Rigatos2016 Chapter SynchronizationAndStabilizatio
Rigatos2016 Chapter SynchronizationAndStabilizatio
8.1 Outline
8.2.1 Outline
scheme for the local power generators can be modified so as to provide the desirable
robustness to the distributed power generation scheme. By redesigning the Kalman
Filter, initially used for estimating the non-measurable elements of the generator’s
state vector, in the form of a disturbances estimator it becomes possible to estimate
in real-time the aggregate disturbance that affects each local generator [98, 99, 169,
292]. Knowing the disturbances’ value it is possible to generate a counter-disturbance
term that serves as auxiliary control input and which finally compensates for the dis-
turbances’ effects. With the previously described control approach each generator
keeps on running at its nominal operating conditions and the distributed generation
units become robust to phenomena such as mechanical input torque variations and
interarea oscillations [15, 439, 474, 483, 504].
The proposed control method is suitable for compensating for both endogenous
and exogenous disturbances that affect the generators’ model. Endogenous distur-
bances are associated with parametric uncertainty, e.g. there can be imprecision in
the values of reactances, stator and rotor circuit coefficients and the rotor’s moment
of inertia. Exogenous disturbances are associated with variations in the mechanical
input torque, as well as with perturbations in the excitation system of each gener-
ator which are due to lack of synchronization with other generators connected to
the grid [50]. The efficient compensation of these disturbances enables to maintain
synchronism between the individual power generators, thus assuring: (i) robustness
improvement of the power grid and maintenance of the synchronization between the
distributed electric power generation units even under adverse conditions in the power
grid, (ii) improvement of the provided power quality (distortions in power quality due
to deviations of the power generation units from their nominal operating conditions
will be avoided), (iii) uninterrupted power flow in the grid, ability to incorporate more
power generation units without affecting stability, and consequently better response
to the increased demand for power supply.
The proposed distributed filtering and control method for the system of the inter-
connected power generators is a highly data-driven one. Actually, to succeed state
estimation and disturbances compensation for a set of n distributed PMSGs it is
necessary that at each sampling period each power generator controller receives n
variables associated with the turn angles of all generators in the power grid. Thus,
the total amount of data that has to be processed consists of n2 variables. This is
a quite intensive computation load which can be managed efficiently thanks to the
distributed control structure that has been implemented (each generator computes
locally its own control signal). Moreover, taking into account that the exchange of
data has to take place in real time one can conclude that communication links of
efficient bit rate have to be established between the generators.
8.2 State Estimation-Based Control of Distributed PMSGs 343
It is considered that the third order model of the PMSG is connected to the power
grid thus forming the model of a Single Machine Infinite Bus (SMIB) system. As
explained in Chap. 3 PMSG mechanical dynamics can be represented as follows
δ̇ = ω
ω0 (8.1)
ω̇ = − 2J
D
(ω − ω0 ) + 2J
(Pm − Pe )
where δ is the turn angle of the generator’s rotor, ω is the rotation speed of the rotor
with respect to synchronous reference, ω0 is the synchronous speed of the generator,
J is the moment of inertia of the rotor, Pe is the active power of the generator, Pm is
the mechanical input torque to the generator which is associated with the mechanical
input power, D is the damping constant of the generator and Te is the electrical torque
which is associated to the generated active power. Moreover, the following variables
are defined: Δδ = δ − δ0 and Δω = ω − ω0 with ω0 denoting the synchronous
speed. The generator’s electrical dynamics has already been described in Chap. 1
and is summarized as follows [215, 249, 266]:
Ėq = 1
Td o
(Ef − Eq ) (8.2)
where Eq is the quadrature-axis transient voltage of the generator, Eq is the quadrature
axis voltage of the generator, Tdo is the direct axis open-circuit transient time constant
of the generator and Ef is the equivalent voltage in the excitation coil. The algebraic
equations of the synchronous generator are given by
xd
Eq = xd
Eq − (xd − xd ) xV s cos(Δδ)
d
Iq = Vs
xd
sin(Δδ)
Eq
Id = xd
− Vs
xd
cos(Δδ)
Vs Eq (8.3)
Pe = xd
sin(Δδ)
Vs Eq V2
Qe = cos(Δδ) − xds
xd
Substituting the electrical equations of the PMSG given in Eq. (8.3) into the equa-
tions of the electrical and mechanical dynamics of the rotor given in Eqs. (8.1) and
(8.2) respectively, the complete model of the Single Machine Infinite Bus model is
obtained
δ̇ = ω − ω0
Vs E
ω̇ = − 2J
D
(ω − ω0 ) + ω0 P2Jm − ω0 2J1 x q sin(Δδ) (8.4)
d
1 xd −xd
Ėq = − T1 Eq + Tdo xd
Vs cos(Δδ) + 1
Td o
Ef
d
x
where Td = xdd Tdo is the time constant of the field winding.
The previously analyzed single-machine infinite-bus model of the PMSG is
described by a nonlinear state-space model of the form
T
g(x) = 0 0 1
Tdo
(8.7)
with control input u = Ef being the field voltage (equivalent voltage in the excitation
coil) and measurable output the turn angle of the rotor
y = h(x) = δ − δ0 (8.8)
The interconnection between several local power generation units which are described
by the previously analyzed SMIB model results into a multi-area multi-machine dis-
tributed power generation system.
A multi-machine power system (MMPS) with n machines, in which the first
machine is chosen as the reference machine can be described by the following non-
linear dynamic model
8.2 State Estimation-Based Control of Distributed PMSGs 345
δ̇i = ωi − ω0
P
ω̇i = − 2JDi
(ωi − ω0 ) + ω0 2Jmii −
i
− ω0 2J1 i [G ii Eqi 2 + Eqi nj=1,j=i Eqj G ij sin(δi − δj − αij )] (8.9)
1 xdi −xdi
Ėq i = − T1 Eq i + Tdo i xd
Vsi cos(Δδi ) + 1
Td o i
Ef i
di i
where the electric torque Pei which is associated with the active power at the i-th
generator is now given by
n
Pei = G ii Eqi 2 + Eqi
j=1,j=i Eqj G ij sin(δi − δj − αij ) (8.10)
for i = 1, 2 . . . , n. For a power grid that consists of n generators the aggregate state
vector comprises the state vectors of the local machines, i.e. x = [x 1 , x 2 , . . . , x n ]T ,
where x i = [x1i , x2i , x3i ]T , with x1i = Δδi , x2i = Δωi and x3i = Eqi are the state
variables for the i-th machine and i = 1, 2, . . . , n.
where ⎛ i⎞ ⎛ ⎞
f1 0
⎝ i⎠ i
f (x) = f2 g (x) =
i ⎝ 0 ⎠ (8.13)
i 1
f3 Td o i
with
f1i (x) = x2i
ω0 2 2
f2i (x) = − 2JDi x2i + 2J Pmi − ω0 2J1 i {G ii x3i + x3i [G ii x3 i +
j
i
j (8.14)
+ x3i nj=1,j=i x3 G ij sin(x1i − x1 − aij )]}
xdi −xd
f3i (x) = − T1 x3i + 1
Td o i xd
i
Vsi cos(x1 i ).
di i
346 8 Synchronization and Stabilization of Distributed Power Generation Units
The procedure for linearization with the use of Lie algebra is as follows: The
following state vector is defined x = [x11 , x21 , x31 , x12 , x22 , x32 , x13 , x23 , x33 ]T while the
8.2 State Estimation-Based Control of Distributed PMSGs 347
state vector of the linearized equivalent of the distributed PMSG model is defined as
z∈R9×1 . It holds that z11 = h1 (x) = Δδ 1 , z21 = Lf h1 (x) = x21 and z31 = Lf2 h1 (x) = f21 .
In a similar manner one obtains z12 = h2 (x) = Δδ 2 , z22 = x22 and z32 = f22 . More-
over, in a similar manner one obtains z13 = h3 (x) = Δδ 3 , z22 = Lf h3 (x) = x22 and
z33 = Lf2 h2 (x) = f23 .
Furthermore, for the vectors defined as z3 = [z31 , z32 , z33 ], f˜ = [f21 , f22 , f23 ] and
h̃ = [h1 , h2 , h3 ] it holds that
or, equivalently
⎛ 1 ⎞ ⎛ ⎞
⎛ 1⎞ ∂f2 ∂f21 ∂f21 ∂f21 ∂f21 ∂f21 ⎛ ⎞
ż3 ∂x1 1
f + ∂x2 2
f + ··· + f
∂xn n ⎟ ∂x1 1
g ∂x2 2
g ··· g
∂xn 3 ⎟ u1
⎜ ⎜ ∂f
⎝ż3 ⎠ = ⎜ 2 f1 +
∂f ∂f22 ∂f22 ⎟ ⎜ 2 g1 ∂f22 ∂f2
g ⎟ ⎝u2 ⎠
2 2 2
2
⎝ ∂x13 ∂x2 2
f + ··· + f
∂xn n ⎠
+ ⎝ ∂x13 ∂x2 2
g ··· ∂xn 3 ⎠
ż33 ∂f2
f +
∂f23
f + ··· +
∂f23
f
∂f2
g
∂f23
g ···
∂f23
g u3
∂x1 1 ∂x2 2 ∂xn n ∂x1 1 ∂x2 2 ∂xn 3
(8.18)
Thus, it holds
ż3 = fa + Mu = v (8.19)
In this manner the initial nonlinear power system is transformed into two decou-
pled linear subsystems which are in the canonical Brunovksy form. For each one of
these subsystems i = 1, 2, 3 the appropriate control law is
i
vi = zd(3) − k3 (z̈i − z̈di ) − k2 (żi − żdi ) − k1 (zi − zdi ) − d̃ i (8.21)
where the feedback control gains k1 , k2 , k3 are chosen such that the associated char-
acteristic polynomial of the linearized system is a Hurwitz one.
348 8 Synchronization and Stabilization of Distributed Power Generation Units
The main principles of the differential flatness theory have been given in the previous
chapters as follows [137, 229, 241, 281, 352, 411, 456]: A finite dimensional system
is considered. This can be written in the general form of an ordinary differential
equation (ODE), i.e. Si (w, ẇ, ẅ, . . . , w(i) ), i = 1, 2, . . . , q. The term w denotes
the system variables (these variables are for instance the elements of the system’s
state vector and the control input) while w(i) , i = 1, 2, . . . , q are the associated
derivatives. Such a system is said to be differentially flat if there is a collection of m
functions y = (y1 , . . . , ym ) of the system variables and of their time-derivatives, i.e.
yi = φ(w, ẇ, ẅ, . . . , w(αi ) ), i = 1, . . . , m satisfying the following two conditions
[137, 229, 241, 281, 352, 411, 456]: (1) There does not exist any differential relation
of the form R(y, ẏ, . . . , y(β) ) = 0 which implies that the derivatives of the flat output
are not coupled in the sense of an ODE, or equivalently it can be said that the flat
output is differentially independent, (2) All system variables (i.e. the elements of the
system’s state vector w and the control input) can be expressed using only the flat
output y and its time derivatives wi = ψi (y, ẏ, . . . , y(γi ) ), i = 1, . . . , s.
Differential flatness for the stand-alone synchronous generator was shown in Chap. 3.
