Next Article in Journal
Interval Constrained Multi-Objective Optimization Scheduling Method for Island-Integrated Energy Systems Based on Meta-Learning and Enhanced Proximal Policy Optimization
Previous Article in Journal
Improved Plasma Etch Endpoint Detection Using Attention-Based Long Short-Term Memory Machine Learning
Previous Article in Special Issue
A Survey on Video Streaming for Next-Generation Vehicular Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

State-of-the-Art Electric Vehicle Modeling: Architectures, Control, and Regulations

by
Hossam M. Hussein
,
Ahmed M. Ibrahim
,
Rawan A. Taha
,
S. M. Sajjad Hossain Rafin
,
Mahmoud S. Abdelrahman
,
Ibtissam Kharchouf
and
Osama A. Mohammed
*
Energy System Research Laboratory, Department of Electrical and Computer Engineering, Florida International University, Miami, FL 33174, USA
*
Author to whom correspondence should be addressed.
Electronics 2024, 13(17), 3578; https://doi.org/10.3390/electronics13173578
Submission received: 31 July 2024 / Revised: 19 August 2024 / Accepted: 28 August 2024 / Published: 9 September 2024
(This article belongs to the Special Issue Featured Review Papers in Electrical and Autonomous Vehicles)

Abstract

:
The global reliance on electric vehicles (EVs) has been rapidly increasing due to the excessive use of fossil fuels and the resultant CO2 emissions. Moreover, EVs facilitate using alternative energy sources, such as energy storage systems (ESSs) and renewable energy sources (RESs), promoting mobility while reducing dependence on fossil fuels. However, this trend is accompanied by multiple challenges related to EVs’ traction systems, storage capacity, chemistry, charging infrastructure, and techniques. Additionally, the requisite energy management technologies and the standards and regulations needed to facilitate the expansion of the EV market present further complexities. This paper provides a comprehensive and up-to-date review of the state of the art concerning EV-related components, including energy storage systems, electric motors, charging topologies, and control techniques. Furthermore, the paper explores each sector’s commonly used standards and codes. Through this extensive review, the paper aims to advance knowledge in the field and support the ongoing development and implementation of EV technologies.

