Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

applied

sciences
Article
Modeling and Control Simulation of Power Converters in
Automotive Applications
Pierpaolo Dini *,† and Sergio Saponara †

Department of Information Engineering, University of Pisa, 56122 Pisa, Italy; sergio.saponara@unipi.it


* Correspondence: pierpaolo.dini@ing.unipi.it
† These authors contributed equally to this work.

Abstract: This research introduces a model-based approach for the analysis and control of an
onboard charger (OBC) system for contemporary electrified vehicles. The primary objective is to
integrate the modeling of SiC/GaN MOSFETs electrothermal behaviors into a unified simulation
framework. The motivation behind this project stems from the fact that existing literature often
relies on finite element method (FEM) software to examine thermal dynamics, necessitating the
development of complex models through partial derivative equations. Such intricate models are
computationally demanding, making it difficult to integrate them with circuit equations in the same
virtual environment. As a result, lengthy wait periods and a lack of communication between the
electrothermal models limit the thorough study that can be conducted during the design stage. The
selected case study for examination is a modular 1ϕ (single phase) onboard computer (OBC). This
system comprises a dual active bridge (DAB) type DC/DC converter, which is positioned after
a totem pole power factor correction (PFC) AC/DC converter. Specifically, the focus is directed
toward a 7 kW onboard computer (OBC) utilizing high-voltage SiC/GaN MOSFETs to ensure optimal
efficiency and performance. A systematic approach is presented for the assessment and selection
of electronic components, employing circuit models for the totem pole power factor correction
(PFC) and dual active bridge (DAB) converter. These models are employed in simulations closely
mimicking real-world scenarios. Furthermore, rigorous testing of the generated models is conducted
across a spectrum of real-world operating conditions to validate the stability of the implemented
control algorithms. The validation process is bolstered by a comprehensive exploration of parametric
variations relative to the nominal case. Notably, each simulation adheres to the recommended
Citation: Dini, P.; Saponara, S.
operational limits of the selected components and devices. Detailed data sheets encompassing
Modeling and Control Simulation of
Power Converters in Automotive
electrothermal properties are provided for contextual reference.
Applications. Appl. Sci. 2024, 14, 1227.
https://doi.org/10.3390/app14031227 Keywords: simulation; modeling; control; AC/DC converters; DC/DC converters; automotive

Academic Editors: Luigi Fortuna and


Mostafa I. Marei

Received: 9 November 2023 1. Introduction


Revised: 23 January 2024 1.1. Motivations
Accepted: 24 January 2024
In recent years, the commercial panorama of electrified cars has experienced unpar-
Published: 1 February 2024
alleled growth. The increasing need for extremely efficient electric and hybrid cars is a
modern challenge that is driving important advances in science. It is noteworthy that the
creation of resilient electronic power flow systems has become a central area of interest
Copyright: © 2024 by the authors.
for innovation [1,2]. For a variety of compelling reasons, it is crucial to use mathematical
Licensee MDPI, Basel, Switzerland.
models and run simulations when designing in-vehicle electronics systems, such as power
This article is an open access article electronics, onboard mechatronics, electric drives, and power electronics. Design optimiza-
distributed under the terms and tion: Using mathematical models and simulations, it is possible to assess several design
conditions of the Creative Commons configurations and choose the most promising one that will provide optimal performance
Attribution (CC BY) license (https:// in terms of effectiveness, dependability, and durability over an extended period of time.
creativecommons.org/licenses/by/ This approach can prevent expensive design mistakes and shorten the time needed to
4.0/). produce a product from start to finish [3,4].

Appl. Sci. 2024, 14, 1227. https://doi.org/10.3390/app14031227 https://www.mdpi.com/journal/applsci


Appl. Sci. 2024, 14, 1227 2 of 27

Cost reduction: The adoption of mathematical models and simulations obviates the
necessity of creating physical prototypes to test various design setups. Consequently,
this leads to a substantial reduction in product development costs and expedites the time
required to bring the product to the market [5–7]. Performance improvement: Power elec-
tronic converters may have their performance carefully assessed under a range of operating
situations, including variations in temperature or load, by using mathematical models
and simulations. This makes it possible to adjust converter control and guarantees top
performance in a variety of operating conditions [8]. Safety assurance: Through the use of
mathematical models and simulations, it is possible to evaluate power electronic converter
behavior in the event of overload or failure. This opens the door for the installation of
strong defenses, guaranteeing the security of both drivers and passengers [9–14]. The
use of mathematical models and a model-based design (MBD) approach is essential to
increase the performance of power electronic systems and to analyze the solution space,
including studying the effects of the variation of circuit parameters, the effects of choices
and simplified hypotheses in the design, and the robustness of control algorithms [15,16]
against unwanted effects such as measurement noise or transduction errors [17,18]. In the
literature, there are many concrete examples of the development and analysis of power
electronic systems and electric drives for automotive applications and the mechatronics
industry. For example, the authors of [19,20] use the MBD approach very extensively to
improve the performance of the control of electric drives in the presence of the Cogging
effect, analyzing the robustness and profiling the computational complexity in simulation.
In [21,22], an analysis of the computational complexity of advanced control algorithms
for power electronic systems based on SW MBD tools is proposed to ensure integration
with real-time constraints in embedded platforms. In [23–26], the use of the MBD approach
for the design and integration of mechatronic systems up to the implementation phases is
proposed. In other works, the authors show how to use MBD and simulation environments
massively, to compensate for the lack of experimental data and address battery health
estimation problems. Works such as [27] show how simulations can create scalable models
of SiC/GaN devices not yet on the market, combining experimental characterization mea-
surements of such devices with complex and scalable models, to be integrated into more
complex electronic systems. Essentially, the use of mathematical models and simulations in
power electronic converter designs for automotive applications helps to ensure the highest
level of user and vehicle safety while also improving product performance and reducing
development costs [28–31]. During these advancements, the growing significance of accu-
rately modeling and simulating complex systems has become evident. A comprehensive
understanding of the intricacies inherent in electric battery charging is particularly crucial
for the design of electric vehicles. Consequently, an exceedingly detailed model is presented,
encompassing the electrothermal dynamics of an onboard computer (OBC) with a modular
design [32–35]. Developing a precise simulation model for onboard electronic systems is
a critical first step in creating innovative, effective gadgets. Model development is often
oriented toward fine-grained attention on certain aspects, such as voltage and current
behavior, thermal patterns, or the subtleties of component magnetism. Still, this method
restricts the scope for a thorough analysis of important scenarios, and ‘destructive’ testing
is, thus, required to investigate difficult operational combinations [36–39]. By utilizing a
multifaceted model that encompasses many components, these constraints are addressed.
While it is understood that not every physical component can be precisely represented
using traditional software methods, finding a balance between model complexity and other
factors is still crucial. This balance makes it easier to quickly and efficiently assess a variety
of scenarios and operational circumstances [40–43]. Hence, the proposed operating protocol
advocates the development of a simulation model that integrates two essential aspects: the
circuit dynamics and the thermal conductivity of switching components. This approach
aims to bridge the divide between comprehensive modeling and practical implementation,
utilizing the 1ϕ OBC circuit configuration as a case study.
Appl. Sci. 2024, 14, 1227 3 of 27

This initiative is poised to usher in a new era of innovation in the realm of electronic
power systems for modern automobiles.

1.2. Related Works


In [44], the description of a unidirectional onboard charger (OBC) configuration in-
cludes modeling and design aspects. This configuration incorporates a step-up DC/DC
converter following an AC/DC diode rectifier stage to achieve power factor correction
(PFC). The system connects to the vehicle’s battery through a phase-shifted full-bridge
(PSFB) type DC/DC converter. The control algorithm relies on electrical circuit modeling
to primarily define transfer functions. In contrast, the current work incorporates a model
that integrates the complexities of a dual active bridge (DAB) circuit and a totem pole
circuit into the control algorithm. Additionally, a detailed examination of the thermal
behavior of SiC/GaN MOSFETs, a critical aspect absent from the cited paper, is included. In
contrast, ref. [45] introduces an alternative control approach employing a step-up converter
following a diode rectifier for the AC/DC conversion stage. However, their simulation
experiments primarily concentrate on diverse operating scenarios. In contrast, the present
study integrates the intrinsic properties of the switching components and conducts thermal
evaluations during simulations, employing a more comprehensive modeling method for
onboard chargers (OBCs). Subsequently, in [46], a thermal study of a ∼7 kW onboard
charger (OBC) is conducted using a multi-stage output DC/DC converter and finite ele-
ment method software. The circuit analysis is based on a rudimentary linear approximation,
while the thermal analysis is thorough. In contrast to the benefits of the proposed model-
based design (MBD) approach, the disadvantage of segregating dynamic and thermal
simulations into separate software environments, resulting in heightened computing costs,
remains significant, particularly for predictive diagnostics. Similarly, unlike the thorough
method employed in this study, ref. [47] does not consider the thermal properties of the
components and instead focuses on modeling the electrical behavior of the onboard charger
(OBC) at lower power and voltage outputs. Furthermore, neither ref. [48] nor ref. [49]
includes comprehensive modeling of the thermal behavior of electronic components, an
essential aspect integrated into the suggested framework in this study; moreover, they do
not provide specific design details of the OBC device. Similar to this, ref. [50] employs an
MBD technique for the validation of control techniques in onboard chargers (OBCs). In
contrast to the all-inclusive technique applied in this study, it solely addresses the analysis
of electrical behavior, disregarding thermal behavior. Likewise, refs. [51,52] primarily
concentrate on energy efficiency and simulations without incorporating a comprehensive
model of the thermal behavior of electronic devices, another crucial aspect of the suggested
work. Based on the literature study, it is evident that the majority of published research
focuses on the dynamics of electrical quantities, ignoring the important connection to the
thermal behavior of the SiC/GaN MOSFETs in the energy conversion stages. Furthermore,
even energy-efficient works neglect to represent temperature elements, which are essen-
tial for confirming the physical boundaries of simulated electronic devices, including the
transistors that make up electronic converters.

