Sustainability 15 08400
Sustainability 15 08400
Article
Grid-Following Inverter-Based Resource: Numerical
State–Space Modeling
Abdullah Alassaf * , Ibrahim Alsaleh , Ayoob Alateeq and Hamoud Alafnan
Department of Electrical Engineering, College of Engineering, University of Hail, Hail 55211, Saudi Arabia;
i.alsaleh@uoh.edu.sa (I.A.); a.alateeq@uoh.edu.sa (A.A.); h.alafnan@uoh.edu.sa (H.A.)
* Correspondence: a.alassaf@uoh.edu.sa
Abstract: In the pursuit of a sustainable electric power system, the integration of renewable energy
sources and distributed energy resources is gradually replacing traditional power generation. These
new resources are integrated into the grid via inverters, which, despite their efficient performance,
present dynamic challenges to the power grid when implemented on a large scale. To maintain grid
stability and ensure effective regulation during abnormal operations, various modeling techniques
are necessary; while the dynamics of inverter-based resources (IBRs) are traditionally modeled by
transfer functions, this paper sheds light on differential-algebraic equations (DAEs) modeling and
numerical integration methods. The inherent limitations of transfer function modeling stem from its
restricted applicability, as it is exclusively suitable for linear and time-invariant systems. In contrast,
the nonlinear DAEs of the IBR system can be converted into a state–space form, which offers a
versatile framework for modeling, evaluating, and designing a diverse array of systems. In addition
to being compatible with time-varying systems and multiple-input multiple-output systems, the
state–space technique may incorporate saturation and dead zone characteristics into the dynamic
model. Our research focuses on IBR modeling in a grid-following scheme, which is current-controlled
and synchronized to the grid by a phase-locked loop (PLL). The presented state–space model consists
of the inverter, grid, control, and designed PLL. Beyond the discussion of its application to IBRs,
the presented method holds the potential to solve a wide range of DAEs. The proposed model is
compared with a benchmarked system.
is
+
Vsabc
Vtabc R + ron L
VDC + VDC
-
iabc
Vabc
abc 𝜌
- dq abc
dq 𝜌
id iq
Vsd Vsq
mabc
abc 𝜌
dq
VDC
md mq Reference
idref Pref
Control in dq-frame iqref Signal
Generator
Qref
2.1. dq-Frame
To control the grid-following converter-interfaced system, it is more practical to reduce
the system’s structure and convert the actual three-phase frame to either αβ-frame or dq-
frame. For large-scale systems, the latter is more convenient as its control is DC and it does
not require a large bandwidth. The following equation converts the three-phase frame to
dq-frame
h i
2π
4π f (t)
2 cos[ρ(t)] cos ρ(t) − 3 cosh ρ(t) − 3 i a
f d (t)
= f b ( t ) , (1)
f q (t) 3 sin[ρ(t)] sin ρ(t) − 2π sin ρ(t) − 4π
3 3 f c (t)
where f () can represent the voltage and the current. ρ is the PCC voltage angle that
allows for the frame conversion and is achieved from PLL. The abc-frame can be retrieved
as follows:
[ρ(t)]2π
cos sin[ρ(t)]2π f (t)
f a (t)
coshρ(t) − 3 i sinhρ(t) − 3 i
f b (t) = d
. (2)
4π 4π f q (t)
f c (t) cos ρ(t) − 3 sin ρ(t) − 3
where Vbs is the peak value of the line-to-neutral voltage, ω0 is the grid angular frequency,
and θ0 is the initial phase angle. The grid is modeled in abc-frame as follows:
i a (t) i a (t) Vta (t) Vsa (t)
d
L ib (t) = −( R + ron ) ib (t) + Vtb (t) − Vsb (t) . (4)
dt
ic (t) ic (t) Vtc (t) Vsc (t)
Sustainability 2023, 15, 8400 5 of 18
For this system, the control inputs Vtd , Vtq , and ω determine the dynamics of the state
variables id , iq , and ρ. It can be noticed that the system includes nonlinear components that
are ωid , ωiq , cos(ω0 t + θ0 − ρ), sin(ω0 t + θ0 − ρ). With the help of the Phase-Locked Loop
(PLL), the system phase angle can be accurately detected. ρ represents the detected angle
from PLL. With ρ = ω0 t + θ0 , the complexity of the nonlinearity disappeared
where Vsd = Vbs and Vsq = 0. In the current form, the system is linear and it is exited by Vs.