It will be proven that the multi-machine power generation system is also a dif-
ferentially flat one. As flat output of the distributed power generation system,
consisting of n PMSGs, the following vector is defined y = [y11 , y12 , . . . , y1n ] or
y = Δδ 1 , Δδ 2 , . . . , Δδ n . For the n-machines power generation system it holds
x11 = y1 , x12 = y2 , x13 = y3 , . . ., x1n = yn and x21 = Δω1 = ẏ1 , x22 = Δω2 = ẏ2 ,
x23 = Δω3 = ẏ3 , . . ., x2n = Δωn = ẏn . Moreover, it holds
ω0
ẋ2i = − 2J x + 2J
Di i
i 2
Pmi −
ω0 2
n j
i
j (8.22)
− 2Ji
[G ii x3i + x3 j=1,j=i [x3 G ij sin(x1i − x1 − αij )]
i
which means that one arrives at a function of the form ż3i = ai (x) + b1i (x)g1 u1 +
b2i (x)g2 u2 + b3i (x)g3 u3 + d̃ i , where in the case of the distributed power generation
that consists of n = 3 machines, and considering for instance i = 1, j = 2, 3 one has
) x2 + Di ω02 [G ii x3i + x3i nj=1,j=i x3 G ij sin(x1i − x1 − αij )]−
Di 2 i 2 j j
ai = ( 2J i (2Ji )
i +
n
ω0 x −x
− 2J [G x x
j
G sin(x i − x j − 4α )(− 1 x i + ( 1 di di V cos(x i ))]−
i
ii 3 j=1,j =i 3 ij 1 1 ij T 3 oi T i
d xd si 1
di
ω0 i
n j xd −xd
− 2J x j=1,j=i G ij sin(x1i − x1 − αij )(− T1 x3i + ( Td1 xid i Vsi cos(x1i ))−
i 3 oi i
ω0 i
n i − x j − α )x i ω0 x i
n
di
j j j j
− 2J x
i 3
x G
j=1,j=i 3 ij cos(x 1 1 ij 2 2Ji 3 j=1,j=i 3 ij cos(x1 − x1 − αij )x2
x G i
(8.25)
ω0
j j
b1i = − 2Ji
[2G ii x3i + nj=1,j=i x3 G ij sin(x1i − x1 − αij )] Td1
oi
ω0
b2i = − 2J i
G i2 sin(x1i − x12 − αi2 ) Td1 (8.26)
o2
ω0
b3i = − 2J i
G i3 sin(x1i − x13 − αi3 ) Td1
o3
Finally for the disturbance term one has d̃ i = − D2Ji ω20 Pmi + ω0 i
Ṗ
2Ji m
i
350 8 Synchronization and Stabilization of Distributed Power Generation Units
Thus, one has the following description of the dynamics of the i-th power generator
ż1i = z2i
ż2i = z3i (8.27)
ż3i = ai (x) + b1 i g1 u1 + b2 i g2 u2 + b3 i g3 u3 + d̃ i
ż31 = a1 (x) + b1 1 g1 u1 + b2 1 g2 u2 + b3 1 g3 u3 + d̃ 1
ż32 = a2 (x) + b1 2 g1 u1 + b2 2 g2 u2 + b3 2 g3 u3 + d̃ 2 (8.28)
ż33 = a3 (x) + b1 3 g1 u1 + b2 3 g2 u2 + b3 3 g3 u3 + d̃ 1
or in matrix form
ż3 = fa (x) + Mu + d̃ (8.29)
where z3 = [z31 , z32 , z33 ]T , u = [u1 , u2 , u3 ]T and d̃ = [d̃1 , d̃2 , d̃3 ]T while
⎛ ⎞ ⎛ 1 ⎞
a1 (x) b1 g1 b2 1 g2 b3 1 g3
fa (x) = ⎝a2 (x)⎠ , M = ⎝b1 2 g1 b2 2 g2 b3 2 g3 ⎠ (8.30)
a3 (x) b1 3 g1 b2 3 g2 b3 3 g3
Setting v = fa (x) + Mu + d̃, one obtains again the linear canonical form for the
i-th generator given by
⎛ i⎞ ⎛ ⎞⎛ i⎞ ⎛ ⎞
ż1 010 z1 0
⎝ż2i ⎠ = ⎝0 0 1⎠ ⎝z2i ⎠ + ⎝0⎠ (vi + d̃ i ) (8.31)
ż3i 000 z3i 1
In this manner the initial nonlinear power system is transformed into three decoupled
linear subsystems which are in the canonical Brunovksy form. For each one of these
subsystems the appropriate control law is
i
vi = zd(3) − k3 (z̈i − z̈di ) − k2 (żi − żdi ) − k1 (zi − zdi ) − d̃ i (8.32)
Since the disturbance term d̃ i , which is due to the time-varying mechanical input
torque at the i-th generator is unknown, in the computation of the control input of Eq.
(8.110) it will be substituted by its estimate d̃ˆ i , which will be provided by a disturbance
observer. For the implementation of the distributed power generation control scheme,
the controller at the i-th power generator makes use of not only its own state vector
X i = [x1i , x2i , x3i ]T , but also of the state vectors of the rest n − 1 power generators,
8.2 State Estimation-Based Control of Distributed PMSGs 351
j j j
i.e. x j = [x1 , x2 , x3 ], j=i, j = 1, 2, . . . , n. The transformation of the dynamical
model of the local power generators into the linear canonical form enables to obtain
estimates of the transformed state vector of the system Y i = [yi , ẏi , ÿi ]T where
yi = Δδ i , through Kalman Filtering, after processing measurements of only the turn
angle difference Δδ i of the i-th power generator. Therefore one has to compute the
estimates X̂ i = [x̂1i , x̂2i , x̂3i ]T after using the estimate provided by the Kalman Filter
Ŷ i = [ŷi , ẏˆ i , ÿˆ i ]T . As already explained, this is a highly data driven application and
requires the exchange and processing in real-time of a large amount of data. It holds
that
x̂1i = ŷi x̂2i = ẏˆ i (8.33)
while for the computation of x̂3i for i = 1, 2, . . . , n one has to solve with respect to
x̂3i the system of equations
D1 ˆ 1
ÿˆ 1 = − 2J ω0 ω0 2 j
1
ẏ + P
2J1 m1
− 2J1
+ +x̂31 nj=1,j=1 [x̂3 G 1j sin(y1 − yj − α1j )]
[G 11 x31
ÿˆ 2 = − 2J
Di ˆ 2 ω0 ω0 2 j
2
ẏ + 2J2
Pm2 − 2J2
[G 22 x32 + x32 nj=1,j=2 [x̂3 G 2j sin(y2 − yj − α2j )]
···
n
Dn ˆ n
ÿˆ n = − 2J ω0 ω0 j
n
ẏ + P
2Jn mn
− 2Jn
[G nn x3n 2 + x̂3n j=1,j=n [x̂3 G nj sin(y
i
− yj − αij )]
(8.34)
The computation of state estimates for the initial MIMO nonlinear system of the
interconnected power generators, out of the estimates obtained for its linearized
equivalent, can be also formulated as an optimization problem. To obtain state esti-
mates for the initial system one comes against a set of coupled nonlinear equations
described in Eq. (8.34). The latter can be solved with respect to the state vector
elements of the initial system using ordinary computation tools for optimization and
nonlinear programming [251, 307]. Moreover, by generalizing the results about state
estimation in stand-alone synchronous generator to the distributed MIMO case of the
interconnected electric power generators, it is possible to compute the state vector
estimate for the initial system, through a filter of the form described in Eq. (3.15) that
uses the inverse Jacobian matrix of the linearization transformation in its observation
gain.
The proposed control scheme, considers as output of the synchronous generator
the rotor’s difference angle Δδ. However, it is possible to avoid the use of encoder
readings about the rotor’s turn angle, and to indirectly estimate this parameter through
the processing of measurement coming from PMUs. Such measurements are the bus
voltage magnitude and the associated phase angle, the line current magnitude and
the associated angle and the electrical output power at the terminal bus [111, 130,
158, 186].
352 8 Synchronization and Stabilization of Distributed Power Generation Units
As already shown in Chap. 3, with the use of differential flatness theory the nonlinear
model of each PMSG can be written in the canonical form of Eqs. (3.11) and (3.54).
Thus one has the linear model
ẏf = Af yf + Bf v
(8.35)
zf = Cf yf
and the measurable variable y1 = δ is associated with the turn angle of the rotor. After
carrying out discretization of matrices Af , Bf and Cf with common discretization
methods one can perform Kalman Filtering on the linear equivalent model. This is
the Derivative-free nonlinear Kalman filtering [352, 353, 371].
The procedure for identifying additive disturbance inputs that affect the individual
power generators have been explained in Chap. 3. In the most generic case it can be
assumed that the mechanical input torque Pm varies in time [143, 208]. In that case
the aggregate disturbance input exerted on each generator’s model is
ω0
Tm = −ω0 (2J)
D
2 Pm + Ṗ
2J m (8.37)
It is also assumed that the dynamics of the disturbance term Tm is defined by its
n-th order derivative Tm(n) . Considering now that after expressing the system’s state
variables as functions of the flat outputs and their derivatives the PMSG’s dynamics
is given by
y(3) = fc (y, ẏ, ÿ) + gc (y, ẏ, ÿ)u + 2H
1
Ṗm or
(3) D 2 ω0
y = v − ω0 (2J) Pm + 2J Ṗm or (8.38)
(3)
y = v − Tm
8.2 State Estimation-Based Control of Distributed PMSGs 353
where
fc (y, ẏ, ÿ) = ( 2JD2 )ẏ − ω0 2J
D Pm
2J
+ D
ω V s x3 sin(ẏ)+
(2J)2 0 xd
ω0 Vs 1 −x
x sin(y) − ω2J0 xV s T1do dx d Vs cos(y)sin(y)−
x
2J xd Td 3 (8.39)
d d
ω0 Vs
− 2H x
x3 cos(y)ẏ
d
D ω0 Vs
gc (y, ẏ, ÿ) = − 2M Tdo x
sin(y) (8.40)
d
⎛ ⎞ ⎛ ⎞⎛ ⎞ ⎛ ⎞
ẏ1 010 y1 0
⎝ẏ2 ⎠ = ⎝0 0 1⎠ ⎝y2 ⎠ + ⎝0⎠ v − Tm (8.41)
ÿ3 000 y3 1
Next, the state vector of the model of Eq. (8.41) is extended to include as additional
state variables the disturbance input Tm [169, 292]. Then, in the new state-space
ω0
description one has z1 = y1 , z2 = y2 , z3 = y3 , z4 = Tm = −ω0 (2J) D
2 Pm + 2J Ṗm ,
z5 = Ṫm , and z6 = T̈m . Without loss of generality, the disturbance input dynamics is
assumed to be described by its third order derivative ż6 = Tm(3) .
Since the dynamics of the disturbance input are taken to be unknown in the design
of the associated disturbances’ estimator one has the following dynamics:
and the measurable state variable is z1 . It can be confirmed that the disturbance
observer model of the PMSG defined in Eq. (8.43) is observable. Defining as Ãd ,
B̃d , and C̃d , the discrete-time equivalents of matrices Ão , B̃o and C̃o respectively, a
Derivative-free nonlinear Kalman Filter can be designed for the aforementioned rep-
resentation of the system dynamics. The associated Kalman Filter-based disturbance
estimator is given by [34, 209, 334, 341, 363]
measurement update:
time update:
To compensate for the effects of the disturbance terms it suffices to use in the control
loop the modified control input, which actually compensates for the effects of the
external disturbance term Tm . This gives v∗ = v − T̂m or v∗ = v − ẑ4 .
To evaluate the performance of the proposed nonlinear control scheme, that uses
Kalman Filtering to estimate the nonmeasurable state vector elements of the distrib-
uted PMSGs and the external disturbances, simulation experiments have been carried
out. Different rotation speed setpoints had been assumed. Moreover, different input
torques (mechanical input power profiles) have been assumed to affect the dynamic
model of each local PMSG. Indicative ratings of the PMSG model are as follows:
direct axis synchronous reactance xd = 2.1 p.u., quadrature axis synchronous reac-
tance xd = 0.4 p.u., infinite bus voltage Vs = 1.0 p.u., direct axis open circuit time
constant Tdo = 5 s, rotor’s moment of inertia J = 50 kgm2 . The nonlinear control
scheme works efficiently even for distributed synchronous generators with uneven
ratings.
The case of distributed PMSG operation under unknown input power (torque)
has been examined. The input power at each local generator was considered to be a
disturbance input to the PMSG model and it was assumed that its change in time was
defined by the third derivative of the associated variable, i.e. Tm(3) where, according
ω0
to Eq. (8.37), Tm = −ω0 (2J) D
2 Pm + 2J Ṗm . The disturbance dynamics was completely
unknown to the controller and its identification was performed in real time by the
disturbance estimator. It is shown that the derivative-free nonlinear Kalman Filter
redesigned as a disturbance observer is capable of estimating simultaneously the
nonmeasurable state variables of the generator (ω and Eq ), as well as the unknown
disturbance input Tm . A first set of results is concerned with the tracking performance
of the control loop in case of piecewise constant disturbance input. The real value of
the state variable of the generator is plotted with blue color, the estimated values is
given with green color, while the associated setpoint is printed in red color. This is
shown in Figs. 8.2a, 8.3a and 8.4a. The estimation of the piecewise constant distur-
bance input for the first generator is shown in Fig. 8.2b, for the second generator is
given in Fig. 8.3b, while for the third generator it is shown in Fig. 8.4b.
Additionally, results about the tracking performance of the control loop in case of
a second piecewise constant disturbance input are shown Figs. 8.3a, 8.5a and 8.7a.
The estimation of the piecewise constant disturbance input for the first generator is
shown in Fig. 8.5b, for the second generator is given in Fig. 8.6b, while for the third
8.2 State Estimation-Based Control of Distributed PMSGs 355
(a) Generator 1
(b) Generator 1
0.15 0.35
0.25
0.05
Δω (p.u.)
0.2
0.15
0
0.1
−0.05
0.05
−0.1 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.2 Sensorless control of the PMSG No 1 under non-measurable (piecewise constant) mechan-
ical input torque in case of speed reference setpoint 1: a convergence of the real and estimated values
of the angular speed difference Δω b estimation of the external mechanical input torque Tm1
(a) Generator 2
(b) Generator 2
0.12 0.25
0.1
Estimation of input torque Tm (p.u.)
0.2
0.08
0.06
0.15
Δω (p.u.)
0.04
0.1
0.02
0
0.05
−0.02
−0.04 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.3 Sensorless control of the PMSG No 2 under non-measurable (piecewise constant) mechan-
ical input torque in case of speed reference setpoint 1: a convergence of the real and estimated values
of the angular speed difference Δω b estimation of the external mechanical input torque Tm2
generator it is shown in Fig. 8.7b. The units of the PMSG state variables have been
expressed again in p.u. (per unit system).