1.3. Authors’ Contributions


This work represents a significant advancement in the design of electronic power
systems for electric and hybrid cars. Advanced methodologies have been introduced in the
design of onboard chargers (OBCs), addressing the electrothermal behavior of SiC/GaN
MOSFETs. Utilizing modern 900 V GaN and 1200 V SiC technologies, a comprehensive
system has been developed to ensure optimal efficiency and performance in a 7 kW OBC
environment. The systematic approach to electrical component selection has contributed
to a deeper understanding of interactions between complex circuit models and realis-
tic simulations.
The rigorous validation process, encompassing various operating scenarios and ro-
bustness evaluations, has yielded an innovative outcome, opening new perspectives in
Appl. Sci. 2024, 14, 1227 4 of 27

research and establishing a solid foundation for the future development of electronic power
systems in the automotive industry.

2. Electrothermal Modeling and Control Algorithm


This section provides a detailed description of operational concepts and control algo-
rithms in energy conversion and explains the OBC circuit being considered. A thorough
evaluation of the thermal performance of the switching components is suggested to verify
the physical limitations of the transistors simulated in Simulink/Simscape (MATLAB 2020b
version in our work). The general configuration of the OBC circuit is shown in Figure 1. It
should be noted that the suggested model-based design (MBD) technique is easily adapt-
able to different circuit topologies, including three-phase circuits, even with only one
single-phase connection to the power supply.

Figure 1. Diagram illustrating the stages of power domain conversion in the reference device.

2.1. AC/DC Converter Model and Control


The study examines the totem pole configuration, a specific type of AC/DC converter
with PFC [53–63]. In Figure 2, there are two SiC transistors (SD1 , SD2 ) and two GaN
transistors (Q1 , Q2 ). The SiC components rectify the current at line frequency, replacing
traditional diodes, while the GaN components perform fast switching at frequencies above
100 kHz. The circuit operates in two modes, depending on the polarization of the AC line
voltage, as shown in the schematic in Figure 3. The activation of SD1 and the deactivation
of SD2 together with the GaNs define the positive AC half-cycle. The modulation logic
handles the negative AC half-cycle, with SD2 active, by deactivating SD1 and activating the
GaNs. This cyclic interaction between phases produces the desired PFC effect, aligning the
current with the AC supply voltage and eliminating reactive power. The SiCs handle the
rectification, while the GaNs orchestrate the switching to obtain the correct voltage on the
DC bus. Figure 4 shows a schematic of the complex control system of a totem pole-type
AC/DC converter. The control is divided into various functional blocks that are crucial
for the operation of the system. A second-order low-pass filter (LPF) separates the in-line
and 90-degree offset components of the AC voltage. A phase-locked loop (PLL) calculates
the phase compensation required to balance the current in the totem pole input connection
with the supply voltage. The feedback controller mediator corrects errors between the
actual voltage, Vo , and the reference, Vre f , for the DC bus; it also handles the AC Iac .

Figure 2. Selected bidirectional PFC active rectifier circuit configuration.


Appl. Sci. 2024, 14, 1227 5 of 27

Figure 3. Current route during the changing operational states of the MOSFETs.

Figure 4. Automatic control of a totem pole PFC converter.

The modulation system, based on an inverse calculation mechanism, synchronizes


∗ and the phase compensation ω to ensure the
the generation of the required AC voltage Vac
effective operation of the system.
    
ω̂ g ωr s
VQ = 2 s2 +ω̂2 s2 + ω 2
+ s2 +ω̂2 Vg
g r g
(1)
 
I Ns
ω̂ g = P + + s D s+ N VQ

Figure 5 highlights the importance of the second-order low-pass filter to obtain a signal
with a 90-degree phase shift relative to the source voltage. This phase shift allows the phase-
locked loop (PLL) block to accurately calculate the frequency required to synchronize
the main voltage and current in the totem pole, achieving the desired PFC effect. The
accurate measurement of the voltage on the DC bus depends on the control logic of the
voltage and current feedback. Next, Figure 6 illustrates the internal configuration of the
PLL block and how it processes input signals, such as α and β from the LPF block. The
result of this complex block is translated into carefully calculated trigonometric functions
related to the ω pulsation, which are fundamental for coordinating the complex dynamics
of the system. Equation (1) shows the transfer functions of VQ and the estimated network
frequency/pulse, which are the outputs of the PLL system.
Appl. Sci. 2024, 14, 1227 6 of 27

Figure 5. Schematic illustration of the reason for using a simple LPF for AC source frequency detection.

Figure 6. The phase-locked-loop shown in a block diagram.

The signal processing system is detailed in Figure 6. It is based on the Park transfor-
mation of the input signals to obtain components along the direct and quadrature axes. The
goal is to cancel the quadrature component to achieve the desired PFC effect in the totem
pole. A PID controller corrects any residual errors. This complex process identifies the
impulse needed to control the current in the totem pole and, after integration, produces the
angle for the Park transformation. The sine and cosine trigonometric functions are crucial
in the feedback control block. Figure 7 depicts the control logic governing the voltage on
the DC bus and the current in the totem pole. Employing PWM modulation concepts, the
objective is to simulate the reverse engineering of an inverter designed for conventional
DC-AC conversion but operating with active devices. The external control loop, supported
by a PID system, stabilizes voltage errors by calculating the magnitude of the current at
the totem pole. With the modulus and the cosine function, adapted to the AC voltage,
the reference value for the internal current control is obtained. In the internal control,
the proportional-resonant (PR) controller is preferred over the traditional PID to handle
sinusoidal systems.

Figure 7. An illustration (in block diagram form) of the voltage/current feedback controller.
Appl. Sci. 2024, 14, 1227 7 of 27

It combines proportional and resonant, offering good performance and simplicity of


implementation, especially in active rectifiers, where standard PI/PID controllers may
have limitations.
"  # 
re f I Ns re f
 s Kr s
Vg = P+ +D Vo − Vo 2 − I g K p + + Vg (2)
s s+N s + ω̂ 2g s 2 + ωr 2

Equation (2) offers a mathematical representation of the active rectifier controller,


linked to the schematic illustration above. The ω resonant pulse plays a vital role by
providing optimized gain precisely tuned to the resonant frequency. This crucial feature
allows the proportional-resonant (PR) controller to effectively reduce steady-state recurring
errors when following or canceling sinusoidal signals; this is consistent with the general
principles of the internal model. Functionally, the PR controller’s output is carefully de-
signed to counteract any disturbances encountered and meet precise DC voltage amplitude
requirements. It is inserted as an additive or subtracted component with respect to the
input AC voltage. A detailed analysis of the modulation block reveals a precise division
into two parts, each of which serves the modulation needs of the circuit. The first part deals
with pulse width modulation (PWM) methods, precisely regulating the branch composed
of GaN-type MOSFETs. On the other hand, the second part generates a “high” signal that
is precisely timed to coincide with the positive half-wave of the cosine signal, carefully
calculated in the previous blocks.

2.2. DC/DC Converter Model and Control


The dual active bridge (DAB) converter, illustrated in Figure 8, is designed with two
active bridges, one on the primary side and one on the secondary side of a high-frequency
transformer [64–73]. Its operation can be described in the following steps:
1 Primary side (DC-AC conversion): The first inverter converts the direct voltage into a
high-frequency alternating voltage. The duty cycle D1 regulates the primary voltage
V1sw . This voltage is applied to the transformer.
2 Power transfer: The high-frequency alternating voltage flows through the transformer,
generating an alternating current on the secondary side and transferring the power
through the inductance.
3 Secondary side (AC-DC conversion): The second inverter converts the high-frequency
alternating current into direct voltage. The duty cycle D2 regulates the secondary
voltage V2sw .
4 Regulation and control: Duty cycles and phase shifts are adjusted to control power
flow and output voltage. Control occurs through modulation, which determines the
duration and synchronization of the work cycles.

Figure 8. Dual active bridge converter reference circuit.

