You are on page 1of 14

JID: JARAP

ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

Annual Reviews in Control 0 0 0 (2016) 1–14

Contents lists available at ScienceDirect

Annual Reviews in Control


journal homepage: www.elsevier.com/locate/arcontrol

Full Length Article

Perspectives in modeling for control of power networks


Arjan van der Schaft a,∗, Tjerk Stegink a,b,c
a
Jan C. Willems Center for Systems and Control, University of Groningen, The Netherlands
b
Johann Bernoulli Institute for Mathematics and Computer Science, University of Groningen, The Netherlands
c
Engineering and Technology Institute Groningen, University of Groningen, The Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: Increasing integration of renewable energy sources in the power grid leads to a complete re-thinking of
Received 19 December 2015 its operation. This necessitates the consideration of new modeling and analysis approaches, which can
Revised 23 February 2016
serve as natural starting points for robust and scalable control and design strategies.
Accepted 8 March 2016
Available online xxx In this ‘vision’ paper we will highlight two topics within the broad area of power networks: the model-
ing and analysis of the synchronous generator, and the modeling and analysis of power networks using
Keywords: the swing equation as an approximate model for the generator. In both cases we will discuss a port-
Synchronous generator
Hamiltonian formulation, which reflects the underlying physics of power flow and energy storage. Al-
Port-Hamiltonian system
Stability
though the port-Hamiltonian model of the synchronous generator reveals a clear structure it still poses
Shifted passivity fundamental challenges for its non-zero steady state stability analysis. It is shown how the swing equa-
Swing equation tion can be directly deduced from the power balance of the port-Hamiltonian model of the synchronous
Networks generator. Under the phasor assumption, this leads to a port-Hamiltonian model of power networks of
generators, which enables a straightforward stability analysis and provides a starting point for control.
© 2016 Published by Elsevier Ltd on behalf of International Federation of Automatic Control.

1. Introduction the operation of the power grid is the concept of a micro-grid. A


micro-grid is generally understood as a small-scale power network
The study of power networks has recently gained much atten- that is able to operate independently of the main (high-voltage)
tion from different angles, including the systems and control point grid, and therefore is less vulnerable to faults and power outages
of view. Fundamental reasons for the growing interest are the in- elsewhere.
creasing share of renewables, such as wind and solar energy, in In the light of these developments it seems relevant to funda-
the electricity production and resulting fundamental challenges in mentally reconsider the basic modeling approaches to power net-
the operation of the network, and at the same time the availabil- works, and to investigate how more accurate models reflecting the
ity of advanced information technology. This has led, or will lead underlying physics, as well as improved analysis and control con-
in the near future, to a re-thinking of the operation of the power cepts, can contribute to novel robust and scalable control method-
grid, replacing the classical top-down structure of energy distribu- ologies. Indeed, while up to now for many of the control problems
tion, from a small number of major power plants to consumers, in the operation of the power network fairly simple models or lin-
by a much more horizontal and distributed structure. In such a earizations of more involved models could be satisfactory, this is
distributed structure many of the consumers are partially or tem- not the case anymore for the control challenges that are posed by
porarily producers of energy as well (called prosumers). The match- the new ways of operating the power grid. In addition, also the
ing of supply and demand in the resulting complex multi-player classical problem of the stability of the large-scale power network
setting, subject to a large degree of uncertainty in the energy pro- after major disruptions (faults) is still largely open.
duction of renewables, leads to an operation of the power grid Instead of providing a general vision on the area, we restrict
that is increasingly often near its capacity constraints and neces- ourselves in this introductory and tutorial paper to two topics
sitates the use of advanced communication and control techniques within the modeling and analysis of power networks. The first
(‘smart grids’). Another important aspect in the restructuring of topic concerns the detailed modeling for control of the synchronous
generator, which is (still) the main working horse of the large
power grid. The synchronous generator constitutes a fascinating

Corresponding author. multi-physics system which can be modeled from first principles
E-mail addresses: a.j.van.der.schaft@rug.nl (A. van der Schaft), t.w.stegink@rug.nl resulting in a classical eight-dimensional model. We will describe
(T. Stegink). the main characteristics and the mathematical structure of this
http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
1367-5788/© 2016 Published by Elsevier Ltd on behalf of International Federation of Automatic Control.

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

2 A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14

model by utilizing the port-Hamiltonian framework. This frame-


work is ideally suited for multi-physics systems and emphasizes
aspects of power flow and energy storage, and therefore provides a
natural setting for the analysis and control of the dynamics of (net-
works of) synchronous generators. Although the structure of the
port-Hamiltonian formulation of the classical eight-dimensional
model is very clear, the analysis and control of this model poses
fundamental challenges. Indeed, while the Hamiltonian (total en-
ergy) is a natural Lyapunov function for stability of the zero equi-
librium, the nonlinearity of the model makes it intrinsically hard to
Fig. 1. The state and port variables of the synchronous generator.
study the stability of steady state values corresponding to non-zero
mechanical torque (the interesting case in applications). In this pa-
⎡ ∂H ⎤
⎡ ⎤ ⎡ ⎤ ∂ψ
per we will discuss some of the available analysis methods, and Is I3 033 031 031 ⎢ ∂ Hs ⎥
⎣I f ⎦ = ⎣013 ⎢ r⎥
end with a number of numerical simulation studies that motivate eT1 0 0 ⎦⎢ ∂ψ ⎥, (1)
further research. ⎣ ∂∂Hp ⎦
The second topic of this paper concerns the modeling of power ω 013 013 1 0
∂H
networks under the assumption of pure sinusoidal currents and ∂θ
voltages of nominal frequency. Under this assumption the cur- where 0lk denotes the l × k zero matrix, I3 denotes the 3 × 3 iden-
rents and voltages can be compactly represented by their phasors, tity matrix, and e1 is the first basis vector of R3 . Eq. (1) is in port-
while the network of linear transmission lines linking the gener- Hamiltonian input-state-output form (van der Schaft, 1996; van der
ators and the loads can be modeled by an admittance matrix. A Schaft & Jeltsema, 2014; van der Schaft & Maschke, 1995)
trait-d’union between the phasor representation of the transmis- ∂H
sion line network and the eight-dimensional model of the syn- x˙ = (J (x ) − R(x )) (x ) + g(x )u,
∂x
chronous generator can be established by a crude approximation of
∂H
the eight-dimensional model by means of the swing equation. We y = gT (x ) ( x ),
present a simple reasoning leading to the swing equation based ∂x
on the power-balance of the port-Hamiltonian model of the eight- J (x ) = −J T (x ), R(x ) = RT (x ) ≥ 0, (2)
dimensional model, which we believe to be fairly novel. It turns with a Poisson structure matrix J(x) given by the constant matrix
out that the resulting system of generators and loads linked by ⎡ ⎤
the network of transmission lines again defines a port-Hamiltonian 066 062
system, which admits a straightforward steady state stability anal- J=⎣ 0 −1 ⎦, (3)
ysis. In particular, the model enjoys the property of shifted passivity 026
1 0
with respect to steady state values, which is an ideal starting point
for robust control. and a resistive structure matrix R(x), which is also constant, having
Disclaimer: This paper is not meant to be an introductory sur- diagonal blocks
vey to modeling and control of power networks, neither a balanced ⎡ ⎤ ⎡ ⎤
rs 0 0 rf 0 0
view on new research directions. Instead it focusses on some of
our personal (and limited) understandings and tastes in this area, Rs = ⎣ 0 rs 0 ⎦, Rr = ⎣ 0 rkd 0 ⎦, d, 0, (4)
and we are sure that others will write about their own. 0 0 rs 0 0 rkq
denoting, respectively, the stator resistances, rotor resistances and
mechanical friction. Here ∂∂Hx (x ) throughout denotes the column
2. Port-Hamiltonian modeling of the synchronous generator
vector of partial derivatives of the function H.
The state variables x of the eight-dimensional model of the syn-
An essential component of the alternating-current power grid
chronous generator comprise of
is the synchronous generator; see e.g. Kundur (1993), Machowski,
Bialek, and Bumby (2008), and Sauer and Pai (1998). Although the • ψs ∈ R3 , the stator fluxes,
contribution of renewables such as solar and wind energy is grow- • ψr ∈ R3 , the rotor fluxes: the first one corresponding to the field
ing, the synchronous generators still are the main working horses winding and the remaining two to the damper windings,
of the large power grid. While less dominant, they are also impor- • p, the angular momentum of the rotor,
tant in micro-grids. • θ , the angle of the rotor.
The standard model for the synchronous generator, as described
Moreover, Vs ∈ R3 , Is ∈ R3 are the three-phase1 stator terminal
e.g. in Kundur (1993), is based on first principles physical model-
voltages and currents, Vf , If are the rotor field winding voltage and
ing. As such it can be naturally written into port-Hamiltonian form
current, and τ , ω are the mechanical torque and angular velocity;
given as, see Fiaz, Zonetti, Ortega, Scherpen, and van der Schaft
see Fig. 1 for a schematic view.
(2013) for details,
The Hamiltonian H (total stored energy of the synchronous gen-
⎡ ⎤ ⎡ ⎤⎡ ∂ H ⎤ erator) is the sum of the magnetic energy of the generator and the
ψ˙ s −Rs 033 031 031 ∂ψs
⎢ ˙ ⎥ ⎢0 ⎢ ∂H ⎥
031 ⎥⎢ ∂ψ
kinetic energy of the rotating rotor, given as the sum of the two
ψ
⎢ r ⎥ ⎢ 33 −Rr 031
⎥⎢ r ⎥ nonnegative terms:
⎢ ⎥=⎣ ⎥
⎣ p˙ ⎦ 013 013 −d −1⎦⎣ ∂∂Hp ⎦

1 T ψs 1 2
θ˙ 013 013 1 0 ∂H H (ψs , ψr , p, θ ) = ψ ψ L (θ )
T −1
+ p
∂θ 2 s r
ψr 2Jr
⎡ ⎤
I3 031 031 ⎡ ⎤ = magnetic energy Hm + kinetic energy Hk , (5)
Vs
⎢033 e1 031 ⎥
+⎢ ⎥⎣V f ⎦
⎣013 0 1 ⎦ 1
τ Note that the vectors Vs , Is are not assumed to be balanced (sum zero); see also
(21) in the next section.
013 0 0