The simulation experiments have confirmed that the proposed state estimation-
based control scheme not only enables implementation of distributed PMSG control
through the measurement of a small number of variables (e.g. of only the rotor’s turn
356 8 Synchronization and Stabilization of Distributed Power Generation Units
(a) Generator 3
(b) Generator 3
0.25 0.7
0.2
0.6
0.05 0.4
0 0.3
−0.05
0.2
−0.1
0.1
−0.15
−0.2 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.4 Sensorless control of the PMSG No 3 under non-measurable (piecewise constant) mechan-
ical input torque in case of speed reference setpoint 1: a convergence of the real and estimated values
of the angular speed difference Δω b estimation of the external mechanical input torque Tm3
0.2
0.4
0.15
0.3
Δω (p.u.)
0.1
0.05
0.2
0
0.1
−0.05
−0.1 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.5 Sensorless control of the PMSG No 1 under non-measurable (piecewise constant) mechan-
ical input torque in case of speed reference setpoint 2: a convergence of the real and estimated values
of the angular speed difference Δω b estimation of the external mechanical input torque Tm1
angle) but also improves the robustness of the PMSG control loop in case of varying
speed setpoints and varying mechanical input torque.
Comparison to the Extended Kalman Filter, which is the most commonly used
nonlinear estimation technique, has shown that the Derivative-free nonlinear Kalman
Filter succeeds estimation of smaller variance for both the nonmeasurable state vector
elements of the PMSG and the term Tm expressing external torques and disturbances
8.2 State Estimation-Based Control of Distributed PMSGs 357
(a) Generator 2
(b) Generator 2
0.2 0.35
0.1
Δω (p.u.)
0.2
0.15
0.05
0.1
0
0.05
−0.05 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.6 Sensorless control of the PMSG no 2 under non-measurable (piecewise constant) mechan-
ical input torque in case of speed reference setpoint 2: a convergence of the real and estimated values
of the angular speed difference Δω b estimation of the external mechanical input torque Tm2
(a) Generator 3
(b) Generator 3
0.4 0.8
Estimation of input torque Tm (p.u.)
0.3 0.7
0.6
0.2
0.5
Δω (p.u.)
0.1
0.4
0
0.3
−0.1
0.2
−0.2 0.1
−0.3 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.7 Sensorless control of the PMSG No 3 under non-measurable (piecewise constant) mechan-
ical input torque in case of speed reference setpoint 2: a convergence of the real and estimated values
of the angular speed difference Δω b estimation of the external mechanical input torque Tm3
inputs. Estimation with the Extended Kalman Filter was based on local linearization
of the PMSG model and on the computation of the Jacobian of Eq. (8.6). The obtained
results are summarized in Table 8.1, and correspond to experiments (a) and (b)
associated with the two turn speed setpoints given in Figs. 8.2, 8.3, 8.4, 8.5, 8.6 and
8.7 under sinusoidal variations of Tm , and considering different noise levels.
358 8 Synchronization and Stabilization of Distributed Power Generation Units
8.3.1 Overview
In this section a new nonlinear H-infinity optimal control method for distributed
synchronous generators is developed and specific results are obtained for the problem
of synchronization and stabilization of a three-area multi-machine power system. The
nonlinear dynamical model of the interconnected synchronous generators undergoes
linearization round temporary operating points which are computed at each iteration
of the control algorithm. These operating points consist of the present value of the
system’s state vector and of the last value of the control inputs vector that was exerted
on it. This linearization procedure is based on the concept of Taylor series expansion
of the system’s states-space model and on the computation of the associated Jacobian
matrices [34, 341, 344]. The modelling error due to this approximate linearization is
considered to be a perturbation that is compensated by the robustness of the control
scheme [352, 376, 380].
The feedback control input is based on the solution of an H-infinity control prob-
lem [352, 376, 380]. This stands for a mini-max differential game between the control
signal and the disturbance terms that affect the model of the distributed generators.
The feedback control gain is found from the solution of an algebraic Riccati equation
at each iteration of the control algorithm [352, 376, 380]. The stability of the con-
trol scheme is demonstrated through Lyapunov analysis. Actually, it is shown that
the control loop satisfies an H-infinity tracking performance criterion which signi-
fies elevated robustness against model uncertainty and perturbation inputs. Besides,
under moderate conditions the asymptotic stability of the control scheme is proven.
Through simulation experiments it is further confirmed that the proposed control
method assures fast and accurate tracking of reference setpoints.
8.3 Nonlinear H-Infinity Control of Distributed Synchronous Generators 359
The state-space model of the multi-machine power system was analyzed in the previ-
ous sections. A multi-machine power system (MMPS) with n machines, in which the
first machine is chosen as the reference machine can be described by the following
nonlinear dynamic model [352, 375]
δ̇i = ωi − ω0
P
ω̇i = − 2Ji (ωi − ω0 ) + ω0 2Jmii −
Di
(8.46)
−ω0 2J1 i [G ii Eqi 2 + Eqi nj=1,j=i Eqj G ij sin(δi − δj − αij )]
xd −x
Ėq i = − T1 Eq i + Td1 xi di Vsi cos(Δδi ) + Td1 Ef i
di oi d i oi
where the electric torque Pei which is associated with the active power at the i-th
generator is now given by
n
Pei = G ii Eqi 2 + Eqi
j=1,j=i Eqj G ij sin(δi − δj − αij ) (8.47)
for i = 1, 2 . . . , n. For a power grid that consists of n generators the aggregate state
vector comprises the state vectors of the local machines, i.e. x = [x 1 , x 2 , . . . , x n ]T ,
where x i = [x1i , x2i , x3i ]T , with x1i = Δδi , x2i = Δωi and x3i = Eqi are the state
variables for the i-th machine and i = 1, 2, . . . , n.
Without loss of generality the three-machine power system of Fig. 8.8 is consid-
ered. The associated dynamic model is given by:
ẋ1 = x2 − ω0 (8.48)
P
ẋ2 = − 2JD1
(x2 − ω0 ) + ω0 2Jm11 −
ω0
1 (8.49)
− 2J1
{G 11 x32 + x3 [x6 G 12 sin(x1 − x4 − a12 ) + x9 G 13 sin(x1 − x7 − a13 )]
1 xd1 −xd1
ẋ3 = − T1 x3 + Tdo1 xd
Vs cos(x1 ) + 1
Td o 1
u1 (8.50)
d1 1
ẋ4 = x5 − ω0 (8.51)
P
ẋ5 = − 2JD2
(x5 − ω0 ) + ω0 2Jm22 −
ω0
2 (8.52)
− 2J2
{G 22 x62 + x6 [x3 G 21 sin(x4 − x1 − a21 ) + x9 G 23 sin(x4 − x7 − a23 )]
360 8 Synchronization and Stabilization of Distributed Power Generation Units
Fig. 8.8 A 3-area distributed power generation model consisting of 3 synchronous machines
1 xd2 −xd2
ẋ6 = − T1 x6 + Tdo2 xd
Vs cos(x4 ) + 1
Td o 2
u2 (8.53)
d2 2
ẋ7 = x8 − ω0 (8.54)
P
ẋ8 = − 2JD2
(x8 − ω0 ) + ω0 2Jm33 −
ω0
3 (8.55)
− 2J3
{G 32 x92 + x9 [x3 G 31 sin(x7 − x1 − a31 ) + x6 G 32 sin(x7 − x4 − a32 )]
1 xd3 −xd3
ẋ9 = − T1 x9 + Tdo3 xd
Vs cos(x7 ) + 1
Td o 3
u3 (8.56)
d3 3
where
f1 (x) = x2 − ω0 (8.58)
P
f2 (x) = − 2JD1
(x2 − ω0 ) + ω0 2Jm11 −
ω0
1 (8.59)
− 2J1
{G 11 x32 + x3 [x6 G 12 sin(x1 − x4 − a12 ) + x9 G 13 sin(x1 − x7 − a13 )]
xd1 −xd
f3 (x) = − T1 x3 + 1
Td o 1 xd
1
Vs cos(x1 ) (8.60)
d1 1
f4 (x) = x5 − ω0 (8.61)
P
f5 (x) = − 2JD2
(x5 − ω0 ) + ω0 2Jm22 −
ω0
2 (8.62)
− 2J2 {G 22 x6 + x6 [x3 G 21 sin(x4 − x1 − a21 ) + x9 G 23 sin(x4 − x7 − a23 )]
2
1 xd2 −xd2
f6 (x) = − T1 x6 + Tdo2 xd
Vs cos(x4 ) (8.63)
d2 2
f7 (x) = x8 − ω0 (8.64)
P
f8 (x) = − 2JD2
(x8 − ω0 ) + ω0 2Jm33 −
ω0
3 (8.65)
− 2J3 {G 32 x9 + x9 [x3 G 31 sin(x7 − x1 − a31 ) + x6 G 32 sin(x7 − x4 − a32 )]
2
1 xd3 −xd3
f9 (x) = − T1 x9 + Tdo3 xd
Vs cos(x7 ) (8.66)
d3 3
Local linearization is performed for the state-space model of the distributed power
generators, though Taylor series expansion, round the operating point (x ∗ , u∗ ) where
x ∗ is the present value of the system’s state vector and u∗ is the last value of the control
input that was exerted on the machine. Thus, one obtains the linearized description
ẋ = Ax + Bu + d̃ (8.68)
For the previous description of the distributed power generators’ model (3-area
3-machine model) by the state-space equation of Eq. (8.57) it holds that A =
∇x ∇x [f (x) + g(x)u] with
⎛ ∂f1 ∂f1 ⎞
∂f1 ∂f1
∂x1
· · · ∂x
∂x29 ∂x3
⎜ ∂f2 ∂f2 ⎟
∂f2
· · · ∂x ∂f2
⎜ ∂x1 ∂x29⎟ ∂x3
⎜ ∂f3 ∂f3 ⎟
∂f3 ∂f3
⎜ ∂x1 · · · ∂x
∂x2 ⎟∂x3
⎜ ∂f4 ∂f4 ⎟
9
A = ∇x [f (x) + g(x)u] = ⎜ ∂f4
· · · ∂x ⎟∂f4 (8.69)
⎜ ∂x1 ∂x29⎟ ∂x3
⎜· · · · · · · · · · · · · · ·⎟
⎜ ⎟
⎝· · · · · · · · · · · · · · ·⎠
∂f5 ∂f5 ∂f5 ∂f5
∂x1 ∂x2 ∂x3
· · · ∂x 9
∂f1
For the first row of the Jacobian matrix A = ∇x [f (x) + g(x)u] one has: ∂x1
= 0,
∂f1 ∂f1 ∂f1 ∂f1 ∂f1 ∂f1 ∂f1 ∂f1
∂x2
= 1, ∂x3
= 0, ∂x4
= 0, ∂x5
= 0, ∂x6
= 0, ∂x7
= 0, ∂x8
= 0, ∂x9
= 0.
For the second row of the Jacobian matrix A = ∇x [f (x) + g(x)u] one has:
∂f2 ω0 ∂f2
∂x1
= − 2J1
x3 [x6 G 12 cos(x1 − x4 − a13 )] + x9 G 13 cos(x1 − x7 − a13 ), ∂x2
= − 2J
D1
1
,
∂f2 −ω0
∂x3
= 2J1
{G 11 2x3 + [x6 ]G 12 sin(x1 − x4 − a12 ) + x9 G 13 sin(x1 − x7 − a13 )]},
∂f2 ω0 ∂f2 ∂f2 ω0
∂x4
= x x G cos(x1
2J1 3 6 12
− x4 − a12 ), ∂x5
= 0, ∂x6
= − 2J1
x3 G 12 sin(x1 − x4 − a12 ),
∂f2 ω0 ∂f2 ∂f2 ω0
∂x7
= x x G cos(x1
2J1 3 9 13
− x7 − a13 ), ∂x8
= 0, ∂x9
= − 2J1
x3 G 13 sin(x1 − x7 − a13 ).
∂f3
For the third row of the Jacobian matrix A = ∇x [f (x) + g(x)u] one has: ∂x1
=
1 xd1 −xd1 ∂f3 ∂f3 ∂f3 ∂f3 ∂f3 ∂f3
−
Tdo1 xd
Vs sin(x1 ), ∂x2
= 0, ∂x3
= − T1 , ∂x4
= 0, ∂x5
= 0, ∂x6
= 0, ∂x7
= 0,
1 d1
∂f3 ∂f3
∂x8
= 0, ∂x9
= 0.
∂f4
For the fourth row of the Jacobian matrix A = ∇x [f (x) + g(x)u] one has: ∂x1
= 0,
∂f4 ∂f4 ∂f4 ∂f4 ∂f4 ∂f1 ∂f1 ∂f1
∂x2
= 0, ∂x3
= 0, ∂x4
= 0, ∂x5
= 1, ∂x6
= 0, ∂x7
= 0, ∂x8
= 0, ∂x9
= 0.
∂f5
For the fifth row of the Jacobian matrix A = ∇x [f (x) + g(x)u] one has: ∂x1
=
ω0 ∂f5 ∂f5
x [x G cos(x4 − x1 − a21 )
2J2 6 3 21
+ x3 G 23 cos(x4 − x7 − a23 )], ∂x2
= 0, ∂x3
=
ω0 ∂f5 ω0
− 2J 2
x6 G 21 sin(x4 −x1 −a21 ), ∂x4
= − 2J2
x6 [x3 G 21 cos(x4 −x1 −a21 )+x9 G 23 cos(x4 −
8.3 Nonlinear H-Infinity Control of Distributed Synchronous Generators 363
∂f5 D2 ∂f5 ω0
x7 −a23 )], ∂x5
= − 2J ,
2 ∂x6
= − 2J2
{G 22 2x6 +[x3 G 21 sin(x4 −x1 −a21 )+x9 G 23 sin(x4 −
∂f5 ω0 ∂f5 ∂f5 ω0
x7 − a23 )]}, ∂x7
= 2J2
x6 x9 G 23 cos(x4 − x7 − a23 ), ∂x8
= 0, ∂x9
= − 2J2
x6 G 23 sin(x4 −
x7 − a23 ).