1. Introduction

Climate change is a worldwide problem that has an enormous influence on people’s lives, upsetting economies in every country. Rising sea levels, more intense weather events, and changing weather patterns are some of the ways that its consequences are still felt today. Since greenhouse gas (GHG) emissions are higher than the desired limits according to the UN’s regular emission gas reports, these emissions are the main cause of climate change. In the absence of any action, it is predicted that the average global surface temperature will climb by more than 3 degrees Celsius this century. Sectors relying on fossil fuel combustion, such as energy and transportation, are considered the main source of the higher rates of GHG levels globally [1,2,3,4]. According to the United States Environmental Protection Agency (EPA), the contribution of the transportation and energy sectors to CO2 emissions, as shown in Figure 1, is 24% and 42%, respectively [5]. To align with the goals of the Paris Agreement held in 2015, the widespread utilization of renewable energy resources became essential. In the same context, replacing the existing internal combustion engine-based (ICE) vehicles with another sustainable and clean energy source is considered a crucial step in the transportation sector.
The shift to electric mobility provides significant advantages, including reduced oil dependence and enhanced environmental health. Electric vehicles (EVs), which exemplify this transition, are known for their lower energy consumption and zero GHG emissions [6]. As a result, many developed nations are promoting the adoption of EVs to reduce air pollution, CO2 emissions, and other greenhouse gases. They support sustainable and efficient transportation through various initiatives, including tax incentives, purchase subsidies, free public parking, and toll-free motorway access [7]. As a result, the global share of electric vehicles is expected to rise from 2% in 2016 to 22% by 2030 [8]. Moreover, EVs present several benefits compared to conventional vehicles regarding cost, reliability, efficiency, and simplicity. The running cost is way lower in terms of maintenance and electricity costs. With fewer and simpler components, EVs experience fewer breakdowns. They also avoid the wear and tear caused by engine explosions, vibrations, and fuel corrosion inherent in traditional vehicles. Moreover, their simple structure eliminates the need for conventional engines’ cooling systems, gearshift, clutch, or noise reduction elements [7,9], Additionally, the EVs, as ancillary services, can support the power grid stability in vehicle-to-grid (V2G) and vehicle-to-home (V2H) operating modes or even minimize the disturbance caused by the dynamic charging profiles while operating in vehicle-to-vehicle (V2V) charging situations.
These benefits, driven by the noticeable progress in EV technologies and models, have motivated public users and countries to rely on EVs as a promising option. Since the mid-2010s, EVs have seen rapid growth, consistently breaking sales records, particularly in light-duty vehicles (LDVs), buses, and smaller vehicles. Global automakers have invested over USD 140 billion in transportation electrification, with 50 LDV EV models currently available in the U.S. and around 130 models anticipated by 2023. Projections for EV market share in LDVs vary significantly, from 10% of sales by 2050 to complete market dominance before 2050. Numerous studies indicate that EVs are becoming, or have already become, economically competitive with internal combustion engine vehicles (ICEVs) in certain applications [10].
Globally, China is leading the EV market. In 2016, six countries had E4W market shares above 1% of passenger light-duty vehicle sales. Norway led with 29% due to effective incentives and tax reductions [11]. The Netherlands and Sweden followed with 6.4% and 3.4%, respectively. The U.K., Netherlands, and Sweden prefer plug-in hybrids (PHEVs), while France and China favor battery electric vehicles (BEVs). Norway, Japan, and other regions have a balanced mix of BEVs and PHEVs [12]. According to [13], EV sales are projected to reach 17 million in 2024, making up over 20% of global car sales. Based on the statistics of the past three years shown in Figure 2, the market share in 2024 could reach 45% in China, 25% in Europe, and over 11% in the U.S., driven by manufacturer competition, decreasing battery and car prices, and continued policy support.
Electric vehicles (EVs) are soon poised to dominate the transportation market. Integrating diverse technologies across multiple engineering fields, including mechanical, electrical, electronics, automotive, and chemical engineering, drives this transformation. By amalgamating these technologies, the overall efficiency of EVs can be significantly enhanced, leading to reduced fuel consumption and lower emissions. As shown in Figure 3, EVs comprise several significant systems: the electric motor, power converter, and energy source, with each contributing to the vehicle’s performance and sustainability. As the adoption of EVs accelerates, the market will witness a substantial shift toward greener and more efficient transportation solutions underpinned by continuous advancements in these integrated technologies.