With Vtd and Vtq as DC, the state variables id and iq are also DC. The conversion angle ρ
is ensured to have the value of ω0 t + θ0 by the PLL mechanism, which is covered by the
following subsection. The active and reactive powers supplied to the PCC node are
As stated earlier, PLL guarantees that, at steady state, Vsq = 0. Thus, the active and reactive
power will be
Ps (t) = 23 Vsd (t)id (t),
(8)
Qs (t) = − 32 Vsd (t)iq (t).
Accordingly, Ps (t) and Qs (t) can simply be controlled by id and iq , respectively. The reference-
controlled signals of the active and reactive power can be set as follows:
With efficient control, the system variables track the reference settings.
- Vtd - 1
idref kd (s) md id
/ ×
Ls + (R + ron )
-
L𝜔 0
id
L𝜔0 VDC
L𝜔 0 2
iq L𝜔0
mq Vtq - 1
iqref kq (s) / × iq
- - Ls + (R + ron )
For this system, a controller is applied to make the inverter’s currents, both d and q
components, follow the change of the reference signals. The difference between the current
state and the reference signal is the error e. The controller k(s) compensates for the error
and generates the signal u, which is scaled by the division by the value of the DC source
and sets the inverter modulation signal m. Next, the inverter processes the modulation
signal m to produce the terminal voltage Vt , as in Equation (10). As the model structure is
DC, the PI controller is sufficient to track the signal reference. The designed controller suits
both d and q loops.
k p s + ki
k(s) = , (11)
s
where proportional and integral gains, respectively, are denoted by k p and k i . As a result,
the system open-loop gain
kp s + k i /k p
`(s) = . (12)
Ls s + ( R + ron )/L
Due to the fact that the system plant pole is inherently close to the origin, the magni-
tude and phase gains of the decline are found at low frequencies. Since the plant pole is
on the left side, stability concerns are not posed. The plant pole can be canceled by setting
kp
the controller zero s = −k i /k p equal to the plant pole. The loop gain will be `(s) = Ls .
Consequently, the closed-loop transfer function of the system evolves to be
Id (s) 1
Idre f (s)
= Gi (s) = τi s+1 ,
k p = L/τi , (13)
k i = ( R + ron )/τi ,
where τ represents the time constant of the closed-loop system. The value of τ should be
both large enough to have a wide bandwidth and small enough to have a fast response. G f f
represents the feed-forward filter, a crucial component employed to predict variations in the
load or system conditions, thereby facilitating the necessary adjustments to the inverter’s
output. The integration of feed-forward filters into inverter control systems offers several
advantages, such as accelerated response times, reduced distortion, and improved efficiency.
The inverter is equipped with the capability to modify its output proactively in anticipation
of fluctuations in the load or system parameters.
Sustainability 2023, 15, 8400 7 of 18
bs cos(ω0 t + θ0 − ρ),
Vsd (t) = V
bs sin(ω0 t + θ0 − ρ). (14)
Vsq (t) = V
dρ
= ω ( t ),
dt
= H ( p)Vsq (t), (15)
bs sin(ω0 t + θ0 − ρ),
= H ( p )V
dρ bs (ω0 t + θ0 − ρ).
= H ( p )V (16)
dt
The difference between the input, ω0 t + θ0 , and the output, ρ, is transmitted through
the transfer function, H ( p)V
bs , to perform the management of this feedback system. The de-
tailed scheme of PLL operation is shown in Figure 3. At the PLL’s termination, a voltage-
controlled oscillator (VCO) is set up to reset the measured angle once it hits 2π.
Va
abc Vd
Vb
Vq
Vc dq
𝜌
𝜌 VCO Limiter Compensator
∫ H(s)
The dynamical performance of the PLL is mostly determined by the design of the
compensator H (s) that is defined as follows:
where H1 (s) handles the system requirements and H2 (s) considers the stability require-
ments. The PLL has a single integrator at the origin, which is from the VCO. For that,
the constant component of the input signal, θ0 , can simply be tracked. However, since the
input signal has a ramp component that is ω0 t, another integrator is required to guarantee
that the steady-state error approaches zero. The compensator part H1 (s) may include this
pole among the other system’s design requirements.