∂f6
For the sixth row of the Jacobian matrix A = ∇x [f (x) + g(x)u] one has: ∂x 1
= 0,
∂f6 ∂f6 ∂f6 xd2 −xd ∂f6 ∂f6 ∂f6 ∂f6
∂x2
= 0, ∂x3
=, ∂x4
= − Td1 xd
2
Vs sin(x4 ), ∂x5
= 0, ∂x6
= − T1 , ∂x7
= 0, ∂x8
= 0,
o2 2 d2
∂f6
∂x9
= 0.
∂f7
For the seventh row of the Jacobian matrix A = ∇x [f (x)+g(x)u] one has: ∂x1
= 0,
∂f7 ∂f7 ∂f7 ∂f7 ∂f7 ∂f7 ∂f7 ∂f7
∂x2
= 0, ∂x3
= 0, ∂x4
= 0, ∂x5
= 0, ∂x6
= 0, ∂x7
= 0, ∂x8
= 1, ∂x9
= 0.
∂f8
For the eight row of the Jacobian matrix A = ∇x [f (x) + g(x)u] one has: ∂x1
=
ω0 ∂f8 ∂f8
x [x G cos(x7 − x1 − a31 ) + x6 G 32 cos(x7 − x4 − a32 )], ∂x
2J3 9 3 31 2
= 0, = ∂x3
ω0 ∂f8 ω0 ∂f8 ∂f8
− 2J3 x9 G 31 sin(x7 − x1 − a31 ), ∂x4 = 2J3 x9 x6 G 32 cos(x7 − x4 − a32 ), ∂x5 = 0, = ∂x6
ω0 ∂f8 ω0
− 2J3 x9 G 23 sin(x7 −x4 −a32 ), ∂x7 = − 2J3 x9 [x3 G 31 cos(x7 −x1 −a31 )+x6 G 32 cos(x7 −
∂f8 D3 ∂f8 ω0
x4 −a32 )], ∂x 8
= − 2J3 , ∂x9 = − 2J 3
{G 33 2x9 +[x3 G 31 sin(x7 −x1 −a31 )+x6 G 32 sin(x7 −
x4 − a32 )]}.
∂f9
For the ninth row of the Jacobian matrix A = ∇x [f (x) + g(x)u] one has: ∂x 1
= 0,
∂f9 ∂f9 ∂f9 ∂f9 ∂f9 x −x ∂f9
= 0, ∂x = 0, ∂x = 0, ∂x = 0, ∂x = − Td1 x3 d3 Vs sin(x7 ), ∂x = 0, and
d
∂x2 3 4 6 7 o3 d 8
3
∂f9
finally ∂x9
= − T1 .
d3
Moreover, it holds that B = ∇u ∇x [f (x) + g(x)u] with
⎛ ⎞
0 0 0
⎜ 0 0 0 ⎟
⎜ ⎟
⎜g1 (x) 0 0 ⎟
⎜ ⎟
⎜ 0 0 0 ⎟
⎜ ⎟
B = ∇u [f (x) + g(x)u] = ⎜
⎜ 0 0 0 ⎟ ⎟ (8.70)
⎜ 0 g2 (x) 0 ⎟
⎜ ⎟
⎜ 0 0 0 ⎟
⎜ ⎟
⎝ 0 0 0 ⎠
0 0 g3 (x)
The design of an H-infinity controller was clearly analyzed in the case of stand-alone
synchronous generators. The initial nonlinear model of the distributed synchronous
generators is in the form
ẋ = f (x, u) x∈Rn , u∈Rm (8.71)
364 8 Synchronization and Stabilization of Distributed Power Generation Units
where matrices A and B are obtained from the computation of the Jacobians
⎛ ∂f ⎞ ⎛ ∂f ∂f ⎞
1 ∂f1 ∂f1 ∂f1
∂x ∂x
· · · ∂x 1
∂u1 ∂u2
1
· · · ∂u
⎜ ∂f21 ∂f22 n
∂f2 ⎟ ⎜ ∂f ∂f ∂f2 ⎟
m
⎜ ∂x1 ∂x2 · · · ∂x ⎟ ⎜ 2 2 · · · ∂u ⎟
A= ⎜ n ⎟ | ∗ ∗ B = ⎜ ∂u1 ∂u2
(x ,u ) m⎟| ∗ ∗
(x ,u ) (8.73)
⎝· · · · · · · · · · · ·⎠ ⎝· · · · · · · · · · · ·⎠
∂fn ∂fn ∂fn ∂fn ∂fn ∂fn
∂x1 ∂x2
· · · ∂xn ∂u1 ∂u2
· · · ∂um
and vector d̃ denotes disturbance terms due to linearization errors. The problem of
disturbance rejection for the linearized model that is described by
ẋ = Ax + Bu + L d̃
(8.74)
y = Cx
where x∈Rn , u∈Rm , d̃∈Rq and y∈Rp , cannot be handled efficiently if the classical
LQR control scheme is applied. This because of the existence of the perturbation
term d̃. The disturbance term d̃ apart from modeling (parametric) uncertainty and
external perturbation terms can also represent noise terms of any distribution.
In the H∞ control approach, a feedback control scheme is designed for trajectory
tracking by the system’s state vector and simultaneous disturbance rejection, con-
sidering that the disturbance affects the system in the worst possible manner. The
disturbances’ effect are incorporated in the following quadratic cost function:
1 T T
J(t) = 2 0
[y (t)y(t) + ruT (t)u(t) − ρ 2 d̃ T (t)d̃(t)]dt, r, ρ > 0 (8.75)
As in the case of nonlinear H-infinity control for stand-alone generators, in the case
of distributed synchronous generators the significance of the negative sign in the
cost function’s term that is associated with the perturbation variable d̃(t) is that the
disturbance tries to maximize the cost function J(t) while the control signal u(t) tries
to minimize it. The physical meaning of the relation given above is that the control
signal and the disturbances compete to each other within a mini-max differential
game. This problem of mini-max optimization can be written as
Again the objective of the optimization procedure is to compute a control signal u(t)
which can compensate for the worst possible disturbance, that is externally imposed
to the system. However, the solution to the mini-max optimization problem is directly
8.3 Nonlinear H-Infinity Control of Distributed Synchronous Generators 365
related to the value of the parameter ρ. This means that there is an upper bound in
the disturbances magnitude that can be annihilated by the control signal.
For the linearized system of the distributed generators given by Eq. (8.74) the cost
function of Eq. (8.75) is defined, where the coefficient r determines the penaliza-
tion of the control input and the weight coefficient ρ determines the reward of the
disturbances’ effects.
It is assumed that (i)The energy that is transferred from the disturbances signal
∞
d̃(t) is bounded, that is 0 d̃ T (t)d̃(t)dt < ∞, (ii) the matrices [A, B] and [A, L] are
stabilizable, (iii) the matrix [A, C] is detectable. Then, the optimal feedback control
law is given by
u(t) = −Kx(t) (8.77)
with
K = 1r BT P (8.78)
AT P + PA + Q − P( 1r BBT − 1
2ρ 2
LL T )P =0 (8.79)
where Q is also a positive definite symmetric matrix. The worst case disturbance is
given by
d̃(t) = ρ12 L T Px(t) (8.80)
The Lyapunov stability analysis for the H-infinity control of the distributed generators
follows the procedure that was explained in the case of the stand-alone generators.
Through Lyapunov stability analysis it will be shown that the proposed nonlinear
control scheme assures H∞ tracking performance for the distributed synchronous
generators, and that in case of bounded disturbance terms asymptotic convergence
to the reference setpoints is achieved.
The tracking error dynamics for the distributed synchronous generators is written
in the form
ė = Ae + Bu + L d̃ (8.81)
366 8 Synchronization and Stabilization of Distributed Power Generation Units
Fig. 8.9 Diagram of the control scheme for the distributed synchronous generators
where in the distributed synchronous generators’ case L = I∈R9×9 with I being the
identity matrix. Variable d̃ denotes model uncertainties and external disturbances of
the distributed generators’ model. The following Lyapunov equation is considered
V = 21 eT Pe (8.82)
V̇ = 21 [eT AT + uT BT + d̃ T L T ]Pe+
(8.84)
+ 21 eT P[Ae + Bu + L d̃]⇒
V̇ = 21 eT AT Pe + 21 uT BT Pe + 21 d̃ T L T Pe+
(8.85)
e PAe + 21 eT PBu + 21 eT PL d̃
1 T
2
8.3 Nonlinear H-Infinity Control of Distributed Synchronous Generators 367
Assumption: For given positive definite matrix Q and coefficients r and ρ there exists
a positive definite matrix P, which is the solution of the following matrix equation
AT P + PA = −Q + P( 2r BBT − 1
ρ2
LL T )P (8.87)
u = − 1r BT Pe (8.88)
V̇ = − 21 eT Qe − 1 T
2ρ 2
e PLL T Pe + eT PL d̃ (8.91)
or, equivalently
V̇ = − 21 eT Qe − 2ρ1 2 eT PLL T Pe+
(8.92)
+ 21 eT PL d̃ + 21 d̃ T L T Pe
Proof : The binomial (ρα − ρ1 b)2 is considered. Expanding the left part of the above
inequality one gets
ρ 2 a2 + 1 2
ρ2
b − 2ab ≥ 0 ⇒ 21 ρ 2 a2 + 1 2
2ρ 2
b − ab ≥ 0 ⇒
(8.94)
ab − 1 2
2ρ 2
b ≤ 21 ρ 2 a2 ⇒ 21 ab + 21 ab − 2ρ1 2 b2 ≤ 21 ρ 2 a2
The following substitutions are carried out: a = d̃ and b = eT PL and the previous
relation becomes
1 T T
2
d̃ L Pe + 21 eT PL d̃ − 1 T
2ρ 2
e PLL T Pe≤ 21 ρ 2 d̃ T d̃ (8.95)
368 8 Synchronization and Stabilization of Distributed Power Generation Units
Equation (8.95) is substituted in Eq. (8.92) and the inequality is enforced, thus giving
V̇ ≤ − 21 eT Qe + 21 ρ 2 d̃ T d̃ (8.96)
Equation (8.96) shows that the H∞ tracking performance criterion is satisfied. The
integration of V̇ from 0 to T gives
T T T
V̇ (t)dt≤ − 21 0 ||e||2Q dt + 21 ρ 2 0 ||d̃||2 dt⇒
0 T T (8.97)
2V (T ) + 0 ||e||2Q dt≤2V (0) + ρ 2 0 ||d̃||2 dt
Robust state estimation with the use of the H-infinity Kalman Filter was already
analyzed in the case of multi-phase machines. The control loop has to be implemented
with the use of information provided by a small number of sensors and by processing
only a small number of state variables. To reconstruct the missing information about
the state vector of the distributed synchronous generators it is proposed to use a
filtering scheme and based on it to apply state estimation-based control [375]. The
recursion of the H∞ Kalman Filter, for the model of the distributed SGs, can be
formulated in terms of a measurement update and a time update part
Measurement update:
Time update:
x̂ − (k + 1) = A(k)x(k) + B(k)u(k)
(8.101)
P− (k + 1) = A(k)P− (k)D(k)AT (k) + Q(k)
8.3 Nonlinear H-Infinity Control of Distributed Synchronous Generators 369
where it is assumed that parameter θ is sufficiently small to assure that the covariance
matrix P− (k) − θ W (k) + C T (k)R(k)−1 C(k) will be positive definite. When θ = 0
the H∞ Kalman Filter becomes equivalent to the standard Kalman Filter. One can
measure only a part of the state vector of the distributed SGs, such as state variables
xi = θi , i = 1, 2, 3 and can estimate through filtering the rest of the state vector
elements.