Several papers have reviewed the technological evolution of EVs from various aspects regarding the charging topologies, the types of motors used, converter configurations, battery management, and standards. In [14], the authors have introduced a review of lithium-ion batteries for EV systems and energy management. They have also discussed challenges, issues, and future research recommendations and highlighted EV energy management system problems and recommendations. However, the research lacks focus on battery safety and recycling processes and the discussion on environmental impacts and memory effects of batteries is limited.
A review of EVs aiming to replace ICE cars and reduce CO2 emissions is carried out by the authors in [15]. They discussed EV drive circuits, battery technology, and future challenges. The EV efficiency, battery types, and motor drives are also addressed. However, the research focus on autonomous driving technology advancements is insufficient and the discussion on the impact of EV adoption on energy infrastructure is limited as the paper gives brief discussion on most of the sections.
Insights into EVs and HEV advancements in current challenges are proposed in [16] Optimization strategies for traction motors, converters, and energy management systems are thoroughly compared. Improved traction motor drives for HEV applications with reliability and control are presented. Moreover, the analysis of power electronic converters for HEVs is unidirectional and bidirectional, but the discussion lacks focus on emerging battery technologies for EVs.
The author’s research work in [17] analyzes EV types, technology, sales evolution, charging modes, and future technologies. It is further extended to reviewing battery technologies, charging standards, power control, and energy management. The research is mainly focused on battery technologies from lead-acid to lithium-ion for EVs. However, the authors have not addressed the converter configurations, the motors employed in EVs, or the standards and regulations needed for operation.
Moreover, the research carried out in [18] presents a comprehensive review of HEV components, architectures, and MPPT techniques. The authors also introduced a comparison of existing hybrid vehicles. Furthermore, the paper focuses on traction motors, battery performance enhancement, and inverter selection. But the whole work is only limited to HEVs while other types of EVs are not addressed.
A comprehensive review is presented in [19] covering EV charging technologies, standards, architectures, and converter configurations broadly. The paper also analyzed charging systems, onboard and offboard chargers, and AC-DC and DC-DC converters and discussed future trends, challenges in EV charging, and grid integration. The limitation of the paper is the lack of focus on EV battery technology advancements and the limited discussion on grid integration challenges with EV charging systems.
The authors in [20] focused on the environmental impact associated with the employment of EVs regarding the reduction in Greenhouse Gases (GHG). They addressed the impact of EVs on reducing carbon footprint in the transport sector through sustainable mobility solutions. Challenges for EVs are proposed in the paper including driving range and charging station availability. However, the paper’s focus was on vehicle emissions and their environmental impact without addressing the vehicle’s key components and their configurations.
In [21], the authors introduced a review of Battery Management Systems (BMSs), charging stations, and traction motors employed in EVs. They concluded that lithium-ion batteries are ideal for EVs based on analysis. Furthermore, for the motor aspect, PMSM is suitable for low power while an induction motor is more convenient for high power. They also addressed that EVs have drawbacks like excessive costs, limited range, and charging time. However, the research work presented lacks the control techniques employed in EVs and the comparison between EVs and HEVs.
This paper comprehensively examines the latest developments in electric vehicle technologies and their regulatory frameworks. It begins by addressing the urgent issue of climate change and the pivotal role EVs play in mitigating its impact. In Section 2, the paper then offers a detailed background and market status on various types of EVs, including battery electric vehicles (BEVs), hybrid electric vehicles (HEVs), plug-in hybrid electric vehicles (PHEVs), and fuel cell electric vehicles (FCEVs). In Section 3, it further explores in Section 3 the spectrum of energy storage options, from lithium-ion batteries to advanced technologies like solid-state batteries and supercapacitors as well as the critical role of battery management systems. The discussion extends in Section 4 on charging technologies, covering both wired and wireless topologies and on-board and off-board chargers, emphasizing the control approaches for each type. Additionally, the paper delves into the state-of-the-art electric motors used in EVs in Section 5, highlighting their advancements and efficiencies. Finally, in Section 6, it examines the standards and regulations that harmonize the EV market, ensuring compatibility and safety across all sectors are discussed. Through this extensive review, the paper aims to provide valuable insights for advancing EV technology and supporting sustainable transportation solutions.

