Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

1004 IEEE JOURNAL OF EMERGING AND SELECTED TOPICS IN POWER ELECTRONICS, VOL. 8, NO.

2, JUNE 2020

Frequency Stability of Synchronous Machines


and Grid-Forming Power Converters
Ali Tayyebi , Dominic Groß , Member, IEEE, Adolfo Anta, Friederich Kupzog,
and Florian Dörfler , Member, IEEE

Abstract— An inevitable consequence of the global power sys- renewable system is the gradual loss of synchronous machines
tem transition toward nearly 100% renewable-based generation (SMs), their inertia, and control mechanisms. This loss of the
is the loss of conventional bulk generation by synchronous rotational inertia changes the nature of the power system to
machines (SMs), their inertia, and accompanying frequency- and
voltage-control mechanisms. This gradual transformation of the a low-inertia network resulting in critical stability challenges
power system to a low-inertia system leads to critical challenges [1]–[3]. On the other hand, low-inertia power systems are char-
in maintaining system stability. Novel control techniques for con- acterized by large-scale integration of generation interfaced by
verters, so-called grid-forming strategies, are expected to address power converters, allowing frequency and voltage regulations
these challenges and replicate functionalities that, so far, have at much faster timescales compared with SMs [1], [4].
been provided by SMs. This article presents a low-inertia case
study that includes SMs and converters controlled under various Indeed, power converters are already starting to provide
grid-forming techniques. In this article, the positive impact of the new ancillary services, modifying their active and reactive
grid-forming converters (GFCs) on the frequency stability of SMs power outputs based on local measurements of frequency and
is highlighted, a qualitative analysis that provides insights into voltage. However, because of the dependence on frequency
the frequency stability of the system is presented, we explore measurements, these grid-following control techniques only
the behavior of the grid-forming controls when imposing the
converter dc and ac current limitations, the importance of the replicate the instantaneous inertial response of SMs after a
dc dynamics in grid-forming control design as well as the critical contingency with a delay and result in degraded performance
need for an effective ac current limitation scheme are reported, on the timescales of interest [5]. To resolve this issue, grid-
and finally, we analyze how and when the interaction between forming converters (GFCs) are envisioned to be the corner-
the fast GFC and the slow SM dynamics can contribute to the stone of future power systems. Based on the properties and
system instability.
functions of SMs, it is expected that GFCs must support load-
Index Terms— Frequency stability, grid-forming converters sharing/drooping, black start, inertial response, and hierarchi-
(GFCs), low-inertia power system, synchronous machines (SMs). cal frequency/voltage regulation. While these services might
not be necessary for a future converter-based grid, a long
I. I NTRODUCTION transition phase is expected, where SMs and GFCs must be

A T THE heart of the energy transition is the change in


generation technology from fossil fuel-based generation
to converter interfaced renewable generation [1]. One of the
able to interact and ensure system stability.
Several grid-forming control strategies have been proposed
in recent years [4]. Droop control mimics the speed droop
major consequences of this transition toward a nearly 100% mechanism present in SMs and is a widely accepted baseline
solution [6]. As a natural further step, the emulation of SM
Manuscript received July 14, 2019; revised October 8, 2019 and
December 5, 2019; accepted December 30, 2019. Date of publication dynamics and control led to so-called virtual SM (VSM)
January 14, 2020; date of current version May 6, 2020. This work was strategies [7]–[9]. Recently, matching control strategies that
supported in part by ETH Zürich Funds, and by the European Union’s Horizon exploit structural similarities of converters and SM and match
2020 Research and Innovation Program under Grant Agreement 691800. This
article reflects only the authors’ views and the European Commission is their dynamic behavior have been proposed [10]–[13]. In con-
not responsible for any use that may be made of the information it con- trast, virtual oscillator control (VOC) uses GFCs to mimic
tains. Recommended for publication by Associate Editor Lennart Harnefors. the synchronizing behavior of the Liénard-type oscillators and
(Corresponding author: Ali Tayyebi.)
Ali Tayyebi is with the Austrian Institute of Technology (AIT), 1210 Vienna, can globally synchronize a converter-based power system [14].
Austria, and also with the Automatic Control Laboratory, ETH Zürich, However, the nominal power injection of VOC cannot be
8092 Zürich, Switzerland (e-mail: ali.tayyebi-khameneh@ait.ac.at). specified. This limitation is overcome by dispatchable VOC
Dominic Groß is with the Department of Electrical and Computer Engi-
neering, University of Wisconsin–Madison, Madison, WI 53706 USA (e-mail: (dVOC) [15]–[17] that ensures synchronization to a prespeci-
dominic.gross@wisc.edu). fied operating point that satisfies the ac power flow equations.
Adolfo Anta and Friederich Kupzog are with the Austrian Institute In this article, the dynamics of the converter dc-link capac-
of Technology (AIT), 1210 Vienna, Austria (e-mail: adolfo.anta@ait.ac.at;
friederich.kupzog@ait.ac.at). itor, the response time of the dc power source, and its current
Florian Dörfler is with the Automatic Control Laboratory, ETH Zürich, limits are explicitly considered. We review four different grid-
8092 Zürich, Switzerland (e-mail: dorfler@ethz.ch). forming control strategies and combine them with standard
Color versions of one or more of the figures in this article are available
online at http://ieeexplore.ieee.org. low-level cascaded control design accounting for the ac volt-
Digital Object Identifier 10.1109/JESTPE.2020.2966524 age control and the ac current limitation and control [18].
2168-6777 © 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See https://www.ieee.org/publications/rights/index.html for more information.

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
TAYYEBI et al.: FREQUENCY STABILITY OF SMs AND GRID-FORMING POWER CONVERTERS 1005

where Cdc denotes the dc-link capacitance, G dc models dc


losses, and, L, C, and R, respectively, denote the filter’s induc-
tance, capacitance, and resistance. Moreover, v dc represents the
dc voltage, i dc is the current flowing out of the controllable dc
current source, mαβ denotes the modulation signal of the full-
bridge averaged switching stage model, i x := (1/2)m αβ i s,αβ
Fig. 1. Converter model in αβ-coordinates with detailed dc energy source
denotes the net dc current delivered to the switching stage,
model based on (1)–(3). i s,αβ and vs,αβ := (1/2)mαβ v dc , respectively, are the ac
switching node current and voltage (i.e., before the output
We explore the various performance aspects of GFC control filter), and i αβ and vαβ are the output current and voltage.
techniques in an electromagnetic transients (EMTs) simulation To obtain a realistic model of the dc energy source,
of the IEEE nine-bus test system: 1) the impact of GFCs on the we model its response time by a first-order system

frequency performance metrics, e.g., nadir and rate of change τdc i̇ τ = i dc − iτ (2)
of frequency (RoCoF) [19]–[22]; 2) the response of GFCs
where is the dc current reference, τdc is the dc source time
i dc
under large load disturbances; 3) their behavior when imposing
dc and ac current limitations; and 4) their response to the constant, and i τ denotes the current provided by the dc source.
loss of SM and performance in a pure converter-based system. Moreover, the dc source current limitation is modeled by the
Furthermore, we provide an insightful qualitative analysis of saturation function

the simulation results. The models used in this article are   iτ , if |i τ | < |i max
dc |
available online [23]. i dc = sat i τ , i max =
dc
(3)
sgn(i τ )i max
dc , if |i | ≥ |i dc |
τ max
This article highlights the positive impact of GFCs on
improving the standard power system frequency stability met- dc
where i max is the maximum dc source current. Note that
rics. Moreover, we observe that limiting the GFCs dc or ac we implicitly assume that some storage element is present
current accompanied by the interaction of fast converters so that the dc source can support bidirectional power flow.
and slow SM dynamics can destabilize some grid-forming In practice, the limit imposed by (3) corresponds to current
controls. This observation highlights the importance of the dc limits of a dc–dc converter, current limits of an energy storage
dynamics in grid-forming control design as well as the critical system, or PV/wind power generation limits. The impact of
need for an ac current limiting mechanism. Furthermore, the dc source limitation (3) is investigated in Section IV. It is
we reveal a potentially destabilizing interaction between the noteworthy that the converter must also limit its ac current
fast synchronization of GFCs and the slow response of SMs to protect its semiconductor switches [26]. This ac current
(see [16], [24] for a similar observation on the interaction of limitation is typically imposed via converter control design
GFCs and line dynamics). Finally, this article shows that an (see Section III for details).
all-GFCs (i.e., no-inertia) system can exhibit more resilience
than a mixed SM-GFCs (i.e., low-inertia) system with respect B. SM Model
to the large load variations.
In this article, we adopt an eighth-order (i.e., including six
The remainder of this article is structured as follows.
electrical and two mechanical states), balanced, symmetrical,
Section II reviews the modeling approach. Section III presents
three-phase SM with a field winding and three damper wind-
system dynamics and adopted grid-forming control techniques.
ings on the rotor [2, Fig. 3.1]
The simulation-based analysis of the study is discussed in
Section IV. Section V presents a qualitative analysis of the θ̇ = ω (4a)
observations made in case studies. The concluding statements J ω̇ = Tm − Te − T f (4b)
and agenda of future work are reported in Section VI. The
choice of control parameters is described in the Appendix. ψ̇s,dq = vs,dq − rs i s,dq − J2 ψs,dq (4c)
ψ̇ f,d = v f,d − r f,d i f,d (4d)
II. M ODEL D ESCRIPTION ψ̇ D = v D − R D i D (4e)
Throughout this study, we use a test system comprised of where θ denotes the rotor angle, J is the inertia constant, ω is
power converters and SMs. This section describes the models the rotor speed, and Tm , Te , and T f w denote the mechanical
of the individual devices and components [23]. torque, electrical torque, and the friction windage torque (see
[2, Sec. 5.7]). Moreover, ψs,dq = [ψs,d ψs,q ] , vs,dq =
A. Converter Model [v s,d v s,q ] , and i s,dq = [i s,d i s,q ] denote the stator winding
flux, voltage, and current in dq-coordinates (with angle θ as
To begin with, we consider the converter model illustrated in (4a)), J2 = [ 01 −1 ◦
0 ] denotes the 90 rotation matrix, and
in Fig. 1 in αβ-coordinates [13], [25] ψ f,d , v f,d , and i f,d denote the d-axis field winding flux,
voltage, and current. Furthermore, rs and r f,d denote the stator
Cdc v̇ dc = i dc −G dc v dc − i x (1a)
and field winding resistances, ψ D = [ψ1d ψ1q ψ2q ] , v D =
L i̇ s,αβ = vsαβ − Ri s,αβ − vαβ (1b) [v 1d v 1q v 2q ] , and i D = [i 1d i 1q i 2q ] are the linkage flux,
C v̇αβ = i s,αβ − i αβ (1c) voltage, and current associated with three damper windings,