Sustainability 2023, 15, 8400 8 of 18
To prevent malfunctioning in the control process, the PLL should also take care of the
distortion, e.g., harmonics, that are embedded in the grid voltage. Otherwise, the distortion
in the grid will be reflected in the converted frame, dq. The double harmonic component is
a significant cause for concern since it has an inherent large magnitude and a frequency
that is the closest to the fundamental frequency. The imbalanced component can also lead
to stability problems in the grid. We must consider the imbalanced component in designing
the PLL mechanism to avoid the PLL malfunctioning. The effect of the double component
distortion can be removed from the PLL when one pair of complex-conjugate zeros located
at ± j2ω0 is added to the compensator. As a result, the loop gain increases. In order to
reduce the loop gain magnitude beyond the frequency of 2ω0 , a double real pole is added
at s = −2ω0 . To date, the compensator taking into account the nominal value of the grid
voltage V bsn and the gain h becomes
s2 + (2ω0 )2
h
H (s) = H2 (s) (18)
Vsn
b s(s + 2ω0 )2
s + ( p/α)
H2 (s) = , (19)
s+p
where √
p = ωc α,
1 + sin δm (20)
α= ,
1 − sin δm
By considering the grid and stability requirements, the compensator becomes
s2 + (2ω0 )2 s + ( p/α)
h
H (s) = . (21)
V
bsn s(s + 2ω0 )2 s+p
that characterizes the loop gain is set to equal 1, which corresponds to the magnitude at
the crossover frequency. By considering V bsn = 391V, the gain of H (s) is h = 2.68 × 105 .
Without considering the compensator gain and the stability requirements, i.e., hH2 (s) = 1,
the phase angle of the system loop gain at the gain crossover (ωc = 200 rad/s) is −210◦ .
To guarantee stability, 60◦ phase margin is added. This criterion can be met using a lead
compensator. As the system needs 90◦ , two identical lead compensators of an angle of 45◦
are used. Following (19), the applied lead compensator
2
s + 83
H2 (s) = . (22)
s + 482
100
50
Phase [degrees]
100
150
200
100 101 102 103 104
Frequency [Hz]
Figure 4. The frequency response of the PLL.
Beyond the crossover frequency, the system gain continues its downward trajectory
with an unchanging slope of −40 dB/dec. This consistent decline in gain beyond the
crossover frequency highlights the system’s inherent ability to suppress higher-frequency
components and maintain stability under a range of operating conditions. The imbalanced
component corresponding to the frequency of 754 rad/s is effectively attenuated by the
designed PLL, thereby aligning Vq to zero. This effective attenuation underpins the efficacy
of the designed PLL, demonstrating its capability to maintain system balance and stability
despite the presence of an imbalanced component.
The single s in the denominator represents ω and as it passes the integrator in the VCO
it becomes ρ, the abc − dq transformation angle. For normal PLL operation, ω has to be
Sustainability 2023, 15, 8400 10 of 18
initialized with the grid frequency. To ease the state–space transformation, this integrator is
separated from the rest of H (s), as follows:
For simplicity, the numerator components are denoted n1 . . . n5 , and the denominator
components are denoted d1 . . . d5 . The process of the state–space transformation first
decomposes the transfer function into several integrators. Then, each integrator represents
a state of the system. Figure 5 clarifies the procedure.
n2 - d2 +
Va n3 - d3 +
abc Vd
Vb Vq n4 - d4 +
+
Vc dq - 1 1 1 1
𝜌 - n5 - d5 +
- x␒ 1 s x␒ 2 s x␒ 3 s x␒ 4 s
-
d2
d3
d4
d5
VCO Limiter
1 𝜔 1
h
s x␒ 6 s x␒ 5
The system states are revealed by breaking down the PLL compensator’s transfer
function, H (s). Initializing the angular frequency, ω, is required for PLL operation, as was
previously explained. For this model, ẋ5 represents ω. Given that the system’s frequency is
60Hz, the initialization of ẋ5 is performed by 60 × 2π rad/s. Furthermore, to guarantee the
normal performance of the PLL, ω must be constrained so that the transformation angle
does not highly deviate from the grid angle. The transformation angle, θ, is represented by
the state ẋ6 . For that, the integrator, which is indicated as VCO, resets itself as it reaches 2π.
The purpose of the VCO is to achieve physical meaning, which is the circular rotation of
the angle. The state–space form of the PLL
While the model can be run to operate by itself as an individual system, it can be extended
to different grid configurations. For both cases, the input is abc-frame and the output is the
regulated dq-frame.