2. Background

There are diversified types of EVs on the market, and they use different technologies and ways of operation. The most common type is battery electric vehicles (BEV) or all-electric vehicles (all-EV). The key concept of this type is to replace conventional combustion engines with electric motors driven by high-capacity rechargeable batteries or more than one energy storage system (ESS). The BEVs are purely electric, with no fuel tank or tailpipe, highlighting them as an emission-free vehicle [22]. However, the Environmental Protection Agency’s (EPA) range is typically between 100 and 250 miles before the next recharging cycle. Several models are available on the market in this category, such as the Nissan LEAF, Hyundai Kona Electric, Tesla Model 3, and Rivian R1T. However, battery-electric vehicles (BEVs) face notable drawbacks, such as shorter driving ranges and longer charging times, compared to other EV types. Over time, these vehicles may experience battery degradation, leading to reduced range and costly replacements. Furthermore, the infrastructure needed for the charging stations and their availability are considered a big challenge for the heavy utilization of BEVs. Additionally, BEVs have a lower energy density than gasoline, affecting their overall efficiency [23,24,25,26]. The BEV is yet to be available in the market, with advancements in research toward overcoming the obstacles to their future dominance.
Hybrid electric vehicles (HEVs) enhance engine flexibility by incorporating an additional propulsion system. Besides the internal combustion engine (ICE), a typical HEV powertrain includes an energy storage system (ESS) and one or more electric motors (EMs). HEVs combine the power of ICE with the emission-free benefits of electric vehicles, offering superior fuel efficiency compared to traditional ICE vehicles [27,28]. They generally operate in charge-sustaining (CS) mode, maintaining the battery’s charge throughout the trip via regenerative braking or the combustion engine. A key limitation of HEVs is that they cannot be charged from the grid. However, the multiple energy sources in HEVs allow various power flow patterns to meet vehicle load requirements through different powertrain topologies and energy management strategies [29,30]. The Honda Civic Hybrid, Kia Sportage Hybrid, and Toyota Crown Signia are examples of the HEV models available on the market. Hybrid cars offer improved fuel economy by using low-speed electric power and assisting the internal combustion engine, leading to fuel savings and reduced emissions. They also have cheaper running costs due to regenerative braking and less brake wear. Additionally, hybrids provide better performance in urban settings and eliminate range anxiety. However, they are more expensive up front, require charging, have battery longevity concerns, and still produce emissions [31].
Unlike HEVs, plug-in hybrid electric vehicles (PHEVs) can be externally charged through power outlets. In PHEVs, most power comes from an electric motor (EM), with the internal combustion engine (ICE) as a backup. When the battery’s state of charge (SOC) drops to a certain level, the PHEV operates like a regular HEV, with the ICE taking over as the primary power source. PHEVs mainly function in charge depletion (CD) mode, depleting the SOC to a threshold level. They extend the all-electric range, enhance local air quality, and can be connected to the grid. Unlike regular HEVs, which must travel 1–2 miles on electric power alone, PHEVs can go 10–40 miles before requesting gas engine assistance [32,33] Additionally, PHEVs support grid tasks by using their onboard batteries as storage in smart homes. Through bidirectional chargers, they enhance energy management by aiding renewable generators, contributing to more efficient energy use in residential settings [34]. The PHEVs are available on the market in different models, such as the BMW i-8, Audi A3 E-Tron, Jeep Wrangler 4xe, and Toyota RAV4.
Fuel Cell Electric Vehicles (FCEVs) are considered a zero-emission type of electric vehicle fueled by hydrogen fuel cells. Toyota Mirai and Hyundai Nexo are available models in this category on the market nowadays. The key concept is to utilize hydrogen to generate electricity through a chemical reaction with oxygen from the air instead of drawing electricity from a battery. Furthermore, FCEVs can recapture the regenerative energy during braking driving situations. Contrary to BEVs, which use batteries to store their energies, FCEVs generate electricity on demand based on the volume of the hydrogen storage tank available onboard [35]. This allows FCEVs to reach longer driving ranges, typically exceeding 300 miles on a single tank of hydrogen, with a refueling time almost like that of conventional gasoline vehicles [36]. Nonetheless, the widespread adoption of FCEVs is impacted by multiple challenges related to the infrastructure needed to install several charging stations, besides the relatively high cost of hydrogen production and storage compared with conventional fuels. Despite these challenges, FCEVs are gaining attention for their potential to provide long driving ranges, typically more than 300 miles. In addition, they have environmentally friendly transportation solutions. These benefits motivated governments and manufacturers to invest more in hydrogen infrastructure and technology development to realize the full potential of FCEVs in achieving sustainable mobility objectives worldwide [37,38]. Other classes, such as Mild Hybrid Electric Vehicles (MHEV), like the Chevrolet Silverado Hybrid, and Range Extended Electric Vehicles (REEV), like the BMW i3, are available on the market. More details regarding the comparison among the commonly available EVs on the market are highlighted in Table 1 [39,40,41,42,43].
The development of control strategies for EVs is a crucial area of research aimed at optimizing energy management, ensuring system stability, improving fuel economy, and reducing emissions. Control strategies in these vehicles typically involve the coordination of energy management systems, power electronic converters, and electric machines. These strategies can be broadly categorized into rule-based methods, optimization-based methods, and learning-based methods. Rule-based strategies, including deterministic and fuzzy logic rules, are known for their simplicity and real-time applicability but often fall short in terms of optimization and adaptability. Optimization-based methods, such as Dynamic Programming (DP), Model Predictive Control (MPC), and Pontryagin’s Minimum Principle (PMP), offer more precise energy management by forecasting future states and optimizing control actions accordingly, though they can be computationally intensive. Recently, learning-based approaches like Artificial Neural Networks (ANN) and Reinforcement Learning (RL) have emerged as promising alternatives due to their ability to learn and adapt to complex nonlinear system behaviors. These strategies are particularly beneficial for handling the dynamic and unpredictable nature of driving environments, contributing to improved fuel economy and reduced emissions. For instance, an adaptive energy management strategy combining ANN and PMP has been shown to improve fuel economy significantly by adjusting to varying vehicle mass, demonstrating the potential of integrating machine learning with traditional optimization techniques [44,45].