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
1006 IEEE JOURNAL OF EMERGING AND SELECTED TOPICS IN POWER ELECTRONICS, VOL. 8, NO. 2, JUNE 2020

Fig. 3. IEEE nine-bus test system with an SM, two large-scale multiconverter
systems (i.e., aggregate GFCs), and constant impedance loads.
TABLE I
Fig. 2. Equivalent model of an individual converter module (left), large-
C ASE S TUDY M ODEL AND C ONTROL PARAMETERS [23]
scale multiconverter system consisting of n identical modules (middle), and
aggregate model (right).

and R D = diag(r1d , r1q , r2q ) denotes the diagonal matrix


of the damper winding resistances. Note that the windage
friction (i.e., modeling friction between the rotor and air)
term is commonly expressed as a speed dependent term, e.g.,
T f = D f ω [2, Sec. 5.7], and it is typically negligible for
system-level studies and is included here for the sake of
completeness and to highlight structural similarities of the
SM and two-level voltage-source converter with the control
presented in Section III-E. Furthermore, the damping torque
associated with the damper windings is included in the SM
model (4) via the damper winding dynamics (4e). For more
details on the SM modeling and parameters’ computation,
the reader is referred to [2, Sec. 3.3] and [3, Chap. 4].
We augment the system (4) with an ST1A type excitation
dynamics with built-in automatic voltage regulator (AVR)
[27, Fig. 21]. To counteract the well-known destabilizing
effect of the AVR on the synchronizing torque, we equip the
system with a simplified power system stabilizer comprised
of a two-stage lead-lag compensator [3, Sec. 12.5]. Finally, RLC dynamics), model the transformers via three-phase linear
the governor and turbine dynamics are, respectively, modeled transformer models, and consider constant impedance loads
by proportional speed droop control and first-order turbine (see Table I for the parameters). We emphasize that the line
dynamics dynamics cannot be neglected in the presence of GFCs due to
potential adverse interactions between their fast response and
p = p  + d p (ω − ω) (5a) the line dynamics [16], [24], [29].
τg ṗτ = p − pτ (5b) Remark 1 (Aggregate Converter Model): In this case study,
each GFC in Fig. 3 is an aggregate model of 200 com-
where p denotes the power set point, p is the governor mercial converter modules (see Table I for the parameters).
output, d p denotes the governor speed droop gain, and ω Each module is rated at 500 kVA, and the aggregate model
and ω denote nominal and measured frequencies, respectively. is rated at 100 MVA, which is equal to the SM rating.
Furthermore, τg denotes the turbine time constant and pτ Each module is interfaced to a medium voltage line via an
denotes the turbine output power. We refer the reader to LV-MV transformer (see Fig. 2). We derive the parameters
[24, Fig. 2] for an illustration of the interplay between the of the aggregate transformer model by assuming a parallel
SM model, the excitation dynamics, the PSS, and governor connection of 100 commercial transformers rated at 1.6 MVA
dynamics. Finally, the MATLAB/Simulink implementation of (see Table I). A detailed presentation and derivation of the
the SM model can be found in [23]. model aggregation are out of the scope of this article, but we
follow developments analogous to [30]–[32] in deriving the
C. Network Model equivalent aggregate converter parameters.
To study the transmission-level dynamics of a low-inertia
power system, we use Sim Power Systems to perform an EMT III. G RID -F ORMING C ONTROL A RCHITECTURES
simulation of the IEEE nine-bus test system shown in Fig. 3 Grid-forming control strategies control (see Fig. 4) a con-
[4], [28]. We model the lines via nominal π sections (i.e., with  for the dc energy
verter through the reference current i dc

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
TAYYEBI et al.: FREQUENCY STABILITY OF SMs AND GRID-FORMING POWER CONVERTERS 1007

Fig. 4. Grid-forming control architecture with reference models described in Sections III-C–III-F.

source (1a) and the modulation signal mαβ for the dc–ac 
where ī s,dq denotes the limited reference current that preserves
conversion stage (1) (see Fig. 1). In the following, we briefly 
the direction of i s,dq ac /i  ). We emphasize
and γi := (i max s,dq
review the low-level cascaded control design (i.e., ac volt- that limiting the ac current can have a strong impact on
age control, and current limitation and control) for two- the stability margins and dynamics of grid-forming power
level voltage-source converters that track a voltage reference converters [34]. While numerous different strategies have been
provided by a reference model (i.e., grid-forming control). proposed to limit the ac current injection of voltage-source
Moreover, we propose a controller for the converter dc voltage converters with grid-forming controls [26], [33], [35]–[39],
that defines the reference dc current. Because their design is the problem of designing a robust ac current limitation strategy
independent of the choice of the reference model, we first that effectively reacts to the load-induced overcurrent and grid
discuss the cascaded voltage/current control and the dc-side faults is an open research problem. Moreover, complex current
control. Subsequently, we review four common grid-forming limitation strategies typically require careful tuning of the con-
control strategies. For each strategy, we describe the angle trollers. Therefore, to provide a clear and concise investigation
dynamics, frequency dynamics, and ac voltage magnitude of the behavior of the existing grid-forming control solutions,
regulation. Throughout this section, we will employ the three we use the simple ac current limiting strategy (7).
phase abc, αβ and dq-coordinates (see [25, Secs. 4.5 and 4.6] 3) AC Current Control: In order to implement this scheme,
for details on the transformations). We remind the reader that a PI controller for the current i s,dq = [i s,d i s,q ] is used to
the Simulink implementation of the controls presented in the 
track ī s,dq
forthcoming subsections is available online [23].

ẋi,dq = ī s,dq − i s,dq, (8a)
A. Low-Level Cascaded Control Design



vs,dq := vdq + Z i s,dq + Ki, p i s,dq − i s,dq + Ki,i xi, dq (8b)
1) AC Voltage Control: We employ a standard converter      
control architecture that consists of a reference model provid- feed-forward terms PI control
ing a reference voltage v̂dq with angle  v̂dq = θ and magnitude    ] is the
v̂dq  (as it is illustrated in Fig. 5). The modulation signal where Z = LωJ2 + RI2 , vs,dq = [v s,d v s,q
mαβ is determined by cascaded proportional–integral (PI) reference for the switching node voltage (i.e., before output
controllers that are implemented in dq-coordinates (rotating filter in Fig. 1), xi,dq = [x i,d x i,q ] denotes the integrator
with the reference angle θ ) and track the voltage reference state, and Ki, p = k p,i I2 and Ki,i = ki,i I2 denote the diagonal
v̂dq (see [18]). The voltage tracking error v̂dq − vdq is used matrices of proportional and integral gains, respectively. Note
to compute the reference i s,dq  i  ] for the switching
= [i sd that the first two terms of the right-hand side of (6b) and (8b)
sq
node current i s,dq are feed-forward terms. Finally, the modulation signal mαβ is
given by
ẋv,dq = v̂dq − vdq (6a)
 
2vs,αβ
i s,dq := i dq + CωJ2 vdq + Kv, p (v̂dq −vdq )+Kv,i xv,dq (6b) mαβ =
      
v dc
(9)
feed-forward terms PI control
where xv,dq = [x v,d x v,q ] denotes the integrator state, 
where vs,αβ 
is the αβ-coordinates image of vs,dq defined in (8)
vdq = [v d v q ] denotes the output voltage measurement,  denotes the nominal converter dc voltage.
and v dc
v̂dq = [v̂ d 0] denotes the reference voltage, i dq = [i d i q ]
denotes the output current, I2 is the 2-D identity matrix, and
Kv, p = kv, p I2 and Kv,i = kv,i I2 denote diagonal matrices of B. DC Voltage Control
proportional and integral gains, respectively.  that is tracked by the control-
2) AC Current Limitation: We assume that the underlying The dc current reference i dc
current controller or low-level protections of the converter lable dc source (2) is given by a proportional control for the
limit the ac current. We model this in abstraction by scaling dc voltage and feed-forward terms based on the nominal ac
down the reference current i s,dq   if it exceeds the predefined active power injection p and the filter losses
ac
converter current limit i max [33, Sec. III], that is
 
   p v dc i x − p
i dc = kdc v dc − v dc +  + G dc v dc + (10)
 i , if i s,dq  ≤ i max
ac
   v 
v dc
ī s,dq := s,dq (7) proportional control 
dc
 
γi i s,dq, if i s,dq  > i max
ac
power injection and loss feed-forward

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
1008 IEEE JOURNAL OF EMERGING AND SELECTED TOPICS IN POWER ELECTRONICS, VOL. 8, NO. 2, JUNE 2020

Fig. 5. Block diagram of the low-level cascaded control design (6)–(8) in Fig. 7. Block diagram of a grid-forming VSM-based (13)–(15).
dq-coordinates rotating with the angle θ provided by the reference model.