Phase shift modulation is a key control technique in DAB, regulating the power trans-
mission between the primary and secondary sides [74–81]. The process can be understood
through the following concepts:
Appl. Sci. 2024, 14, 1227 8 of 27

- Phase and work cycles: The phase ϕ represents the time shift between the voltages
V1sw and V2sw . The phase shift approach often requires that D1 = D2 = D = 0.5,
simplifying the control to just one parameter, ϕ.
- Power equation: The power equation (see Equation (3)) relates the phase shift ϕ to the
power transfer between the primary and secondary sides. Maximum power transfer
occurs when ϕ = π/2.  
nV1 V2 ϕ
P= 2π 2 f sw Ltot
ϕ 1− π
(3)
8P∗ f sw Ltot
 q 
ϕ∗ = π
2 1− 1− nV1 V2∗

- Advantages and limitations: Phase shift modulation is simple and requires only one
control parameter. However, it has limitations such as a narrow working range and
higher currents in the transformer.
The concept of the phase shifting modulation is illustrated in Figure 9. In summary,
DAB offers flexible bidirectional power control through the transformer, with phase-shift
modulation as the key mechanism for efficient control of power flow.

Figure 9. DAB switching voltage and inductance current in different operating conditions of the
PS modulation.
 h   i
Iv Ii
t∗ϕ = Pv + s Pi + s V2 re f − V2 − I2 T2π
sw

(4)
PW M2 = PW M1 (t − t∗ϕ )
As schematized in Figure 10, the objective of the feedback control system is to auto-
matically obtain the temporal shift between the modulation signal of the first stage PW M1
and the one that manages the second conversion stage PW M2 . In particular, the control
algorithm must manage the voltage request V2 re f , and manage the current feedback; this
process obtains the phase shift ϕ∗ between the primary and secondary voltage of the
transformer. In Equation (4), the control expressions are explicitly reported.
Appl. Sci. 2024, 14, 1227 9 of 27

Figure 10. Schematic illustration of the phase shifting control for the selected DC/DC converter.

2.3. Switching Device Thermal Modeling and Power Losses


The transition from the physical device of a power MOSFET to its equivalent circuit
model, based on the Foster model for thermal modeling, involves the representation of the
thermal layers that extend from the junction (J) to the surrounding environment (A) (as
schematized in Figure 11). Imagine the MOSFET as a system of stacked thermal layers. The
junction (J) is the internal region where heat is generated due to power losses. Surrounding
it is the case layer (C), which is the body of the MOSFET that wraps around the junction.
The case is connected to the heat sink (S), an external component that absorbs heat from the
case. Finally, the heat sink is in contact with the environment (A), which is the surrounding
temperature influencing the heat dissipation process. The Foster model simplifies this
thermal complexity by using thermal resistors to connect the layers. Thermal resistances,
such as RthJC , RthCS , and RthSA , act as thermal bridges between layers, allowing heat to flow
through the model, analogous to the flow of current in an electrical circuit. The conceptual
step then involves assigning thermal resistances to the different thermal layers, reflecting
the propagation of heat through the MOSFET during its operation. The Foster model
equations describe the change in temperature in each of the layers in response to heat flow.
The utilization of this equivalent circuit representation enables simulations to comprehend
the thermal behavior of the MOSFET under diverse operating conditions. In summary,
Foster’s model simplifies the intricate thermal behavior of the MOSFET, facilitating the
analysis and simulation of device heating in real-world scenarios.

Figure 11. From the internal layers of the devices to the Foster model.

3. Technical Specifications and Components: Sizing and Selection


In this section, the rationale employed for selecting the values of passive and active cir-
cuit components is presented, referencing the design specifications for an onboard charger
(OBC) device for electric vehicles. In this study, reference is made to the commercial solu-
tion presented by Infineon [82], and for completeness, the specific operational techniques
are detailed. The technical specifications of the active rectifier are outlined in Table 1,
while Table 2 provides the characteristics of the DC/DC converter. The first step is to size
the circuit components of the LCL filter input to the active rectifier. For such a filter, its
main objective involves attenuating high-frequency harmonics with respect to the mains
frequency; it aims to meet the specifications for current ripple at the input of the controlled
rectifier (the part consists of the active switching components). The maximum value of the
filter capacitance can be derived by Equation (5).
Appl. Sci. 2024, 14, 1227 10 of 27

Table 1. Active rectifier key specifications.

Parameters Specifications
Single Phase
Voltage: 90–265Vrms
Input AC frequency: f g = 50–60 Hz
Current: 32Arms @240 V
Power Factor: ≥0.99
PFC output: 400Vdc (typical)
Output Max output power: 7.4 kW@400Vdc
peak efficiency: ηmax = 0.985
PFC stage for HV Lion Battery OBC
switching frequency: f sw = 120 kHz
Performance isolation: reinforced
input AC sensing
PFC output voltage sensing
over-temperature protection
short-circuit-protection
Protection over-current protection
under-voltage protection
over-voltage protection

Table 2. Dual Active Bridge Key Specifications.

Parameters Specifications
Prim. voltage 400–450Vdc
Sec. voltage 250–450Vdc
Power rating forward 7.4 kW
Output current 20 A
Efficiency ηmax = 0.98
Switching frequency f sw ∈ [200; 800] kHz
resonance at f res = 500 kHz

The value must take into account the reactive power Qmax associated with a desired
power factor value PF, as well as take into account the value of the AC supply voltage Vg
(rms) and the supply frequency f g .

Qmax tan[ arc(cos(1 − PF ))] Pin ∼


C f ,max = = = 1.1 mF (5)
2π f g Vg 2 2π f g Vg 2

Ensuring that the resonant frequency of the LCL filter meets the constraint in
Equation (6), it must necessarily be much higher than the supply frequency, and much
lower than the switching frequency of the rectifier stage.

f g << f LCL << f sw


lower + fupper
lower = 10 f = 500 Hz −→ f
f LCL LCL ∼
f LCL g LCL = = 5 kHz (6)
2
upper
f LCL = f sw /10 = 10 kHz

From the circuit analysis, it is possible to derive the expression of the filter resonant
frequency as an explicit function of the circuit components, and derive the value of the
series inductance of the LCL filter, as in Equation (7).

1 1 ∼
f LCL = q −→ L f ,min = 2 = 5.8 µH (7)
2πL f C f 2πC f ,max f LCL
Appl. Sci. 2024, 14, 1227 11 of 27

The relationships reported above are obviously useful as guidelines for choosing the
values of the components, but on their own, they do not represent systematic and reliable
procedures. In fact, starting from the values obtained, a simulation analysis was carried out
to find the values that make the dynamic behavior of the filter connected to the AC/DC
converter as satisfactory as possible, in compliance with the specifications. This involved
calculating the PF and the ripple (in terms of THD) for each combination of L f and C f
values. The values chosen at the end of the evaluation were L f = 190 µH and C f = 525 µF.
Regarding the selection of active components using SiC/GaN technology, for each of the
three power conversion stages, practical considerations must be made that guarantee the
safety of the device itself, and they must be in line with the technical specifications. For
the choice of switching devices, the most relevant aspects concern the maximum voltage
between the drain and source Vds , the maximum current absorbed by the drain channel
Id , and the switching frequency f sw . Once commercial devices that respect these three
fundamental aspects are identified, it will be necessary to select the device that can be
driven with the gate density that can be provided via a switching system, possibly equipped
with a gate driver unit (GDU); this device should also dissipate the lowest possible power
per joule effect and have the lowest-declared resistance, Rds(on) .

Vds,max ≥ (1 + αv )Vdc,max
Pout,max (8)
Id,max ≥ (1 + αi )
Vg,min

Referring to the conditions in Equation (8), it is possible to obtain the “minimum”


values for the search for marketed devices, in terms of the maximum drain source voltage
and maximum drain current. Coefficients αv and αi represent robustness parameters, and
serve to make the selection of components “oversized”. Obviously, these parameters
cannot be too high to avoid excessive sizing compared to the design specifications of the
converter. The above choice refers to the devices for the PFC AC/DC (SiC) stage and the
DC/AC (GaN) module of the DAB converter. Assuming that the AC/AC conversion by
the transformer from galvanic isolation is as ideal as possible, for power conservation,
GaN devices could be selected for smaller voltages and higher currents (consistent with
the transformer gain), and to meet the input–output voltage specifications of the DAB
itself. For the PFC active rectifier and the initial DC/AC conversion phase of the DAB, the
minimum considered AC supply voltage is Vg,min = 110Vrms , while the maximum value
for the DC-bus voltage is Vdc,max . By incorporating safety coefficients of 10%, the minimum
thresholds for component selections are derived: Vds,max ≃ 900 V and Id,max ≃ 60 A. These
GaN devices are regarded as suitable for the second AC/DC conversion phase of the DAB
converter, although in this instance, the Vds,max constraint may be reduced.

4. Software-in-the-Loop Simulation Analysis


In the article, the utilization of a C-caller in Simulink for implementing code equivalent
to the previously explained controller models will be described. This approach enables the
translation of transfer functions into executable code, leveraging the discretization process
to obtain recursive functions. The C-caller in Simulink serves as a bridge connecting the
Simulink modeling environment and code written in the C language. This enables the
integration of more complex control models, derived from transfer functions, within the
Simulink environment, facilitating more advanced simulations. The process begins with
extracting transfer functions from control models. These functions are then translated into C
code, incorporating the discretization process to suit the needs of the simulation in Simulink.
Discretization is critical when adapting the continuum model to the characteristics of
Simulink discrete-time simulation. Once the equivalent C code is obtained, the C-caller in
Simulink is used to integrate this implementation into the overall model. This component
allows Simulink to call C code during simulations, allowing the implemented control
functions to be executed (see Figure 12).
Appl. Sci. 2024, 14, 1227 12 of 27

Figure 12. Schematic representation of the software-in-the-loop simulation environment.