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14 3

where Jr is the rotational inertia of the rotor (not to be confused are the vectors of stator currents and rotor currents respectively,
with the Poisson structure matrix J), and L(θ ) is a 6 × 6 inductance and
matrix.
∂H
In the round rotor case (no saliency; cf. Kundur, 1993; Ma- ω= (10)
chowski et al., 2008)
∂p

is the angular velocity of the rotor. Furthermore,
Lss Lsr (θ )
L (θ ) = (6) ∂ H 1 T ∂ −1
LTsr (θ ) Lrr τm = = ψ L ( θ )ψ (11)
∂θ 2 ∂θ
where
is the air-gap torque, i.e., the mechanical torque acting on the rotor
⎡ ⎤ ⎡ ⎤
Laa −Lab −Lab Lf fd Lakd 0 due to rotating magnetic field.
Lss = ⎣−Lab −Lab ⎦, Lrr = ⎣ Lakd ⎦ It follows from (1) that the torque balance in the mechanical
Laa Lkkd 0 (7)
domain is given by
−Lab −Lab Laa 0 0 Lkkq
p˙ = −dω − τm + τ , (12)
while
⎡ ⎤ where τ is the external torque exerted on the rotor (e.g., by a tur-
Cos θ Cos θ −Sin θ bine), dω is the torque due to friction in the rotational motion of
Lsr (θ ) = ⎣Cos(θ − 2π
3
) Cos(θ − 23π ) −Sin(θ − 23π )⎦ · the rotor (for simplicity assumed to be linear in the angular veloc-
2π ity ω), and τ m is the air-gap torque.
Cos(θ + 3
) Cos(θ + 23π ) −Sin(θ + 23π )
The system, being port-Hamiltonian, satisfies the following
⎡ ⎤
La f d 0 0 power-balance with respect to its three ports
×⎣ 0 Lakd 0 ⎦ (8) d
H = IsT Vs + I f V f + ωτ − IsT Rs Is − IrT Rr Ir − dω2 (13)
0 0 Lakq dt

Remark 2.1. This can be extended to the saliency case, and to sat- Here ωτ defines the mechanical power delivered to the syn-
uration. Saliency corresponds to a more complicated expression for chronous generator (by a prime mover such as a turbine),
L(θ ) (see Kundur, 1993), while saturation can be captured by re- −IsT Vs denotes the electrical power delivered to the network, If Vf
placing the purely quadratic dependence of the magnetic energy is the power delivered by the excitation system, and finally
on the flux variables ψ s , ψ r by a more involved one. IsT Rs Is , IrT Rr Ir , dω2 denote the power losses due to resistance (in the
stator, respectively rotor) and mechanical friction. In particular, the
Remark 2.2. Many synchronous generators, in particular turbo- synchronous generator (1) is passive (van der Schaft, 1996) with
generators, have a solid-steel rotor body which acts as a screen respect to its three ports, with storage function H.
in the q-axis. This effect is often taken into account by an ad- In normal circumstances, the magnitude of the terms
ditional damper winding along the q-axis, leading to a nine- ωτ and −IsT Vs will be large as compared to the terms
dimensional model (Machowski et al., 2008) and corresponding I f V f , IsT Rs Is , IrT Rr Ir , dω2 . In fact, in steady-state operation one
port-Hamiltonian formulation (Stegink, De Persis, & van der Schaft, would expect that ωτ ≈ −IsT Vs (the ideal scenario being that all
in preparation). mechanical power is converted into electrical power).
Although typically the power at the excitation port (Vf , If ) will
A crucial feature of the magnetic energy term Hm in the Hamil- be small, its presence is crucial for the magnetization of the ro-
tonian H is its dependency on the mechanical rotor angle θ ; see tor and thus for the functioning of the synchronous generator; see
the formula (8) for Lsr (θ ). This dependence is responsible for the also Section 4.1.1. The excitation port is also important for control
interaction between the mechanical domain of the generator (the purposes, in particular for regulating the voltage amplitude at the
mechanical motion of the rotor) and the electro-magnetic do- stator terminals.
main (the dynamics of the magnetic fields in the rotor and sta- We finally note that, as is commonly the case for physical mod-
tor), and thus for the functioning of the synchronous generator as els, the model (1) is bi-directional. In particular, it also models a
an energy-conversion device, transforming mechanical power into synchronous motor. In this case, IsT Vs would be the electrical power
electrical power. delivered to the motor, and −ωτ would be the mechanical power
The synchronous generator is a fascinating case of a multi- delivered by the motor. Thus the above port-Hamiltonian model
physics system, connected to its environment by three types of (1) is really the model of a synchronous machine, covering both syn-
ports; see Fig. 1. First, there is the scalar mechanical port with chronous generators and motors.
power variables τ , ω, to be interconnected to a prime mover, such
as a turbine. This port is also used for control purposes, e.g., via so-
called droop control. Second, there are three stator terminal ports, 3. Analysis of the synchronous generator by the Blondel–Park
with vectors of power variables Vs , Is , which denote the vectors of transformation
three-phase voltages and currents to be delivered to the electrical
network. Third, there is the port with scalar power variables Vf , If , The dynamics of the eight-dimensional model (1) is difficult to
which is responsible for the magnetization of the rotor, and which analyze; especially for non-zero mechanical torque τ (from an ap-
is controlled by the excitation system (Kundur, 1993; Machowski plication perspective the only interesting situation). A main rea-
et al., 2008; Pai, 1989). son for this is the same fact that is responsible for the func-
Note that as always in a port-Hamiltonian formulation the par- tioning of the synchronous
  generator: the magnetic energy Hm =
ψs is depending on the mechanical angle θ .
tial derivatives of the Hamiltonian H have a direct physical inter- ψsT ψrT L−1 (θ )
pretation (Fiaz et al., 2013; van der Schaft, 1996; van der Schaft & ψr
Jeltsema, 2014; van der Schaft & Maschke, 1995). In fact, There is a way, however, to eliminate the dependence of
the magnetic energy on θ , by transforming the stator flux vari-
∂H ∂H ables ψ s depending on θ . This is known as the Blondel–Park
Is = , Ir = (9)
∂ψs ∂ψr or dq0 transformation, defined as the coordinate transformation

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

4 A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14

⎡ ⎤
(Fiaz et al., 2013; Kundur, 1993; Machowski et al., 2008) I3 031 031 ⎡ ⎤
⎢0 Vdq0
⎡ ⎤
ψd ⎢ 33 e1 031 ⎥

+ ⎢ ⎥⎣ V f ⎦
ψdq0 = ψq ⎦ := Tdq0 (θ )ψs
⎣ ⎣013 0 1 ⎦
τ
ψ0 013 0 0
⎡ ⎤ ⎡ ∂H ⎤
 Cos θ Cos(θ − 23π ) Cos(θ + 2π
3
) ⎡ ⎤ ⎡ ⎤ ∂ψdq0
2⎢ Idq0 I3 033 031 031⎢ ∂H ⎥
:= 2π
⎣ Sin θ Sin(θ − 3 ) Sin(θ + 2π
)⎥
⎦ψs (14) ⎣ I f ⎦=⎣013 ⎢ ⎥
3 3 eT1 0 0 ⎦⎢ ∂ψr ⎥ (18)
√1 √1 √1 ⎣ ∂∂Hp ⎦
2 2 2 ω 013 013 1 0
0
Then the vector Vs transforms in the same way to Vdq0 :=
Tdq0 (θ )Vs , while the vector of currents Is transforms to Idq0 := Here the resistive matrix Rs is the same as before (due to the or-
T−T
dq0
(θ )Is , implying that IsT Vs = Idq
T V
0 dq0
; and thus power is thonormality of Tdq0 ), but the constant Poisson structure matrix J
preserved. given by (3) has been transformed into the state dependent Poisson
Since the Blondel–Park transformation satisfies2 the orthonor- structure
mality property ⎡ ⎡ ⎤ ⎤
−ψq
T−1 (θ ) = TTdq0 (θ ), (15) ⎢ ⎣ ψd ⎦ 031 ⎥
dq0 ⎢ 033 033 ⎥
⎢ ⎥
⎢ 0 ⎥
J (xdq0 ) := ⎢
031 ⎥
Is and Vs actually transform in the same way; that is, Idq0 = (19)
Tdq0 (θ )Is , Vdq0 := Tdq0 (θ )Vs . ⎢ 033 033 031 ⎥
⎢ ⎥
Note that I0 and V0 measure the ‘unbalance’ in the three-phase ⎣ ψq −ψd 0 013 0 −1⎦
current and voltage; in ideal operation I0 and V0 are zero.
013 013 1 0
The Hamiltonian H transforms to a Hamiltonian H expressed in
the new coordinates xdq0 := (ψ dq0 , ψ r , p, θ ), whose magnetic part Hence, in the new coordinates the expression ∂∂θH equals zero (no
is given by dependence on θ ), but on the other hand, the air-gap torque τ m

now shows up as the product of the seventh row of the trans-
1 T ψdq0
Hm = ψ ψ L T
r
−1
, (16) formed Poisson structure matrix J (xdq0 ) with ∂∂x H . Indeed, the
2 dq0 ψr dq0
air-gap torque is equivalently given as
where


τm = ψd Iq − ψq Id (20)
Tdq0 (θ ) 0 TTdq0 (θ ) 0
L= L (θ ) Thus from a mathematical point of view we have reduced the com-
0 I3 0 I3 plexity of the expression of the magnetic energy, however sacrific-
⎡   ⎤
Ld 0 0 3
L 3
L 0 ing the simplicity of the Poisson structure matrix.
2 afd 2 akd
⎢ 3 ⎥ An important advantage of the transformed model (18) is the
⎢ 0 Lq 0 0 0 − L ⎥ fact that we can split off the dynamics of θ , and consider steady
⎢ 2 akq ⎥
⎢ 0 0 L0 0 0 0 ⎥ states of the reduced seven-dimensional model, instead of periodic
⎢ ⎥
= ⎢ ⎥, motions of the original eight-dimensional model. Furthermore, it is
⎢ 32 La f d 0 0 Lf fd Lakd 0 ⎥ clear that the dynamics of ψ 0 is given by
⎢ ⎥
⎢ 3 ⎥
⎣ 2 Lakd 0 0 Lakd Lkkd 0 ⎦ ∂H
3 ψ˙ 0 = −rs + V0 , (21)
0 − L
2 akq
0 0 0 Lkkq ∂ψ0
(17) which in view of the form of L in (17) is independent of the rest
of the dynamics. Moreover, for a constant V0 , the dynamics (21) is
and we conclude that L is independent of θ . Thus the Blondel–
asymptotically stable if and only if rs > 0.
Park transformation results in an expression of the magnetic en-
ergy Hm , given by (16), which is independent of the mechanical
Remark 3.1. Notice that L0 ≈ 0 as stated on Kundur (1993, p. 65).
angle θ .
In fact, if the so-called leakage flux not crossing the air-gap would
On the other hand, the elimination of θ from the magnetic
be neglected then L0 = 0. In this case (21) reduces to the algebraic
energy expression Hm has other consequences. In fact, in the
equation −rs I0 + V0 = 0.
new coordinates xdq0 = (ψdq0 , ψr , p, θ ) the eight-dimensional sys-
tem takes the form (Fiaz et al., 2013) Hence for analysis purposes we can reduce the model (18) to
⎡ ⎡ ⎤ ⎤ the six-dimensional port-Hamiltonian system
−ψq
⎡ ˙ ⎤ ⎡ ∂H ⎤
ψdq0 ⎢ −Rs ⎣ ψd ⎦ 0⎥
⎡ ˙ ⎤ ⎡