−1
Vsd
0 0 0 0 0
gf gf
x7 x7
−1
Vsq
0 0 0 0 0
x8 gf
x8
gf
d x9 0
= 0 0 0 − Ki 0 x9
+ Ki idre f
. (26)
dt
x10 0
0 1 0 0 − Ki
x10
Ki iqre f
x11 −(K p + Rt ) x11 −(Vsd −K p idre f )
1
1
0 0 0
L L L
L
x12 x12
1 1 −(K p + Rt ) −(Vsq −K p iqre f )
0 L 0 L 0 L L
1 1
s gf Vsd
Kp x␒ 7 -
ron + R
- Vtd - -
idref 1 md 1 1
id
Ki
s
/ × L s
x␒ 9 x␒ 11
-
L𝜔0
id L𝜔0 VDC
L𝜔0 2
iq L𝜔0
mq Vtq -
iqref Ki
1
/ ×
1
L
1
iq
x␒ 10 s
- x␒ 12 s
- -
1 1
Kp ron + R
s
x␒ 8
gf Vsq
-
ẋ = ( x, u), (27)
where x and u denote the state and the input of the system, respectively. FEM is a first-order
technique, and the process of its integration is straightforward. The state step is mainly
based on the preceding state step. FEM is generally formulated as follows:
x (i+1) = x (i) + ∆t f x (i) , u(i) , (28)
Sustainability 2023, 15, 8400 12 of 18
where x (i) is available from the prior step and ∆t is the time step. Despite its reputation for
simplicity and stability, FEM is the least precise method.
BEM is an implicit method and is known for its high stability. BEM integrates the
previous state with the current state, which is attained by numerical approaches, such
as the Newton method if the system contains nonlinear components. The BEM standard
form is
x (i+1) = x (i) + ∆t f x (i+1) , u(i+1) . (29)
which is equivalent to
0n,1 = h x (i) , x (i+1) , u(i+1) , ∆t = h(i+1) . (31)
where
( i +1) ∂h(i+1) ∂ f ( i +1)
H(k) = ( i +1)
= ∆t ( i +1)
− In . (33)
∂x(k) ∂x(k)
( i +1)
The initial value of the state iteration is x(0) = x (i) .
The implicit Trapezoidal Method (ITM) combines the BEM and the FEM to provide
excellent stability and precision in the calculation. The ITM is attained by combining the
FEM and BEM as follows:
1 1
h(i+1) = x (i) + ∆t f x (i) , u(i) + ∆t f x (i+1) , u(i+1) − x (i+1) , (34)
2 2
and
( i +1) ∂h(i+1) 1 ∂ f ( i +1)
H(k) = ( i +1)
= ∆t − In . (35)
∂x(k) 2 ∂x (i+1)
(k)
The ITM is one of the most trustworthy and efficient techniques. In fact, it is the solver of
preference for commercial and noncommercial power system software products. The ITM
standard form is
∆t (i+1) (i+1) ∆t (i) (i)
x ( i +1) = x ( i ) + f x ,u + f x ,u . (36)
2 2
5. Case Study
5.1. Robust PLL Performance under Voltage Imbalance Conditions
In the design of modern power systems, an essential consideration is the ability to
maintain a stable operation in the presence of voltage imbalances. This case study presents a
thorough evaluation of the proposed PLL model (26), demonstrating its robust performance
under various conditions, including the presence of voltage imbalances. The PLL model,
grounded in the natural behavior of the operating system, is specifically designed to address
the challenges posed by imbalanced components, a common occurrence in distribution
systems. A comprehensive simulation is conducted, assessing the model’s response to
three distinct scenarios, as illustrated in Figure 7.
Sustainability 2023, 15, 8400 13 of 18
Initially, the PLL operates under ideal conditions, with pure voltage composed solely
of the fundamental component. This baseline evaluation serves to validate the model’s
performance when no imbalances are present. Following this, at t = 0.02 s, an imbalance
is introduced to the system to assess the PLL’s ability to adapt and maintain accurate and
stable operation. The simulation results show that the PLL continues to function effec-
tively, with the distortion caused by the imbalance manifesting primarily in the direct- and
quadrature-axis voltage components, Vd and Vq . This observation confirms the model’s
ability to handle voltage imbalances while producing minimal distortion in the output
voltages. Finally, at t = 0.08 s, the imbalance is removed from the system, allowing for
the evaluation of the PLL’s response to the restoration of balanced conditions. The sim-
ulation demonstrates that the PLL promptly returns to smooth operation, reaffirming its
adaptability and resilience in the face of varying operating conditions.