The integration enables the evaluation of the system performance in greater detail,
taking into account the effects of the implemented controllers. Furthermore, employing
the C-caller allows flexibility to experiment with different control implementations di-
rectly within the Simulink environment, streamlining the analysis and optimization of the
control system.

4.1. AC/DC Converter Simulation Analysis


The response of the AC/DC converter to a voltage step of magnitude Vre f = 400 V
is shown in Figure 13. In particular, the DC bus’s transient voltage is compared with
various AC source circumstances. The following specifications are typical of single-phase
AC power supply standards: (i) Vg = 230Vrms & f g = 50 Hz (standard in the EU); and
(ii) Vg = 110Vrms & f g = 60 Hz (standard in the USA). Under both power supply scenarios,
it is evident that the control system can manage the demand with a comparatively small
steady-state mean error. As anticipated, the voltage transients are noticeably different,
nevertheless. Since the controller must make up for the AC/DC conversion from drastically
different absolute voltage levels, the overshoot is greater in the case of lower AC voltage.
The transient voltage simulation results are shown in Figures 14–17. All of these results
show comparable trends, and the intended DC-bus voltages are Vre f = 500 V, Vre f = 600 V,
Vre f = 700 V, and Vre f = 800 V, respectively. It is crucial to remember that the simulations
are run by first setting beginning conditions for the voltage VC (0) of the active rectifier’s
output capacitor, which is always set to be around 50 V lower than the given Vre f value.

Figure 13. Totem pole voltage control at Vre f = 400 V.


Appl. Sci. 2024, 14, 1227 13 of 27

Figure 14. Totem pole voltage control at Vre f = 500 V.

Figure 15. Voltage control of the totem pole with Vre f = 600 V.

Figure 16. Totem pole voltage control at Vre f = 700 V.


Appl. Sci. 2024, 14, 1227 14 of 27

Figure 17. Totem pole voltage control at Vre f = 800 V.

Furthermore, it can be observed that the difference between the AC voltage instances
that meet EU and USA requirements becomes less significant as the needed DC-bus voltage
rises. This pattern highlights the role that greater DC-bus voltage requirements play in
mitigating the effects of the differences in AC voltage between EU and USA standards. The
current IL (t) behavior under the EU-standard AC supply is depicted in Figure 18, which
overlays the scenarios with the various needed DC-bus voltages.
The current behavior is clearly influenced by the fluctuating Vre f it makes sense, based
on the form of the current, in that the total harmonic distortion (THD) varies noticeably at
lower DC-bus voltages. The alignment of IL with Vg , however, is still mostly steady despite
the fluctuation in Vre f . This illustrates the control mechanism’s resilience and accomplishes
the desired power factor adjustment goal. It is evident that the transient voltage behavior
also has an indirect impact on the current’s adjustment to align with the supply voltage.
Specifically, for Vre f = 400 V, the effective current is in line with Vg , and its effective value
only makes sense in relation to the needed power flow when the transient voltage becomes
closer to the end of the steady-state operation (see Figure 19).

Figure 18. Under Vre f , current control is applied in various circumstances, where Vgrid = 230Vrms &
50 Hz.
Appl. Sci. 2024, 14, 1227 15 of 27

Figure 19. Under Vre f , current control is applied in various circumstances, where Vgrid = 110Vrms &
60 Hz.

Figure 20 displays the outcomes of a comparable study, with the current profiles
corresponding to simulations run with the supply circumstances, Vg / f g = 110Vrms /60 Hz.
It is immediately noticed that the inductor current dynamics are far less sensitive to
alignment with Vg with this setup.

Figure 20. Robustness analysis of the active rectifier control with respect to the initial conditions of
the output capacitor.

The transient voltage response for the scenario when Vre f = 600 V is shown in
Figure 20, which also provides a robustness analysis of the variance in the beginning
voltage of the active rectifier’s output capacitor. While this change affects the transient
voltage directly, the behavior is essentially the same, especially in the steady-state operation.
The transient current behavior as it aligns with Vg , as seen in Figure 21, is more
sensitive to this fluctuation. The inductor current exhibits a high peak before aligning with
Vg due to the substantial voltage differential during the transient, which is ascribed to
VC (0). Since the DC/DC module would be linked to the battery during a bench experiment,
which naturally imposes a continuous voltage (depending on its state of charge), it is
also important to note that some starting circumstances are just simulated. Therefore,
the current peak seen in this circumstance might be caused by the voltage differential
Appl. Sci. 2024, 14, 1227 16 of 27

applied across the AC/DC converter’s input inductor. Such an analysis may be used to
validate design decisions or to develop the project’s hardware and software monitoring
and preventive logic.

Figure 21. Robustness analysis of the PFC with respect to the variation of the initial condition of the
output capacitor.

Figure 22 shows an additional potential robustness study, showing the impact of


changing the AC/DC converter’s output capacitor value. The difference in the voltage
ripple is undoubtedly the most noticeable consequence. Considering the capacitor’s char-
acteristic to store charge and, consequently, energy, the rationale holds. Consequently, a
larger capacitance can absorb a greater quantity of energy related to oscillations, leading to
a reduction in the voltage ripple. As no discernible impact of the supply voltage on current
alignment was identified, this aspect is not presented in the study.

Figure 22. Effect of the output capacitor value variation on transient and steady-state behaviors in
the active rectifier’s output voltage.
Appl. Sci. 2024, 14, 1227 17 of 27

4.2. Dual Active Bridge Simulation Analysis


As the input voltage V1 supplied by the AC/DC converter varies within the operational
range of interest V1 ∈ [200; 500], Figure 23 shows the analysis of the output voltage transient
of the DC/DC converter for a control voltage of Vre f = 250 V (battery voltage connected to
the second stage of the DAB conversion). Figures 24 and 25 also show a completely similar
analysis. The step response gravitates toward an (almost) linear first-order transient for
lower voltage levels and displays second-order behavior for higher V1 voltage levels under
all operating circumstances that have been investigated and demonstrated. The more DAB
output voltages used, the more obvious this discovery becomes. A thorough examination
of the robustness of the phase-shifting control method is provided in the previously stated
Figure 26. This study takes into account possible fluctuations in the values of circuit
components, with a particular emphasis on the variations in leakage inductance and output
capacitor. The analysis is given for the intermediate working state, with the goal of reaching
a steady battery voltage of 350 V, to guarantee clarity and comprehensiveness. It is clear
that adjustments to the values of both circuit components affect the transient behavior more
than the steady-state value, with the residual voltage ripple being partially impacted.

Figure 23. Impact of the Vin fluctuation on the DAB converter’s output voltage control, when
Vre f = 250 V.

Figure 24. Impact of the Vin fluctuation on the DAB converter’s output voltage control, when
Vre f = 350 V.

It was found that, in particular, a greater value for the inductance tends to enhance
the voltage step response’s overshoot, whereas a value that is too low makes the dynamic
response resemble a first-order linear system, which also affects the steady-state error. It is
evident that the output capacitance influences the maximum overshoot and the settling
time, but it appears to have little effect on the steady-state value (at least within the
examined range of variation). This allows the desired voltage step response to be achieved
with almost no steady-state error.
Appl. Sci. 2024, 14, 1227 18 of 27

Figure 25. Impact of the Vin fluctuation on the DAB converter’s output voltage control, when
Vre f = 450 V.

Figure 26. Variations in the output capacitor and the primary-side inductance of the transformer
have an impact on the DAB converter’s voltage regulation.

4.3. Switching Devices Thermal Simulation Analysis


The thermal behavior of the switching components of the previously examined
AC/DC converter is depicted in Figure 27, with reference to the “intermediate” work-
ing state, which calls for an AC supply of 230 V & 50 Hz and a DC bus of 650 V. One branch
of the active rectifier operates at a high switching frequency, while the other operates at
the grid frequency, as was previously mentioned in relation to converter operation. Since
the two branches are typically powered by the same current, their conduction losses are
comparable; in contrast, the high-frequency branch has much larger conduction losses. The
simulation results align perfectly with this observation.
Appl. Sci. 2024, 14, 1227 19 of 27

Figure 27. Thermal transient of the high-frequency and low-frequency switching devices in the
active rectifier.

Consequently, for the two branches operating at distinct frequencies, a stable state
value of TjHP, f inal ∼= 75 [°C] is identified for the high-frequency branch devices, while a
regime value of TjLP, f inal ∼ = 61 [°C] is observed for the grid-frequency devices within the
branch. Additionally, two transients with differing rise times are identified for each branch.
The thermal transient of the SiC-integrated devices is displayed in Figure 28 for the chosen
DAB converter. The switching losses (as well as those for switching) differ greatly because,
on average, the primary and secondary stage branches are not crossed by the same current
due to the galvanic isolation transformer. Specifically, the average current flowing through
the active rectifier’s components crosses the first conversion stage (DC/AC).
In fact, the operating temperature of the SiC primary components is comparable to
that of the rectifier’s high-frequency components, with a steady state average value of
Tjp , f inal ∼
= 82 [°C]. Since the resonance frequency is higher than that of high-frequency
components, the switching losses weigh much more, leading to overheating at higher
speeds. On the other hand, because the transformer is employed to decrease the effective
voltage value, a greater current passes through the components of the AC/DC secondary
circuit, making them susceptible to noticeably higher conduction losses.
The average junction temperature in secondary SiC devices has a steady-state value of
Tjs , f inal ∼
= 109 [°C]. It is noteworthy that the steady-state values are within the manufac-
turer’s stated operating limitations and are consistent with the design and size decisions
made during the early modeling stages.