ψd
⎤⎡ ∂ H ⎤
⎢ 0 ⎥ ∂ψdq0 rs 0 −ψq ∂ψ
⎢ ψ˙ ⎥ ⎢ ⎥⎢ ∂ H ⎥ ⎢ψ˙ ⎥ ⎢ − 023 ⎢ ∂ Hd ⎥
⎢ r ⎥ ⎢ 0 ⎥⎢ ∂ψr ⎥ ⎢ q⎥ ⎢ ψd ⎥ ⎢ ⎥
⎢ ⎥=⎢ −Rr 0⎥⎢ ⎥ ⎢ ⎥=⎣
0 rs ⎥⎢ ∂ψq ⎥
⎣ p˙ ⎦ ⎢ 0

0 ⎥⎣ ∂∂Hp ⎦ ⎣ ψ˙ r ⎦ 032 −Rr 031 ⎦⎢ ∂ H ⎥
⎢ ⎥  ⎣ ∂ψr ⎦
θ˙ ⎣ ψq −ψd 0 0 −d −1⎦ 0 ψ −ψd 013 −d 
p˙ q ∂H
0 0 1 0 ∂p
⎡ ⎤⎡ ⎤
I2 021 021 Vdq
+ ⎣032 e1 031 ⎦⎣ V f ⎦
2
Actually, the choice of the constants in (14) is dictated by this orthonormality
property.
012 0 1 τ

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14 5

⎡ ∂ H ⎤
following port-Hamiltonian system
⎡ ⎤ ⎡ ⎤ ∂ψd
021 ⎢ ∂ H ⎥
⎤⎡ ∂ Haug ⎤
Idq I2 023 ⎡ ˙ ⎤ ⎡

⎢ ∂ψq ⎥ ψd

⎣ I f ⎦ = ⎣012 eT1 0 ⎦⎢ ⎥, (22) rs 0 −ψq ∂ψd


⎢ ∂ H ⎥ ⎢ψ˙ ⎥ ⎢ − 023 ⎢ ∂ Haug ⎥
ω 012 013 1 ⎣ ∂ψr ⎦ ⎢ q⎥ ⎢ 0 rs ψd ⎥⎢ ∂ψq ⎥
⎥ ⎢


∂H ⎢ ⎥=⎣ 032 −Rr 031 ⎦⎢ ∂ Haug ⎥
∂p ⎣ ψ˙ r ⎦  ⎣ ∂ψr ⎦
p˙ ψ q −ψd 013 −d aug
∂H
where Vdq , Idq ∈ R2 are the first two components of Vdq0 , respec- ∂p
tively Idq0 . Furthermore, the Hamiltonian H  of the six-dimensional ⎡ ⎤⎡ ⎤
I2 021 021 Vdq − V̄dq
port-Hamiltonian system is equal to the Hamiltonian H of the orig-
⎢ ⎥
inal eight-dimensional model minus the term 21L ψ02 . + ⎣032 e1 031 ⎦⎣ V f − V̄ f ⎦
0
Note that also the six-dimensional model is passive with respect 012 0 1 τ
to the inputs Vf , Vdq , τ and outputs If , Idq , ω, with storage function ⎡ ∂ Haug ⎤
given by H . In fact, the following power balance is derived from ⎡ ⎤ ⎡ ⎤ ∂ψd
(13) Idq − I¯dq I2 023 021 ⎢ ∂ Haug ⎥
⎢ ⎥ ⎢ ∂ψq ⎥
⎣ I f − I¯f ⎦ = ⎣012 eT1 0 ⎦⎢ ⎥
⎢ ∂ Haug ⎥, (27)
d  T
H = Idq Vdq + I f V f + ωτ − rs Idq
T
Idq − IrT Rr Ir − dω2 (23)
ω 012 0 1 ⎣ ∂ψ r

dt 13
aug
∂H
∂p
Hence, for zero inputs V f = 0, Vdq = 0, τ = 0 the zero equilibrium
ψd = ψq = 0, ψr = 0, p = 0 is globally asymptotically stable. where I¯dq , I¯f are the steady-state values. This immediately im-
This can be extended to shifted passivity with respect to plies that the system is passive with respect to the shifted inputs
any constant V̄ f , V̄dq and τ̄ = 0 as follows. The standard method Vdq − V̄dq , V f − V̄ f and shifted outputs Idq − I¯dq , I f − I¯f , and the orig-
(Maschke, Ortega, & van der Schaft, 20 0 0; van der Schaft, 1996; inal input τ and output ω at the mechanical port, that is
van der Schaft & Jeltsema, 2014) of studying shifted passivity for
d 
a general port-Hamiltonian system (2) is based on adding to its Haug ≤ (Idq − I¯dq )T (Vdq − V̄dq ) + (I f − I¯f )(V f − V̄ f ) + ωτ
Hamiltonian H a function determined by the input matrix g(x) and dt
the complete structure matrix J (x ) − R(x ), provided additional in- This also implies that for Vdq = V̄dq , V f = V̄ f and τ = 0 the steady
tegrability conditions are satisfied (Fiaz et al., 2013; Maschke et al., state is globally asymptotically stable.
20 0 0). Unfortunately, as shown in Fiaz et al. (2013), this analysis can-
In case of constant matrices J − R and g (not depending on x) not be extended to non-zero constant τ̄ , since in this case the in-
these integrability conditions are automatically satisfied, and the tegrability conditions are not satisfied for rs = 0. Hence, the case of
method is equivalent to considering the shifted Hamiltonian (some- non-zero τ̄ (the case of interest for applications) remains open. In
times called Bregman or availability function) the next Section 3.1 we will study the steady-state stability of the
six-dimensional model for non-zero constant τ̄ , V̄ f , V̄dq by an al-
(x ) := H (x ) − ∂T H
H (x̄ )(x − x̄ ) − H (x̄ ), (24) ternative, partial, approach based on Caliskan and Tabuada (2014).
∂x Furthermore, in Section 4 we will discuss the stability of the syn-
with respect to a steady state x̄ corresponding to ū, that is chronous generator coupled to a linear resistive load for non-zero
constant τ̄ , V̄ f .
∂H
(J − R ) (x̄ ) + gū = 0
∂x 3.1. The shifted Hamiltonian as storage function for non-zero
Then constant torque

d  
∂T H 
∂H As concluded above, the integrability conditions for shifted pas-
H=− (x )R (x ) + (y − ȳ )T (u − ū )
dt ∂x ∂x sivity of the six-dimensional port-Hamiltonian system (22) for
≤ (y − ȳ )T (u − ū ), (25) non-zero constant τ̄ are not satisfied. In this subsection we
will discuss another approach, utilized in Caliskan and Tabuada
where ȳ = gT ∂∂Hx (x̄ ) is the corresponding steady state output value. (2014) for a simplified model of the synchronous generator, which
In case the original Hamiltonian H is convex, the shifted Hamil- starts from the definition of the shifted Hamiltonian as in (24), and
tonian H˜ will have a minimum at x̄, thus proving passivity with then derives additional conditions under which this shifted Hamil-
respect to the shifted inputs u − ū and outputs y − ȳ, with storage tonian serves as a storage function guaranteeing passivity with re-
function H. This is called shifted passivity. spect to the shifted inputs and outputs.
Although the Poisson structure matrix in (22) is state- Denote for compactness of notation
dependent, the integrability conditions mentioned above turn out
s = (ψd , ψq , ψr , p), u = (Vdq , V f , τ ), y = (Idq , I f , ω )
to be satisfied for any constant V̄dq , V̄ f and τ̄ = 0 (Fiaz et al., 2013).
In fact, continuing on Fiaz et al. (2013), we can define the aug- Let ū be a constant input, and consider any corresponding steady-
mented Hamiltonian state value s̄ of the system (22), resulting in a steady state output
ȳ. Exploiting, as in Caliskan and Tabuada (2014), the linear depen-
H (ψdq , ψr , p) − 1 V̄ f ψ f − 1 V̄ T ψdq
aug (ψdq , ψr , p) := H (26) dence of the Poisson structure matrix in (22) on ψ d , ψ q , neces-
rf rs dq
sary and sufficient conditions on the system parameters for shifted
(with ψ f the first component of the rotor flux vector ψ r ), which passivity of (22) with respect to the shifted Hamiltonian3 can be
obtained as follows.
has a minimum at the steady state values (ψ̄dq , ψ̄r , p̄), and is in
fact, up to a constant, equal to the shifted Hamiltonian derived
from H  as in (24). With the aid of this augmented Hamiltonian 3
Note that these conditions may not be necessary for shifted passivity with re-

Haug the six-dimensional model (22) can then be rewritten as the spect to another storage function.

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

6 A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14

 corresponding to
As in (24), define the shifted Hamiltonian H
the Hamiltonian H of the six-dimensional model (22) by

(s ) := H
H (s ) − (∇ H
(s̄ ))T (s − s̄ ) − H
(s̄ ), (28)

where we are using the shorthand notation ∇ H for ∂∂Hx .


Denote furthermore the 6 × 6 Poisson structure matrix of
(22) by J(s ), the resistive structure matrix by R
, and the input
matrix by 
g. Then Fig. 2. The synchronous generator SG coupled to a linear resistive load rl .