3. Energy Storage System (ESS)

Energy storage systems are considered key elements in real-world applications. The high technologies, power, and energy capacities offered by the ESS make it crucial in several applications, such as power utilities, renewable energy systems, and transportation electrification. For instance, the ESS is used for grid support from the generation levels downstream to the end users in voltage and frequency regulations, stability and resiliency enhancement, and load balancing [44,45]. Regarding the EV industry, ESS plays a pivotal role as the main source of energy for types like the BEV and the backup source for others, such as the HEV and PHEV. Furthermore, it is expected that the short-term grid storage demand can be fulfilled by EVs only by 2030 [46]. ESSs come in various forms: chemical, electrical, electrochemical, thermal, and mechanical. With their distinct features of cyclability, efficiency, power, and energy densities, electrochemical ESSs are highly valued in industrial, commercial, and utility applications. Furthermore, its integration with other ESSs is applicable, forming hybrid energy storage systems (HESS). This category encompasses a wide array of devices, as shown in Figure 4, including batteries, supercapacitors (SCs), and fuel cells (FC), making them a preferred choice for EVs [47,48]. Nonetheless, several obstacles must be considered when integrating the ESS, including its life cycle, dependability, cost, and environmental and safety concerns. Therefore, advanced management systems are needed to address these challenges. This section is organized to delve into details on the different ESS in the EV market, the main objectives of the BMS systems, and the common management strategies for the HESS.

3.1. Battery Energy Storage System (BESS)

The driving range of EVs, which is heavily dependent on battery capacity, is crucial to their success. The driving range of EVs, which is heavily dependent on battery capacity, is crucial to their success. EV batteries must be compact, flexible, and capable of rapid and frequent recharging. A high energy density, low maintenance profile, long lifespan, and cost-effectiveness are also imperative [49]. Broadly, different commercial batteries are used in the EV market. Besides the traditional lead-acid batteries, a broader array of battery types is now extensively used, such as ZEBRA batteries, nickel-metal hydride (NiMH) batteries, and lithium-ion batteries (LIBs). It is worth mentioning that battery utilization, defined as C-rate, is measured by the ratio between the charging and discharging currents of the battery concerning its capacity. This rate provides a detailed indication regarding the charging and discharging process of the battery and how heavily it is being used [50]. More details regarding each model and a summary are provided at the end of this section.

3.1.1. Lithium-Ion Batteries (LIBs)

LIBs are leading EVs in the meantime due to their remarkable features in terms of sustainability, high efficiency, extended life cycle, excellent high-temperature performance, and noticeable high energy densities [51,52]. Although some concerns about their safety and environmental impacts during their production and recycling processes are debated, LIBs are still the winning horse in the EV market [53,54]. Moreover, advanced research is going to minimize these challenges and improve their overall performance [55,56,57].
The operation concept of LIBs is performed by shuttling the lithium ions through an electrolyte between the anode, typically made of graphite, and the lithium metal cathode. A separator prevents physical contact between the anode and the cathode, facilitating efficient energy storage and release [58]. Different lithium-ion chemistries are used in commercial EVs; each has advantages and disadvantages. To date, Li-iron-phosphate (LFP), Li-ion manganese oxide (LMO), Li-titanate-oxide (LTO), Li-nickel-cobalt-aluminum-oxide (NCA), and Li-nickel-manganese-cobalt-oxide (NMC) batteries are employed by carmakers such as Toyota, Nissan, Tesla, BMW, etc. In 2016, 83% of the EV market relied on LFP and NMC batteries. Reaching 2025, NMC batteries are expected to reach 41% of the market share compared to LFP [59]. Table 2 summarizes the key differences among all these types [60,61,62,63].
Long life spans, enhanced safety, and remarkable thermal stability are all offered by LFP batteries. Compared to the other LIBs, they offer improved cycle performance and structural stability. Furthermore, their cost and durability make them perfect for heavy-duty purposes like buses and vehicles [59,64]. However, they have a low nominal voltage and less electronic conductivity, reducing energy density [65]. The LMO batteries are renowned for being non-toxic, having a solid construction, and being less expensive. In addition, they are providing three-dimensional diffusion channels for lithium, which enhances their rate capabilities. Despite their benefits, their low performance in high-temperature conditions and their low capacities are the main drawbacks of this type. However, merging the LMO with NMC has achieved noticeable enhancements in specific energy and longevity [59].
Known for their outstanding overall performance, low self-heating levels, and high energy density, NMC batteries are considered the perfect option for the automotive industry. In addition, NMC chemistry can be tailored to achieve specific performance for several applications that require frequent cycling. However, relying on expensive materials such as cobalt and nickel and the complex manufacturing process are challenging the widespread use of NMC batteries [66,67]. Replacing manganese with aluminum introduced NCA batteries with improved performance and a longer life span. Additionally, NCA batteries are highly valued for their exceptional power and energy densities, making them an optimal option for long-range and high-performance applications. However, the primary challenges for NCA batteries are their low thermal stability and the high cost of cobalt [68,69,70]. Regarding the LTO, they are distinguished by their fast charging and discharging patterns, robust spinel structure, and unique safety performance. Furthermore, their high operating potential prevents the formation of lithium dendrites and allows the use of cost-effective aluminum foil. Additionally, LTO batteries excel in extreme temperature conditions and boast an exceptionally long cycle life. However, their low energy density and low electrical conductivity present challenges that necessitate advancements in stable electrolytes and high-voltage cathode materials [68,71,72,73]
Because they are reliable and reasonably priced, LFP batteries are ideal for heavy-duty applications; nevertheless, their energy density is low. When combined with NMC for EVs, LMO batteries offer good rate capability at a reasonable cost. High specific capacities and energy densities are provided by NMC and NCA batteries, making them appropriate for various automotive applications. Although LTO batteries offer exceptional lifespan and safety, further research and development are required to surpass their energy density constraints. Because each type of battery has distinct advantages, it may be used in various ways in the developing EV industry. Figure 5 highlights the main differences among the selected common LIBs in the EV market [60,61,62,63].