Figure 28. Thermal transient of the primary and secondary side switching devices in the DC/DC
converter.
Appl. Sci. 2024, 14, 1227 20 of 27

4.4. Battery Charging Scenario


To fully validate the proposed OBC model, in this section, we show the simulation
results obtained by applying three different charging protocols of a Li-ion battery. The sim-
ulations carried out refer to the use of a model in the Simscape environment (MathWorks),
which, starting from high-level specifications, identifies the intrinsic characteristics of the
battery, taking into account the chemistry. The Li-ion battery models made available by the
simulation environment are created by integrating the following Equation (9):
(
Vb,0 − RIb − VQ,d + Vexp in discharge mode
Vbattery = − Bq
Vb,0 − RIb − VQ,c + Ae in discharge mode
KQ
VQ,d = (q + I ∗ t)
Q−q (9)
KQ KQ
VQ,c = I∗t − q
Q − 0.1q Q−q
dVexp 
= B| Ib | −Vexp + Asign( Ib )
dt
where Vb denotes the measured battery voltage charge/discharge) [V]; Vb,0 denotes the
battery’s constant voltage [V]; K denotes the polarization constant [V/Ah]; Q denotes the
Rt
battery capacity [Ah]; q = 0 Ib (τ )dτ the battery charge [A]; A denotes the exponential
zone amplitude [V]; I ∗ denotes the filtered current [A]; B denotes the exp. zone time
constant inverse; R denotes the internal resistance [Ω]; and Ib denotes the measured battery
current [A]. A realistic characteristic inherent to the discharge behavior of the battery for
the different currents supplied is automatically generated, and the nominal characteristic is
explained by highlighting the nominal and exponential zones (see Figure 29). Regarding
the proposed simulation results, the data relating to the simulated battery are presented in
Table 3. Figure 30 shows the different behaviors of the battery in terms of the voltage and
charging current supplied by the electronic converters, in three different modes.

Figure 29. Discharging characteristics of Li-ion battery generated from Simscape model.
Appl. Sci. 2024, 14, 1227 21 of 27

Table 3. Battery parameters of Simscape model.

Characteristic Value
Maximum capacity 150 [Ah]
Cut-off voltage 187.5 [V]
Fully charged voltage 291 [V]
Nominal discharge current 65.2 [A]
Internal resistance 16.67 [mΩ]
Capacity at nominal voltage 135.65 [Ah]
Exp. zone voltage 270 [V]
Exp. zone capacity 7.37 [Ah]

Figure 30. Behavior of the battery during the charging phase, in three different protocols.

In particular, the results relate to the following charging protocols:


• Constant current protocol (CCP): In this case, the nominal voltage of the battery is
taken as reference and the current value is imposed to obtain the desired output power.

Pout
Ire f ,CCP = (10)
Vb,nominal
Appl. Sci. 2024, 14, 1227 22 of 27

• Adaptive current protocol (ACP): In this case, the current is calculated based on the
instantaneous battery voltage to keep the power flow between the OBC and the battery
constant.
P∗
Ire f ,ACP (t) = out (11)
Vb (t)
• Pulse current protocol (PCP): In this case, a current greater than that necessary to
obtain a constant desired power flow is used, providing the battery with current
pulses followed by zero-current phases.

Pout
Ire f ,PCP = 2 PW M (0.1 Hz, 50%d.c.) (12)
Vb,min

As expected from engineering practices, the PCP protocol provides the fastest charging,
compared to the other two cases, for the same power supplied to the battery for charging
(see Figure 31).

Figure 31. SOC behavior of the battery during the charging phase, in the three different protocols.

This is due to the intrinsic behavior of the battery, which is more reactive in the high
current phase than it is when it discharges in the zero current phase (in fact, batteries
are systems with memory). Obviously, the aim of this work is not to propose an efficient
charging protocol, but to show that the design and control systems obtained in simulation
for the two converters individually can provide realistic information on the behavior of the
Li-ion battery, with typical specifications of automotive applications.

5. Final Discussion and Conclusions


In the pursuit of a scientific article, simulations were conducted to evaluate the perfor-
mance of control algorithms and the sizing of passive components within the framework of
an onboard charger (OBC) power electronic system. The investigation comprised separate
simulations for the AC/DC conversion stage, implemented via a single-phase PFC circuit.
Notably, a detailed analysis of the voltage step response for various DC-link reference
values was undertaken, yielding noteworthy outcomes. For Vre f = 400 V, the steady-state
error was 0, with settling times ranging between 1.36 and 1.65 s, contingent on the AC
voltage being 230 V/50 Hz or 110 V/60 Hz. Similarly, for Vre f = 550 V, a steady-state
error of 0 was achieved, with nearly identical settling times for both AC input configura-
tions (approximately 1.2 s). For Vre f = 600 V, Vre f = 700 V, and Vre f = 800 V, behaviors
in both AC input configurations were practically overlapping, exhibiting a steady-state
error trending toward 0 and a voltage ripple consistently below 2%. The evaluation of the
AC/DC converter included an assessment of the power factor (PF), reaching approximately
97% after the voltage transient for 230 V/50 Hz AC input and even reaching 98% for
Appl. Sci. 2024, 14, 1227 23 of 27

110 V/60 Hz. This performance aligns with market supplier declarations and surpasses the
declared design in some industrial products. Furthermore, in the 230 V/50 Hz configura-
tion, current harmonic distortion was lower compared to the 110 V/60 Hz configuration,
never exceeding 5%. Robustness analysis of the feedback system concerning variations
in circuit parameters within a 20% range from the nominal case demonstrated acceptable
performance degradation. This involved maintaining a voltage ripple consistently below
2.5%, a current ripple below 10%, a steady-state error below 2%, a PF consistently above
90%, and total harmonic distortion (THD) consistently below 6.5%. A parallel analysis was
performed on the DC/DC conversion stage, employing a DAB converter. The output volt-
age response, evaluated at a constant output power of 7 KW, yielded distinct performances
for various Vre f values. For Vre f = 250 V, there was a maximum overshoot of 15%, with
zero steady-state error and a maximum settling time of 3.2 ms. Similarly, for Vre f = 350 V, a
maximum overshoot of 15% was observed, along with zero steady-state error and a settling
time below 6 ms. For Vre f = 450 V, the overshoot was below 10%, the steady-state error
was below 1%, and the settling time was below 52.5 ms. These performances were assessed
while varying the input voltage at the DC/DC primary. Additionally, feedback robustness
was evaluated concerning variations in passive parameters within a 50% uncertainty range
from the nominal value, revealing slight performance degradation limited to a worst-case
25% overshoot. Thermal performances were also scrutinized to ensure that the junction
temperature of switching components remained within the operational limits declared
in the supplier data sheets. This comprehensive investigation into the intricacies of OBC
system design and control, encompassing both AC/DC and DC/DC conversion stages,
sets the stage for future advancements. The study highlights the adaptability and resilience
of control algorithms, emphasizing the critical importance of precision in component sizing
and algorithmic design. The findings pave the way for optimizing OBC efficiency and
reliability through advanced control strategies and thermal management. Future devel-
opments could explore the implementation of feedback from thermal models, leveraging
adaptive or predictive controllers for enhanced precision. However, careful consideration of
computational challenges, especially in resource-limited embedded systems, is imperative
for practical implementation in automotive environments.
The balance between control precision and computational practicality must be carefully
evaluated in ongoing efforts to enhance OBC performance and contribute to the continued
evolution of electric vehicle technology.
In conclusion, our comprehensive investigation into the intricacies of the onboard
charger (OBC) system design and control, covering both AC/DC and DC/DC conversion
stages, aligns with the contemporary emphasis on nonlinear circuits and systems. The
nonlinear nature of power systems is becoming increasingly crucial in the age of complexity,
with a growing focus on advanced power plant technologies. Our study underscores the
adaptability and resilience of control algorithms, emphasizing the critical importance of
precision in component sizing and algorithmic design within the context of nonlinear
systems. The findings not only contribute to the current state of knowledge in power
electronics but also align with the broader trends in nonlinear technology highlighted in
the referenced editorial. As we navigate the nonlinear landscape of power systems, our
results showcase the robustness of the OBC system under various conditions, demonstrat-
ing acceptable performance degradation within a specified parameter range. This aligns
with the challenges posed by nonlinear dynamics in power systems, as discussed in the
editorial [83]. Moreover, our study paves the way for future advancements by highlight-
ing the potential for optimizing OBC efficiency and reliability through advanced control
strategies and thermal management. The editorial emphasizes the multidisciplinary nature
of research in nonlinear technology, mirroring our approach that considers both AC/DC
and DC/DC conversion stages. Looking ahead, future developments could explore the
implementation of feedback from thermal models, leveraging adaptive or predictive con-
trollers for enhanced precision. However, as the editorial suggests, careful consideration of
computational challenges, especially in resource-limited embedded systems, is imperative
Appl. Sci. 2024, 14, 1227 24 of 27

for practical implementation in automotive environments. The delicate balance between


control precision and computational practicality must be thoroughly evaluated in ongoing
efforts to enhance OBC performance and contribute to the continued evolution of electric
vehicle technology. In summary, our study not only adds valuable insights to the field of
OBC design and control but also aligns with the overarching themes of nonlinear tech-
nology discussed in the referenced editorial. By recognizing the interconnected nature of
nonlinear components and systems, our research contributes to the ongoing dialogue on
advanced power plant technologies, paving the way for future interdisciplinary research in
the realm of nonlinear power systems.