 (s ) − ∇ H
˙ = (∇ H
H (s̄ ))T ([J(s ) − R (s ) + 
]∇ H gū + g(u − ū ))
 function, if
(s ) − ∇ H
= (∇ H (s̄ )) [J(s ) − R
T ]∇ H
(s ) − [J(s̄ )
 ⎡ ⎤
−R (s̄ ) + 
]∇ H g(u − ū ) −2rs 0 0 0 kω̄Lakq −Ld I¯q
 ⎢ 0 ⎥
(s ) − ∇ H
= (∇ H (s̄ )) J(s − s̄ )∇ H
T (∇ H
(s̄ ) − R (s ) −2rs kω̄La f d kω̄Lakd Lq I¯d
⎢ 0 ⎥
 ⎢ ⎥
(s̄ )) + 
− ∇H g(u − ū ) (29) ⎢ 0 kω̄La f d −2r f 0 0 −kI¯q La f d ⎥
⎢ ⎥
⎢ 0 kω̄Lakd −2rkd −kI¯q Lakd ⎥
⎢ 0 0 ⎥
where we have used, guided by Caliskan and Tabuada (2014), ⎢ ⎥
⎣kω̄Lakq 0 0 0 −2rkq −kI¯d Lakq ⎦
(s ) − ∇ H
(∇ H (s ) − J(s̄ )∇ H
(s̄ ))T (J(s )∇ H (s̄ )) −Ld I¯q Lq I¯d −kI¯q La f d −kI¯q Lakd −kI¯d Lakq −2d
(s ) − ∇ H
= (∇ H (s ) − J(s )∇ H
(s̄ ))T (J(s )∇ H (s̄ ) <0 (30)

(s̄ ) − J(s̄ )∇ H
+ J(s )∇ H (s̄ )) Corollary 3.3. Let u = ū with steady state value s̄. If (30) holds then
(s ) − ∇ H
(s̄ ))T (J(s )∇ H
(s̄ ) − J(s̄ )∇ H
(s̄ )) s̄ is a globally asymptotically stable steady state of (22).
= (∇ H
(s ) − ∇ H
= (∇ H (s̄ ))T J(s − s̄ )∇ H
(s̄ ), Basically, the condition (30) is satisfied if the damping constant
d and the resistances rs , rf , rkd , rkq are large enough and the steady
exploiting skew-symmetry of J. state s̄ is located in a sufficiently small neighborhood around the
Define  Id := Id − I¯d and use likewise notation for the remaining origin.
variables with respect to their steady state values. Then the term
(s ) − ∇ H
(∇ H (s̄ ))T J(s − s̄ )∇ H
(s̄ ) in the last line of (29) can be
written in the form
4. Stability analysis of the synchronous generator coupled to a
⎡  ⎤T ⎡ 
⎤⎡ ⎤ linear load
Id 0 0 0 0 0 −ψ q I¯d
⎢ ⎥ ⎢  ⎥⎢ I¯q ⎥
⎥⎢
⎢ Iq ⎥ ⎢0 0 0 0 0 ψ d ⎥ By far the simplest scenario of a power network containing syn-
⎢ ⎥ ⎢ ⎥⎢ ⎥ chronous generators is the case of a single synchronous genera-
⎢ If ⎥ ⎢0 0 0 0 0 0 ⎥⎢ I¯f ⎥
⎢ ⎥ ⎢ ⎥⎢ ⎥ tor connected via its stator terminals to a three-phase linear re-
⎢ ⎥ ⎢0 0 ⎥ ⎢ ⎥
⎢Ikd ⎥ ⎢ 0 0 0 0 ⎥⎢I¯kd ⎥ sistive load, with identical resistances (rl , rl , rl ), cf. Fig. 2. We will
⎢ ⎥ ⎢ ⎥⎢ ⎥ see that already in this simplest case a full dynamical analysis be-
⎣Ikq ⎦ ⎣0 0 0 0 0 0 ⎦⎣I¯kq ⎦
comes rather complicated for non-zero torque τ . Substitution of
ω
 
ψ 
−ψ 0 0 0 0 ω̄ the characteristics of the three-phase linear load in (22), using the
q d
orthonormality of Tdq0 in (15), leads to the following equations (al-
= −  ω̄ + 
Id ψ  ω̄ + ω
Iq ψ  (ψ  I¯ )
 I¯ − ψ
q d q d d q most identical to (22))

 = L
Since ψ I, this term can be further rewritten as ⎡ ˙ ⎤ ⎡

ψd
⎤⎡ ∂ H ⎤
r 0 −ψq ∂ψ
⎢ψ˙ ⎥ ⎢ − 023 ⎢ ∂ Hd ⎥
⎡ ⎤T⎡ ⎤⎡ ⎤ ⎢ q⎥ ⎢ 0 r ψd ⎥⎢ ∂ψq ⎥
⎥⎢

I˜d kω̄Lakq −Ld I¯q I˜d
⎢ ⎥=⎣ 0 −R 0 31 ⎦⎢ ∂ H⎥
0 0 0 0 ⎣ ψ˙ r ⎦ 
32

r
⎣ ∂ψr ⎦
⎢ ˜ ⎥⎢ ⎥⎢ ⎥
⎢ Iq ⎥ ⎢ 0 0 kω̄La f d kω̄Lakd 0 Lq I¯d ⎥⎢ I˜q ⎥ ψq −ψd 013 −d ∂H 
⎢ ⎥⎢ ⎥⎢ ⎥ p˙ ∂p
⎢ I˜ ⎥ ⎢ 0 kω̄La f d −kI¯q La f d ⎥ ⎢ ⎥
1⎢ f ⎥⎢ 0 0 0 ⎥⎢ I˜f ⎥ ⎡ ⎤
⎢ ⎥⎢ ⎥⎢ ⎥ 021 021

2 ⎢I˜ ⎥ ⎢ 0 kω̄Lakd 0 0 0 −kI¯q Lakd ⎥⎢I˜kd ⎥
⎢ kd ⎥ ⎢ ⎥⎢ ⎥ Vf
⎢˜ ⎥ ⎢ ⎥⎢ ⎥ + ⎣ e1 031 ⎦
⎣Ikq ⎦ ⎣kω̄Lakq 0 0 0 0 −kI¯d Lakq ⎦⎣I˜kq ⎦
τ
ω˜ −Ld I¯q Lq I¯d −kI¯q La f d −kI¯q Lakd −kI¯d Lakq 0 ω˜ 0 1
⎡ ∂ H ⎤
 ∂ψ


⎢ ∂ H ⎥
d
3
where k = 2. By substitution into (29) we obtain the following If 012 eT1 0 ⎢ ∂ψ ⎥
= ⎢ q ⎥, (31)
result.
ω 012 013 ⎢ ∂ H⎥
1 ⎣ ∂ψ ⎦
r

∂H
Proposition 3.2. Consider a constant ū, and a corresponding steady ∂p
state value s̄ of the system (22), resulting in a steady state output
ȳ. Then the system (22) is passive with respect to the shifted inputs with r := rs + rl . Furthermore, consider constant values τ̄ of the
 as storage
and outputs u − ū, y − ȳ, using the shifted Hamiltonian H mechanical torque and V̄ f of the excitation voltage. The resulting

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14 7

steady state values of (31), denoted by ¯, are given as Among others, this shows the (physically obvious) fact that a

¯   nonzero excitation voltage V̄ f is required for a nonzero steady-state
Id −ψ̄q current.
0 = −r + ω̄
I¯q ψ̄d
⎡ ⎤ 4.1.2. Desired equilibria
V̄ f Now consider the problem of designing the constant mechani-
0 = −Rr I¯r + ⎣ 0 ⎦ cal torque τ̄ and excitation voltage V̄ f in such a way that the steady
state frequency ω̄ is equal to a desired value ωdes , and that in ad-
0
dition the voltage magnitude V̄d2 + V̄q2 is equal to a desired value
0 = −τ̄m − dω̄ + τ̄ , (32)
Vdes . It can be noted, cf. Fig. 2, that Vl = rl Idq , where Vl is the volt-
while in view of (21) I¯0 = 0. As proved in Fiaz et al. (2013), given
age across the load. Hence, it should hold that
a steady state value ω̄ of the frequency ω the steady state values 
I¯d , I¯q , I¯r , ψ̄dq0 , ψ̄r of the rest of the state variables are uniquely de-
rl I¯d2 + I¯q2 = Vdes . (35)
termined. The complication is therefore in the determination of the
steady state values of ω̄, which are solutions of a polynomial equa- By solving equations (33)–(35) we obtain the following equations
tion of degree 3 as detailed in the next section.4 for the required mechanical torque and excitation voltage:
Note that a direct consequence of this analysis is that steady
τ̄ = dωdes + Vdesr2(ωrldes
+rs )
2
states of (31) correspond to sinusoidal solutions of the original sys-
l√
tem (1), with frequency equal to a steady-state value ω̄.  2 Vdes r f L2 ω2 + ( rl +rs )2
(36)
V̄ f = ± 3
d des
La f d rl ωdes
4.1. Equilibrium frequencies
Suppose that the voltage amplitude over the load is constant and
is equal to Vdes . Then, because of equation (35), a lower load re-
In this subsection we look more carefully at the steady state
sistance rl implies that the current flowing through the load has a
frequencies ω̄. They satisfy the equation (Fiaz et al., 2013)
  larger magnitude and therefore the power consumption of the load
k2 L2a f d rV̄ f2 increases. Observe from equation (36) that a larger power consum-
τ̄ = ω d +   , ing load (a lower rl ) implies a higher required mechanical torque
r 2f L2d ω2 + r 2
and excitation voltage magnitude, as we would expect.
and thus are solutions of the polynomial equation
  Remark 4.2. One should bear in mind that if the constant inputs
p(ω ) = 3V̄ f2 L2a f d r ω − 2r 2f (τ̄ − dω ) L2d ω2 + r 2 = 0 (33) τ̄ , V̄ f are chosen as in (36), then this does not necessarily mean
that there is only one steady state frequency ω̄, nor that the steady
Recall the following standard result concerning polynomials of de-
state corresponding to ω̄ is stable. These aspects are investigated
gree 3; see e.g. Irving (2004).
more thoroughly in the following subsections.
Proposition 4.1. Given a polynomial p(x ) = ax3 + bx2 + cx + d with
discriminant 4.2. Steady state stability analysis based on linearization
 = b2 c2 − 4ac3 − 4b3 d − 27a2 d2 + 18abcd
In this section we linearize the dynamics of the synchronous
Then the following holds true: generator coupled to a linear resistive load around a given steady
• if  > 0 then p(x ) = 0 has 3 distinct real roots. state, denoted by ¯. Rewrite (31) as
• if  < 0 then p(x ) = 0 has 1 real root and 2 complex conjugate ⎡ ˙ ⎤ ⎡ ⎤
ψd −rId − ψq ω
roots.
⎢ ψ˙ ⎥ ⎢ −rIq + ψd ω ⎥
• if  = 0 then at least two 2 roots coincide and all roots are real. ⎢ q⎥ ⎢ ⎥
⎢ ⎥ ⎢ ⎥
By considering the discriminant  of (33) it thus follows that ⎢ ψ˙ f ⎥ ⎢ −r f I f + V f ⎥
⎢ ⎥=⎢ ⎥,
(33) has only one real zero ω̄ if and only if ⎢ ˙ ⎥ ⎢ −rkd Ikd ⎥
  ⎢ψkd ⎥ ⎢ ⎥
24 dr L2a f d r 4f V̄ f2 3d2 r 2 − 5τ̄ 2 L2d
⎢ ⎥ ⎣ ⎦
⎣ψ˙ kq ⎦ −rkq Ikq
  ψq Id − ψd Iq − dω + τ
+ 9L4a f d r 2f V̄ f4 12d2 r 2 − τ̄ 2 L2d p˙
 2 whereas before r = rs + rl , and
+ 54 dr L6a f dV̄ f6 + 16r 6f d2 r 2 + τ̄ 2 L2d >0
⎡ψ ⎤ ⎡ L I + kL I + kL I ⎤
dd d afd f akd kd
4.1.1. Equilibrium currents ⎢ ψq ⎥ ⎢ Lq Iq − kLakq Ikq ⎥
From Proposition 4.2 of Fiaz et al. (2013), given a steady state ⎢ ⎥ ⎢ ⎥
⎢ ψ ⎥ ⎢kL I + L I + L I ⎥
frequency ω̄, there is a unique flux vector ψ̄d , ψ̄q , ψ̄r , and therefore ⎢ f ⎥ ⎢ afd d f fd f akd kd ⎥
a unique steady state current vector I¯d , I¯q , I¯r associated to it; see
⎢ ⎥=⎢ ⎥
⎢ψkd ⎥ ⎢ kLakd Id + Lakd I f + Lkkd Ikd ⎥
(32). More explicitly, the steady state currents are given by ⎢ ⎥ ⎢ ⎥
  ⎣ψkq ⎦ ⎣ −kLakq Iq + Lkkq Ikq ⎦
3 La f d Ld ω̄2V̄ f 3 La f d r ω̄V̄ f p 1
ω
I¯d = −   , I¯q =   M
2 r 2 + L2 ω̄2 r f 2 r 2 + L2 ω̄2 r f 
d d (34) 3
V̄ f and k = 2. Furthermore, rewrite the system in terms of the cur-
I¯f = , I¯ = I¯kq = 0
r f kd rents, instead of the fluxes. Then its linearization around (I¯, ω̄ ) is
given by (37), where