500
Vabc
−500
391
Vdq
377
ω
2π
ρ
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Time (s)
Figure 7. Illustrating the performance of the designed PLL model under various conditions, including
the presence and removal of voltage imbalances. The first subplot illustrates the three-phase voltage
signals, with the imbalance component evident in the blue phase. The second subplot represents the
Vd and Vq variables, depicted in blue and orange, respectively. The third subplot reveals the grid’s
angular frequency, while the fourth subplot discloses the detected grid’s angle.
The case study under consideration was also subjected to tests involving the or-
dinary Phase-Locked Loop (PLL) to evaluate its performance and effectiveness under
circumstances of imbalanced operational conditions. This meticulous examination was
orchestrated to assess the resilience of the PLL when it encounters conditions that veer
from the balanced state, thus testing the robustness of its design. The outcomes of the
simulation are visually represented in Figure 8. The graphic illustration provides a clear
representation of the system’s behavior during and after the intrusion of the imbalanced
component. A careful inspection of the figure reveals that the performance of the PLL
begins to exhibit signs of deterioration immediately following the onset of the imbalanced
component. This indicates that the system’s stability and efficiency are adversely affected
by such disruptions.
In the specific context of operations under the influence of the imbalanced component,
it is noteworthy that the PLL encounters difficulty in accurately detecting the fundamental
Sustainability 2023, 15, 8400 14 of 18
frequency of the grid. This is a significant observation as it highlights the limitations of the
PLL under non-ideal conditions. Furthermore, it is observed that, once the imbalanced
component is eliminated from the system, the PLL is unable to revert to its normal operation.
This suggests a lack of adaptive recovery mechanisms in the PLL to restore balance and
resume normal functioning after the system experiences instability. This inability to self-
correct and re-establish normal operating parameters underscores the need for further
improvements in the design and control strategy of the PLL to enhance its resilience to
imbalances and other operational anomalies.
500
Vabc
−500
391
Vdq
377
ω
360
2π
ρ
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Time (s)
Figure 8. The performance of the ordinary PLL model under various conditions, including the
presence and removal of voltage imbalances. The first subplot displays the three-phase voltage
signals, wherein the blue phase contains the imbalance component. The second subplot illustrates
the Vd and Vq signals, represented in blue and orange, respectively. The third subplot exhibits the
grid’s angular frequency, while the fourth subplot presents the detected angle of the grid.
Parameter Value
ω0 377 rad/s
Vs d 391 V
L 100 µL
R 0.75 Ω
ron 0.88 Ω
VDC 1250 V
fs 3420 Hz
gf 8 × 10−6
τi 2 ms
kp 0.05
ki 0.815
391
Vabc
−391
391
Vdq
377
ω
2π
ρ
1000
id ,idref
2000
iq ,iqref
0
0.1 0.2 0.3 0.4 0.5
Time (s)
Figure 9. The simulation of an inverter connected to the grid with changes in the references. The
first subplot illustrates the three-phase voltage signals, with the imbalance component present in the
blue phase. The second subplot represents the Vd and Vq variables, indicated in blue and orange,
respectively. The third subplot displays the grid’s angular frequency, while the fourth subplot
discloses the detected grid’s angle. The fifth subplot depicts id and its control signal, and, the sixth
subplot exhibits iq and its associated control signal.
The third subplot illustrates the grid angular frequency measured by the PLL. In the
state–space model, the angular frequency is denoted by x5 . The grid angle, represented by
x6 , is computed by integrating the limited angular frequency. While regular integrators
can perform this integration, resettable integrators are necessary to restrict the angle range
between 0 and 2π.
The fifth subplot demonstrates the inverter current aligned to the d-axis and its
reference. The active power, as explained in Equation (8), is related to this current and
can be primarily regulated using it. To assess the effectiveness of the control applied to
id , the reference current, idre f , starts with a value of 0 and is changed to 1000 at t = 0.1 s.
As observed in the plot, the response of id adequately follows its reference.
Sustainability 2023, 15, 8400 16 of 18
Similarly, the sixth subplot shows the inverter current that is aligned to the q-axis and
its reference. The reactive power is associated with it, iq , as in Equation (8). Furthermore, its
response is tested to validate the system control. The reference current, iqre f , starts with a
value of 0, and is then changed at t = 0.2 s to 2000. As displayed, the response appropriately
changes and matches its reference.