Fig. 6. Droop control frequency and ac voltage control block diagrams based
on (11) and (12). Fig. 8. Matching control block diagram based on (16)–(18).

where v dc i x is the dc power flowing into the switches, p is the where D p (ω − ω) is commonly referred to as (virtual)
ac power injected into the grid, and the last term on the right- damping [8, Sec. II-B] and is inspired by the speed droop
hand side of (10) implements a feed-forward power control response of an SM. Note that the speed dependent term in
that compensates the filter losses. The loss compensation is (13b) has no exact analog in an SM. The response of SM
required to ensure exact tracking of the power reference by damper windings and its turbine governor are on different
matching control (see Section III-E) and also improves the dc timescales, and the SM damper windings do not act relative
voltage regulation for the other control strategies considered to the nominal frequency. In contrast, the speed dependent
in this study. Thus, to employ an identical dc voltage control, term in (13b) provides both instantaneous damping and at
we apply (10) for all control strategies discussed throughout steady state (i.e., equivalent to SM turbine governor droop)
this article. relative to the nominal frequency. Moreover, Jr is the virtual
rotor’s inertia constant. Note that the dynamics (13) reduce to
C. Droop Control droop control (11) when using Jr /D p ≈ 0 as recommended in
Droop control resembles the speed droop property (5a) of [8]. These angle dynamics capture the main salient features of
the SM governor [6] and trades off deviations of the power VSMs but do not suffer from drawbacks of more complicated
injection (from its nominal value p ) and frequency deviations implementations (see [4] for a discussion). The three-phase
(from ω ) voltage induced by the VSM is given by

θ̇ = ω (11a) 2π 4π
v̂abc = 2ωM f i f sin (θ ) sin θ − sin θ −
ω = ω + dω ( p  − p) (11b) 3 3
(14)
where dω denotes the droop gain. To replicate the service
provided by the AVR of SMs, we use a PI controller acting where M f and i f are, respectively, the virtual mutual induc-
on the output voltage error tance magnitude and excitation current. Similar to (12), we uti-
t lize input i f to achieve exact ac voltage regulation via PI
v̂d = k p (v  − vdq ) + ki (v  − vdq (τ )) d τ. (12) control and thereby replicate the function of the AVR of an SM
0
kp  ki t
to obtain the direct axis reference v̂ d for the underlying voltage if = (v − vdq ) + (v  − vdq (τ )) d τ. (15)
loop (v  and vdq  are the reference and measured voltage Mf Mf 0
magnitudes). We remark that v̂ q = 0 and the reactive power Transforming v̂abc to dq-coordinates with θ and ω as in (13),
injection varies such that exact voltage regulation is achieved. the voltage and current loops and modulation signal generation
The droop control block diagram is shown in Fig. 6. remain the same as (6)–(9).

D. VSM
E. Matching Control
Many variations of VSMs have been proposed [7], [8].
Matching control is a grid-forming control strategy that
In this article, we consider the frequency dynamics induced
exploits structural similarities between power converters and
by the synchronverter [8] (see Fig. 7)
SMs [10]–[13], [40] and is based on the observation that
θ̇ = ω (13a) the dc-link voltage—similar to the SM frequency—indicates
1 power imbalances. Hence, the dc voltage, up to a constant
Jr ω̇ =  ( p  − p) + D p (ω − ω) (13b)
ω factor, is used to drive the converter frequency. This control
Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
TAYYEBI et al.: FREQUENCY STABILITY OF SMs AND GRID-FORMING POWER CONVERTERS 1009

technique structurally matches the differential equations of


a converter to those of an SM. Furthermore, analogous to
the machine input torque, the dc current is used to control
the ac power. The angle dynamics of matching control are
represented by
θ̇ = kθ v dc (16)
where kθ := ω /v dc
 . Finally, the ac voltage magnitude is

controlled through the modulation magnitude μ by a PI


controller
t
μ = k p (v  − vdq ) + ki (v  − vdq (τ )) d τ. (17)
0
The reference voltage for the voltage controller in
αβ-coordinates is given by
v̂αβ = μ[− sin θ cos θ ] . (18) Fig. 9. Block diagram of dVOC in polar coordinates corresponding to (22).
Note that singularity at v̂dq  = 0 only appears in the dVOC implementation
Transforming v̂αβ to dq-coordinates with θ and ω as in (16), in polar coordinates but not in the implementation in rectangular coordinates,
i.e., (21).
the voltage and current loops and modulation signal generation
remain the same as (6)–(9). Fig. 8 depicts the block diagram
associated with the matching control. the converter, the angle κ := tan−1 (lω /r ) models the network
To further explain the matching concept, we replace v dc in inductance to resistance ratio, and η and α are the positive
(1a) by ω/kθ from (16) resulting in control gains. Furthermore, we have
θ̇ = ω (19a) cos κ − sin κ 1 p q 
R2 (κ) := , K :=  2 R2 (κ)
sin κ cos κ v −q  p
Cdc ω̇ = kθ i dc − kθ i x − G dc ω. (19b)
where R2 (κ) is the 2-D rotation by κ. As shown in [16],
Recalling the SM’s angle and frequency dynamics (4a)–(4b)
the dynamics (21) reduce to a harmonic oscillator if phase
and replacing T f by D f ω
synchronization is achieved (i.e., Kv̂αβ − R2 (κ)i αβ = 0) and
θ̇ = ω (20a) v̂αβ  = v  (i.e., (v 2 − v̂αβ 2 )v̂αβ = 0). Rewriting (21) in
J ω̇ = Tm − Te − D f ω. (20b) polar coordinates for an inductive network (i.e., κ = π/2)
reveals the droop characteristics (see [15]–[17]) of dVOC as
Comparing (19) and (20) reveals the structural matching of
GFC dynamics to that of SM. Dividing (19b) by kθ2 to obtain p p
θ̇ = ω + η − (22a)
the same units as in (20b) and matching variables result in Jr = v 2 v̂dq 2
Cdc /kθ2 , D f = G dc /kθ2 , Tm = i dc /kθ , and Te = i x /kθ . In other q q
v̂˙ dq  = η 2
− v̂dq  + ηφ(v  , v̂dq )(22b)
words, using matching control, the inertia constant of the GFC v v̂dq 2
is linked to its internal energy storage, the dc-side losses G dc ω
are linked to the machine windage friction losses D f ω, and the where φ(v  , v̂dq ) = (α/v 2 )(v 2 − v̂dq 2 )v̂dq . In other
frequency droop response is provided through the proportional words, for a high-voltage network and near the nominal steady
dc voltage control kdc (v dc −v ) = (k /k )(ω −ω) (see (10)). state (i.e., v̂dq  ≈ v  ), the relationship between the frequency
dc dc θ
The structural matching induced by (16) also extends to the and active power resemble that of standard droop control given
converter ac filter and generator stator dynamics (see [11], [13] in (11) with dω = η/v 2 . Moreover, when choosing the control
for a detailed derivation). gain α to obtain postfault voltages consistent with the other
control algorithms described earlier, the first term in (22b) is
F. dVOC negligible and (22b) reduces to the voltage regulator v̂˙ dq  ≈
dVOC [15]–[17] is a decentralized grid-forming control −2ηα(v̂dq  − v  ) near the nominal steady state. The block
strategy that guarantees almost global asymptotic stability for diagram of dVOC grid-forming strategy is presented in Fig. 9.
interconnected GFCs with respect to nominal voltage and
power set points [15], [16]. The analytic stability conditions IV. N ETWORK C ASE S TUDY
for dVOC explicitly quantify the achievable performance and In this section, we explore various performance aspects of
include the dynamics and transfer capacity of the transmission the grid-forming control techniques in the presence of SMs.
network [16]. In the forthcoming discussion, we use the test system shown
The dynamics of dVOC in αβ-coordinates are represented in Fig. 3. The parameters and control gains are given in Table I.
by The implementation in Simulink is available online [23].
v̂˙ αβ = ω J2 v̂αβ + η(Kv̂αβ − R2 (κ)i αβ + φ(v  , v̂αβ )) (21) In order to avoid the delay associated with the fre-
quency measurement and signal processing introduced
where φ(v  , v̂αβ )
= (α/v 2 )(v 2 −v̂αβ 2 )v̂ , v̂
αβ αβ = [v̂ α v̂ β ] by standard synchronous reference frame phase-locked
is the reference voltage, i αβ = [i α i β ] is current injection of loop (SRF-PLL) [22], we use the mechanical frequency of the

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
1010 IEEE JOURNAL OF EMERGING AND SELECTED TOPICS IN POWER ELECTRONICS, VOL. 8, NO. 2, JUNE 2020

sharing behavior. The Appendix presents the tuning criteria


and derivation of some control parameters.