3.1.2. Nickel-Metal Hydride (NiMH) Batteries

NiMH battery stands among the most promising ESSs. The commercial use of NiMH batteries started in the early 1990s. The early chemistry of NiMH batteries employed a metal alloy for the cathode of a mixture of lanthanum, nickel, cobalt, and silicon. Later, advancements occurred to the battery featuring it with more energy densities than lead-acid batteries [74]. In addition, Nickel Metal Hydride (NiMH) batteries offer a three times higher specific energy compared with the standard Nickel Cadmium (NiCd) batteries [75]. These rechargeable batteries are widely used for various electric loads, including mobile applications, stationary ESSs, and especially transportation systems. Over one billion cylindrical NiMH cells are produced annually to replace toxic NiCd batteries and primary alkaline batteries, providing the same rated voltage and higher energy capacity [76,77].
Several benefits make nickel-metal hydride (NiMH) batteries a desirable option for a range of applications. They are perfect for devices that need continuous power because of their longer runtimes due to their higher energy density as compared to nickel-cadmium (NiCd) batteries. NiMH batteries pose fewer threats to the environment and human health since they do not contain harmful cadmium. Compared to lithium-ion batteries, they are also considered safer because of their decreased danger of thermal runaway and high cycle life, which allows them to withstand several charge-discharge cycles [78,79,80]. However, NiMH batteries do have certain drawbacks. Their increased self-discharge rate causes them to lose charge more quickly while not in use [81]. Even though it is not as bad as with NiCd batteries, the memory effect might nonetheless happen and eventually reduce capacity if the battery is not periodically fully drained. Furthermore, although still superior to NiCd, their energy density is still lower than that of lithium-ion batteries, which means they are less appropriate for small lightweight applications. Lastly, extreme weather conditions might cause NiMH batteries to function less well [78,79,82].
The NiMH industry has experienced rapid growth since mass production started in Japan in 1990, with well-known battery makers accelerating their development. This growth occurred in the volumetric and gravimetric energy densities, which are aimed at enhancing the overall power density. By 2000, Japan had produced 1 billion small Ni-MH batteries annually. Moreover, China, which participated early in Ni-MH battery development by building numerous production lines, such as the Shenzhen BYD and Shenyang Sanpu Companies, now leads global production and export, surpassing Japan. However, the market has declined due to the global economic crisis and the rise of lithium-ion battery utilization. Despite this, 85% of HEVs still rely on Ni-MH batteries. With increasing demands for energy density, high power, and long cycle life, ongoing efforts focus on enhancing Ni-MH battery properties [83,84]. Nowadays, different car makers use NiMH batteries in their HEV models, including the Toyota Prius, Honda Insight, Ford Escape Hybrid, and Lexus Hybrid.