4
This holds for the round rotor case considered here. Without the round rotor Lred 0
assumption one needs to consider a polynomial equation of degree 5, cf. Fiaz et al. E= , Ld = Lq
(2013). 0 M

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

8 A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14

Fig. 3. The mechanical torque τ̄ (blue) and excitation voltage V̄ f (orange) as func-
tions of the load resistance rl . (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.) Fig. 4. The real part of the eigenvalues of the matrix E −1 A. The blue lines corre-
spond to stable eigenvalues (these eigenvalues should be multiplied by -1), whereas
the red line correspond to unstable eigenvalues. This plot shows that the linearized
system is asymptotically stable if rl ≥ 1.8. (For interpretation of the references to
color in this figure legend, the reader is referred to the web version of this article.)
Here Lred is the 5 × 5 inductance matrix obtained from the 6 × 6
inductance matrix L defined in (17) by deleting the third row and
third column (corresponding to the 0-coordinate).
A sufficient condition for asymptotic stability of the linearization
is given below.

Proposition 4.3. The linearized system (37) is asymptotically stable


if A + AT < 0.

Proof. Since E > 0 it follows from Theorem 3.1 of Duan and Patton
(1998) that E −1 A is Hurwitz. 

Remark 4.4. Remarkably, there is a close connection between the


matrix A in (37) and the matrix displayed in (30). In fact, the ma-
trix A + AT is equal to the matrix in (30) if rs = r and Ld = Lq .

Fig. 5. The steady state frequencies as functions of the load resistance rl . The blue
lines are the real parts of the steady state frequencies whereas the red line corre-
sponds to the imaginary part. Remarkably, the point where the linearized system
4.3. Numerical analysis of the stability of the linearized system
becomes (un)stable (at rl ≈ 1.8, see Fig. 4), is the point where two of the three
steady state frequencies coincide. (For interpretation of the references to color in
Proposition 4.3 only states a sufficient condition for the stabil- this figure legend, the reader is referred to the web version of this article.)
ity of the linearized system. In this subsection we will perform a
numerical analysis of the stability of the linearization of the syn-
chronous generator coupled with a resistive load. We first consider states associated to the minimum and maximum real steady state
the case of constant inputs τ̄ , V̄ f given by (36). Afterwards, we con- frequencies are locally asymptotically stable.
sider the case of constant inputs plus proportional control.
We throughout take ωdes = 1 and Vdes = 1. Furthermore, we ⎡ −r −Lq ω̄ 0 0 kLakq ω̄ −Ld I¯q

take rs = 0.003, d = 0.01. The other parameters correspond to a ⎢ ⎥
⎡ ⎤ ⎢ ⎥
555 MVA, 24 kV, 0.9 p.f. 60 Hz, 3600 RPM turbine-generator, see I˜˙d ⎢ ⎥⎡ ⎤
⎢ Ld I¯d + kLa f d I¯f ⎥
⎢ ⎥ ⎢Ld ω̄ kLa f d ω̄ kLakd ω̄
−r 0 I˜
Kundur (1993, p. 102). ⎥ d
⎢ I˜˙q ⎥ ⎢ ⎥ ⎢ I˜ ⎥
⎢ ⎥ ⎢ ⎥⎢ q ⎥
⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢ I˜˙ ⎥ ⎢ 0 −r f ⎥ ⎢ I˜ ⎥
⎢ f⎥ ⎢ 0 0 0 0 ⎥⎢ f ⎥
E⎢ ⎥ = ⎢ ⎥⎢ ⎥
⎢I˜˙ ⎥ ⎢ ⎥ ⎢I˜kd ⎥
⎢ kd ⎥ ⎢ ⎥⎢ ⎥
4.3.1. Constant control ⎢ ⎥ ⎢ ⎥ ⎢˜ ⎥
⎢I˜˙ ⎥ ⎢ 0 0 0 −rkd 0 0 ⎥ ⎣Ikq ⎦
The constant input values (36) as a function of the resistive load ⎣ kq ⎦ ⎢ ⎥
⎢ ⎥ ω˜
rl are depicted in Fig. 3. The stability of the linearized system is ω˜˙ ⎢ ⎥
⎢ ⎥
⎣ 0 0 0 0 −rkq 0 ⎦
determined by the eigenvalues of E −1 A, which are plotted as func-
tions of the resistive load rl in Fig. 4. 0 −kLa f d I¯f −kLa f d I¯q −kLakd I¯q −kLakq I¯d −d
  
As noted before, there are three steady state frequencies ω̄, A
which are plotted as functions of rl in Fig. 5. Remarkably, the point (37)
where there are three real steady state frequencies with two of
them coinciding (rl ≈ 1.8) happens at the same point as where the
linearized system becomes (un)stable, see Fig. 4. This suggests that
the steady state associated with the middle steady state frequency 4.3.2. Proportional control
is unstable if there are three distinct real steady state frequencies. In case a frequency proportional controller (plus a constant part
In fact, as suggested in the following subsection, the two steady as before) is used then there is additional freedom in choosing

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14 9

Stegink et al. (in preparation) it will be shown how also these in-
termediate steps can be understood from a port-Hamiltonian point
of view, emphasizing the expressions of the Hamiltonian function
(stored energy) for each of the steps. See also the recent papers
(Natarajan & Weiss, 2014a; 2014b) for an alternative approach.
The simplification of the eight-dimensional model to the swing
equation as followed in this section is directly based on the en-
ergy balance (13) of the original eight-dimensional model (1). As
before, split the total Hamiltonian as H = Hm + Hk , where the mag-
netic energy Hm (ψ s , ψ r , θ ) and the kinetic energy Hk (p) are given
respectively as


1 T ψs 1 2
Hm := ψ ψrT L−1 (θ ) , Hk := p (38)
2 s ψr 2Jr

Using this splitting we can decompose the original power balance


(13) into a power balance for the magnetic energy
Fig. 6. The equilibrium frequencies as function of rl in case dcl = 10. The black line
shows the real part of the (unstable) eigenvalues. The red line shows the (non- d
negative) imaginary part of the equilibrium frequencies. In this case the linearized Hm = −IsT Rs Is − IrT Rr Ir + IsT Vs + I f V f + ωτm , (39)
system is asymptotically stable if rl ≥ 0.1. This plot confirms that the two outer
dt
equilibrium frequencies are stable and the middle one is unstable. (For interpreta- together with a power balance for the kinetic energy
tion of the references to color in this figure legend, the reader is referred to the
web version of this article.) d
H = −dω2 + ωτ − ωτm , (40)
dt k

d. This can be done by the introduction of a controller of the by simply adding the term ωτ m to the power balance for the mag-
form τ = −dˆ(ω − ωdes ) + τ̄ , where τ̄ is as in (36). In this way the netic energy, and subtracting the same term from the power bal-
damping coefficient dcl of the closed-loop system becomes equal ance for the kinetic energy. Equivalently, we are thus looking at the
to dcl = d + dˆ; and thus can be chosen as desired. It can be shown 8-dimensional model as the interconnection of a seven-dimensional
that if dcl is larger, then the stability margin in terms of rl is larger, magnetic subsystem (involving the variables ψ s , ψ r , θ appearing
see Fig. 6. By manipulating dcl , numerical results even suggest that in the expression for the magnetic energy) together with a one-
for any given rl > 0 there exists a sufficiently large dcl such that the dimensional mechanical subsystem (just involving the angular mo-
linearized system is asymptotically stable. This claim is strength- mentum p). This interconnection takes place via the port variables
ened by the following result which holds for all positive parameter ω, τ m , with τ m the air-gap torque (cf. (13)), and with ωτ m the
values. power flowing between the mechanical subsystem and the mag-
netic subsystem.
Proposition 4.5. Given constant rl > 0 and let the polynomial p be The swing equation is now obtained by neglecting the seven-
given by (33), where r = rs + rl . If the constant input values (36) (de- dimensional dynamics of the magnetic subsystem, and replacing it
pending on d) are substituted in p, then the discriminant  of the by a static relationship around a reference value ω̄ (e.g., 50Hz) of
resulting polynomial p satisfies the frequency (angular velocity) ω. This is accomplished by making
the key assumption that the increase in magnetic energy is, apart
lim  = −∞.
d→∞ from losses and the power delivered by the excitation system, pro-
portional to the deviation of the frequency ω with respect to its
Proposition 4.5 states that if dcl is sufficiently large, then  <
nominal value ω̄, that is,
0 and thus there exists only one real equilibrium. As numerical re-
sults show, if there is only one real equilibrium, then the linearized d
Hm = −IsT Rs Is − IrT Rr Ir + A(ω − ω̄ ) + I f V f , (41)
system is asymptotically stable. The result of Proposition 4.5 is dt
in line with Proposition 4.3, in the sense that the conditions of for some positive constant A. Physically, this corresponds to the
Proposition 4.3 also imply that rl is sufficiently large. fact that once a deviation of the frequency from a given value ω̄
occurs, then this will lead to the creation of a magnetic field asso-
5. From the eight-dimensional model to the swing equation ciated with the damper windings and a resulting restoring torque
(Lenz’ force). Therefore, A is also called the damping constant corre-
In this section we will discuss an approach to drastically sim- sponding to the damper windings (although it does not correspond
plify the eight-dimensional model of the synchronous generator to dissipation of energy).
into another standard model, called the swing equation. This sim- In view of the balance equation (13) the assumption (41) is
plification is based on neglecting the dynamics of the magnetic equivalent to the assumption that the port variables ω, τ on the
flux variables and replacing it by a static relationship between me- mechanical side of the seven-dimensional system are related with
chanical power and electrical power supplied at the stator termi- the port variables on the electrical network side by the expression
nals. This leads to a one-dimensional model, still capturing the
mechanical dynamics of the system. The resulting swing equa- ωτm = −IsT Vs + A(ω − ω̄ ) (42)
tion is again in port-Hamiltonian form; however, with respect to
a completely different set of variables than the original eight- This has a clear interpretation: around steady state ω̄ the mechan-
dimensional system. ical power ωτ m delivered by the rotor is directly transformed into
The simple approach taken in this section seems to be relatively electrical power −IsT Vs at the stator terminals transmitted to the
new. In the literature, see e.g. Kundur (1993), Machowski et al. network, plus the term A(ω − ω̄ ) due to the restoring torque of the
(2008), and Galaz, Ortega, and Bazanella (2002), the simplifica- damper windings.
tion towards the swing equation is typically undertaken through a With regard to the one-dimensional system in the angular mo-
number of intermediate steps; each of which has its own merits. In mentum p this leads to the following expression, directly relating