C. Impact of Grid-Forming Control on Frequency Metrics


In this section, we test the system behavior for different
load disturbances pi . The network base load pl is constant
and uniformly distributed between nodes 5, 7, and 9, while
pi is only applied at node 7. For each disturbance input,
we calculate ||ωi ||∞ and |ω̇i | for the SM at node 1 and
normalize these quantities by dividing by |pi |. Figs. 11 and
12 illustrate the distribution of system disturbance input/output
gains associated with introduced frequency performance met-
rics. Note that the network base load pl is 2 p.u., and the ele-
Fig. 10. Postevent frequency nadir and RoCoF.
ments of the load disturbance sequence pi ∈ [0.2, 0.9], i =
1, . . . , 100 are uniformly increasing by 0.007 p.u. starting
from p1 = 0.2 p.u.
SM at node 1 in Fig. 3 to evaluate the postdisturbance system All the grid-forming controls presented in Section III are
frequency (e.g., in Figs. 11 and 12). For the GFCs, we use the originally designed without consideration of the converter dc
internal controller frequencies defined by (11), (13), (16), and and ac current limitation. Thus, to explore the intrinsic behav-
(22). We remark that in a real-world system and in an EMT ior of the GFCs and their influence on frequency stability,
simulation (in contrast to a phasor simulation), there is no the network loading scenarios described earlier are selected
well-defined frequency at the voltage nodes during transients, such that the GFCs dc and ac currents do not exceed the limits
whereas the internal frequencies of the GFCs are always well- imposed by (3) and (7). However, in practice, GFCs are subject
defined [1, Sec. II-J], [41]. Finally, we note that in all the to strict dc and ac current limits and are combined with current
forthcoming case studies, we assume that the system is in limiting strategies in a modular fashion. The impact of the dc
steady state at t = 0. and ac current limits on the performance of GFCs is explored
in Sections IV-D–IV-F1.
Figs. 11 and 12 suggest that regardless of the choice of
A. Performance Metrics control strategy, the presence of GFCs improves the metrics
We adopt the standard power system frequency perfor- compared to the all-SM system. This possibly observation
mance metrics, i.e., maximum frequency deviation ||ω||∞ can be explained by the fast response of converters compared
(i.e., frequency nadir/zenith) and RoCoF |ω̇| (i.e., the slope of with the slow turbine dynamics, i.e., τg in (5) is larger than
line tangent to the postevent frequency trajectory) defined by τdc in (2). Because of this, the converters reach frequency
synchronization at a faster timescale and then synchronize
||ω||∞ := max |ω − ω(t)| (23a)
t ≥t0 with the SM (see Fig. 13). Overall, for any given disturbance
|ω(t0 + T ) − ω(t0 )| input, the converters are able to react faster than the SM, and
|ω̇| := (23b) the remaining power imbalance affecting the SM is smaller
T
where t0 > 0 is the time when the disturbance is applied than in the all-SM system. This result highlights that the fast
to the system, and T > 0 is the RoCoF calculation window response of GFCs should be exploited instead of designing the
[1], [22]. See Fig. 10 for visual representation of the metrics controls of a converter (fast physical system) to emulate the
described by (23). In this article, we use T = 250 ms, which slow response of SMs [1].
is in line with values suggested for protection schemes (see We observe that droop control and dVOC exhibit very sim-
[21, Table 1]). Dividing the metrics (23) by the size of the ilar performance confirming the droop-like behavior of dVOC
magnitude of the disturbance results in a measure of the system in predominately inductive networks [see (22)]. Moreover,
disturbance amplification. the difference between droop control and VSM arises from
the inertial (derivative control) term in (13), and the RoCoF
is considerably higher when using matching control. This
B. Test Network Configuration and Tuning Criteria can be explained by the fact that VSM, droop control, and
In order to study the performance of the control approaches dVOC ignore the dc voltage and aggressively regulate the ac
introduced in Section III, we apply the same strategy (with quantities to reach angle synchronization, thus requiring higher
identical tuning) for both converters (i.e., at nodes 2 and 3 in transient peaks in dc current to stabilize the dc voltage (see
Fig. 3), resulting in four different SM-GFC-paired models. Section IV-D). Although improving RoCoF, this approach can
As a benchmark, we also consider an all-SMs system with lead to instability if the converter is working close to the rated
three identical SMs (i.e., at nodes 1-3). Selecting fair tuning power of the dc source (but far away from its ac current limit),
criteria for the different control strategies is a challenging as will be shown in Section IV-D. On the other hand, matching
task. For this study, we tune the control parameters such control regulates the dc-link voltage both through the dc source
that all generation units exhibit identical proportional load and by adjusting its ac signals.

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
TAYYEBI et al.: FREQUENCY STABILITY OF SMs AND GRID-FORMING POWER CONVERTERS 1011

Fig. 11. Normalized distribution of the RoCoF |ω̇i |/|pi | of the SM frequency at node 1 for load disturbances pi ranging from 0.2 to 0.9 p.u. at node 7.
For each load disturbance, |ω̇i |/|pi | is normalized by the maximum value corresponding to the all-SMs configuration.

Fig. 12. Normalized distribution of the nadir ||ωi ||∞ /|pi | of the SM frequency at node 1 for load disturbances pi ranging from 0.2 to 0.9 p.u. at
node 7. For each load disturbance, ||ωi ||∞ /|pi | is normalized by the maximum value corresponding to the all-SMs configuration.

We selected the RoCoF calculation window according to


the guideline [21], which accounts for noise and possible
oscillations in the frequency signal. However, these guidelines
were derived for a power system fully operated based on
SMs. Given that GFCs introduce faster dynamics, machines
are expected to reach the frequency nadir faster. Hence, smaller
RoCoF windows might need to be considered in a low-inertia Fig. 13. Frequency of the system with two VSMs after a 0.75-p.u. load
power system to properly assess system performance (the increase. The converters quickly synchronize with each other and then slowly
reader is referred to [42, Sec. III-C] for further insights on the synchronize with the machine.
choice of RoCoF window in a low-inertia system). We note
that the performance of the different grid-forming control
strategies shown in Figs. 11 and 12 is sensitive to the tuning is solely presented to emphasize the need for an ac current
of control gains and choice of RoCoF computation window. limitation mechanism to implicitly stabilize the dc voltage.
However, due to the comparably slow response of conventional To begin with, we set the network base load pl and load
generation technology, the performance improvements for the change p to 2.25 and 0.9 p.u., respectively (i.e., a total net-
system with GFCs over the all-SM system persist for a wide work load after the disturbance of 3.15 p.u.), which is equally
range of parameters. Moreover, using comparable tuning (see shared by the SM and the GFCs. We expect a postdisturbance
Section IV-B), the differences between the different grid- converter power injection of 1.05 p.u. and i dc close to the dc
dc = ±1.2 p.u. Fig. 14 shows the dc voltage
current limit i max
forming techniques observed in this section are expected to
remain the same. and delayed dc current before saturation for the converter at
node 2. For sufficiently large disturbance magnitude, i dc and,

consequently, i τ exceed the current limit, i.e., the dc current


D. Response to a Large Load Disturbance i dc = i max
dc is saturated.

In this section, we analyze the response of the GFCs to large Figs. 14 and 15 highlight that the matching control succeeds
disturbances when the dc source is working close to its maxi- to stabilize the dc voltage despite the saturation of the dc
mum rated values. Specifically, we focus on the implications of source. The nature of matching control, which accounts for the
the dc current limit (3) to highlight the response of GFCs when dc-side dynamics while regulating the ac dynamics, i.e., (16),
interfacing curtailed renewable generation with low headroom results in increased robustness with respect to large distur-
(i.e., PV/wind) or a converter system with an undersized dc–dc bances. From a circuit-theoretic point of view, this is possible
converter stage. In such scenarios, the dc current limit can only if the sum of the ac power injection and filter losses
be activated independently of the ac current limit. For clarity equals the approximately constant dc power inflow v dc i maxdc .

of representation, we first focus exclusively on the matching The converter can inject constant ac power into the network
control as its angle dynamic (16) explicitly considers the dc only if its angle difference with respect to the remaining
voltage. Droop control, which does not consider the dc voltage, devices in the network is constant. In the presence of the

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
1012 IEEE JOURNAL OF EMERGING AND SELECTED TOPICS IN POWER ELECTRONICS, VOL. 8, NO. 2, JUNE 2020

Fig. 14. DC current demand of the converter at node 2 (top) and its dc Fig. 17. DC voltage (top) and dc current demand (bottom) of the converter
voltage (bottom) after a 0.9-p.u. load disturbance. at node 2 after a 0.9-p.u. load disturbance when all the limitation schemes,
i.e., (3), (7), and (24) are active and τg = 5 s.

frequency dynamics [i.e., (11), (13), and (22a)] of droop


control, VSM, and dVOC purely rely on ac measurements,
they are agnostic to the dc dynamics and source saturation.
Consequently, a prolonged duration of dc source saturation
results in a collapse of the dc voltage. A potential remedy is
to incorporate an ac current limitation scheme to prevent the
GFC from depleting the dc-link capacitor. This observation
motivates the discussion in Section IV-E with restricted focus
on droop control, VSM, and dVOC techniques.