391
Vabc
−391
391
Vdq
377
ω
200
id ,idref
50
iq ,iqref
0
0.1 0.2 0.3 0.4 0.5
Time (s)
Figure 10. Simulation with voltage and frequency change events that occur in the grid. The first
subplot shows the three-phase voltage signals, with the imbalance component appearing in the
blue phase. The second subplot displays the variables Vd and Vq , represented in blue and orange,
respectively. The third subplot illustrates the grid’s angular frequency. The fourth subplot presents
the detected angle of the grid. In the fifth subplot, we can see id and its control signal and the sixth
subplot exhibits iq and its corresponding control signal.
In the conducted simulation, the inverter initially operates under normal conditions.
However, at t = 0.08 s, we introduced a disruption by reducing the grid voltage to 85% of
its nominal value. This sudden drop persisted until 0.2 s, at which point we allowed the
grid to recover to its nominal voltage. The response of the designed PLL to this sudden
change was prompt and effective. The PLL swiftly tracked the alteration in grid voltage
and mirrored it in its DC values, thereby demonstrating its ability to respond to voltage
variations. As depicted in Figure 10, the PLL was also capable of restoring normal operation
in sync with the grid recovery, further illustrating its adaptability and resilience.
Sustainability 2023, 15, 8400 17 of 18
6. Conclusions
The penetration of IBRs that interface renewables and DERs changes the power system
infrastructure. The importance of modeling can never be overstated. Transfer functions are
frequently used to model the grid connection of IBRs. On the other hand, this paper models
it with differential-algebraic equations which can be converted into state–space form.
The state–space modeling is superior to the transfer function modeling in addition to being
in the time domain. Multiple-input, multiple-output systems, and time-variant systems can
both be handled by the state–space approach. State–space models can also be employed to
implement Kalman filters. We also develop the state–space model of a PLL customized to
address imbalanced components. Reaching improved methods of analysis and simulation
is the purpose of achieving a different sort of modeling. Furthermore, preparing the system
for the deployment of state–space optimal control strategies is another objective of the grid-
following IBR state–space model development. By employing primary integration methods
capable of addressing nonlinear systems, our proposed model has been numerically solved,
demonstrating a precise correspondence between the transfer function model and the
state–space model. Through the case studies presented, the validity and robustness of the
state–space approach have been further substantiated, providing a strong foundation for
future research and development in the field of power electronics and grid integration.
Author Contributions: Conceptualization, A.A. (Abdullah Alassaf) and I.A.; methodology, A.A.
(Abdullah Alassaf) and I.A.; software, A.A. (Abdullah Alassaf) and I.A.; validation, A.A. (Ayoob
Alateeq) and H.A.; writing—original draft preparation, A.A. (Abdullah Alassaf); formal analysis,
A.A. (Ayoob Alateeq) and I.A.; investigation, A.A. (Ayoob Alateeq) and H.A.; writing—review and
editing, A.A. (Ayoob Alateeq) and H.A.; project administration, I.A.; funding acquisition, I.A. All
authors have read and agreed to the published version of the manuscript.
Funding: This research has been funded by Scientific Research Deanship at University of Ha’il, Saudi
Arabia through project number BA-2110.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data used to perform the case studies are available upon request from
the corresponding author.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Chow, J.H.; Sanchez-Gasca, J.J. Power System Modeling, Computation, and Control; John Wiley & Sons: Hoboken, NJ, USA, 2020.
2. Pepermans, G.; Driesen, J.; Haeseldonckx, D.; Belmans, R.; D’haeseleer, W. Distributed generation: Definition, benefits and issues.
Energy Policy 2005, 33, 787–798. [CrossRef]
3. Geng, S.; Hiskens, I.A. Unified Grid-Forming/Following Inverter Control. IEEE Open Access J. Power Energy 2022, 9, 489–500.
[CrossRef]
4. Li, Y.; Gu, Y.; Green, T.C. Revisiting grid-forming and grid-following inverters: A duality theory. IEEE Trans. Power Syst. 2022,
37, 4541–4554. [CrossRef]
5. Huang, L.; Xin, H.; Li, Z.; Ju, P.; Yuan, H.; Lan, Z.; Wang, Z. Grid-synchronization stability analysis and loop shaping for
PLL-based power converters with different reactive power control. IEEE Trans. Smart Grid 2019, 11, 501–516. [CrossRef]
6. Cao, Y.; Yu, J.; Xu, Y.; Li, Y.; Yu, J. An Efficient Phase-Locked Loop for Distorted Three-Phase Systems. Energies 2017, 10, 280.