Author Contributions: Conceptualization, P.D. and S.S.; Methodology, P.D. and S.S.; Software, P.D.;
Validation, P.D.; Formal analysis, P.D.; Investigation, P.D. and S.S.; Resources, P.D. and S.S.; Data
curation, P.D. and S.S.; Writing—original draft, P.D.; Writing—review & editing, S.S.; Supervision,
S.S.; Project administration, S.S.; Funding acquisition, S.S. All authors have read and agreed to the
published version of the manuscript.
Funding: The work has been partially supported by Centro Nazionale di Ricerca in High-Performance
Computing Big Data and Quantum Computing SPOKE 6 Multiscale modelling & Engineering
applications; by the ECSEL JU project Hiefficient n. 101007281 (EU ECSEL-2020-2-RIA call); and by
MIUR FoReLab Project Dipartimenti di Eccellenza.
Data Availability Statement: The data that support the findings of this study are available from the
corresponding author, P.D., upon reasonable request.
Conflicts of Interest: The authors declare no conflicts of interest.

References
1. Mo, T.; Li, Y.; Lau, K.t.; Poon, C.K.; Wu, Y.; Luo, Y. Trends and emerging technologies for the development of electric vehicles.
Energies 2022, 15, 6271. [CrossRef]
2. Li, J.; Wang, F.; He, Y. Electric vehicle routing problem with battery swapping considering energy consumption and carbon
emissions. Sustainability 2020, 12, 10537. [CrossRef]
3. Neshat, M.; Sergiienko, N.Y.; Mirjalili, S.; Majidi Nezhad, M.; Piras, G.; Astiaso Garcia, D. Multi-mode wave energy converter
design optimisation using an improved moth flame optimisation algorithm. Energies 2021, 14, 3737. [CrossRef]
4. Tran, D.D.; Chakraborty, S.; Lan, Y.; Baghdadi, M.E.; Hegazy, O. NSGA-II-based codesign optimization for power conversion and
controller stages of interleaved boost converters in electric vehicle drivetrains. Energies 2020, 13, 5167. [CrossRef]
5. Sikorski, W.; Wielewski, A. Low-Cost Online Partial Discharge Monitoring System for Power Transformers. Sensors 2023, 23, 3405.
[CrossRef]
6. Zhou, L.; Eull, M.; Preindl, M. Optimization-Based Estimation and Model Predictive Control for High Performance, Low Cost
Software-Defined Power Electronics. IEEE Trans. Power Electron. 2023, 38, 1022–1035. [CrossRef]
7. Safayatullah, M.; Elrais, M.T.; Ghosh, S.; Rezaii, R.; Batarseh, I. A Comprehensive Review of Power Converter Topologies and
Control Methods for Electric Vehicle Fast Charging Applications. IEEE Access 2022, 10, 40753–40793. [CrossRef]
8. Ramezanzadeh, S.; Ozbulut, M.; Yildiz, M. A numerical investigation of the energy efficiency enhancement of oscillating water
column wave energy converter systems. Energies 2022, 15, 8276. [CrossRef]
9. Ghimire, P.; Park, D.; Zadeh, M.K.; Thorstensen, J.; Pedersen, E. Shipboard Electric Power Conversion: System Architecture,
Applications, Control, and Challenges [Technology Leaders]. IEEE Electrif. Mag. 2019, 7, 6–20. [CrossRef]
10. Liu, S.; Shen, Y.; Huang, J.; Li, Y.; Zhao, Y.; Tuo, H. Design Method of Output Intrinsic Safety Boost Converter Based on Minimum
Frequency and Considering Temperature Effects. IEEE J. Emerg. Sel. Top. Power Electron. 2023, 11, 4233–4244. [CrossRef]
11. Kotb, R.; Chakraborty, S.; Tran, D.D.; Abramushkina, E.; El Baghdadi, M.; Hegazy, O. Power Electronics Converters for Electric
Vehicle Auxiliaries: State of the Art and Future Trends. Energies 2023, 16, 1753. [CrossRef]
12. Liu, M.; Cao, X.; Cao, C.; Wang, P.; Wang, C.; Pei, J.; Lei, H.; Jiang, X.; Li, R.; Li, J. A Review of Power Conversion Systems and
Design Schemes of High-Capacity Battery Energy Storage Systems. IEEE Access 2022, 10, 52030–52042. [CrossRef]
13. Savio Abraham, D.; Verma, R.; Kanagaraj, L.; Giri Thulasi Raman, S.R.; Rajamanickam, N.; Chokkalingam, B.; Marimuthu Sekar,
K.; Mihet-Popa, L. Electric vehicles charging stations’ architectures, criteria, power converters, and control strategies in microgrids.
Electronics 2021, 10, 1895. [CrossRef]
14. Ozdemir, S.; Altin, N.; Nasiri, A.; Cuzner, R. Review of Standards on Insulation Coordination for Medium Voltage Power
Converters. IEEE Open J. Power Electron. 2021, 2, 236–249. [CrossRef]
15. Cosimi, F.; Dini, P.; Giannetti, S.; Petrelli, M.; Saponara, S. Analysis and design of a non-linear MPC algorithm for vehicle
trajectory tracking and obstacle avoidance. In Applications in Electronics Pervading Industry, Environment and Society: APPLEPIES
2020 8 ; Springer: Cham, Switzerland, 2021; pp. 229–234.
Appl. Sci. 2024, 14, 1227 25 of 27