E. Incorporating the AC Current Limitation


In this section, we investigate if the ac current limitation (7)
Fig. 15. DC current demand and saturated dc current (top), frequency of
presented in Section III-A can mitigate the instabilities of the
the converter (using matching control) at node 2 and SM after a 0.9-p.u. load GFCs controlled by droop control, VSM, and dVOC under
disturbance (bottom). dc source saturation. To this end, we consider the same base
load and disturbance as in previous test case, i.e., pl = 2.25
and p = 0.9 p.u. For this scenario, the GFCs dc transient
current demand i τ and the switching node current magnitude
i s,dq  exceed the limits (i.e., i max
dc = i ac = 1.2 p.u.) imposed
max
by (3) and (7).
We observe that using the ac current limiter does not
stabilize the dc voltage of the GFCs controlled by droop
control, VSM, and dVOC. Fig. 16 illustrates the behavior of
Fig. 16. DC voltage of the converter at node 2 after a 0.9-p.u. load disturbance the GFC at node 2. Specifically, the current limitation imposed
when both dc and ac limitation schemes (3) and (7) are active and τg = 5 s. by (7) results in integrator windup, a loss of ac voltage control,
and, ultimately, the instability of the grid-forming control
[34] and dc voltage. We remark that GFCs exhibit the same
slow SM angle and frequency dynamics, this implies that the
instability behavior when the ac current limit is smaller than
GFCs need to synchronize their frequencies to the SM so ac < i dc .
that of the dc side, i.e., i max
that the relative angle θGFC − θSM = θmax is constant (see max
To mitigate this load-induced instability, we explore a
the frequency of GFC 1 following the dc source saturation at
current limitation scheme that modifies the active power set
t = 0.5 s in Fig. 15). Note that the behavior for the matching
point when i s,dq  exceeds a certain threshold value, that is
controlled GFCs is similar to that of the SM (see Section III-E 
and [10], [11], [13]), i.e., it achieves synchronization both 0, if i s,dq  ≤ i th
ac

under controlled or constant mechanical input power (i.e., dc p := (24)
γ p (i s,dq  − i th
ac ), if i s,dq  > i th
ac
current injection). Fig. 14 shows that for the GFCs controlled
by droop control, the dc-link capacitors discharge to provide and p is added to p in (10), (11), (13b), and (21), γ p
i τ − i max
dc (i.e., the portion of current demand, which is not denotes a proportional control gain, and i th ac < i ac is the
max
provided due to the saturation (3)). Because the angle and activation threshold. Note that the control law (24) implicitly

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
TAYYEBI et al.: FREQUENCY STABILITY OF SMs AND GRID-FORMING POWER CONVERTERS 1013

Fig. 18. DC voltage (top) and dc current demand (bottom) of the converter Fig. 19. DC current demand (top) and dc voltage (bottom) of the converter
at node 2 after a 0.9-p.u. load disturbance when both dc and ac limitation at node 2 after loss of the SM at node 1.
schemes (3) and (7) are active and τg = 1 s.
complementary benefits by using both dc and ac measurements
manipulates the grid-forming dynamics through their set points in grid-forming control.
such that the ac current magnitude stays within the admissible
limits for large increases in load. We emphasize that this F. Interaction of GFC and SM Dynamics
strategy aims at mitigating instabilities induced by large load The forthcoming discussion highlights that the interplay
increases and that the resulting GFC response to grid faults between the fast GFC and slow SM dynamics influences the
needs to be carefully studied. system stability. The following case studies are tightly related
For a 0.9-p.u. load increase, the current limitation strategy to the scenarios in Sections IV-D and IV-E.
(24) is able to stabilize the system with i thac = 0.9 p.u. and
1) Loss of SM: We study the response of GFCs when
γ p = 2.3( pb /i b ), where pb and i b denote the converter base
ac ac
disconnecting the SM at node 1, that is, the system turns into
power and current, respectively. Fig. 17 depicts the response an all-GFCs network. The implications of such a contingency
of the same GFC as in Fig. 16. Note that (24) effectively are threefold. First, the power injected by the machine, which
stabilizes the dc voltage for droop control, VSM, and dVOC. partially supplies the base load, is no longer available. Sec-
Moreover, in contrast to Fig. 14, after the postdisturbance ond, the stabilizing dynamics associated with the machine’s
transient, the dc source is no longer in saturation. Broadly governor, AVR, and PSS are removed from the system. Third,
speaking, this strategy succeeds to stabilize the system by the slow dynamics of the SM no longer interact with the fast
steering the GFC power injection away from the critical limits. dynamics of the GFCs.
However, this also influences the postdisturbance operating For this test, we set the base load to 2.1 p.u., and the SM and
point of the GFCs due to the threshold value being below GFCs set points are set to 0.6 and 0.75 p.u., respectively. Note
the rated value. that when the SM at node 1 is disconnected, the converters
Finally, we observe that the different timescales in a increase their power output according to the power sharing
low-inertia system contribute to the instabilities observed in behavior inherent to all four grid-forming controls (see the
Section IV-D. In particular, if the SM’s turbine responds Appendix). The resulting increase in the converter power
faster, GFCs with the standard limitation strategy (7) pre- injection to roughly 1.05 p.u. is similar to the load disturbance
serve stability—without the need to implement (24)—despite scenario used to illustrate the instability behavior of droop
the fact that transient dc and ac currents exceed the limits. control in Fig. 14. Fig. 19 shows i τ and v dc for the converter
Fig. 18 shows the GFCs responses when the SM turbine at node 2. Although the disturbance magnitude affecting the
delay τg is 1 s (see Fig. 16 where τg = 5 s). It can be converters is similar to the one in studied in Section IV-D, all
seen that the slow SM turbine dynamics again contribute GFCs remain stable after the loss of the SM without the need
to the system instability when dc and ac currents are satu- to incorporate (24). In particular, due to the absence of the slow
rated. We conclude that the presence of different timescales turbine dynamics and fast synchronization of the converters,
in a low-inertia system—often neglected in the literature dc for around 50 ms, while it remains
i τ exceeds the limit i max
[26], [33]–[39]—must be considered in designing a robust ac above the limit for a prolonged period of time in Fig. 14.
current limitation mechanism for the GFCs. This highlights the problematic interaction between the fast
Remark 2 (DC and AC Measurements): We observe that response of the GFCs and the slow response of the SM. While
using ac measurements to drive the angle dynamics, e.g., the SM perfectly meets classic power system control objectives
VSMs use active power measurements in (13), improves on slower timescales, the dominant feature of GFCs is their
the frequency performance of GFCs (see Figs. 11 and 12). fast response. However, the fast response of GFCs can also
On the other hand, using dc voltage measurements, e.g., result in unforeseen interactions with other parts of the system,
matching control (16), results in robustness with respect to dc such as the slow SM response (shown here), line dynamics (see
current limits. Further research is required to combine these [16], [29]), and line limits [43].

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
1014 IEEE JOURNAL OF EMERGING AND SELECTED TOPICS IN POWER ELECTRONICS, VOL. 8, NO. 2, JUNE 2020

ac current control [see (6) and (8)] achieve perfect tracking


(i.e., v dq = v̂ dq ). Assuming that the system operates near the
nominal steady state (i.e., v dq  ≈ v  , ω ≈ ω ) and p  = 0,
q  = 0, we rewrite the remaining dynamics in terms of the
voltage angle and power injection at every bus. This results
in a simplified model of the angle θ and the frequency ω
of a GFC or SM relative to a frame rotating at the nominal
frequency ω .
1) Droop Control and dVOC: For a converter controlled by
droop control or dVOC, we obtain
θ̇ = −dω p (25)
where p is the power flowing out of the converter and dω is
the droop control gain and given by dω = η/v  2 for dVOC.
Fig. 20. DC current demand (top) and dc voltage (bottom) after a 0.9-p.u.
load disturbance in an all-GFC system.
2) SM, VSM, and Matching Control: For an SM, a VSM,
and a converter controlled by matching control, we obtain
2) Low-Inertia Versus No-Inertia Systems: Considering the θ̇ = ω (26a)
same load disturbance scenario as in Section IV-D, we observe
2H ω̇ = −Dω + sat( pτ , pmax ) − p (26b)
the same instability of droop control when the test system
contains one GFC and two SMs, i.e., the instability cannot be τ ṗτ = − pτ − d p ω. (26c)
prevented by adding more inertia to the system. For an SM, the parameters directly correspond to the parame-
Fig. 20 shows the dc current demand i τ (i.e., before ters of the machine model presented in Section II-B, i.e., H ,
saturation) and dc voltage in an all-GFCs (i.e., no-inertia) d p , and τ = τg are the machine inertia constant, governor gain,
system for a load increase of p = 0.9 p.u. In this case, and turbine response time, and the model does not capture SM
the GFCs quickly synchronize to the postevent steady state, losses (i.e., D = 0) or damper winding torques. Throughout
which does not exceed the maximum dc current, saturate this article, we have not considered a limit on the turbine
the dc source for only approximately 200 ms, and remain power output (i.e., pmax = ∞) because an SM, in contrast
stable. In contrast, in the system with two GFCs and one to a GFC, typically has sufficient reserves to respond to the
SM, the SM does not reach its increased postevent steady- load changes and faults considered in this article. For the
state power injection for several seconds. During this time, VSM presented in Section III-D, we obtain τ = 0, pτ = 0,
the response of droop control results in a power injection H = 1/2 J ω , and D = D p ω , i.e., the VSM does not
that exceeds the limits of the dc source and collapses the dc emulate a turbine and implements no saturation of the damping
voltage. In other words, the slow response of the SM due to the term in its frequency dynamics (13b). Finally, for matching
large turbine time constant prolongs the duration of dc source  i dc , H = 1/2 C /k 2 ,
control, we obtain τ = τdc , pmax = v dc max dc θ
saturation for the GFCs, which results in depleting the dc link d p = kdc /kθ (see Section III-E), and D = 0, i.e., by linking
if the converter ac current is not limited. We conclude this frequency and dc voltage, matching control clarifies that the
section by remarking that in the mixed SM-GFCs system in dc source plays the role of the turbine in a machine, and the
Sections IV-D and IV-E, it is vital to account for dc side limits proportional dc voltage control plays the role of a governor.
(either through ac current limits or dc voltage measurements)
to stabilize the GFCs in presence of SMs. However, in an all-
B. Reduced-Order Model
GFC system—either by default or due to the loss of SM—the
dc side is only limited briefly and the system remains stable. For brevity of the presentation, we will now restrict our
This observation highlights that the interaction of the fast GFC attention to the case of one SM and one GFC. The equivalent
dynamics and slow SM dynamics can potentially contribute to inertia constants and turbine time constants for the different
system instability. GFC control strategies are either zero or negligible compared
with typical inertia constants and turbine time constants for
V. Q UALITATIVE A NALYSIS
machines (see Table I). We, therefore, assume that the states of
In this section, we provide a qualitative but insightful analy- the SM are slow variables, while the states of the GFC are fast
sis that explains the results observed in Sections IV-C and IV- and apply ideas from singular perturbation theory [12], [44].
D. To this end, we develop simplified models that capture the Using the dc power flow approximation and pd,GFC and
small-signal frequency dynamics of SMs and GFCs. Applying pd,SM to denote a disturbance input, we obtain pGFC =
arguments from singular perturbation theory [12], [44] results b(θ GFC − θ SM ) + pd,GFC and pSM = b(θ SM − θ GFC ) + pd,SM,
in a model that highlights the main salient features of the where b is the line susceptance. Neglecting the frequency
interaction of SMs and GFCs. dynamics and dc source dynamics, i.e., letting τ → 0 and
H → 0, the dynamics of the relative angle δ = θ SM − θ GFC
A. Frequency Dynamics Incorporating GFCs are given by δ̇ = ωSM − (DGFC + d p,GFC )(bδ − pd,GFC) if
To obtain a simplified model of the frequency dynamics |d p,GFC | < | pmax |, δ̇ = ωSM − DGFC (bδ − pd,GFC ± pmax )
of the GFCs, we assume that the cascaded ac voltage and if the dc source is saturated, and DGFC denotes the damping