[CrossRef]
Sustainability 2023, 15, 8400 18 of 18
7. Blaabjerg, F.; Teodorescu, R.; Liserre, M.; Timbus, A.V. Overview of control and grid synchronization for distributed power
generation systems. IEEE Trans. Ind. Electron. 2006, 53, 1398–1409. [CrossRef]
8. Wang, X.; Taul, M.G.; Wu, H.; Liao, Y.; Blaabjerg, F.; Harnefors, L. Grid-synchronization stability of converter-based resources—An
overview. IEEE Open J. Ind. Appl. 2020, 1, 115–134. [CrossRef]
9. Blaabjerg, F. Control of Power Electronic Converters and Systems; Academic Press: Cambridge, MA, USA, 2018; Volume 2.
10. Hsieh, G.C.; Hung, J.C. Phase-locked loop techniques. A survey. IEEE Trans. Ind. Electron. 1996, 43, 609–615. [CrossRef]
11. Ortega, A.; Milano, F. Comparison of different PLL implementations for frequency estimation and control. In Proceedings of the
2018 18th International Conference on Harmonics and Quality of Power (ICHQP), Ljubljana, Slovenia, 13–16 May 2018; pp. 1–6.
12. Milano, F.; Manjavacas, Á.O. Frequency Variations in Power Systems: Modeling, State Estimation, and Control; John Wiley & Sons:
Hoboken, NJ, USA, 2020.
13. Rodríguez, P.; Pou, J.; Bergas, J.; Candela, J.I.; Burgos, R.P.; Boroyevich, D. Decoupled double synchronous reference frame PLL
for power converters control. IEEE Trans. Power Electron. 2007, 22, 584–592. [CrossRef]
14. Alturki, M.; Abbassi, R.; Albaker, A.; Jerbi, H. A New Hybrid Synchronization PLL Scheme for Interconnecting Renewable
Energy Sources to an Abnormal Electric Grid. Mathematics 2022, 10, 1101. [CrossRef]
15. Filipović, F.; Petronijević, M.; Mitrović, N.; Banković, B.; Kostić, V. A novel repetitive control enhanced phase-locked loop for
synchronization of three-phase grid-connected converters. Energies 2019, 13, 135. [CrossRef]
16. Ciobotaru, M.; Teodorescu, R.; Blaabjerg, F. A new single-phase PLL structure based on second order generalized integrator. In
Proceedings of the 2006 37th IEEE Power Electronics Specialists Conference, Jeju, Republic of Korea, 18–22 June 2006; pp. 1–6.
17. Nicastri, A.; Nagliero, A. Comparison and evaluation of the PLL techniques for the design of the grid-connected inverter systems.
In Proceedings of the 2010 IEEE International Symposium on Industrial Electronics, Bari, Italy, 4–7 July 2010; pp. 3865–3870.
18. Cataliotti, A.; Cosentino, V.; Nuccio, S. A phase-locked loop for the synchronization of power quality instruments in the presence
of stationary and transient disturbances. IEEE Trans. Instrum. Meas. 2007, 56, 2232–2239. [CrossRef]
19. Glazunova, A.; Semshikov, E.; Negnevitsky, M. Real-Time Flexibility Assessment for Power Systems with High Wind Energy
Penetration. Mathematics 2021, 9, 2056. [CrossRef]
20. Gude, S.; Chu, C.C. Three-phase PLLs by using frequency adaptive multiple delayed signal cancellation prefilters under adverse
grid conditions. IEEE Trans. Ind. Appl. 2018, 54, 3832–3844. [CrossRef]
21. Milano, F.; Dassios, I.; Liu, M.; Tzounas, G. Eigenvalue Problems in Power Systems; CRC Press: Boca Raton, FL, USA, 2020.
22. Kroutikova, N.; Hernandez-Aramburo, C.A.; Green, T.C. State-space model of grid-connected inverters under current control
mode. IET Electr. Power Appl. 2007, 1, 329–338. [CrossRef]
23. Yazdani, A.; Iravani, R. Voltage-Sourced Converters in Power Systems: Modeling, Control, and Applications; John Wiley & Sons:
Hoboken, NJ, USA, 2010.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.