16. Wang, D.; Shen, Z.J.; Yin, X.; Tang, S.; Liu, X.; Zhang, C.; Wang, J.; Rodriguez, J.; Norambuena, M. Model Predictive Control
Using Artificial Neural Network for Power Converters. IEEE Trans. Ind. Electron. 2022, 69, 3689–3699. [CrossRef]
17. Dini, P.; Begni, A.; Ciavarella, S.; De Paoli, E.; Fiorelli, G.; Silvestro, C.; Saponara, S. Design and Testing Novel One-Class Classifier
Based on Polynomial Interpolation With Application to Networking Security. IEEE Access 2022, 10, 67910–67924. [CrossRef]
18. Xiao, Z.; Jiang, Y.; Sun, T.; Wu, Y.; Tang, Y. Refining Power Converter Loss Evaluation: A Transfer Learning Approach. IEEE Trans.
Power Electron. 2024, 1–12. [CrossRef]
19. Bernardeschi, C.; Dini, P.; Domenici, A.; Saponara, S. Co-simulation and Verification of a Non-linear Control System for Cogging
Torque Reduction in Brushless Motors. In Proceedings of the Software Engineering and Formal Methods: SEFM 2019 Collocated
Workshops: CoSim-CPS, ASYDE, CIFMA, and FOCLASA, Oslo, Norway, 16–20 September 2019; Revised Selected Papers 17;
Springer: Cham, Switzerland, 2020; pp. 3–19.
20. Dini, P.; Saponara, S. Review on model based design of advanced control algorithms for cogging torque reduction in power drive
systems. Energies 2022, 15, 8990. [CrossRef]
21. Dini, P.; Saponara, S.; Colicelli, A. Overview on Battery Charging Systems for Electric Vehicles. Electronics 2023, 12, 4295.
[CrossRef]
22. Dini, P.; Ariaudo, G.; Botto, G.; Greca, F.L.; Saponara, S. Real-time electro-thermal modelling & predictive control design of
resonant power converter in full electric vehicle applications. IET Power Electron. 2023, 16, 2045–2064.
23. Pacini, F.; Matteo, S.D.; Dini, P.; Fanucci, L.; Bucchi, F. Innovative Plug-and-Play System for Electrification of Wheel-Chairs. IEEE
Access 2023, 11, 89038–89051. [CrossRef]
24. Bernardeschi, C.; Dini, P.; Domenici, A.; Mouhagir, A.; Palmieri, M.; Saponara, S.; Sassolas, T.; Zaourar, L. Co-simulation of
a model predictive control system for automotive applications. In Proceedings of the International Conference on Software
Engineering and Formal Methods, Virtual Event, 6–10 December 2021; Springer: Cham, Switzerland, 2021; pp. 204–220.
25. Benedetti, D.; Agnelli, J.; Gagliardi, A.; Dini, P.; Saponara, S. Design of a Digital Dashboard on Low-Cost Embedded Platform in a
Fully Electric Vehicle. In Proceedings of the 2020 IEEE International Conference on Environment and Electrical Engineering and
2020 IEEE Industrial and Commercial Power Systems Europe (EEEIC / ICPS Europe), Madrid, Spain, 9–12 June 2020; pp. 1–5.
[CrossRef]
26. Dini, P.; Saponara, S. Processor-in-the-Loop Validation of a Gradient Descent-Based Model Predictive Control for Assisted
Driving and Obstacles Avoidance Applications. IEEE Access 2022, 10, 67958–67975. [CrossRef]
27. Dini, P.; Saponara, S.; Chakraborty, S.; Hosseinabadi, F.; Hegazy, O. Experimental Characterization and Electro-Thermal Modeling
of Double Side Cooled SiC MOSFETs for Accurate and Rapid Power Converter Simulations. IEEE Access 2023, 11, 79120–79143.
[CrossRef]
28. Zghaibeh, M.; Belgacem, I.B.; Baloch, M.H.; Chauhdary, S.T.; Kumar, L.; Arıcı, M. Optimization of green hydrogen production in
hydroelectric-photovoltaic grid connected power station. Int. J. Hydrogen Energy 2024, 52, 440–453. [CrossRef]
29. Gaspar, J.F.; Pinheiro, R.F.; Mendes, M.J.; Kamarlouei, M.; Soares, C.G. Review on hardware-in-the-loop simulation of wave
energy converters and power take-offs. Renew. Sustain. Energy Rev. 2024, 191, 114144. [CrossRef]
30. Zheng, J.; Zeng, Y.; Zhao, Z.; Liu, W.; Xu, H.; Wang, H.; Mou, D. MPSoC-Based Dynamic Adjustable Time-Stepping Scheme with
Switch Event Oversampling Technique for Real-time HIL Simulation of Power Converters. IEEE Trans. Transp. Electrif. 2023.
[CrossRef]
31. Hassan, M.A.; Li, E.p.; Li, X.; Li, T.; Duan, C.; Chi, S. Adaptive Passivity-Based Control of dc–dc Buck Power Converter With
Constant Power Load in DC Microgrid Systems. IEEE J. Emerg. Sel. Top. Power Electron. 2019, 7, 2029–2040. [CrossRef]
32. Mahdizadeh, S.; Gholizadeh, H.; Gorji, S.A. A Power Converter Based on the Combination of Cuk and Positive Output Super Lift
Lou Converters: Circuit Analysis, Simulation and Experimental Validation. IEEE Access 2022, 10, 52899–52911. [CrossRef]
33. Shaker, D.H.; Shneen, S.W.; Abdullah, F.N.; Aziz, G.A. Simulation Model of Single-Phase AC-AC Converter by Using MATLAB.
J. Robot. Control (JRC) 2022, 3, 656–665. [CrossRef]
34. Razzhivin, I.; Askarov, A.; Rudnik, V.; Suvorov, A. A hybrid simulation of converter-interfaced generation as the part of a
large-scale power system model. Int. J. Eng. Technol. Innov. 2021, 11, 278. [CrossRef]
35. Fusheng, Z.; Naayagi, R. Power Converters for DC Microgrids—Modelling and Simulation. In Proceedings of the 2018 IEEE
Innovative Smart Grid Technologies—Asia (ISGT Asia), Singapore, 22–25 May 2018; pp. 994–999. [CrossRef]
36. Şuşcă, M.; Mihaly, V.; Stănese, M.; Morar, D.; Dobra, P. Unified cacsd toolbox for hybrid simulation and robust controller synthesis
with applications in dc-to-dc power converter control. Mathematics 2021, 9, 731. [CrossRef]
37. Bai, H.; Liu, C.; Breaz, E.; Al-Haddad, K.; Gao, F. A Review on the Device-Level Real-Time Simulation of Power Electronic
Converters: Motivations for Improving Performance. IEEE Ind. Electron. Mag. 2021, 15, 12–27. [CrossRef]
38. Milton, M.; O, C.D.L.; Ginn, H.L.; Benigni, A. Controller-Embeddable Probabilistic Real-Time Digital Twins for Power Electronic
Converter Diagnostics. IEEE Trans. Power Electron. 2020, 35, 9850–9864. [CrossRef]
39. Imaoka, J.; Okamoto, K.; Shoyama, M.; Ishikura, Y.; Noah, M.; Yamamoto, M. Modeling, Magnetic Design, Simulation Methods,
and Experimental Evaluation of Various Powder Cores Used in Power Converters Considering Their DC Superimposition
Characteristics. IEEE Trans. Power Electron. 2019, 34, 9033–9051. [CrossRef]
40. Estrada, L.; Vázquez, N.; Vaquero, J.; de Castro, Á.; Arau, J. Real-time hardware in the loop simulation methodology for power
converters using labview FPGA. Energies 2020, 13, 373. [CrossRef]
Appl. Sci. 2024, 14, 1227 26 of 27

41. Pico, H.N.V.; Johnson, B.B. Transient Stability Assessment of Multi-Machine Multi-Converter Power Systems. IEEE Trans. Power
Syst. 2019, 34, 3504–3514. [CrossRef]
42. Bai, H.; Luo, H.; Liu, C.; Paire, D.; Gao, F. A Device-Level Transient Modeling Approach for the FPGA-Based Real-Time Simulation
of Power Converters. IEEE Trans. Power Electron. 2020, 35, 1282–1292. [CrossRef]
43. Rosso, R.; Andresen, M.; Engelken, S.; Liserre, M. Analysis of the Interaction Among Power Converters Through Their
Synchronization Mechanism. IEEE Trans. Power Electron. 2019, 34, 12321–12332. [CrossRef]
44. Xu, R.; Fang, W.; Liu, X.d.; Liu, Y.; Hu, Y.; Liu, Y.f. Design and experimental verification of on-board charger for electric vehicle.
In Proceedings of the 2014 International Power Electronics and Application Conference and Exposition, Shanghai, China, 5–8
November 2014; IEEE: Piscataway, NJ, USA, 2014; pp. 1422–1427.
45. Cao, L.; Li, H.; Zhang, H. Model-free power control of front-end PFC AC/DC converter for on-board charger. In Proceedings of
the 2016 IEEE 8th International Power Electronics and Motion Control Conference (IPEMC-ECCE Asia), Hefei, China, 22–26 May
2016; IEEE: Piscataway, NJ, USA, 2016; pp. 2719–2723.
46. Noh, J.H.; Song, S.i.; Hur, D.J. Numerical Analysis of the Cooling Performance in a 7.2 kW Integrated Bidirectional OBC/LDC
Module. Appl. Sci. 2020, 10, 270. [CrossRef]
47. Ramesh, P.; Patra, A.; Kastha, D. Design and Simulation of an On-Board Integrated Charger using Cell Bypass Balancing Circuit
for Electric Vehicles. In Proceedings of the IECON 2018-44th Annual Conference of the IEEE Industrial Electronics Society,
Washington, DC, USA, 21–23 October 2018; IEEE: Piscataway, NJ, USA, 2018; pp. 2032–2037.
48. Zhu, X.; Kong, L.; Yang, X.; Xu, Y. Design of Vehicle Charger for Pure Electric Vehicle Based on MATLAB Simulation. J. Phys.
Conf. Ser. 2020, 1635, 012020. [CrossRef]
49. Nguyen, H.V.; Lee, D.C. Advanced single-phase onboard chargers with small DC-link capacitors. In Proceedings of the 2018
IEEE International Power Electronics and Application Conference and Exposition (PEAC), Shenzhen, China, 4–7 November 2018;
IEEE: Piscataway, NJ, USA, 2018; pp. 1–6.
50. Tao, H.; Zhang, G.; Zheng, Z. Onboard charging DC/DC converter of electric vehicle based on synchronous rectification and
characteristic analysis. J. Adv. Transp. 2019, 2019, 2613893. [CrossRef]
51. Kim, I.; Lee, S.; Park, J.W. Design and Control of OBC-LDC Integrated Circuit with Variable Turns Ratio for Electric Vehicles. In
Proceedings of the 2020 IEEE Energy Conversion Congress and Exposition (ECCE), Detroit, MI, USA, 11–15 October 2020; IEEE:
Piscataway, NJ, USA, 2020; pp. 885–890.
52. Gaurav, A.; Gaur, A. Modelling of Hybrid Electric Vehicle Charger and Study the Simulation Results. In Proceedings of the 2020
International Conference on Emerging Frontiers in Electrical and Electronic Technologies (ICEFEET), Patna, India, 10–11 July
2020; IEEE: Piscataway, NJ, USA, 2020, pp. 1–6.
53. Yuan, J.; Poorfakhraei, A.; Emadi, A. A Novel Phase Shift Control for Single-Stage Bidirectional Isolated Totem-Pole AC/DC
Onboard Electric Vehicle Chargers. In Proceedings of the 2023 IEEE Transportation Electrification Conference & Expo (ITEC),
Detroit, MI, USA, 21–23 June 2023; pp. 1–6. [CrossRef]
54. Ma, H.; Pan, Y.; Lu, Y.; Chen, X.; Huang, Y. Research on Single-Phase Three-Level Pseudo Totem-Pole Rectifiers With a Unified
Pulsewidth Modulation. IEEE J. Emerg. Sel. Top. Power Electron. 2023, 11, 5052–5061. [CrossRef]
55. Kumar, V.; Yi, K. Single-Phase, Bidirectional, 7.7 kW Totem Pole On-Board Charging/Discharging Infrastructure. Appl. Sci. 2022,
12, 2236. [CrossRef]
56. Zhao, T.; Burgos, R.; Wen, B.; McLean, A.; Mattos, R.F. Design of Three-Level Flying Capacitor Totem Pole PFC in USB Type-C
Power Delivery for Aircraft Applications. In Proceedings of the 2022 IEEE 9th Workshop on Wide Bandgap Power Devices &
Applications (WiPDA), Redondo Beach, CA, USA, 7–9 November 2022; pp. 254–258. [CrossRef]
57. Le, T.T.; Hakim, R.M.; Park, J.; Choi, S. A Single-stage Four-Phase Totem-Pole AC-DC Converter with Wide Voltage Range and
Compact Integrated Magnetic Component. In Proceedings of the 2021 IEEE Energy Conversion Congress and Exposition (ECCE),
Vancouver, BC, Canada, 10–14 October 2021; pp. 2208–2212. [CrossRef]
58. Wei, Y.; Luo, Q.; Alonso, J.M.; Mantooth, A. A Magnetically Controlled Single-Stage AC–DC Converter. IEEE Trans. Power
Electron. 2020, 35, 8872–8877. [CrossRef]
59. Itoh, K.; Ishigaki, M.; Kikuchi, N.; Harada, T.; Sugiyama, T. A Single-Stage Rectifier with Interleaved Totem-pole PFC and Dual
Active Bridge (DAB) Converter for PHEV/BEV On-board Charger. In Proceedings of the 2020 IEEE Applied Power Electronics
Conference and Exposition (APEC), New Orleans, LA, USA, 15–19 March 2020; pp. 1936–1941. [CrossRef]
60. Belkamel, H.; Kim, H.; Choi, S. Interleaved Totem-Pole ZVS Converter Operating in CCM for Single-Stage Bidirectional AC–DC
Conversion with High-Frequency Isolation. IEEE Trans. Power Electron. 2021, 36, 3486–3495. [CrossRef]
61. Yu, Z.; Xia, Y.; Ayyanar, R. A Simple ZVT Auxiliary Circuit for Totem-Pole Bridgeless PFC Rectifier. IEEE Trans. Ind. Appl. 2019,
55, 2868–2878. [CrossRef]
62. Gong, X.; Wang, G.; Bhardwaj, M. 6.6 kW Three-Phase Interleaved Totem Pole PFC Design with 98.9 Efficiency for HEV/EV
Onboard Charger. In Proceedings of the 2019 IEEE Applied Power Electronics Conference and Exposition (APEC), Anaheim, CA,
USA, 17–21 March 2019; pp. 2029–2034. [CrossRef]
63. He, Q.; Luo, Q.; Ma, K.; Sun, P.; Zhou, L. Analysis and Design of a Single-Stage Bridgeless High-Frequency Resonant AC/AC
Converter. IEEE Trans. Power Electron. 2019, 34, 700–711. [CrossRef]
64. Alemanno, A.; Morici, R.; Pretelli, M.; Florian, C. Design of a 7.5 kW Dual Active Bridge Converter in 650 V GaN Technology for
Charging Applications. Electronics 2023, 12, 1280. [CrossRef]
Appl. Sci. 2024, 14, 1227 27 of 27