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
TAYYEBI et al.: FREQUENCY STABILITY OF SMs AND GRID-FORMING POWER CONVERTERS 1015

provided by the GFC. For typical droop gains and network


parameters, the relative angle dynamics are fast compared to
the machine dynamics. Letting δ̇ → 0, we obtain the reduced-
order model
2H ω̇SM = −sat(DGFC ωSM , pmax )+ pτ,SM + pd (27a)
τ ṗτ,SM = − pτ,SM − d p ωSM (27b)
where pd = pd,GFC + pd,SM , H , d p , and τ are the inertia
constant, governor gain, and turbine time constant of the SM,
DGFC is the damping provided by the GFC, i.e., DGFC = 1/dω
(droop, dVOC, VSM), or DGFC = d p (matching). Moreover,
the droop control, dVOC, and VSM implement no saturation
of their power injection in their respective angle/frequency
dynamics [i.e., (11b), (22a), and (13b)], resulting in pmax =
∞. In contrast, for matching control, the saturation of the dc Fig. 21. Change in averaged RoCoF and frequency nadir when transitioning
 i dc .
source results in pmax = v dc from a system with three SMs to a system with one SM and two GFCs.
max
We note that for the case of underdamped dynamics (27)
and without saturation, a closed-form expression for the step and frequency nadir according to (27) are shown in Fig. 21.
response and frequency nadir (27) can be found in [19, The case with pmax = 1.2 corresponds to matching control,
Sec. V-A]. However, even for the seemingly simple model the one with pmax = ∞ to droop control, dVOC, and VSMs.
(27), the dependence of the nadir on the parameters is very It can be seen that the GFCs result in an improvement
involved and does not provide much insight. By neglecting compared with the all-SM scenario, the implicit saturation of
the damping term in (27a) and the feedback term − pτ,SM the power injection by matching control results in a smaller
in (27b), an insightful expression for the nadir is obtained improvement compared with the droop control, dVOC, and
in [45]. However, the key feature of the GFCs is that they VSMs, and the reduction in the average RoCoF and frequency
contribute damping, which is not captured by analysis in [45]. nadir is line with the corresponding results in Figs. 11 and 12.
Nonetheless, the model (27) provides several insights that we
discuss in Section V-C.
D. Instability in the Presence of Large Load Disturbance
C. Impact of Grid-Forming Control on Frequency Stability The instabilities of the droop control, dVOC, and the VSM
observed in Section IV-D can qualitatively be investigated
A system with three SMs as in Section IV can be modeled
using a simplified model of the dc side. To compute the power
by (27) with H = 3 HSM , d p = 3 d p,SM , τ = τSM , and
v dc i x flowing out of the dc-link capacitor, we assume that the
DGFC = 0. This corresponds to the well-known center-of-
controlled converter output filter dynamics are fast and can
inertia frequency model with first-order turbine dynamics (see
be neglected (i.e., i̇ s,αβ = 0, v̇αβ = 0) and that the ac output
[3], [19]). In contrast, if two SMs are replaced by GFCs with  i
filter losses are negligible. This results in v dc i x = vs,αβ s,αβ =
equal droop setting, we obtain H = HSM , d p = d p,SM , τ = 
τSM , and DGFC = 2 d p,SM . vαβ i αβ = p. Moreover, we neglect the dc source dynamics to
obtain the simplified dc voltage dynamics
This highlights that on the timescales of the SM, the GFCs
    dc  p
provide fast-acting frequency control. After an increase in Cdc v̇ dc = −G dc v dc + sat kdc v dc − v dc , i max − (28)
load, the machine inertia serves as buffer until the relatively v dc
slow turbine provides additional power to the machine. In con- that is, the active power p flowing into the grid is drawn from
trast, the converters respond nearly instantaneously to any the dc-link capacitor that is stabilized by a proportional control
imbalance, and therefore, the need for inertia is decreased. [see (10)] if the dc current i τ is not saturated. For a large
Intuitively, this increase in fast primary frequency control enough constant perturbation p > 0, the dc current in (28)
should result in lower nadir values. Similarly, the additional becomes saturated, the controllability of the voltage v dc is lost,
damping provided by the converter in (27) can be interpreted and the dc voltage becomes unstable.
as a filter acting on the power imbalance, i.e., the GFCs In other words, if the dc source is saturated, the power p has
are providing DGFC ωSM or pmax , and the power imbalance to be controlled to stabilize the dc voltage. Matching control
affecting the machine is reduced, therefore resulting in smaller achieves this through the angle dynamics θ̇ = kθ v dc , which
average RoCoF. converges to a constant angle difference (i.e., ωSM = ωGFC )
To validate the model (27) and our interpretation, we com- and power injection when the dc source is saturated and the
pute the frequency nadir and averaged RoCoF [see (23)] for SM has enough reserves to maintain ωSM ≈ ω with the
the machine parameters and disturbance used in Section IV, GFC providing its maximum output power. Moreover, for
and H = ν3 HSM , d p = ν3 d p,SM , τ = τSM , and DGFC = matching control, it can be verified that (ωGFC − ω )/ω =
(1 − ν)d p,SM and D = 0, where ν ∈ [1/3, 1] is a scalar (v dc − v dc )/v  . Therefore, the dc voltage deviation is propor-
dc
parameter that interpolates the parameters between the two tional to the frequency deviation and the GFC with matching
cases (all SM, one SM and two GFCs). The average RoCoF control remains stable for the scenario shown in Section IV-D.

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
1016 IEEE JOURNAL OF EMERGING AND SELECTED TOPICS IN POWER ELECTRONICS, VOL. 8, NO. 2, JUNE 2020