(a) (b)
1 1
ω1 (p.u,)
ω 1 (p.u,)
0.5 0.5
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
1 1
ω2 (p.u,)
ω 2 (p.u,)
0.5 0.5
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
1 1
ω3 (p.u,)
ω 3 (p.u,)
0.5 0.5
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
Fig. 8.10 Tracking of reference setpoints by the rotational speeds ωi i = 1, 2, 3 of the distributed
synchronous generators a setpoint 1 (p.u) b setpoint 2 (p.u)
(a) (b)
1 1
ω 1 (p.u,)
ω 1 (p.u,)
0.5 0.5
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
1 1
ω 2 (p.u,)
ω (p.u,)
0.5 0.5
2
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
1 1
ω 3 (p.u,)
ω 3 (p.u,)
0.5 0.5
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
Fig. 8.11 Tracking of reference setpoints by the rotational speeds ωi i = 1, 2, 3 of the distributed
synchronous generators a setpoint 3 (p.u) b setpoint 4 (p.u)
properties should hold (despite modeling uncertainties that may affect the power
generators) and synchronization between the individual power generators should
be succeeded (despite their exposure to external perturbations) [88, 102, 116, 198,
213, 237, 238, 495]. This section proposes adaptive fuzzy control based on differen-
tial flatness theory and using exclusively output feedback, for solving the problem
of synchronization between distributed synchronous generators and for succeeding
stabilization of a distributed power generation system. The model of the individual
8.4 Flatness-Based Adaptive Control of Distributed PMSGs 371
(a) (b)
1 1
ω 1 (p.u,)
ω (p.u,)
0.5 0.5
1
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
1 1
ω (p.u,)
ω (p.u,)
0.5 0.5
2
2
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
1 1
ω 3 (p.u,)
ω (p.u,)
0.5 0.5
3
0 0
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time (sec) time (sec)
Fig. 8.12 Tracking of reference setpoints by the rotational speeds ωi i = 1, 2, 3 of the distributed
synchronous generators a setpoint 5 (p.u) b setpoint 6 (p.u)
function for the proposed control scheme comprises three quadratic terms: (i) a term
that describes the tracking error of the generators’ state variables, (ii) a term that
describes the error in the estimation of the non-measurable state vector elements
with respect to the reference setpoints, and (iii) a sum of quadratic terms associated
with the distance of the weights of the neurofuzzy approximators from the values that
give the best approximation of the unknown dynamics of the generators. It is proven
that an adaptive (learning) control law can be found assuring that the first derivative
of the Lyapunov function will remain negative, thus assuring that the stability of
the control loop will be preserved and that accurate tracking of the setpoints by the
systems state variables will be succeeded (H-infinity tracking performance) [117,
215, 226, 263].
In the previous sections it was shown that, for the case of the N distributed synchro-
nous generators, the dynamic model of the i-th input-output linearized model of the
synchronous machine is given by
i
y(3) = ai (x) + b1i (x)g1 u1 + b2i (x)g2 u2 + b3i (x)g3 u3 (8.102)
which means that one arrives at a function of the form ż3i = ai (x) + b1i (x)g1 u1 +
b2i (x)g2 u2 + b3i (x)g3 u3 + d̃ i , where in the case of the distributed power generation
that consists of n = 2 machines, and considering for instance i = 1, j = 2 (Fig. 8.13)
one has
) x2 + Di ω02 [G ii x3i + x3i nj=1,j=i x3 G ij sin(x1i − x1 − αij )]−
Di 2 i 2 j j
ai = ( 2J i (2Ji )
i +
n
ω0 x −x
− 2J [G x x
j
G sin(x i − x j − α )(− 1 x i + ( 1 di di V cos(x i ))]−
i
ii 3 j=1,j =i 3 ij 1 1 ij T 3oi T i
d xd si 1
di
(8.103)
ω0 i
n j xd −xd
− 2J x j=1,j=i G ij sin(x1i − x1 − αij )(− T1 x3i + ( Td1 xid i Vsi cos(x1i ))−
i 3 oi i
ω0 i
n ω0 i
n
di
j j j j j
− 2J x j=1,j=i x3 G ij cos(x1i − x1 − αij )x2i 2J
i 3
x j=1,j=i x3 G ij cos(x1i − x1 − αij )x2
i 3
ω0
j j
b1i = − 2J [2G ii x3i + nj=1,j=i x3 G ij sin(x1i − x1 − αij )] Td1
i
ω0
oi (8.104)
b2i = − 2J i
G i2 sin(x1i − x12 − αi2 ) Td1
o2
Finally for the disturbance term one has d̃ i = − D2Ji ω20 Pmi + 2J
ω0 i
i
Ṗm . Thus, one has the
i
following description of the dynamics of the i-th power generator
8.4 Flatness-Based Adaptive Control of Distributed PMSGs 373
ż1i = z2i
ż2i = z3i (8.105)
ż3i = ai (x) + b1 i g1 u1 + b2 i g2 u2 + d̃ i
ż31 = a1 (x) + b1 1 g1 u1 + b2 1 g2 u2 + d̃ 1
(8.106)
ż32 = a2 (x) + b1 2 g1 u1 + b2 2 g2 u2 + d̃ 2
or in matrix form
ż3 = fa (x) + Mu + d̃ (8.107)
Setting v = fa (x) + Mu + d̃, one obtains again the linear canonical form for the
i-th generator given by
⎛ i⎞ ⎛ ⎞⎛ i⎞ ⎛ ⎞
ż1 010 z1 0
⎝ż2i ⎠ = ⎝0 0 1⎠ ⎝z2i ⎠ + ⎝0⎠ (vi + d̃ i ) (8.109)
ż3i 000 z3i 1
374 8 Synchronization and Stabilization of Distributed Power Generation Units
In this manner the initial nonlinear power system is transformed into two decoupled
linear subsystems which are in the canonical Brunovksy form. For each one of these
subsystems the appropriate control law is
i
vi = zd(3) − k3 (z̈i − z̈di ) − k2 (żi − żdi ) − k1 (zi − zdi ) − d̃ i (8.110)
It is assumed now that after defining the flat outputs of the initial MIMO nonlinear
system of the distributed synchronous generators, and after expressing the system
state variables and control inputs as functions of the flat output and of the associated
derivatives, the system can be transformed in the Brunovsky canonical form [375]:
ẋ1 = x2
ẋ2 = x3
···
ẋr1 −1 = xr1
y1 = x1
p
ẋr1 = f1 (x) + j=1 g1j (x)uj + d1 y2 = xr1 −1
(8.111)
ẋr1 +1 = xr1 +2 ···
ẋr1 +2 = xr1 +3 yp = xn−rp +1
···
ẋp−1 = xp
p
ẋp = fp (x) + j=1 gpj (x)uj + dp
p
yi(ri ) = fi (x) + j=1 gij (x)uj + dj (8.112)
Equivalently, in vector form, one has the following description for the system
dynamics
y(r) = f (x) + g(x)u + d (8.113)
where matrix A has the MIMO canonical form, i.e. with elements
⎛ ⎞
0 1 ··· 0
⎜0 0 · · · 0 ⎟
⎜ ⎟
⎜ ⎟
Ai = ⎜ ... ... · · · ... ⎟
⎜ ⎟
⎝0 0 · · · 1 ⎠ (8.115)
0 0 · · · 0 ri ×ri
BiT = 0 0 · · · 0 1 1×ri
Ci = 1 0 · · · 0 0 1×ri
ẋ = Ax + Bv + Bd̃
(8.117)
y = CT x
where v = f (x) + g(x)u. The reference setpoints for the system’s outputs y1 , . . . , yp
are denoted as y1m , . . . , ypm , thus for the associated tracking errors it holds
e1 = y1 − y1m
e2 = y2 − y2m
(8.118)
···
ep = yp − ypm
376 8 Synchronization and Stabilization of Distributed Power Generation Units
The error vector of the outputs of the transformed MIMO system is denoted as
E1 = [e1 , . . . , ep ]T
ym = [y1m , . . . , ypm ]T
··· (8.119)
(r)
ym(r) = [y1m (r) T
, . . . , ypm ]
(r)
where yim denotes the r-th order derivative of the i-th reference output of the MIMO
dynamical system. Thus, one can also define the following vectors: (i) a vector
containing the state variables of the system and the associated derivatives, (ii) a
vector containing the reference outputs of the system and the associated derivatives
r −1
x = [x1 , . . . , x1r1 −1 , . . . , xp , . . . , xpp ]T (8.120)
r1 −1 r −1
Ym = [y1m , . . . , y1m , . . . , ypm , . . . , ypm
p
]T (8.121)
while in a similar manner one can define a vector containing the tracking error of the
system’s outputs and the associated derivatives
r −1
e = Ym − x = [e1 , . . . , e1r1 −1 , . . . , ep , . . . , epp ]T (8.122)
It is assumed that matrix g(x) is a nonsingular one, i.e. g −1 (x) exists and is bounded
for all x∈Ux , where Ux ⊂Rn is a compact set. In any case, the problem of singularities
in matrix g(x) can be handled by appropriately modifying the state feedback-based
control input.
The objective of the adaptive fuzzy controller, denoted as u = u(x, e|θ ) is: all the
signals involved in the controller’s design are bounded and it holds that limt→∞ e = 0,
(ii) the H∞ tracking performance criterion is succeeded for a prescribed attenuation
level.
In the presence of non-gaussian disturbances wd , successful tracking of the refer-
ence signal is denoted by the H∞ criterion [352, 375]:
T T
0 eT Qedt ≤ ρ 2 0 wd T wd dt (8.123)
where ρ is the attenuation level and corresponds to the maximum singular value of
the transfer function G(s) of the linearized model associated to Eqs. (8.116) and
(8.117).
The control signal of the MIMO nonlinear system of the distributed synchronous
generators which has been transformed into the Brunovsky form as described by Eq.
(8.117) contains the unknown nonlinear functions f (x) and g(x). In case that the
8.4 Flatness-Based Adaptive Control of Distributed PMSGs 377
where
Φf (x) = (ξf1 (x), ξf2 (x), . . . ξfn (x))T (8.125)
with ξfi (x), ı = 1, . . . , n being the vector of kernel functions (e.g. normalized fuzzy
Gaussian membership functions), where
thus giving ⎛ ⎞
φf1,1 (x) φf1,2 (x) · · · φf1,N (x)
⎜ 2,1 2,2 2,N ⎟
⎜φ (x) φf (x) · · · φf (x)⎟
Φf (x) = ⎜ f ⎟ (8.127)
⎝ ··· ··· ··· ··· ⎠
n,1 n,2 n,N
φf (x) φf (x) · · · φf (x)
with ξgi (x), ı = 1, . . . , N being the vector of kernel functions (e.g. normalized fuzzy
Gaussian membership functions), where
ξgi (x) = φgi,1 (x), φgi,2 (x), . . . , φgi,N (x) (8.130)
thus giving ⎛ ⎞
φg1,1 (x) φg1,2 (x) · · · φg1,N (x)
⎜φg2,1 (x) φg2,2 (x) · · · φg2,N (x)⎟
Φg (x) = ⎜
⎝ ···
⎟ (8.131)
··· ··· ··· ⎠
φgn,1 (x) φgn,2 (x) · · · φgn,N (x)
It holds that ⎛ ⎞ ⎛ 1 p⎞
g1 g1 g12 · · · g1
⎜ g2 ⎟ ⎜ g 1 g22 · · · g2 ⎟
p
⎜ ⎟
g=⎝ ⎠=⎝ 2 ⎜ ⎟ (8.135)
··· ··· ··· ··· · · ·⎠
p
gn gn1 gn2 · · · gn
Using the above, one finally has the relation of Eq. (8.124), i.e. ĝ(x|θg ) = Φg (x)θg .
If the state variables of the system are available for measurement then a state-feedback
control law can be formulated as
where fˆ (x|θf ) and ĝ(x|θg ) are neurofuzzy models to approximate f (x) and g(x),
respectively. uc is a supervisory control term, e.g. H∞ control term that is used
to compensate for the effects of modelling inaccuracies and external disturbances.
Moreover, K T is the feedback gain matrix that assures that the characteristic poly-
nomial of matrix A − BK T will be a Hurwitz one.
The control of the system described by Eq. (8.113) becomes more complicated when
the state vector x is not directly measurable and has to be reconstructed through a
state observer. The following definitions are used
• error of the state vector e = x − xm
• error of the estimated state vector ê = x̂ − xm
• observation error ẽ = e − ê = (x − xm ) − (x̂ − xm )
When an observer is used to reconstruct the state vector, the control law of Eq.
(8.136) is written as
Applying Eq. (8.137) to the nonlinear system described by Eq. (8.113), results
into
and equivalently
e1 = C T e (8.141)
ê1 = C T ê (8.143)
The feedback gain matrix is denoted as K∈Rn×p . The observation gain matrix
is denoted as Ko ∈Rn×p and its elements are selected so as to assure the asymptotic
elimination of the observation error.
For the model of the 2-area distributed power generation system it holds that
(3)
x1,1 = f1 (x) + g1 (x)u
(3) (8.144)
x1,2 = f2 (x) + g2 (x)u
380 8 Synchronization and Stabilization of Distributed Power Generation Units
Therefore, the considered power generator system is a differentially flat one. Next,
taking into account also the effects of additive disturbances the dynamic model
becomes
(3)
x1,1 = f1 (x, t) + g1 (x, t)u + d1
(3) (8.147)
x1,2 = f2 (x, t) + g2 (x, t)u + d2
(3)
x1,1 f1 (x, t) g (x, t) d
(3) = + 1 u+ 1 (8.148)
x1,2 f2 (x, t) g2 (x, t) d2
where [uc1 uc2 ]T is a robust control term that is used for the compensation of the
model’s uncertainties and of external disturbances and KiT = [k1i , k2i , . . . , kn−1
i
, kni ].
