China University of Technology, Guangzhou, China, in 2008 and 2011, respectively, and the Ph.D. degree from the University of Newcastle, Australia, in
2013. He is currently a University Postdoctoral Fellow with the School of
Electrical and Information Engineering, University of Sydney, Australia. He
was a Research Fellow at the Center for Intelligent Electricity Networks,
University of Newcastle, from 2013 to 2014. His research interests include
power system stability and control, power system planning, smart grid, and
intelligent system applications to power engineering.

Jin Ma (M06) received the B.S. and M.S. degrees from Zhejiang University,
Hangzhou, China, in 1997 and 2000, respectively, and the Ph.D. degree from
Tsinghua University, Beijing, China, in 2004, all in electrical engineering.
From 2004 to 2013, he was a Faculty Member of North China Electric
Power University. Since 2013, he has been with the School of Electrical
and Information Engineering, University of Sydney. His major research interests are load modeling, nonlinear control system, dynamic power system, and
power system economics. He is a Member of CIGRE W.G. C4.605 Modeling
and Aggregation of Loads in Flexible Power Networks, and a Corresponding
Member of CIGRE Joint Workgroup C4-C6/CIRED Modeling and Dynamic
Performance of Inverter Based Generation in Power System Transmission and
Distribution Studies. He is a registered Chartered Engineer in the U.K.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
XU et al.: ROBUST TSCOPF WITH UNCERTAIN DYNAMIC LOADS

Zhao Yang Dong (M99SM06) received the Ph.D. degree from the
University of Sydney, Australia, in 1999. He is currently a Professor
and the Head of the School of Electrical and Information Engineering,
University of Sydney. He was the Ausgrid Chair and the Director of
the Centre for Intelligent Electricity Networks, University of Newcastle,
Australia. He also held academic and industrial positions with the Hong Kong
Polytechnic University and Transend Networks (now TasNetworks), Australia.
His research interest includes smart grid, power system planning, power
system security, renewable energy systems, electricity market, load modeling, and computational intelligence and its application in power engineering.
He is an Editor of the IEEE T RANSACTIONS ON S MART G RID, the IEEE
T RANSACTIONS ON S USTAINABLE E NERGY, IEEE P OWER E NGINEERING
L ETTERS, and IET Renewable Power Generation.

11

David J. Hill (S72M76F93LF14) received the B.E. degree in electrical engineering and the B.Sc. degree in mathematics from the University of
Queensland, Australia, in 1972 and 1974, respectively, and the Ph.D. degree
in electrical engineering from the University of Newcastle, Australia, in 1976.
He is the Chair of Electrical Engineering with the Department of Electrical
and Electronic Engineering, University of Hong Kong. He is the Director
of the Centre for Electrical Energy Systems and the Program Coordinator
for the multiuniversity RGC Theme-Based Research Scheme Project on
Sustainable Power Delivery Structures for High Renewables. He is also a
part-time Professor with the Centre for Future Energy Networks, University
of Sydney, Australia. From 2005 to 2010, he was an Australian Research
Council Federation Fellow at Australian National University. Since 2006, he
has been the Theme Leader and the Deputy Director with the ARC Centre of
Excellence for Mathematics and Statistics of Complex Systems. He has held
various positions at the University of Sydney since 1994, including as the
Chair of Electrical Engineering until 2002 and again from 2010 to 2013, along
with an ARC Professorial Fellowship. He has also held academic and substantial visiting positions at the universities of Melbourne, California (Berkeley),
Newcastle (Australia), Lund (Sweden), Munich, and Hong Kong (City and
Polytechnic). He currently holds Honorary Professorships with the City
University of Hong Kong, South China University of Technology, Wuhan
University, and Northeastern University, China. From 1996 to 1999 and from
2001 to 2004, he served as the Head of the respective departments in Sydney
and Hong Kong. His general research interests are in control systems, complex networks, power systems and stability analysis, control and planning of
future energy networks, and basic stability and control questions for dynamic
networks.
Prof. Hill is a Fellow of the Society for Industrial and Applied
Mathematics, USA; the Australian Academy of Science; and the Australian
Academy of Technological Sciences and Engineering. He is also a Foreign
Member of the Royal Swedish Academy of Engineering Sciences.

You might also like