3.1.3. ZEBRA Batteries

Developed since the 1980s, sodium-nickel chloride (Na-NiCl2) batteries are high-temperature rechargeable batteries called ZEBRA (Zero Emission Battery Research Activity). Featured by better safety, reliability, and life span profiles, it is considered one of the most valuable electrochemical devices for stationary energy storage and EV applications [85]. Melted sodium and nickel chloride undergo electrochemical processes that underpin the ZEBRA battery’s operation. The solid ceramic electrolyte, composed of beta-alumina, separates the positive electrode of nickel chloride from the negative electrode of molten sodium in the cell. As a result of the reaction between sodium ions and nickel chloride during discharge, nickel and sodium chloride (table salt) are produced. Electrical energy is produced by releasing electrons in this process [86]. Since all the sodium is added as salt, ZEBRA cells are created in a discharged condition, contrary to LIBs, assembled in a charging state, preventing the processing of metallic sodium. Generally, battery packs, including up to 500 ZEBRA cells, with cooling plates and vacuum insulation for thermal control, can be configured in parallel or series [87]
There are multiple noteworthy benefits to ZEBRA batteries. For instance, their high energy density makes them perfect for applications needing small energy storage solutions. In addition, these batteries are intrinsically safe. Since it employs a solid-state ceramic electrolyte, the possibility of short circuits and inhibiting dendrite development has decreased. Furthermore, the lack of harmful chemicals highlighted it as safe for the environment compared to the other available batteries. ZEBRA batteries’ high working temperature guarantees superior thermal stability and reliable performance. However, more insulation and heating components are needed to achieve that. This will add to the complexity of the battery design and increase energy consumption. Another negative aspect is the starting time, which delays instant utilization as the battery warms from a cold condition to its operational temperature. Moreover, ZEBRA batteries are more expensive than alternative technologies due to the high cost of the materials and production techniques. The battery system becomes more complicated and has more potential sites of failure because of the demand for thermal control devices [88,89,90,91,92]. It is worth mentioning that ZEBRA batteries can achieve robust performance for EV operations when aligned with additional storage devices, such as supercapacitors [93,94].

3.1.4. Solid-State Batteries (SSB)

Since LIBs were commercialized in the 1990s, significant research has been dedicated to improving this technology. Although high-voltage cathode materials can modestly boost LIB energy density, this method is impractical due to the decomposition of the liquid electrolyte at the negative electrode under high-voltage conditions. Moreover, the flammability and volatility of liquid electrolytes and additives raise safety concerns [95]. In this context, solid-state batteries (SSBs) represent a major advancement in battery technology, offering several advantages over traditional LIBs. Unlike LIBs, which use a liquid electrolyte for ion transport between the cathode and anode, SSBs employ a solid-state electrolyte (SSE). This advancement promises significant enhancements in safety, energy density, and lifespan compared to conventional lithium-ion batteries [96]. For instance, SSBs have better energy densities than lithium-ion batteries. This is largely because they can employ lithium metal as an anode, which has a larger theoretical capacity than graphite anodes in typical lithium-ion cells [97]. In addition, SSEs are more chemically stable and resistant to deterioration over time than liquid electrolytes. This contributes to a longer cycle life. Furthermore, the possibility of dendrite development is minimized, which is a typical problem in LIBs, resulting in short circuits and possible failures [98]. Moreover, SSBs can function safely at elevated temperatures without the threat of thermal runaway, a significant safety issue for lithium-ion batteries.
Generally, SSBs can be categorized into three types based on their chemical composition. Firstly, the inorganic solid ceramic electrolytes, usually made from lithium ceramics, provide high ion conductivity and thermal stability. However, they are fragile and challenging to make [99]. Secondly, organic solid polymer electrolytes, made from polymers like polyethylene oxide (PEO) or polyvinylidene fluoride (PVDF), offer good mechanical flexibility and processability but have lower ion conductivity. Lastly, composite solid electrolytes (CSEs) combine inorganic ceramics with organic polymers to achieve high ion conductivity and good mechanical properties, allowing for tailored properties through varied compositions and structures [100].