Substituting Eq. (8.149) into Eq. (8.148) the closed-loop tracking error dynamics is
obtained −1
(3)
x1,1 f1 (x, t) g1 (x, t) ĝ1 (x, t)
(3) = + ·
x1,2 f2 (x, t) g2 (x, t) ĝ2 (x, t)
(3) (8.150)
xd1 fˆ1 (x, t) K1T uc1 d1
·{ − ˆ − e+ }+
xd2(3) f2 (x, t) K2T uc2 d2
When the estimated state vector x̂ is used in the feedback control loop, equivalently
to Eq. (8.140) one has
382 8 Synchronization and Stabilization of Distributed Power Generation Units
f1 (x, t) − fˆ1 (x̂, t)
ė = Ae − BK ê + Buc + B{
T
+
f2 (x, t) − fˆ2 (x̂, t)
(8.156)
g (x, t) − ĝ1 (x̂, t)
+ 1 u + d̃}
g2 (x, t) − ĝ2 (x̂, t)
The associated state observer will be described again by Eqs. (8.142) and (8.143).
or equivalently
ẽ1 = C T ẽ (8.160)
ẽ1 = C T ẽ (8.162)
8.4 Flatness-Based Adaptive Control of Distributed PMSGs 383
Fig. 8.14 Neurofuzzy approximator used for estimating the unknown dynamics of the distributed
generators
Next, the following approximators of the unknown dynamics of the distributed power
generators are defined
fˆ (x̂|θf ) x̂∈R6×1 fˆ1 (x̂|θf ) ∈ R1×1
fˆ (x̂) = ˆ1 (8.163)
f2 (x̂|θf ) x̂∈R6×1 fˆ2 (x̂|θf ) ∈ R1×1
where N is the number of fuzzy rules, n is the number of fuzzy sets in the antecedent
part of each rule, x̂ is the estimate of the state vector and μAij (x̂) is the i-th membership
function of the antecedent (IF) part of the i-th fuzzy rule. Similarly, the following
approximators of the unknown system dynamics are defined (Fig. 8.14)
ĝ (x̂|θg ) x̂∈R6×1 ĝ1 (x̂|θg ) ∈ R1×2
ĝ(x̂) = 1 (8.165)
ĝ2 (x̂|θg ) x̂∈R6×1 ĝ2 (x̂|θg ) ∈ R1×2
The value of the approximation error defined in Eq. (8.153) that corresponds to the
optimal values of the weights vectors θf∗ and θg∗ one has
w = f (x, t) − fˆ (x̂|θf∗ ) + g(x, t) − ĝ(x̂|θg∗ ) u (8.168)
where
wa = {[f (x, t) − fˆ (x̂|θf )] + [g(x, t) − ĝ(x̂|θg )]}u (8.171)
θ̃f = θf − θf∗
(8.173)
θ̃g = θg − θg∗
In case of the distributed synchronous generators, the adaptation law of the neuro-
fuzzy approximators weights θf and θg as well as the equation of the supervisory
control term uc are derived from the requirement for negative definiteness of the
Lyapunov function
V = 21 êT P1 ê + 21 ẽT P2 ẽ + θ̃ θ̃
1 T
2γ1 f f
+ 1
2γ2
tr[θ̃gT θ̃g ] (8.174)
The selection of the Lyapunov function is based on the following principle of indirect
adaptive control ê : limt→∞ x̂(t) = xd (t) and ẽ : limt→∞ x̂(t) = x(t). This yields
limt→∞ x(t) = xd (t). Substituting Eqs. (8.142), (8.143) and (8.159), (8.160) into
Eq. (8.174) and differentiating results into
8.4 Flatness-Based Adaptive Control of Distributed PMSGs 385
(8.178)
Assumption 1: For given positive definite matrices Q1 and Q2 there exist positive
definite matrices P1 and P2 , which are the solution of the following Riccati equations
[375]
(A − BK T )T P1 + P1 (A − BK T ) + Q1 = 0 (8.179)
T
(A − Ko C T ) P2 + P2 (A − Ko C T )−
(8.180)
−P2 B( 2r − ρ12 )BT P2 + Q2 = 0
The conditions given in Eqs. (8.179) to (8.180) are related to the requirement that
the systems described by Eqs. (8.142), (8.143) and (8.159), (8.160) have stable poles.
Substituting Eqs. (8.179) to (8.180) into V̇ yields
V̇ = 21 êT {(A − BK T )T P1 + P1 (A − BK T )}ê + ẽT CKoT P1 ê + 21 ẽT {(A − Ko C T )T P2 +
P (A − K C T )}ẽ + ẽT P B(u + w + d̃) + 1 θ̃˙ T θ̃ + 1 tr[θ̃˙ θ̃ ]
T
2 o 2 c γ1 f f γ2 g g
i.e.
V̇ = − 21 êT Q1 ê + ẽT CKoT P1 ê−
− 21 ẽT {Q2 − P2 B( 2r − ρ12 )BT P2 }ẽ + ẽT P2 B(uc + w + d̃)+ (8.181)
+ 1 θ̃˙ T θ̃ + 1 tr[θ̃˙ θ̃ ]
T
γ1 f f γ2 g g
386 8 Synchronization and Stabilization of Distributed Power Generation Units
1
ua = − ẽT P2 B + Δua (8.182)
r
where assuming that the measurable elements of vector ẽ are {ẽ1 , e˜3 , . . . , e˜k }, the
term Δua is such that
⎛ ⎞
p11 ẽ1 + p13 ẽ3 + · · · + p1k ẽk
⎜p13 ẽ1 + p33 ẽ3 + · · · + p3k ẽk ⎟
− 1r ẽT P2 B + Δua = − 1r ⎜
⎝
⎟
⎠ (8.183)
··· ······
p1k ẽ1 + p3k ẽ3 + · · · + pkk ẽk
which finally stands for a product between the state vector elements {ẽ1 , e˜3 , . . . , e˜k }
and the elements of matrix P2 which is obtained from the solution of the previous
Riccati equation.
• The control term ub is given by
or equivalently,
8.4 Flatness-Based Adaptive Control of Distributed PMSGs 387
Fig. 8.15 The proposed output feedback-based adaptive fuzzy control scheme for distributed syn-
chronous generators
V̇ = − 21 êT Q1 ê − 21 ẽT Q2 ẽ − 1 T
2ρ 2
ẽ P2 BBT P2 ẽ+
(8.186)
1 ˙T
θ̃ θ̃ + γ12 tr[θ̃˙ g θ̃g ]
T
+ ẽT P2 B(w + d̃ + Δua ) + γ1 f f
It holds that θ̃˙f = θ̇f − θ˙f∗ = θ˙f and θ̃˙g = θ̇g − θ˙g∗ = θ˙g . The following weights
adaptation laws are considered:
where assuming N fuzzy rules and associated kernel functions the matrices dimen-
sions are θf ∈RN×1 , θg ∈RN×2 , Φ(x)∈R2×N , B∈R6×3 , P∈R6×6 and ẽ∈R6×1 .
The update of θf is a gradient type algorithm. The update of θg is also a gra-
dient type algorithm, where uc implicitly tunes the adaptation gain γ2 [34, 344].
Substituting Eq. (8.187) in V̇ gives
388 8 Synchronization and Stabilization of Distributed Power Generation Units
V̇ = − 21 êT Q1 ê − 21 ẽT Q2 ẽ −
1 T
2ρ 2
ẽ P2 BBT P2 ẽ + BT P2 ẽ(w + d + Δua )+
(8.188)
+ γ11 (−γ1 )ẽT P2 BΦ(x̂)(θf − θf∗ ) + γ12 (−γ2 )tr[uẽT P2 BΦ(x̂)(θg − θg∗ )]
or
V̇ = − 21 êT Q1 ê − 21 ẽT Q2 ẽ − 1 T
2ρ 2
ẽ P2 BBT P2 ẽ + BT P2 ẽ(w + d̃ + Δua )+
+ γ11 (−γ1 )ẽT P2 BΦ(x̂)(θf − θf∗ ) + γ12 (−γ2 )tr[uẽT P2 B(ĝ(x̂|θg ) − ĝ(x̂|θg∗ )]
(8.189)
Taking into account that u ∈ R3×1 and ẽT PB(ĝ(x|θg ) − ĝ(x|θg∗ )) ∈ R1×3 it holds
V̇ = − 21 êT Q1 ê − 21 ẽT Q2 ẽ − 1 T
2ρ 2
ẽ P2 BBT P2 ẽ + BT P2 ẽ(w + d̃ + Δua )+
+ γ11 (−γ1 )ẽT P2 BΦ(x̂)(θf − θf∗ ) + γ12 (−γ2 )tr[ẽT P2 B(ĝ(x̂|θg ) − ĝ(x̂|θg∗ ))u]
(8.190)
Since ẽT P2 B(ĝ(x̂|θg ) − ĝ(x̂|θg∗ ))u∈R1×1 it holds
V̇ = − 21 êT Q1 ê − 21 ẽT Q2 ẽ − 1 T
2ρ 2
ẽ P2 BBT P2 ẽ + BT P2 ẽ(w + d̃ + Δua )+
(8.192)
+ γ11 (−γ1 )ẽT P2 BΦ(x̂)(θf − θf∗ ) + γ12 (−γ2 )ẽT P2 B(ĝ(x̂|θg ) − ĝ(x̂|θg∗ ))u
V̇ = − 21 êT Q1 ê − 21 ẽT Q2 ẽ − 1 T
2ρ 2
ẽ P2 BBT P2 ẽ+
(8.194)
+BT P2 ẽ(w + d̃) + ẽT P2 Bwα
w1 = w + d̃ + wα + Δua (8.195)
V̇ = − 21 êT Q1 ê − 21 ẽT Q2 ẽ − 1 T
2ρ 2
ẽ P2 BBT P2 ẽ + ẽT P2 Bw1 (8.196)
V̇ = − 21 êT Q1 ê − 21 ẽT Q2 ẽ − 1 T
2ρ 2
ẽ P2 BBT P2 ẽ+
(8.197)
+ 21 ẽT PBw1 + 1 T T
w B P2 ẽ
2 1
Proof : The binomial (ρa − ρ1 b)2 ≥ 0 is considered. Expanding the left part of the
above inequality one gets
The following substitutions are carried out: a = w1 and b = ẽT P2 B and the
previous relation becomes
1 T T
w B P2 ẽ
2 1
+ 21 ẽT P2 Bw1 − 1 T
2ρ 2
ẽ P2 BBT P2 ẽ
(8.200)
≤ 21 ρ 2 w1T w1
The above relation is used in V̇ , and the right part of the associated inequality is
enforced
1 1 1
V̇ ≤ − êT Q1 ê − ẽT Q2 ẽ + ρ 2 w1T w1 (8.201)
2 2 2
Thus, Eq. (8.201) can be written as
1 1
V̇ ≤ − E T QE + ρ 2 w1T w1 (8.202)
2 2
where
ê Q1 0
E= , Q= = diag[Q1 , Q2 ] (8.203)
ẽ 0 Q2
limt→∞ E(t) = 0 ⇒
limt→∞ ê(t) = 0 (8.206)
limt→∞ ẽ(t) = 0
The performance of the proposed adaptive fuzzy control scheme for distributed syn-
chronous generators was tested through simulation experiments. The state feedback
gain was K∈R6×2 . The basis functions used in the estimation of fi (x̂, t), i = 1, . . . , 3
x̂−cj 2
and gij (x̂, t), i = 1, . . . , 3 j = 1, . . . , 3 were μAj (x̂) = e( σ ) , j = 1, . . . , 2. Since
for each unknown function fi (x̂, t) and gij (x̂, t) there are 6 inputs and each one of them
consists of 3 fuzzy sets, for the approximation of the functions fi (x̂, t) i = 1, . . . , 3,
and gij (x̂, t) i = 1, . . . , 3, j = 1, . . . , 3 there will be 729 fuzzy rules of the form:
The estimation of the control input gain functions ĝij (x̂, t) i = 1, . . . , 3 was
derived in a similar way. The sampling period was taken to be 0.01 sec. In the begin-
ning of the training of the neuro-fuzzy approximators their weights were initialized
to zero. Moreover, the elements of the system’s state vector were also initialized to
zero. The positive definite matrices P1 ∈R6×6 and P2 ∈R6×6 stem from the solution of
the algebraic Riccati equations Eqs. (8.179) and (8.180), for Q1 and Q2 also positive
definite.
The approximations fˆ and ĝ were used in the derivation of the control law, given
by Eq. (8.137). To show the disturbance rejection capability of the proposed adaptive
8.4 Flatness-Based Adaptive Control of Distributed PMSGs 391
(a) (b)
0.9 0.55
0.8
ω 1 (p.u.)
u1 (p.u.)
0.7 0.5
0.6
0.5
0 2 4 6 8 10 12 14 16 18 20 0.45
t (sec) 0 5 10 15 20
t (sec)
0.6
0.9
0.8
u2 (p.u.)
0.55
ω 2 (p.u.)
0.7
0.5
0.6
0.5 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
Fig. 8.16 a Tracking of reference set-point 1 by the state variables ωi , i = 1, . . . , 2 of the distrib-
uted power generators’ model b Control inputs ui , i = 1, . . . , 2 applied to the individual power
generators
fuzzy controller, at the beginning of the second half of the simulation time additive
sinusoidal disturbances of amplitude equal to 10 % of the mean value of the control
inputs were applied to the model of the coupled synchronous generators.