VI. S UMMARY AND F URTHER W ORK GFC control techniques and SM perform equal-load sharing.
In this article, we provide an extensive review of different Moreover, by selecting kdc based on this criteria, the dc voltage
grid-forming control techniques. Subsequently, we used the control gain in (10) is automatically set, which is identical for
IEEE nine-bus test system incorporating high-fidelity GFC all GFC implementation [23].
and SM models to investigate the performance of different Regarding the ac voltage regulation, the control gains in
control techniques and their interaction with SM. Our case (12), (15), (17), and (22b) are selected to regulate the ac
studies revealed that: 1) the presence of the GFCs improves voltage at approximately equal timescales. We refer to [46]
the frequency stability metrics compared to the baseline all- for details on tuning the cascaded inner loops presented in
SMs system; 2) under a sufficiently large load disturbance, it is Section III-A. It is noteworthy that the timescale of the
vital to implement an ac current limiting scheme to stabilize reference model (i.e., grid-forming dynamics) must be slower
the grid-forming techniques that only rely on the ac mea- than ac voltage control shown in Fig. 4 to ensure the optimal
surements; 3) matching control exhibits robustness to the dc performance. Similarly, the ac current control must be faster
source saturation since its angle dynamics takes into account than the outer voltage controller. Finally, the choice of virtual
the dc quantities; 4) we explored the stabilizing influence of inertia constant in (13b) can largely influence VSM’s dynamic
the ac current limitation; and 5) we investigated the behavior behavior. We adopted the recommendation J/D p = 0.02
of GFCs in response to the loss of SM and in all-GFCs system proposed in [8]. The parameters used in the implementation
highlighting a potentially destabilizing interaction between the [23] are reported in Table I.
GFCs and SMs dynamics. Moreover, we provided a qualitative
R EFERENCES
analysis of GFCs impact on frequency stability. Topics for the
[1] F. Milano, F. Dörfler, G. Hug, D. J. Hill, and G. Verbic, “Foundations
future works include: 1) further exploration of the impacts and challenges of low-inertia systems (invited paper),” in Proc. Power
of dc and ac current limitations on the GFCs behavior and Syst. Comput. Conf. (PSCC), Jun. 2018.
providing a formal stability analysis; 2) proposing a complete [2] P. W. Sauer and M. A. Pai, Power System Dynamics and Stability.
Upper Saddle River, NJ, USA: Prentice-Hall, 1998.
ac current limitation strategy that is robust to load-induced [3] P. Kundur, Power System Stability and Control. New York, NY, USA:
overcurrent as well as grid faults; 3) the seamless transition McGraw-Hill, 1994.
between grid-forming and grid-following operations; and 4) [4] A. Tayyebi, F. Dörfler, F. Kupzog, Z. Miletic, and W. Hribernik, “Grid-
forming converters-inevitability, control strategies and challenges in
blending of the different control strategies into a controller future grids application,” in Proc. CIRED Workshop, 2018.
that achieves their complementary benefits. [5] High Penetration of Power Electronic Interfaced Power Sources
(HPOPEIPS) ENTSO-E Guidance Document for National Implemen-
A PPENDIX tation for Network Codes on Grid Connection, ENTSO-E, Brussels,
T UNING C RITERIA Belgium, 2017.
[6] M. Chandorkar, D. Divan, and R. Adapa, “Control of parallel connected
The load-sharing capability of the control techniques pre- inverters in standalone AC supply systems,” IEEE Trans. Ind. Appl.,
sented in Section III is investigated in [4], [7], and [15]. vol. 29, no. 1, pp. 136–143, Jan. 1993.
[7] S. D’Arco, J. A. Suul, and O. B. Fosso, “A virtual synchronous
Considering a heterogeneous network consisting of several machine implementation for distributed control of power converters in
GFCs (with different control) and SMs, we tune the control smartgrids,” Electr. Power Syst. Res., vol. 122, pp. 180–197, May 2015.
parameters such that all the units exhibit identical proportional [8] Q.-C. Zhong and G. Weiss, “Synchronverters: Inverters that mimic
synchronous generators,” IEEE Trans. Ind. Electron., vol. 58, no. 4,
load-sharing in steady state. For the SM- and droop-controlled pp. 1259–1267, Apr. 2011.
GFCs, (5a) and (11b) can be rearranged to [9] Q.-C. Zhong, P.-L. Nguyen, Z. Ma, and W. Sheng, “Self-synchronized
synchronverters: Inverters without a dedicated synchronization unit,”
1
ω − ω = ( p − p ) (29a) IEEE Trans. Power Electron., vol. 29, no. 2, pp. 617–630, Feb. 2014.
dp [10] I. Cvetkovic, D. Boroyevich, R. Burgos, C. Li, and P. Mattavelli, “Mod-
eling and control of grid-connected voltage-source converters emulating
ω − ω = dω ( p − p  ). (29b) isotropic and anisotropic synchronous machines,” in Proc. IEEE 16th
Workshop Control Modeling Power Electron. (COMPEL), Jul. 2015.
For VSM, assuming steady-state frequency and setting [11] C. Arghir, T. Jouini, and F. Dörfler, “Grid-forming control for power
θ̈ = ω̇ = 0 in (13b) results in converters based on matching of synchronous machines,” Automatica,
vol. 95, pp. 273–282, Sep. 2018.
1
ω − ω = ( p − p  ). (30) [12] S. Curi, D. Gross, and F. Dörfler, “Control of low-inertia power grids:
D p ω A model reduction approach,” in Proc. IEEE 56th Annu. Conf. Decis.
Control (CDC), Dec. 2017, pp. 5708–5713.

For matching control, we assume that in steady state, i dc ≈ i dc [13] C. Arghir and F. Dorfler, “The electronic realization of synchronous

and v dc /v dc ≈ 1. Setting ω̇ = 0 in (19b) and replacing i dc by machines: Model matching, angle tracking, and energy shaping tech-
niques,” IEEE Trans. Power Electron., vol. 35, no. 4, pp. 4398–4410,
the expression from (10) yield Apr. 2020.
kθ [14] M. Sinha, F. Dorfler, B. B. Johnson, and S. V. Dhople, “Uncovering
ω − ω = ( p − p ). (31) droop control laws embedded within the nonlinear dynamics of Van
kdc v dc der Pol oscillators,” IEEE Trans. Control Netw. Syst., vol. 4, no. 2,
pp. 347–358, Jun. 2017.
Finally, for dVOC, assuming v̂ dq  ≈ v  in steady state, [15] M. Colombino, D. Groz, J.-S. Brouillon, and F. Dorfler, “Global phase
the angle dynamics (22a) becomes and magnitude synchronization of coupled oscillators with application
η to the control of grid-forming power inverters,” IEEE Trans. Autom.
ω − ω =  2 ( p − p  ). (32) Control., vol. 64, no. 11, pp. 4496–4511, Nov. 2019.
v [16] D. Gros, M. Colombino, J.-S. Brouillon, and F. Dorfler, “The effect
Hence, for any given droop gain d p , if dω , D p , kdc and η are of transmission-line dynamics on grid-forming dispatchable virtual
oscillator control,” IEEE Trans. Control Netw. Syst., vol. 6, no. 3,
selected such that the slopes of (29)–(32) are equal, all the pp. 1148–1160, Sep. 2019.

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
TAYYEBI et al.: FREQUENCY STABILITY OF SMs AND GRID-FORMING POWER CONVERTERS 1017