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

10 A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14

the change in kinetic energy of the rotor to the power delivered to 6. Networks
the electrical network:
In principle, it is rather straightforward to formulate time-
d
H = −dω2 + ωτ − ωτm domain models of power networks. One considers the graph with
dt k the basic components such as synchronous generators and loads
= −dω2 − A(ω − ω̄ ) + IsT Vs + ωτ as (grounded) nodes, and with edges corresponding to the trans-
≈ −A(ω − ω̄ ) + IsT Vs + ωτ mission lines; similar to electrical circuit modeling. Various time-
domain models may be used to describe the transmission lines,
= −A(ω − ω̄ ) + Pelectric + Pmech , (43)
ranging from the telegrapher’s equations to simple RLC-circuits
where we additionally assume that the mechanical losses dω2 due (such as ladder networks). In particular, the systematic approach
to friction can be neglected. as proposed in Fiaz et al. (2013) for time-domain modeling of
Furthermore, assuming that ω is close to ω̄ we can make the power networks (approximating the transmission lines by standard
approximation models) has the advantage that the number of algebraic con-
straints in the obtained model is minimized.
d d 1 2 d 1 On the other hand, time-domain models of large power net-
H = p = Jr ω2 = (Jr ω ) ω˙ ≈ Mω˙ , (44)
dt k dt 2Jr dt 2 works necessarily are high-dimensional, while the dynamics of the
transmission lines is often not the main source of complicated dy-
with M := Jr ω̄. Eqs. (43) and (44) lead to the following equation namical behavior. As a consequence, most of the models for power
for the deviation ω˜ = ω − ω̄ of the frequency with respect to the networks are based on modeling the transmission line network as
nominal value ω̄ a (complex-valued) admittance matrix; based on the assumptions
of linearity and pure sinusoidal voltage and current signals with

˜˙ = −A ω
˜ + Pelectric + Pmech , (45)
nominal fixed frequency. Of course, such models are not suited for
called the swing equation. the analysis of transient behavior.
The assumption (42) leading to the swing equation can be only
justified on a relatively short time-interval after the deviation ω ˜ 6.1. Phasor model of the transmission line network
from the nominal value ω̄ occurs. Indeed, after some time no fur-
ther increase of the magnetic energy in the damper windings will The quasi-steady state modeling of a transmission line network
take place. This observation regarding the underlying assumptions by an admittance matrix is done as follows. Consider the set of
and validity of the model is well-known in the power engineer- sinusoids ECos(ω̄t + δ ) with amplitude E and phase shift δ for a
ing literature; see e.g. (Kundur, 1993; Machowski et al., 2008; Pai, given (nominal) frequency ω̄ (e.g., 50Hz). Its phasor representation
1989). In fact, the terminology ‘swing equation’ is based on this ob- is given by the complex√number Ee jδ , with E ≥ 0 and δ ∈ [0, 2π )
servation: the swing equation is thought to be reasonably accurate real variables, and j := −1. Conversely, the sinusoid correspond-
for the ‘first swing’ of the synchronous generator after a deviation ing to the phasor Ee jφ is given as ECos(ω̄t + δ ) = Re Ee jδ e jω̄t .
of the nominal value ω̄ occurs. An asymptotic analysis of the dy- The usefulness of the phasor representation of sinusoids stems
namics of the synchronous generator based on the swing equation from two facts: (1) the map from sinusoid ECos(ω̄t + δ ) to phasor
model, as is sometimes done in the (control) literature, is therefore Ee jδ is linear, (2) the phasor of the derivative of a sinusoid is the
debatable. We also refer to the recent paper (Caliskan & Tabuada, phasor of the original sinusoid multiplied by jω̄, and similarly the
2015) for an evaluation of the swing equation model in comparison phasor of the integral of a sinusoid is equal to the phasor of the
with the eight-dimensional model (1). original sinusoid divided by jω̄.
Thus from a modeling point of view we may conclude that the Now consider a network composed of linear transmission lines,
swing equation has only limited validity. On the other hand, from a where the voltages and currents are all assumed to be pure sinu-
control point of view an extra term of the same form as the damp- soids with equal frequency ω̄. Then the phasor I of the current I
ing Aω ˜ appears in case a standard droop control scheme is used. through a transmission line is related to the phasor V of the volt-
In droop control the deviation of the mechanical torque τ with re- age across the line as
spect to the nominal value τ̄ is set to be proportional to the devia-
tion ω˜ . As a consequence, the swing equation may be a reasonable I = τV (47)
model for a droop-controlled synchronous generator.
Here the complex number τ captures the characteristics of the
The swing equation (8) is formally equivalent to the equation
transmission line: it is equal to the conductance G > 0 in case of
of motion for an actuated rotating mass subject to linear friction.
a purely resistive line, equal to jω̄C, C > 0, in case of a purely ca-
However, from a physical point of view it should be emphasized 1
pacitive line, equal to jω̄ , L > 0, in case of a purely inductive line,
that ω˜ denotes deviation from the nominal value ω̄ (instead of fre- L
and combinations of these for the general case.
quency itself), and that A ω˜ is a restoring force; not dissipation.
The transmission line network defines a graph, with m edges
Furthermore notice that the physical dimensions (units) used in
corresponding to the transmission lines, and n nodes correspond-
the swing equation are different from the physical dimensions for
ing to the buses (where components such as generators and loads
an actuated rotating mass. In fact, the physical dimension of the
are attached). Endowing the graph with an arbitrary orientation
input in the swing equation is power instead of torque. This means
(specifying in which direction the currents flowing through the
that in the following input-state-output port-Hamiltonian system
edges are considered to be positive, as well as specifying the signs
based on (45), given as
of the voltage drops), we describe the resulting oriented graph by
p˜˙ = −A ω ˜ + u, u = Pelectric + Pmech , its n × m incidence matrix5 D. Denoting the voltage potentials at
the nodes by E1 , . . . , En , then the m-dimensional vector V of voltage
dHs 1 2
y= , Hs ( p˜ ) = p˜ , (46)
d p˜ 2M
5
The incidence matrix D of a directed graph with n nodes and m edges is an n ×
where p˜ := Mω˜ , the output y is necessarily dimensionless, ensur- m matrix, with kth column the vector with −1 at the position of the tail vertex of
ing that the product uy = Pelectric y + Pmech y still has dimension of the kth edge, 1 at the position of the head vertex of this edge, and zeros elsewhere.
power. Obviously 1T D = 0, where 1 is the vector of all ones.

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14 11

drops across the transmission lines is given by the m-dimensional 6.2. Networks of generators
vector V = DT E.
Now assume that the voltage potentials are sinusoids with A simple model of a network of synchronous generators6 is ob-
phasors Ek = Ek e jδk , k = 1, . . . , n. Then the phasors of the currents tained by combining the above quasi-static model of a transmis-
through the edges are equal to the phasors of the voltage drops sion line network with the swing equation model of a synchronous
across these edges, multiplied by the complex numbers γ1 , . . . , γm generator, under the following assumptions:
capturing the characteristics of the transmission lines. It follows (1) The amplitudes Ek of the voltage potentials Ek at the nodes
that the vector I of phasors of the currents injected at the nodes k = 1, . . . , n are all constant.
of the network is related to the phasors of the voltage potentials at (2) All transmission lines are purely inductive, i.e., Y = − jB,
the nodes by the compact formula, see e.g. van der Schaft (2010), with B an n × n real matrix, called the susceptance matrix of the
Dörfler and Bullo (2013), and Simpson-Porco, Dörfler, and Bullo purely inductive transmission line network.7
(2013a) Under these assumptions the active power transfer Pij from
node i with voltage phasor angle δ i to node j with voltage phasor
I = DT DT E, T = diag(τ1 , . . . , τm ) (48)
angle δ j is computed from (48) as
The n × n complex matrix Y := DT DT
is called the admittance ma-
Pi j = i j Sin(δi − δ j ), (54)
trix of the transmission line network. From the definition it is ob-
vious that it is symmetric (Y = Y T ) and has zero column and row where i j := Bi j Ei E j , with Bij the susceptance of the transmission
sums (since the column sums of D are zero). Furthermore, splitting line between nodes i and j.
between the real and imaginary part This can be combined with the dynamics of the swing equation
(45) for each generator as follows. For ease of notation denote the
Y = D Re(T )DT + jD Im(T )DT =: Re(Y ) + j Im(Y ), (49)
frequency deviation ω − ω̄ in the sequel simply by ω. With this
it follows that both Re(Y ) and Im(Y ) are symmetric matrices abuse of notation, the power balance at the ith generator node
with row and column sums zero. In fact, see e.g. van der Schaft within the network can be written as
(2010) and Simpson-Porco et al. (2013a), Re(Y ) is a Laplacian ma- 
Mi ω˙ i = −Ai ωi − Pi j + ui , (55)
trix in case of a purely resistive transmission line network, while
j∈Ni
Im(Y ) is a Laplacian matrix in case of a purely capacitive network
and minus a Laplacian matrix in case of a purely inductive net- with ui denoting the generated/consumed power at the ith node,
work. and Ni denoting the set of neighbors of node i in the transmission
In case the transmission lines are modeled by RLC-circuits the line network graph. Taking all these equations together in vector
expressions for Re(Y ) and Im(Y ) are easily found by using the notation we arrive at the following n-dimensional system
aforementioned properties of phasors of derivatives, reducing the Mω˙ = −Aω − D Sin (DT δ ) + u, (56)
solution of a set of linear differential equations for the RLC-circuit
to the solution of complex algebraic equations. where ω is the n-dimensional vector of frequency deviations, u
Eq. (48) is the starting point for power flow analysis. For exam- is the n-dimensional vector of produced/consumed power at all
ple, split up the nodes into a subset corresponding to energy pro- nodes, A is the diagonal matrix of damping constants, and M is
duction (generators g) and energy consumption (corresponding to the diagonal matrix with diagonal elements M1 , . . . , Mn model-
loads l). Then (48) is correspondingly split into ing the inertia of the generators. Furthermore, Sin z, z ∈ Rm , de-



notes the element-wise sinus function with ith element given as
Ig Ygg Ygl Eg Sin zi , i = 1, . . . , m, and is the m × m diagonal matrix with kth
= (50)
Il Ylg Yll El diagonal element given by ij , if the kth transmission line is be-
tween node i and j.
The analysis of this equation is depending on the model for the
Furthermore, it is assumed that the voltage phasor angles δ i
loads. Two simple possibilities are: (1) linear resistive loads, (2)
corresponding to the ith generator satisfy
constant power loads. In the first case
δ˙ i = ωi , i = 1, . . . , n, (57)
Il = −GEl , (51)
with ωi the frequency deviation of the ith generator.
for some Laplacian matrix G modeling the conductances of the The coupled differential equations (56) and (57) can be com-
loads. In this case, equations (50) reduce to the power flow gen- pactly written as follows. Denote the phase difference across the kth
eration equations transmission line from node i to node j by
 