The performance of the differential flatness theory-based adaptive fuzzy control
loop was tested in the case of tracking of different reference setpoints. The obtained
results are depicted in Figs. 8.16, 8.17, 8.18, 8.19, 8.20 and 8.21. The state variables
were measured in the per unit (p.u.) system. It can be observed that the proposed
adaptive fuzzy control scheme succeeded fast and accurate tracking of all these
setpoints.
The RMSE (root mean square error) of the examined control loop is also cal-
culated (assuming the same parameters of the controller) in the case of tracking of
the previous setpoints 1 to 5. The results are summarized in Table 8.2. The tracking
accuracy of the control method was remarkable despite the fact that (i) the dynamic
model of the power generators was completely unknown, (ii) only output feedback
was used in the implementation of the control scheme. It can be also seen that the
transient characteristics of the control scheme are quite satisfactory.
8.5.1 Outline
(a) (b)
0.6
1
0.9
u1 (p.u.)
0.55
ω (p.u.)
0.8
0.5
1
0.7
0.6 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
1 0.6
0.9 0.55
u2 (p.u.)
ω (p.u.)
0.8 0.5
2
0.7 0.45
0.6 0.4
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
Fig. 8.17 a Tracking of reference set-point 2 by the state variables ωi , i = 1, . . . , 2 of the distrib-
uted power generators’ model b Control inputs ui , i = 1, . . . , 2 applied to the individual power
generators
(a) (b)
1 0.55
0.9
ω 1 (p.u.)
0.8
u1 (p.u.)
0.7 0.5
0.6
0.5
0 2 4 6 8 10 12 14 16 18 20 0.45
0 5 10 15 20
t (sec) t (sec)
1 0.55
0.9
ω 2 (p.u.)
u2 (p.u.)
0.8
0.5
0.7
0.6
0.5 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
Fig. 8.18 a Tracking of reference set-point 3 by the state variables ωi , i = 1, . . . , 2 of the distrib-
uted power generators’ model b Control inputs ui , i = 1, . . . , 2 applied to the individual power
generators
power generation units in microgrids, such as photovoltaics and fuel cells requires the
interfacing of such devices with the electricity network through inverters [28, 259,
317, 324, 392, 393]. In this direction, several results on inverters’ control have been
obtained. In [386], control of photovoltaic inverters is analyzed in fault ride-through
situations and voltage sags conditions. In [223], predictive control is proposed for
voltage source inverters. A discrete-time model of the inverters is used to compute
8.5 Control and Synchronization of Distributed Inverters 393
(a) (b)
1 0.55
0.9
u1 (p.u.)
ω 1 (p.u.)
0.8
0.5
0.7
0.6
0.5 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
1 0.6
0.9
ω 2 (p.u.)
u2 (p.u.)
0.55
0.8
0.7
0.5
0.6
0.5 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
Fig. 8.19 a Tracking of reference set-point 4 by the state variables ωi , i = 1, . . . , 2 of the distrib-
uted power generators’ model b Control inputs ui , i = 1, . . . , 2 applied to the individual power
generators
(a) (b)
1 0.55
0.9
ω 1 (p.u.)
u1 (p.u.)
0.8
0.5
0.7
0.6
0.5 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
1 0.6
0.9
0.55
ω 2 (p.u.)
u2 (p.u.)
0.8
0.7
0.5
0.6
0.5 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
Fig. 8.20 a Tracking of reference set-point 5 by the state variables ωi , i = 1, . . . , 2 of the distrib-
uted power generators’ model b Control inputs ui , i = 1, . . . , 2 applied to the individual power
generators
the control law while Lyapunov stability analysis is also provided. In [164], a con-
trol method is proposed for the amplitude and frequency of the output of voltage
inverters, in the case of distributed (grid-tied) functioning of the inverters, as well
as in the case of stand-alone functioning (islanding). In [413], sliding mode con-
trol is proposed for active and reactive power regulation of a three-phase inverter
394 8 Synchronization and Stabilization of Distributed Power Generation Units
(a) (b)
1 0.55
0.9
u1 (p.u.)
ω 1 (p.u.)
0.8
0.5
0.7
0.6
0.5 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
1 0.6
0.9
u2 (p.u.)
ω 2 (p.u.)
0.55
0.8
0.7
0.5
0.6
0.5 0.45
0 2 4 6 8 10 12 14 16 18 20 0 5 10 15 20
t (sec) t (sec)
Fig. 8.21 a Tracking of reference set-point 6 by the state variables ωi , i = 1, . . . , 2 of the distrib-
uted power generators’ model b Control inputs ui , i = 1, . . . , 2 applied to the individual power
generators
The dynamic, model of the three-phase inverter, given in Chap. 1 is reviewed here
⎛ ωLf VLq (iI2 +iI2q ) ⎞
1 pf VLd +qf VLq
⎛ ⎞ ωVLq + 1
i
Cf Id
− Cf VL2 +VL2q
+ ωCf VLq − (VL2 +VL2q )
d
⎜ ⎟
⎜ ⎟
d d
VLd
⎜ VLq ⎟ ⎜−ωVLd + − 1 pf VLq −qf VLd
− ωCf VLd +
ωLf VLd (iI2 +iI2q ) ⎟
⎜ ⎟=⎜ ⎟
1
i d
d
⎝ iId ⎠ ⎜
Cf Iq Cf VL2 +VL2q (VL2 +VL2q ) ⎟+
dt ⎜ d d
⎟
⎜ ωiIq − 1
VLd ⎟
iIq ⎝ Lf ⎠
− ωiId − L1f VLq
⎛ ⎞
0 0
⎜ 0 0 ⎟ VId
⎜
+ ⎝ 1 0⎠ ⎟
Lf VIq
0 L1f
(8.209)
while the measurement equation of the inverter’s model is
⎛ ⎞
VLd
y1 VLd 1000 ⎜ ⎜ VLq ⎟
⎟
= = (8.210)
y2 VLq 0 1 0 0 ⎝ i Ld ⎠
i Lq
and by using the state variables notation x1 = VLd , x2 = VLq , x3 = iLd and x4 = iLq
one has
⎛ p x +q x ωL x (x 2 +x 2 )
⎞
⎛ ⎞ ωx2 + C1f x3 − C1f f x12 +x2f 2 + ωCf x2 − f(x22 +x3 2 ) 4
x1 ⎜ 1 2 1 2 ⎟
⎜ −qf x1 ωLf x1 (x32 +x42 ) ⎟
⎜x2 ⎟ ⎜−ωx1 + C1 x4 − C1 pf xx22 +x − ωC x + ⎟
⎜ ⎟=⎜ 2 f 1 (x12 +x22 ) ⎟
⎟+
d f f
⎝x3 ⎠ ⎜
1 2
dt
⎜ ωx − 1
x ⎟
x4 ⎝ 4 Lf 1 ⎠
− ωx3 − Lf x2 1 (8.211)
⎛ ⎞
0 0
⎜ 0 0 ⎟ u1
+⎜⎝ Lf 0 ⎠ u2
1 ⎟
0 L1f
ẋ = f (x) + G(x)u
(8.213)
y = h(x)
where Pim is the measured active power of the i-th power generation unit and Pid
is the desirable active power. Coefficient kpi is a “droop” gain which is practically
computed by dividing the range of variation of the inverter’s frequency (ωmax –ωmin )
by the maximum active power Pimax that the inverter can produce. Equation (8.214)
denotes that when representing the inverter as an equivalent synchronous generator
the turn speed should be proportional to the active power that the inverter provides.
Next, a relation is obtained for the measured active power Pim and the real active
power Pi of the inverter Pi . It holds that Pim (s) = e−sτpi P(s) (delay relation in fre-
quency domain) which after intermediate operations gives the differential equation
Δδ̇i = Δωi
(8.216)
τpi Ṗim = −Pim + Pi
kpi kpi
Δω̇i = τ pi
Pim − τ pi
Pi (8.220)
or equivalently
Ji Δω̇i = Pim − Pi (8.221)
where the virtual moment of inertia is defined as Ji = τpi /kpi . Additionally, from Eq.
(8.214) one has
By substituting Eq. (8.222) into Eq. (8.221) the following relation is obtained
The damping coefficient Dpi = k1p is defined. Using this coefficient in Eq. (8.222)
i
gives
Ji Δω̇i = −Dpi Δωi + Pid − Pi (8.224)
which is the equation that describes the rotational motion of a synchronous genera-
tor. In ideal conditions there is no interaction (power exchange) between distributed
power units connected to the same electricity grid. However, frequently such inter-
action exists and in the latter case Eq. (8.224) should be enhanced by including an
interaction term
n
Ji Δω̇i = −Dpi Δωi + (Pid − Pi ) + j=1,j=i G ij sin(δi − δj ) (8.225)
where δi is the virtual turn angle that is associated with the i-th power generation unit
(inverter). About the coupling coefficients G ij these are functions of the conductance
of the grid line which connects the i-th to the j-th power generation unit, as well as
of the grid voltage that is measured at points i and j respectively.
Thus, finally the dynamics of the i-th power generation unit is described as
8.5 Control and Synchronization of Distributed Inverters 399
In the design of the control and synchronization system for the distributed power
generation units, it is considered that the parameters Ji , Dpi and G ij are either known
or can be computed from measurements. Moreover, it is considered that the i-th
local controller not only processes measurements coming from the associated power
generation unit, but also uses measurements coming from the other power units which
are connected to the grid (that is the virtual turn angles δj ).
In Eq. (8.226), it has be shown that under certain conditions, the dynamics of the
inverter becomes equivalent to that of the synchronous power generator. Then one
has that the dynamics of the inverter is composed of two parts (i) the rotation part
and (ii) the electrical part.
The virtual rotation part is given by:
Ji is the virtual inertia, Dpi is the virtual damping coefficient, Pid is the desirable
active power of the i-th unit and Pi and the measured active power, G ij is a coupling
coefficient between the i-th and the j-th power generation unit and sin(δi − δj ) is
a synchronization index between the i-th and the j-th power generation unit [49,
396–398, 511, 512].
The electrical part has been defined in Chap. 6 and is given by:
⎛ ωLf VLq (iI2 +iI2q ) ⎞
⎛ ⎞ 1 pf VLd +qf VLq
ωVLq + 1
i
Cf Id
− Cf VL2 +VL2q
+ ωCf VLq − d
(VL2 +VL2q )
VLd ⎜ ⎟
⎜ VLq ⎟ ⎜ ωLf VLd (iI2 +iI2q ) ⎟
d d
1 pf VLq −qf VLd
⎜ ⎟=⎜ −ωVLd + − − ωCf VLd + ⎟
1
i d
⎟+
⎝ iId ⎠ ⎜
d
Cf Iq Cf VL2 +VL2q (VL2 +VL2q )
dt ⎜ d d
⎟
⎝ ωiIq − L1f
VLd ⎠
iIq
− ωiId − L1f VLq
⎛ ⎞
0 0
⎜ 0 0 ⎟ VId
⎜
+ ⎝ 1 0⎠ ⎟
Lf VIq
0 L1f
(8.228)
The synchronizing control approach for the i-th inverter makes use of Eq. (8.227)
and of the linearized inverter model given in Eq. (8.228). First, the value of Pi , that
is the active power that the i-th inverter should inject to the grid, is found from
the solution of the control problem of Eq. (8.227). Subsequently Pi is used in the
computation of the solution of the control problem of Eq. (8.228). This is depicted
in Fig. 8.22.
400 8 Synchronization and Stabilization of Distributed Power Generation Units
Fig. 8.22 Control loops for the virtual synchronous generator model and for the electrical part of
the inverter
⎛ i⎞ ⎛ ⎞⎛ i⎞ ⎛ ⎞
ż1 0100000 z1 000
⎜żi ⎟ ⎜0 0 0 0 0 0 0⎟ ⎜zi ⎟ ⎜1 0 0⎟ ⎛ i ⎞
⎜ 2i ⎟ ⎜ ⎟ ⎜ 2⎟ ⎜ ⎟
⎜ż ⎟ ⎜0 0 0 1 0 0 0⎟ ⎜zi ⎟ ⎜0 0 0⎟ v1i
⎜ i⎟ = ⎜
3 ⎟ ⎜ 3i ⎟ + ⎜ ⎟⎝ ⎠
⎜ż ⎟ ⎜0 0 0 0 0 0 0⎟ ⎜z ⎟ ⎜0 1 0⎟ v2i (8.229)
⎜ 4i ⎟ ⎜ ⎟ ⎜ 4i ⎟ ⎜ ⎟ v
⎝ż5 ⎠ ⎝0 0 0 0 0 1 0⎠ ⎝z5 ⎠ ⎝0 0 0⎠ 3
ż6i 0000000 z6i 001
and using Eq. (6.53) it holds for the i-th inverter’s model
8.5 Control and Synchronization of Distributed Inverters 401
Fig. 8.23 Distributed DC power generation units connected through inverters to the grid
A state estimator for each local power generation unit can be also designed in the form
of a disturbance observer. It is considered that the linearized model of the i-th inverter
is affected by additive input disturbances (which are considered to be modelling
uncertainties and external perturbations such as load changes, power injected due to
402 8 Synchronization and Stabilization of Distributed Power Generation Units
the connection of other power generation units to the grid, faults in the grid etc.).