[17] G.-S. Seo, M. Colombino, I. Subotic, B. Johnson, D. Gros, and [38] I. Sadeghkhani, M. E. Hamedani Golshan, J. M. Guerrero, and
F. Dorfler, “Dispatchable virtual oscillator control for decentralized A. Mehrizi-Sani, “A current limiting strategy to improve fault ride-
inverter-dominated power systems: Analysis and experiments,” in Proc. through of inverter interfaced autonomous microgrids,” IEEE Trans.
IEEE Appl. Power Electron. Conf. Expo. (APEC), Mar. 2019. Smart Grid, vol. 8, no. 5, pp. 2138–2148, Sep. 2017.
[18] N. Pogaku, M. Prodanovic, and T. C. Green, “Modeling, analysis and [39] A. D. Paquette and D. M. Divan, “Virtual impedance current limiting
testing of autonomous operation of an inverter-based microgrid,” IEEE for inverters in microgrids with synchronous generators,” IEEE Trans.
Trans. Power Electron., vol. 22, no. 2, pp. 613–625, Mar. 2007. Ind. Appl., vol. 51, no. 2, pp. 1630–1638, Aug. 2015.
[19] F. Paganini and E. Mallada, “Global analysis of synchronization per- [40] L. Huang et al., “A virtual synchronous control for voltage-source
formance for power systems: Bridging the theory-practice gap,” IEEE converters utilizing dynamics of DC-link capacitor to realize self-
Trans. Autom. Control., to be published. synchronization,” IEEE J. Emerg. Sel. Topics Power Electron., vol. 5,
[20] M. Pirani, J. W. Simpson-Porco, and B. Fidan, “System-theoretic no. 4, pp. 1565–1577, Dec. 2017.
performance metrics for low-inertia stability of power networks,” in [41] F. Milano and A. Ortega, “Frequency divider,” IEEE Trans. Power Syst.,
Proc. IEEE 56th Annu. Conf. Decis. Control (CDC), Dec. 2017, vol. 32, no. 2, pp. 1493–1501, May 2017.
pp. 5106–5111. [42] A. Crivellaro et al., “Beyond low-inertia systems: Massive integra-
[21] Frequency Measurement Requirements and Usage, RG-CE Syst. Protec- tion of grid-forming power converters in transmission grids,” 2019,
tion Dyn. Sub Group, Brussels, Belgium, 2018. arXiv:1911.02870. [Online]. Available: https://arxiv.org/abs/1911.02870
[43] “Queensland and South Australia system separation on 25 August 2018,”
[22] B. K. Poolla, D. Gros, and F. Dorfler, “Placement and implementation
Austral. Energy Market Operator, Melbourne, VIC, Australia, Tech.
of grid-forming and grid-following virtual inertia and fast frequency
Rep., 2019. [Online]. Available: https: //www.aemo.com.au/-/media/
response,” IEEE Trans. Power Syst., vol. 34, no. 4, pp. 3035–3046,
Files/Electricity/NEM/Market_Notices_and_Events/Power_System_
Jul. 2019.
Incident_Reports/2018/Qld—SA-Separation-25-August-2018-Incident-
[23] A. Tayyebi, D. Groß, and A. Anta. (2019). GridFormingConvert- Report.pdf
ers: Implementation of Grid-Forming Control Techniques in IEEE 9- [44] G. Peponides, P. Kokotovic, and J. Chow, “Singular perturbations and
Bus System. Git Repository. [Online]. Available: https://github.com/ time scales in nonlinear models of power systems,” IEEE Trans. Circuits
ATayebi/GridFormingConverters Syst., vol. CS-29, no. 11, pp. 758–767, Nov. 1982.
[24] U. Markovic, O. Stanojev, E. Vrettos, P. Aristidou, and G. Hug, [45] L. Sigrist, I. Egido, and L. Rouco, “Principles of a centralized UFLS
“Understanding stability of low-inertia systems,” 2019, engrXiv preprint. scheme for small isolated power systems,” IEEE Trans. Power Syst.,
[Online]. Available: http://engrxiv.org/jwzrq vol. 28, no. 2, pp. 1779–1786, May 2013.
[25] A. Yazdani and R. Iravani, Voltage-Sourced Converters in Power Sys- [46] T. Qoria, F. Gruson, F. Colas, X. Guillaud, M.-S. Debry, and T. Prevost,
tems: Modeling, Control, and Applications. Hoboken, NJ, USA: Wiley, “Tuning of cascaded controllers for robust grid-forming voltage source
2010. converter,” in Proc. Power Syst. Comput. Conf. (PSCC), Jun. 2018.
[26] G. Denis, T. Prevost, M.-S. Debry, F. Xavier, X. Guillaud, and A. Menze,
“The Migrate project: The challenges of operating a transmission grid
with only inverter-based generation. A grid-forming control improve-
ment with transient current-limiting control,” IET Renew. Power Gener.,
vol. 12, no. 5, pp. 523–529, Apr. 2018.
[27] IEEE Recommended Practice for Excitation System Models for Power Ali Tayyebi received the B.Sc. degree in elec-
System Stability Studies, IEEE Standard 421.5, 2016. trical engineering from the University of Tehran,
[28] R. D. Zimmerman, C. E. Murillo-Sanchez, and R. J. Thomas, “MAT- Tehran, Iran, in 2012, the M.Sc. degree in engi-
POWER: Steady-state operations, planning, and analysis tools for power neering mathematics (Joint MATHMODS Program)
systems research and education,” IEEE Trans. Power Syst., vol. 26, no. 1, from the University of L’Aquila, L’Aquila, Italy,
pp. 12–19, Feb. 2011. and the University of Hamburg, Hamburg, Germany,
[29] P. Vorobev, P.-H. Huang, M. Al Hosani, J. L. Kirtley, and K. Turitsyn, in 2014, and the M.Sc. degree in sustainable trans-
“High-fidelity model order reduction for microgrids stability assess- portation and electric power systems (Joint STEPS
ment,” IEEE Trans. Power Syst., vol. 33, no. 1, pp. 874–887, Program) from the Sapienza University of Rome
Jan. 2018. (La Sapienza), Rome, Italy, the University of Not-
[30] V. Purba, S. V. Dhople, S. Jafarpour, F. Bullo, and B. B. Johnson, tingham, Nottingham, U.K., and the University of
“Reduced-order structure-preserving model for parallel-connected three- Oviedo, Oviedo, Spain, in 2016. He is currently pursuing the Ph.D. degree
phase grid-tied inverters,” in Proc. IEEE 18th Workshop Control Mod- with the Austrian Institute of Technology (AIT), Vienna, Austria, and the
eling Power Electron. (COMPEL), Jul. 2017. Automatic Control Laboratory, Swiss Federal Institute of Technology (ETH),
[31] M. M. S. Khan, Y. Lin, B. Johnson, V. Purba, M. Sinha, and S. Dhople, Zürich, Switzerland, with a focus on the design of grid-forming converter
“A reduced-order aggregated model for parallel inverter systems with control for low-inertia power systems.
virtual oscillator control,” in Proc. IEEE 19th Workshop Control Mod- In 2016, he joined AIT as an M.Sc. thesis candidate and then as a Research
eling Power Electron. (COMPEL), Jun. 2018. Assistant. His main research interests are the nonlinear systems and control
[32] V. Purba, B. B. Johnson, M. Rodriguez, S. Jafarpour, F. Bullo, and theory with applications to power systems.
S. V. Dhople, “Reduced-order aggregate model for parallel-connected Mr. Tayyebi was a recipient of the EU Scholarship for master studies from
single-phase inverters,” IEEE Trans. Energy Convers., vol. 34, no. 2, 2014 to 2016.
pp. 824–837, Jun. 2019.
[33] M. G. Taul, X. Wang, P. Davari, and F. Blaabjerg, “Current limiting
control with enhanced dynamics of grid-forming converters during fault
conditions,” IEEE J. Emerg. Sel. Topics Power Electron., to be published.
[34] H. Xin, L. Huang, L. Zhang, Z. Wang, and J. Hu, “Synchronous Dominic Groß (Member, IEEE) received the
instability mechanism of P-f droop-controlled voltage source converter Diploma degree in mechatronics and the Ph.D.
caused by current saturation,” IEEE Trans. Power Syst., vol. 31, no. 6, degree in electrical engineering from the University
pp. 5206–5207, Nov. 2016. of Kassel, Kassel, Germany, in 2010 and 2014,
[35] X. Wang, Y. W. Li, F. Blaabjerg, and P. C. Loh, “Virtual- respectively.
impedance-based control for voltage-source and current-source convert- He was with the Research Division, Volkswa-
ers,” IEEE Trans. Power Electron., vol. 30, no. 12, pp. 7019–7037, gen Group, Wolfsburg, Germany, from 2014 to
Dec. 2015. 2015. From 2016 to 2019, he was a Post-Doctoral
[36] Q.-C. Zhong and G. C. Konstantopoulos, “Current-limiting droop control Researcher with the Automatic Control Laboratory,
of grid-connected inverters,” IEEE Trans. Ind. Electron., vol. 64, no. 7, ETH Zürich, Zürich, Switzerland. He is currently
pp. 5963–5973, Jul. 2017. an Assistant Professor with the Department of Elec-
[37] A. Gkountaras, S. Dieckerhoff, and T. Sezi, “Evaluation of current trical and Computer Engineering Department, University of Wisconsin–
limiting methods for grid forming inverters in medium voltage micro- Madison, Madison, WI, USA. His research interests include distributed control
grids,” in Proc. IEEE Energy Convers. Congr. Expo. (ECCE), Sep. 2015, and optimization of complex networked systems with applications in power
pp. 1223–1230. systems dominated by power electronic devices.

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.
1018 IEEE JOURNAL OF EMERGING AND SELECTED TOPICS IN POWER ELECTRONICS, VOL. 8, NO. 2, JUNE 2020

Adolfo Anta received the Licenciatura degree Florian Dörfler (Member, IEEE) received the
from the ICAI Engineering School, Madrid, Spain, Diploma degree in engineering cybernetics from the
in 2002, and the M.Sc. and Ph.D. degrees from University of Stuttgart, Stuttgart, Germany, in 2008,
the University of California at Los Angeles, Los and the Ph.D. degree in mechanical engineering from
Angeles, CA, USA, in 2007 and 2010, respectively. the University of California at Santa Barbara, Santa
From 2002 to 2005, he was a Design Engineer with Barbara, CA, USA, in 2013.
Airbus, Madrid. From 2010 to 2012, he was a Post- From 2013 to 2014, he was an Assistant Professor
Doctoral Researcher with the Technical University with the University of California at Los Angeles,
of Berlin, Berlin, Germany, and the Max Planck Los Angeles, CA, USA. He is currently an Associate
Institute, Berlin. From 2012 to 2018, he was a Professor with the Automatic Control Laboratory,
Lead Researcher with GE Global Research Europe, ETH Zürich, Zürich, Switzerland.
Münich, Germany. He is currently a Research Engineer with the Austrian Dr. Dörfler’s students were the winners or finalists for the Best Student Paper
Institute of Technology (AIT), Vienna, Austria. His research interests cover Awards at the 2013/2019 European Control Conference, the 2016 American
a wide range of control applications, in particular stability issues in power Control Conference, and the 2017 PES PowerTech Conference. His articles
systems. received the 2010 ACC Student Best Paper Award, the 2011 O. Hugo Schuck
Dr. Anta received the Fulbright Scholarship in 2005, the Alexander von Best Paper Award, the 2012–2014 Automatica Best Paper Award, and the
Humboldt Fellowship in 2011, was a Finalist for the Student Best Paper Award 2016 IEEE Circuits and Systems Guillemin-Cauer Best Paper Award. He was a
at the IEEE Conference on Decision and Control in 2008, and received the recipient of the 2009 Regents Special International Fellowship, the 2011 Peter
2010 EMSOFT Best Paper Award and the IEEE CSS George S. Axelby Award J. Frenkel Foundation Fellowship, and the 2015 UCSB ME Best PhD Award.
in 2011.

Friederich Kupzog received the Diploma degree


in electrical engineering and information technology
from RWTH Aachen, Aachen, Germany, and the
Ph.D. degree from the Institute of Computer Tech-
nology, TU Wien, Vienna, Austria, in 2008.
In 2006, he joined the Institute of Computer
Technology, TU Wien. Until 2012, he stayed at TU
Wien as a Post-Doctoral Researcher and built up
the research group Energy & IT at the Institute of
Computer Technology. Since 2012, he has been a
Senior Scientist at the Austrian Institute of Technol-
ogy GmbH (AIT), Vienna. He has been the Head of the Competence Unit
Electrical Energy Systems, Vienna, since August 2018. His research interest
lies in verification methods for networked smart grid systems.

Authorized licensed use limited to: SHIBAURA INSTITUTE OF TECHNOLOGY. Downloaded on April 04,2024 at 13:51:05 UTC from IEEE Xplore. Restrictions apply.

You might also like