Ig = Ygg − Ygl (Yll + G )−1Ylg Eg =: Yg Eg , (52) qk := δ j − δi , k = 1, . . . , m (58)
where the admittance matrix Yg is a Schur complement, which in- In vector notation q = DT δ, with D the n × m incidence matrix of
herits the Laplacian properties from Y and G; see van der Schaft the network. Therefore q˙ = DT ω, and equations (56) and (57) result
(2010) and Dörfler and Bullo (2013) for the basic ideas. in the (n + m )-dimensional system
In the second case, the complex power S := Il∗Vl , with ∗ denoting
transpose and complex conjugate, is a prescribed constant vector q˙ = DT ω
S̄l , that is Mω˙ = −Aω − D Sin q + u, (59)
Il∗ Vl = S̄l (53) This model is often taken as a fruitful starting point for
control purposes, especially for control of micro-grids;
The vector Re(S̄l ) is called the vector of active power consumed
at the loads, while Im(S̄l ) is called the vector of reactive power
consumed at the loads. See e.g. Willems (2010) for an insightful 6
Instead of a network of single synchronous generators and synchronous ma-
treatment of these notions; in particular, the active power can be chines this may also represent a network of aggregated generator areas. Further-
more, it may also represent a micro-grid, where (some of) the generators actually
equated (up to a constant) with the average power consumed at the
correspond to inverters (coupled to other power sources) which are controlled in
loads. In the case of constant power consumption, the resulting re- such a way as to emulate synchronous generators.
lation between Ig and Eg is quadratic, and much more difficult to 7
Usually the inductance of the transmission line starting from a generator node
analyze. is increased by adding an inductor representing the inductance of the generator.

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

12 A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14

see e.g. Dörfler, Simpson-Porco, and Bullo (2016), Simpson- This can be summarized in the following proposition:
Porco et al. (2013a), Simpson-Porco, Dörfler, and Bullo (2013b),
Proposition 6.1. The set of steady states of (60) is given as
Trip, Bürger, and De Persis (2016), De Persis and Monshizadeh
(2015), Schiffer (2015), and Schiffer, Seel, Raisch, and Sezi (2015). E := {(q̄, p̄) | p̄ = M1ω∗ , D Sin q̄ = −A1ω∗ + ū} (66)
It is readily seen that the model (59) can be written in port- If E is non-empty, then it has dimension equal to  := dim ker D,
Hamiltonian form as while the dynamics (60) has  independent conserved quantities


 ∂ H (q, p)

C (q ) = kT q with k ∈ ker D. Furthermore, p̄ is unique, and determined
q˙ 0 DT ∂q 0
= ∂ H (q, p)
+ , p = Mω by ū.
p˙ −D −A ∂p u
In the sequel let us assume that E is non-empty. By the exis-
∂H
y= (q, p) = ω, (60) tence of conserved quantities (or equivalently, from the first line
∂p of (59)) it follows that the solution (q(·), p(·)) of (60) starting from
with input u the vector of generated/consumed power at the an arbitrary initial condition (q0 , p0 ) at time 0 will satisfy
nodes, output y the vector of frequency deviations at the nodes, q(t ) − q0 ∈ im DT , t≥0 (67)
and with Hamiltonian
In the so-called ‘DC-approximation’, where we replace Sin q by q
1 T −1
H (q, p) := p M p − 1T Cos q, (61) the set of steady states EDC is given as the affine space
2
EDC = {(q̄, p̄) | p̄ = M1ω∗ , D q̄ = −A1ω∗ + ū}
where Cos denotes the component-wise cosinus function.
The Poisson structure matrix of this port-Hamiltonian system is It follows that for any q0 there exists a unique v0 ∈ im DT such that
given by the constant matrix (determined by the topology of the q0 + v0 = q̄∞ , with (q̄∞ , p̄) ∈ EDC (recall that p̄ is uniquely deter-
network) mined). Hence, q̄∞ is the unique steady state value corresponding

to initial value q0 .
0 DT
J= , (62) One may expect that this can be generalized9 to the original
−D 0 equations involving the sin mapping, in the sense that for any
whereas the resistive structure matrix is R = diag(0, A ). initial condition (q0 , p0 ) there exists a unique (modulo π -shifts)
The first term in the Hamiltonian (61) is the kinetic energy, steady state value (q̄∞ , p̄) ∈ E.
shifted with respect to the nominal angular momentum p̄ := Mω̄ Since the Poisson structure matrix J and the resistive structure
of the generators. The second term can be interpreted as a potential matrix R in (60) are independent of the state (q, p), we may use
energy corresponding to the phasor approximation of the inductive the general construction of the shifted Hamiltonian as in (24) for
transmission line network. the stability analysis of the steady states in E. With respect to the
Hamiltonian H given by (61) the shifted Hamiltonian with respect
6.2.1. Dynamical analysis to a steady state (q̄, p̄) is computed as (modulo a constant)
The dynamical analysis of the port-Hamiltonian system (60) for (q, p)
H
a constant vector of inputs u = ū turns out to be rather straight- 1
forward, and can be performed along the lines of van der Schaft := ( p − p̄)T M−1 ( p − p̄) − 1T Cos q − (q − q̄ )T Sin q̄ (68)
2
and Maschke (2013). Consider the set of steady state values (q̄, p̄ =
By construction of the shifted Hamiltonian the dynamics (60) can
Mω̄ ), given as the solutions of
 ∂ H  then be rewritten as




 ∂ H 

∂ q (q̄, p̄)
T
0 D 0 q˙ 0 DT ( q, p ) 0
0= + (63) ∂q
−D −A ∂ H (q̄, p̄) ū = 
+ , (69)
∂H
∂ p (q, p)
∂p p˙ −D −A u − ū
Assuming without loss of generality that the graph is connected, where the constant input value ū has been incorporated in the def-
and thus ker DT = span 1, it follows from the upper half of the . It follows that
inition of H
equations that the steady state value ω̄ is determined as ω̄ = d 
∂ H (q̄, p̄) = 1ω for a certain common scalar frequency deviation
∗ H (q, p) = −( p − p̄)T M−1 AM−1 ( p − p̄) + (y − ȳ )T (u − ū )
∂p dt
ω∗ . Furthermore, by pre-multiplication of the lower half of the
≤ (y − ȳ )T (u − ū ) (70)
equations by 1T we obtain the equation

In fact, the shifted Hamiltonian H defines a storage function for
1T A1ω∗ = 1T ū, (64)
all p and all q with qk ∈ (− π2 , π2 ) if q̄k ∈ (− π2 , π2 ), k = 1, . . . , m. In

uniquely determining ω∗ . (Note that 1T A1
= i Ai , i.e., the sum of fact, by convexity of minus the cosinus function on the interval
the damping constants of all the generators.) (− π2 , π2 ) it follows that
By substitution of (64) into the second set of equations of (q̄, p̄) ≤ H
(q, p)
H (71)
(63) we obtain the following equation in the steady state value q̄:
Hence shifted passivity with respect to the shifted inputs and out-
0 = −D Sin q̄ − A1ω∗ + ū (65) puts u − ū, y − ȳ results, with the steady state output value ȳ =
∂ H (q̄, p̄) = M1ω determined by ū.
∂p ∗
Eq. (65) does not necessarily have a solution q̄ (since the mapping
Furthermore, for u = ū it follows that the largest invariant set
Sin : RM → RM is bounded), and if there exists a solution then this d 
within the set {(q, p) | dt H (q, p) = 0} is equal to E, and thus by
solution is unique if and only if ker D = 0 (no cycles in the trans-
LaSalle’s Invariance principle the dynamics will converge to E,
mission line network graph). At the same time, if ker D = 0, also
where it is to be expected (see the above discussion) that it will
conserved quantities are present. Indeed, all functions C (q ) = kT q
converge to a steady state value which is solely determined by the
d
with k ∈ ker D are conserved quantities8 for (60); that is, dt C (q ) =
initial condition.
k D ω = 0 along solutions of (60).
T T

9
See van der Schaft, Rao, and Jayawardhana (2013) for a very similar statement
8
In fact, being conserved quantities for all possible Hamiltonians they are in the context of equilibria of mass-action kinetics chemical reaction networks, and
Casimirs; solely determined by the Poisson structure matrix. van der Schaft and Maschke (2013) for a general theory in the linear case.

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14 13

6.3. Networks of generators and frequency dependent loads 7. Conclusions

A direct extension of the previous analysis is to networks con- Power networks constitute an important research area with
sisting of generators (denoted by g), and frequency dependent and many challenges to the systems and control community. In this pa-
constant loads10 (denoted by l). This leads to a bipartite directed per we have concentrated on two topics within the broad area of
graph with nodes representing either a generator or a load. The modeling for control of power networks: the dynamics of the syn-
incidence matrix of the graph is split correspondingly into chronous generator and a simple model of power networks based

on the quasi-steady state approximation of the transmission line
Dg
D= (72) network, the assumption of constant voltage potentials, and the
Dl swing equation model of the synchronous generator. Both cases
Modeling as before the generators by swing equations the total lead to an insightful port-Hamiltonian formulation, with the dy-
system takes the form namical analysis of the eight-dimensional synchronous generator