With reference to the state-space description given in Eq. (8.229), the disturbance
terms can describe both modelling uncertainties and external perturbations.
The disturbances’ dynamics can be represented by the n-th order derivative of the
disturbances variables together with the associated initial conditions (however, since
the disturbances are going to be estimated by the Kalman Filter, the knowledge of
the initial conditions finally becomes obsolete). Thus the additive disturbances are
equivalently described in the form d̃1(n) = fd1 , d̃2(n) = fd2 and d̃3(n) = fd3 .
Without loss of generality it is assumed that in the relation d̃i(n) = fdi , the deriv-
ative’s order is n = 2. The state vector is extended by including as additional state
variables the disturbances and their derivatives. Thus, one has z1i = x1 , z2i = ẋ1 ,
zi = x , zi = ẋ , zi = x , zi = ẋ , zi = d̃ , zi = d̃˙ , zi = d̃ , zi = d̃˙ , zi = d̃ and
3 2 4 2 5 3 6 3 7 1 8 1 9 2 10 2 11 3
i
z12 = d̃˙ 3 .
It holds that ż1i = z2i , ż2i = z7 + v1i , ż3i = z4i , ż4i = z9 + v2i , ż5i = z6i , ż6i = z11
i
+ v3i ,
ż7 = z8 , ż8 = fd1 , ż9 = z10 , ż10 = fd2 , ż11 = z12 , ż12 = fd3 . Therefore, one has the
i i i i i i i i i
where the extended control inputs vector is ve = [v1i , v2i , v3i , fd1 , fd2 , fd3 ]T and
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
010000000000 000000 100
⎜0 0 0 0 0 0 1 0 0 0 0 0 ⎟ ⎜1 0 0 0 0 0 ⎟ ⎜0 0 0 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 0 1 0 0 0 0 0 0 0 0 ⎟ ⎜0 0 0 0 0 0⎟ ⎜0 1 0 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 0 0 0 0 0 0 1 0 0 0 ⎟ ⎜0 1 0 0 0 0 ⎟ ⎜0 0 0 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 0 0 0 1 0 0 0 0 0 0 ⎟ ⎜0 0 0 0 0 0⎟ ⎜0 0 1⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 0 0 0 0 0 0 0 0 1 0 ⎟ ⎜ 0 0 1 0 0 0 ⎟ T ⎜0 0 0 ⎟
⎜
Ae = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎟ Be = ⎜0 0 0 0 0 0⎟ Ce = ⎜0 0 0⎟ (8.235)
⎜0 0 0 0 0 0 0 1 0 0 0 0 ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 0 0 0 0 0 0 0 0 0 0⎟ ⎜0 0 0 1 0 0 ⎟ ⎜0 0 0 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 0 0 0 0 0 0 0 1 0 0 ⎟ ⎜0 0 0 0 0 0⎟ ⎜0 0 0 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 0 0 0 0 0 0 0 0 0 0⎟ ⎜0 0 0 0 1 0 ⎟ ⎜0 0 0 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝0 0 0 0 0 0 0 0 0 0 0 1⎠ ⎝0 0 0 0 0 0⎠ ⎝0 0 0 ⎠
000000000000 000001 000
For the extended state-space description of the system the state observer becomes
where for matrices Ae and Ce holds Ao = Ae and Co = Ce while for matrix Bo differs
from Be in the elements of the 10th and 12th row, where all elements are set to zero.
For the extended state-space description of the system it is possible to perform
state estimation using the Derivative-free nonlinear Kalman Filter. This filter con-
sists of the standard Kalman Filter recursion on the linearized equivalent model
of the i-th inverter, and of an inverse transformation based on differential flatness
theory (as explained in Sect. 6.2.3) which enables to obtain estimates for the state
variables of the initial nonlinear model of the inverter (if the latter is necessary). In
the filter’s algorithm, the previously defined matrices Ae ,Be and Ce are substituted
by their discrete-time equivalents Aed , Bed and Ced . This is done through common
discretization methods. The recursion of the Kalman Filter in this case is [34, 341,
344, 367]:
measurement update:
time update:
P− (k + 1) = Aed P(k)ATed + Q(k)
(8.238)
x̂ − (k + 1) = Aed x̂(k) + Bed v(k)
After identifying the disturbance terms, the control input of the inverter is modified
as follows:
v1i = z̈1 − kd1 (ż1 − ż1d ) − kp1 (z1 − z1d ) − ẑ7
v2i = z̈3 − kd2 (ż3 − ż3d ) − kp2 (z3 − z3d ) − ẑ9 (8.239)
v3i = z̈5 − kd3 (ż5 − ż5d ) − kp3 (z5 − z5d ) − ẑ11
The inclusion of the disturbance estimation terms ẑ7 , ẑ9 and ẑ11 in the feedback
control inputs enables to compensate for effects of the perturbations d̃1 , d̃2 and d̃3 .
The performance of the proposed distributed control scheme for the synchronization
of parallel inverters was tested through simulation experiments. A model of N = 3
distributed DC power generation units was considered, while each one of these units
was connected to the grid through an inverter (see Fig. 8.24). The power exchange
between each inverter and the grid was considered to be described by the model of
a synchronous generator.
The three interconnected inverters, shown in Fig. 8.24, are assumed to have differ-
ent model parameters which are described in Table 8.3. Moreover, the three invert-
ers are considered to be subjected to different perturbation inputs. Synchronization
404 8 Synchronization and Stabilization of Distributed Power Generation Units
means that given the desirable reference rotation speed ωi , one can compute the
amount of active power Pi contributed to the grid by each local inverter which in
turn results into ωi . Next, knowing Pi one can compute for the inner control loop of
Eq. (8.228) the associated voltage reference setpoints VLd and VLq and can solve the
control problem for the electrical part of the inverter’s dynamics. This means that
the synchronization problem of each local inverter is finally turned into a problem
of nonlinear feedback control for the associated electrical model of the inverter.
The main relation that the section considers for synchronization of the invert-
ers with a reference frequency of the grid is Eq. (8.227). According to this relation
the objective is that all inverters finally attain the same frequency ωi . By know-
ing the active power Pi which results into virtual rotation speed ωi , one can com-
pute also the setpoints for the output voltages VLd and VLq (see Fig. 8.22) which
have to be used in the control problem of the inverter’s electrical part described by
Eq. (8.228). One can compute these setpoints through Eqs. (1.98) and (1.99). In
8.5 Control and Synchronization of Distributed Inverters 405
(a) (b)
1.5 1
0.8
0.4
1
Gen 1ω (p.u.)
0.2
−0.2
0.5
−0.4
−0.6
−0.8
0 −1
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.25 a Angular speed of power generation unit 1 (red line setpoint, blue line real value), b
synchronization error between power generation units 1 and 2
(a) (b)
d/dt d1 (p.u.)
1 1 1
1
d1 (p.u.)
d/dt V L (p.u.)
0.6
q
0 0 0
0 10 20 30 40 0 10 20 30 40
d
0.4
−0.5 time time
0.2
d/dt d2 (p.u.)
1 1
d2 (p.u.)
0 −1
0 2 4 0 2 4
0.5 0.5
time time
1 1 0 0
0 10 20 30 40 0 10 20 30 40
d/dt V L (p.u.)
d/dt d3 (p.u.)
1 1
q
d3 (p.u.)
0 0
d
0.5 0.5
−0.5 −0.5
−1 −1 0 0
0 2 4 0 2 4 0 10 20 30 40 0 10 20 30 40
time time time time
Fig. 8.26 Inverter of power generation unit 1 a Voltage vector components VLd and VLq and their
derivatives, (red line setpoint, blue line real value, green line estimated value) b estimation (green
line) of disturbance inputs (blue line) and of their derivatives with the use of the Derivative-free
nonlinear Kalman Filter
Eq. (1.98) the active power should be given the value that is computed from the
solution of the control problem for the virtual synchronous generator model of Eq.
(8.227). Additionally, in Eq. (1.99) a reference value for the reactive power can be
used. The currents iL,d and iL,q are considered to be measurable. Thus one finally
has a set of two equations with unknowns the output voltages VLd and VLq . Solving
this system with respect to VLd and VLq provides reference setpoints for the output
voltages which finally result in the synchronization of the inverters with the grid.
406 8 Synchronization and Stabilization of Distributed Power Generation Units
(a) (b)
300 1
pf
0.9 qf
200
0.8
0.7
100
pf − qf (p.u.)
0.6
Vabc
0 0.5
0.4
−100
0.3
0.2
−200
0.1
−300 0
1 1.2 1.4 1.6 1.8 2 5 10 15 20 25 30 35 40
time (sec) t
Fig. 8.27 Inverter of power generation unit 1 a Three-phase voltage variables VL , (red line Va , blue
line Vb , green line Vc ) b active and reactive power of the inverter
(a) (b)
1.5 1
0.8
sync error DG2 −DG3 (p.u.)
0.6
0.4
1
Gen 2 ω (p.u.)
0.2
−0.2
0.5
−0.4
−0.6
−0.8
0 −1
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.28 a Angular speed of power generation unit 2 (red line setpoint, blue line real value), b
synchronization error between power generation units 2 and 3
(a) (b)
d/dt d1 (p.u.)
1 1
1
d1 (p.u.)
1
d/dt V L (p.u.)
0.8 0.5 0.5
0.5
V L (p.u.)
0.6 0 0
d
0 0 10 20 30 40 0 10 20 30 40
q
0.4
−0.5
time time
0.2
d/dt d2 (p.u.)
1 1
d2 (p.u.)
0 −1
0 2 4 0 2 4
0.5 0.5
time time
1 1 0 0
d/dt V L (p.u.)
0 10 20 30 40 0 10 20 30 40
0.5 0.5
V L (p.u.)
time time
d/dt d3 (p.u.)
q
0 0 1 1
d3 (p.u.)
q
−1 −1 0 0
0 2 4 0 2 4 0 10 20 30 40 0 10 20 30 40
time time time time
Fig. 8.29 Inverter of power generation unit 2 a Voltage vector components VLd and VLq and their
derivatives, (red line setpoint, blue line real value, green line estimated value) b estimation (green
line) of disturbance inputs (blue line) and of their derivatives with the use of the Derivative-free
nonlinear Kalman Filter
(a) (b)
300 1
pf
0.9 qf
200
0.8
0.7
100
pf − qf (p.u.)
0.6
V abc
0 0.5
0.4
−100
0.3
0.2
−200
0.1
−300 0
1 1.2 1.4 1.6 1.8 2 5 10 15 20 25 30 35 40
time (sec) t
Fig. 8.30 Inverter of power generation unit 2 a Three-phase voltage variables VL , (red line Va , blue
line Vb , green line Vc ) b active and reactive power of the inverter
the per unit (p.u.) system. It can be noticed that the proposed control and state estima-
tion scheme achieved both satisfactory transients and good tracking performance of
the reference setpoints. Moreover, fast synchronization between the distributed power
generation units was achieved. It is noted that the proposed control and synchroniza-
tion approach is scalable and can be applied to a larger number of N interconnected
408 8 Synchronization and Stabilization of Distributed Power Generation Units
(a) (b)
1.5 1
0.8
0.4
1
Gen 3 ω (p.u.)
0.2
−0.2
0.5
−0.4
−0.6
−0.8
0 −1
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
time time
Fig. 8.31 a Angular speed of power generation unit 3 (red line setpoint, blue line real value), b
synchronization error between power generation units 3 and 1
(a) (b)
d/dt d1 (p.u.)
1 1 1 1
d1 (p.u.)
d/dt V L (p.u.)
0.6
d
0 0 0
0.4 0 10 20 30 40 0 10 20 30 40
d
1 1
d2 (p.u.)
0 −1
0 2 4 0 2 4
0.5 0
time time
1 1 0 −1
0 10 20 30 40 0 10 20 30 40
d/dt V L (p.u.)
d/dt d3 (p.u.)
1 1
q
0 0
d3 (p.u.)
q
0.5 0.5
−0.5 −0.5
−1 −1 0 0
0 2 4 0 2 4 0 10 20 30 40 0 10 20 30 40
time time time time
Fig. 8.32 Inverter of power generation unit 3 a Voltage vector components VLd and VLq and their
derivatives, (red line setpoint, blue line real value, green line estimated value) b estimation (green
line) of disturbance inputs (blue line) and of their derivatives with the use of the Derivative-free
nonlinear Kalman Filter
DC power generation units. The simulation diagrams show convergence of the virtual
turn speed ωi of each inverter, to the associated reference value ωi∗ .
Additionally, it is noted that the section’s results confirm the usefulness of the
synchronverters concept for control and stabilization of distributed power generation
units, e.g. in the case of renewable energy systems [511]. The methodology developed
in this section enables a power corporation to control a synchronverter in the same
8.5 Control and Synchronization of Distributed Inverters 409
(a) (b)
300 1
pf
0.9 qf
200
0.8
0.7
100
pf − qf (p.u.)
0.6
Vabc
0 0.5
0.4
−100
0.3
0.2
−200
0.1
−300 0
1 1.2 1.4 1.6 1.8 2 5 10 15 20 25 30 35 40
time (sec) t
Fig. 8.33 Inverter of power generation unit 3 a Three-phase voltage variables VL , (red line Va , blue
line Vb , green line Vc ) b active and reactive power of the inverter