65. Fernandez-Hernandez, A.; Gonzalez-Hernando, F.; Garcia-Bediaga, A.; Villar, I.; Abad, G. Design Space Analysis of the
Dual-Active-Bridge Converter for More Electric Aircraft. Energies 2022, 15, 9503. [CrossRef]
66. Xu, F.; Liu, J.; Dong, Z. Minimum Backflow Power and ZVS Design for Dual-Active-Bridge DC–DC Converters. IEEE Trans. Ind.
Electron. 2023, 70, 474–484. [CrossRef]
67. Koszel, M.; Grzejszczak, P.; Nowatkiewicz, B.; Wolski, K.; Szymczak, M.; Czaplicki, A. Design of dual active bridge for DC
microgrid application. In Proceedings of the 2022 Progress in Applied Electrical Engineering (PAEE), Koscielisko, Poland, 27
June–1 July 2022; pp. 1–6. [CrossRef]
68. Saha, J.; Gorla, N.B.Y.; Subramaniam, A.; Panda, S.K. Analysis of Modulation and Optimal Design Methodology for Half-Bridge
Matrix-Based Dual-Active-Bridge (MB-DAB) AC–DC Converter. IEEE J. Emerg. Sel. Top. Power Electron. 2022, 10, 881–894.
[CrossRef]
69. Henao-Bravo, E.E.; Ramos-Paja, C.A.; Saavedra-Montes, A.J.; González-Montoya, D.; Sierra-Perez, J. Design method of dual
active bridge converters for photovoltaic systems with high voltage gain. Energies 2020, 13, 1711. [CrossRef]
70. Das, D.; Mishra, S.; Singh, B. Design Architecture for Continuous-Time Control of Dual Active Bridge Converter. IEEE J. Emerg.
Sel. Top. Power Electron. 2021, 9, 3287–3295. [CrossRef]
71. Jeung, Y.C.; Lee, D.C. Voltage and Current Regulations of Bidirectional Isolated Dual-Active-Bridge DC–DC Converters Based on
a Double-Integral Sliding Mode Control. IEEE Trans. Power Electron. 2019, 34, 6937–6946. [CrossRef]
72. Das, D.; Basu, K. Optimal Design of a Dual-Active-Bridge DC–DC Converter. IEEE Trans. Ind. Electron. 2021, 68, 12034–12045.
[CrossRef]
73. Liu, T.; Yang, X.; Chen, W.; Li, Y.; Xuan, Y.; Huang, L.; Hao, X. Design and Implementation of High Efficiency Control Scheme
of Dual Active Bridge Based 10 kV/1 MW Solid State Transformer for PV Application. IEEE Trans. Power Electron. 2019,
34, 4223–4238. [CrossRef]
74. Zhang, W.; Xu, Y.; Ren, J.; Su, J.; Zou, J. Synchronous random switching frequency modulation technique based on the carrier
phase shift to reduce the PWM noise. IET Power Electron. 2020, 13, 892–897. [CrossRef]
75. Li, Y.; Tian, H.; Li, Y.W. Generalized Phase-Shift PWM for Active-Neutral-Point-Clamped Multilevel Converter. IEEE Trans. Ind.
Electron. 2020, 67, 9048–9058. [CrossRef]
76. Rahman, S.; Meraj, M.; Iqbal, A.; Reddy, B.P.; Khan, I. A Combinational Level-Shifted and Phase-Shifted PWM Technique for
Symmetrical Power Distribution in CHB Inverters. IEEE J. Emerg. Sel. Top. Power Electron. 2023, 11, 932–941. [CrossRef]
77. Moradisizkoohi, H.; Elsayad, N.; Shojaie, M.; Mohammed, O.A. PWM Plus Phase-Shift-Modulated Three-Port Three-Level
Soft-Switching Converter Using GaN Switches for Photovoltaic Applications. IEEE J. Emerg. Sel. Top. Power Electron. 2019,
7, 636–652. [CrossRef]
78. Arazm, S.; Vahedi, H.; Al-Haddad, K. Generalized Phase-Shift Pulse Width Modulation for Multi-Level Converters. In
Proceedings of the 2018 IEEE Electrical Power and Energy Conference (EPEC), Toronto, ON, Canada, 10–11 October 2018; pp. 1–6.
[CrossRef]
79. Li, S.; Liang, S.; Li, Z.; Zheng, S. A Phase-Shift-Modulated Resonant Two-Switch Boosting Switched-Capacitor Converter and Its
Modulation Map. IEEE Trans. Ind. Electron. 2023, 70, 7783–7795. [CrossRef]
80. Ghosh, S.; Singh, B. Effect of Phase Shift Definition on Power Characterization of Isolated DC-DC Converter Under PWM Plus
Phase Shift Modulation. IEEE Trans. Circuits Syst. II Express Briefs 2023, 70, 4454–4458. [CrossRef]
81. Calderon, C.; Barrado, A.; Rodriguez, A.; Alou, P.; Lazaro, A.; Fernandez, C.; Zumel, P. General analysis of switching modes in a
dual active bridge with triple phase shift modulation. Energies 2018, 11, 2419. [CrossRef]
82. Instruments, T. 7.4-kW EV/HEV Bidirectional On-Board Charger Reference Design with GaN—Application Note. 2022 Available
online: https://www.ti.com/lit/ug/tiduf18/tiduf18.pdf?ts=1695714670313&ref_url=https%253A%252F%252Fwww.google.it%
252F (accessed on 5 November 2023).
83. Fortuna, L.; Buscarino, A. Nonlinear Technologies in Advanced Power Systems: Analysis and Control. Energies 2022, 15, 5167.
[CrossRef]

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.

You might also like