 ∂ H 

T model still posing fundamental questions. We believe both cases
q˙ 0 DTg ∂q 0 D illustrate the relevance of a reconsideration of the modeling ap-
= ∂H
+ vg + l ωl proaches to power networks for control purposes, and motivate a
p˙ −Dg 0 ∂p I 0
further research focus on this. Especially the property of shifted
∂H passivity with respect to steady state values has turned out to be
ωg =
∂p a very useful starting point for control (De Persis & Monshizadeh,
∂H 2015; Stegink, De Persis, & van der Schaft, 2015a; 2015b; Trip et al.,
vl = Dl , (73) 2016).
∂q
Next to the modeling, analysis and control of the physical power
together with damping relations, between the delivered power vg , network a major challenge is the design and control of the com-
vl and the frequencies ωg , ωl at the generator and load nodes, munication and control layer (‘smart grids’). The overall aim is to
given by produce and consume energy in such a way that there is balance
vg = −Ag ωg + ug generators between supply and demand, while the physical network is oper-
(74) ating within a desired range (frequency regulation, no voltage col-
vl = −Al ωl + ul loads lapse, sharing of power, regulation of ‘reactive power’, satisfaction
Here Ag , Al are the diagonal damping matrices of the generators, of flow and storage constraints). Furthermore, there is a need to
respectively, loads. Furthermore, H is the same Hamiltonian as de- achieve this in an optimal manner, while at the same time guaran-
fined before in (61), with p corresponding to the angular momenta teeing robustness with respect to e.g. variations in uncontrollable
p = Mωg of the generators (there is no inertia associated to the energy supply/demand and faults.
loads). This is also linked to economic considerations and dynamic pric-
Solving for ωl = −A−1l
vL + A−1
l
ul this leads to ing of energy supply and demand, for which there is a large liter-


 ∂ H 

ature. From a port-Hamiltonian perspective it was recently shown
q˙ −DTl A−1
l
Dl DTg ∂q 0 DTl A−1
l
ug in Stegink et al., (2015a, 2015b) how the primal-dual gradient al-
= ∂H
+ (75) gorithm for the maximization of the social welfare can be intrin-
p˙ −Dg −Ag ∂p I 0 ul
sically formulated as another port-Hamiltonian system, which can
Note that this again is a port-Hamiltonian model, with Poisson be coupled to the physical network so as to obtain an overall port-
structure matrix J and resistive structure matrix R respectively Hamiltonian system composed of a physical and a cyber-economic
given by part. The combination between the physical network and the dy-


namical control architecture, together with a strong optimization
0 DTg DTl A−1 Dl 0
J= , R= l
(76) and economic perspective, poses many interesting research ques-
−Dg 0 0 Ag tions.
The dynamical analysis can be performed in the same way as for
the network of generators. Indeed, a steady state analysis for con-
Acknowledgments
stant generated power ūg , ūl leads to the requirement


 ω̄g We would like to thank Claudio De Persis, Jacquelien Scherpen
DTg DTl = 0, (77) and Nima Monshizadeh of the University of Groningen for many
ω̄l
inspiring discussions and collaborations closely related to the con-
which again, under the standing assumption that the transmission tents of this paper. We also thank the reviewers for their feedback
line network is connected (ker DT = span 1), implies on the presentation of the paper.
Large parts of the current paper are based on a talk by the
ω̄g = 1ω∗ , ω̄l = 1ω∗ , (78)
first author presented at the 5th IFAC Workshop on Lagrangian and
where the steady state frequency deviation ω∗ is determined by Hamiltonian Methods for Non Linear Control, Lyon, France, July 2015.
(1T Ag 1 + 1T Al 1 )ω∗ = 1T ūg + 1T ūl (79)
Furthermore, the steady state value q̄ satisfies References

Dg Sin q̄ = −Ag 1ω∗ + ūg (80) Caliskan, S. Y., & Tabuada, P. (2014). Compositional transient stability analysis of
multimachine power networks. IEEE Transactions on Control of Network Systems,
By using as Lyapunov function the same shifted Hamiltonian 1(1), 4–14.
(68) as in the previous subsection shifted passivity and asymptotic Caliskan, S. Y., & Tabuada, P. (2015). Uses and abuses of the swing equation model.
In IEEE 54th conference on decision and control (CDC), Osaka, Japan, December
stability of the steady states results. 15–18 (pp. 6662–6667).
De Persis, C., & Monshizadeh, N. (2015). A modular design of incremental Lyapunov
functions for microgrid control with power sharing. in press.arXiv: 1510.05811
10
Also many classes of inverters can be modeled this way. See De Persis and Mon- Dörfler, F., & Bullo, F. (2013). Kron reduction of graphs with applications to electrical
shizadeh (2015) for a related set-up. networks. IEEE Transactions on Circuits and Systems I, 60(1), 150–163.

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017
JID: JARAP
ARTICLE IN PRESS [m5G;May 11, 2016;21:16]

14 A. van der Schaft, T. Stegink / Annual Reviews in Control 000 (2016) 1–14

Dörfler, F., Simpson-Porco, J. W., & Bullo, F. (2016). Breaking the hierarchy: Dis- Trip, S., Bürger, M., & De Persis, C. (2016). An internal model approach to frequency
tributed control and economic optimality in microgrids. IEEE Transactions on regulation in power grids. Automatica, 64, 240–253.
Control of Network Systems. in press van der Schaft, A. J. (1996). L2 -gain and passivity techniques in nonlinear con-
Duan, G.-R., & Patton, R. J. (1998). A note on Hurwitz stability of matrices. Automat- trol, Lecture Notes in Control and Information Sciences: 218 p. 168. Springer-Ver-
ica, 34(4), 509–511. lag, Berlin. (2nd revised and enlarged edition, Springer-Verlag, London, 20 0 0
Fiaz, S., Zonetti, D., Ortega, R., Scherpen, J. M. A., & van der Schaft, A. J. (2013). A (Springer Communications and Control Engineering series), p.xvi+249.)
port-Hamiltonian approach to power network modeling and analysis. European van der Schaft, A. J. (2010). Characterization and partial synthesis of the behavior of
Journal of Control, 19(6), 477–485. resistive circuits at their terminals. Systems & Control Letters, 59, 423–428.
Galaz, M., Ortega, R., & Bazanella, A. S. (2002). Transient stabilization of power sys- van der Schaft, A. J., & Jeltsema, D. (2014). Port-Hamiltonian systems theory: An intro-
tems via total energy shaping: a comparative simulation study with the classical ductory overview. NOW Publishers, Boston-Delft.(Originally published as Foun-
scheme. In Proceedings of CLCA. dations and Trends in Systems and Control, 1(2–3), 173–378, 2014.)
Irving, R. S. (2004). Rings integers, polynomials, and rings: A course in algebra. van der Schaft, A. J., & Maschke, B. (2013). Port-Hamiltonian systems on graphs.
Springer. SIAM Journal on Control and Optimization, 51(2), 906–937.
Kundur, P. (1993). Power system stability and control. Mc-Graw-Hill Engineering. van der Schaft, A. J., & Maschke, B. M. (1995). The hamiltonian formulation of en-
Machowski, J., Bialek, J. W., & Bumby, J. R. (2008). Power system dynamics: Stability ergy conserving physical systems with external ports. International Journal of
and control (2nd ed.). John Wiley & Sons, Ltd. Electronics and Communications, 49, 362–371.
Maschke, B. M., Ortega, R., & van der Schaft, A. J. (20 0 0). Energy-based lyapunov van der Schaft, A. J., Rao, S., & Jayawardhana, B. (2013). On the mathematical struc-
functions for forced hamiltonian systems with dissipation. IEEE Transactions on ture of balanced chemical reaction networks governed by mass action kinetics.
Automatic Control, 45, 1498–1502. SIAM Journal on Applied Mathematics, 73, 953–973.
Natarajan, V., & Weiss, G. (2014a). Almost global asymptotic stability of a con- Willems, J. L. (2010). Reflections on power theories for poly-phase nonsinusoidal
stant field current synchronous machine connected to an infinite bus. In Pro- voltages and currents. In International school on nonsinusoidal currents and com-
ceedings of the 53rd IEEE conference on decision and control, Los Angeles, CA pensation, Lagow, Poland, June 15–18 (pp. 5–16).
(pp. 3272–3279).
Natarajan, V., & Weiss, G. (2014b). A method for proving the global stability of a Arjan van der Schaft received the undergraduate (cum laude) and Ph.D. degrees in
synchronous generator connected to an infinite bus. In Proceedings of IEEE 28th mathematics from the University of Groningen, The Netherlands. In 1982, he joined
convention of electrical & electronics engineers, Israel, Eilat. the Department of Applied Mathematics, University of Twente, where he was ap-
Pai, E. (1989). Energy function analysis for power systems stabilization. Kluwer Aca- pointed as full professor in Mathematical Systems and Control Theory in 20 0 0. In
demic Publishers, Boston. September 2005, he returned to his Alma Mater as a full professor in mathematics.
Sauer, P., & Pai, M. (1998). Power system dynamics and stability. Prentice Hall. He is fellow of the Institute of Electrical and Electronics Engineers (IEEE). He was
Schiffer, J. (2015). Stability and power sharing in microgrids. Germany: TU Berlin invited speaker at the International Congress of Mathematicians, Madrid, 2006. He
(Ph.D. thesis). was the 2013 recipient of the Three-yearly awarded Certificate of Excellent Achieve-
Schiffer, J., Seel, T., Raisch, J., & Sezi, T. (2015). Voltage stability and reactive power ments of the IFAC Technical Committee on Nonlinear Systems. He is (co-)author
sharing in inverter-based microgrids with consensus-based distributed voltage of the following books: System Theoretic Descriptions of Physical Systems (1984),
control. In IEEE Transactions on Control Systems Technology. in press. Variational and Hamiltonian Control Systems (1987, with P.E. Crouch), Nonlinear Dy-
Simpson-Porco, J. W., Dörfler, F., & Bullo, F. (2013a). Synchronization and power namical Control Systems (1990, with H. Nijmeijer), L2 -Gain and Passivity Techniques
sharing for droop-controlled inverters in islanded microgrids. Automatica, 49(9), in Nonlinear Control (1996, 20 0 0), An Introduction to Hybrid Dynamical Systems
2603–2611. (20 0 0, with J.M. Schumacher), and Port-Hamiltonian Systems Theory: An Introduc-
Simpson-Porco, J. W., Dörfler, F., & Bullo, F. (2013b). Voltage stabilization in mi- tory Overview (2014, with D. Jeltsema).
crogrids via quadratic droop control. In IEEE conference on decision and control
(CDC), Florence, Italy (pp. 7582–7589). Tjerk Stegink received the M.Sc. degree in applied mathematics from the University
Stegink, T. W., De Persis, C., & van der Schaft, A. J. in preparation. of Groningen, The Netherlands, in 2014, where he has been a Ph.D. student since
Stegink, T. W., De Persis, C., & van der Schaft, A. J. (2015a). A port-Hamiltonian ap- August 2014. His research interests include the modeling and distributed optimal
proach to optimal frequency regulation in power grids. In IEEE 54th conference control of power systems.
on decision and control (CDC), Osaka, Japan, December 15–18 (pp. 3224–3229).
Stegink, T. W., De Persis, C., & van der Schaft, A. J. (2015b). A unifying energy-based
approach to optimal frequency and market regulation in power grids. in press.
arXiv:1510.05420.

Please cite this article as: A. van der Schaft, T. Stegink, Perspectives in modeling for control of power networks, Annual Reviews in
Control (2016), http://dx.doi.org/10.1016/j.arcontrol.2016.04.017

You might also like