3.1.5. Lead Acid Batteries (LABs)

Lead-acid batteries, invented by Gaston Planté in 1859, are the earliest rechargeable batteries still used today due to their durability, dependability, and low cost [101]. They are structurally made up of lead dioxide and sponge lead electrodes bathed in a sulfuric acid electrolyte, separated by a porous substance to prevent short circuits. While discharging, chemical reactions produce lead sulfate, water, and electrical energy, whereas charging reverses these processes. In general, they are classified as either flooded or sealed batteries. Regarding the flooded LAB, the plates are fully submerged by a liquid electrolyte [102]. This type is commonly used in vehicle starters, industrial forklifts, and backup power systems. However, regular maintenance is required to restore the water lost through electrolysis and evaporation. In contrast, sealed LABs have immobilized electrolytes in gel form or absorbed in a glass mat. This type, featuring a maintenance-free profile, can be found in motorized scooters, uninterruptible power supplies (UPSs), and backup lighting systems [103].
Although LABs have a relatively small energy density (30–40)% of the theoretical limit compared to 90% for lithium-ion batteries (LIBs), their production methods and recyclability significantly reduce their ecological footprint due to the use of readily available, inexpensive materials, and an inflammable water-based electrolyte [104]. Nonetheless, they are heavier and bulkier than other advanced batteries. Additionally, their low energy density, the need for frequent maintenance, and their short cycle life stand against their wide utilization in the transportation sector [105]. Even though LABs are among the earliest battery technologies, they are still widely used in many applications. Sustained progress endeavors to mitigate their constraints and enhance their efficacy, guaranteeing their pertinence within the dynamic terrain of energy storage. In this context, improved performance and more excellent durability are provided by enhanced flooded batteries (EFBs), especially for applications involving start-stop vehicles. Moreover, advanced lead-acid batteries use innovative chemicals and production processes to boost longevity and efficiency. Research is also focused on developing lead-acid battery recycling procedures to improve sustainability. It is worth mentioning that LAB’s low cost and great recyclability make them more attractive for use in grid storage. In the future, these developments might be combined to produce more competitive LABs that satisfy the needs of contemporary applications such as energy storage without sacrificing their dependability and affordability [106,107,108].
To conclude, the market for electric vehicles (EVs) has witnessed the use of multiple battery technologies, each with distinct features. The first kind of batteries were LABs. Despite their lower energy density and heavy weight, they are dependable and cost-effective. Since its introduction in the 1980s, nickel-metal hydride (NiMH) batteries have been extensively utilized in hybrid automobiles due to their increased energy density. ZEBRA batteries, developed in the 1990s, are noted for their high working temperatures and were early candidates in electric vehicles. However, the complexity regarding their production besides their costs stands against their broad use. In the meantime, LIBs are dominating the EV industry because of their high energy density and efficiency while the SSBs are expecting to revolutionize the electric vehicle industry. It is worth mentioning that there are three types of cell housing for these batteries: cylindrical, prismatic, and pouch batteries. Today’s EVs include all three, and each has advantages and disadvantages. For instance, Tesla utilizes cylindrical cases for their reliability and high energy density, despite their bulkiness. Volkswagen favors prismatic housings, which are lighter and fit well in compact spaces, although they have a shorter life cycle. Regarding the type of pouch used by GM and Hyundai, it offers flexibility to fit small irregular spaces but is prone to swelling and potential fire risks. Table 3 provides a comparison of the most common types of batteries in the EV market [109,110]

3.1.6. Other Energy Storage Types

Supercapacitors (SCs)

Recently, there has been a noticeable growth in the power needs in specific applications beyond the capacity of typical design batteries. Hence, supercapacitors (SCs), or ultracapacitors, are becoming a more popular alternative storage medium due to the increasing need for instantaneous energy storage solutions, especially in the renewable and electric car industries. Unlike batteries, which rely on chemical reactions, capacitors, in general, store energy physically, allowing for rapid charge and discharge cycles. However, SCs, compared to traditional capacitors, have a different structure and working mechanism. In addition, SCs have significantly higher capacitance and energy density due to their larger surface area and very small separation distance [111].
Three categories may be used to categorize SCs based on the various charge storage techniques they use [112]. The standard type is the electric double layer (Helmholtz layer) SC. Typically, they are structured as two porous electrodes immersed in an electrolyte with a separator. Due to porous materials like activated carbon frequently used to make electrodes, the surface area used for charge storage has dramatically increased. Electrostatic and electrochemical reactions are critical to the SC concept. Supercapacitors retain energy in the electric dual layers that develop at the interface between the electrode material and the electrolyte, in contrast, conventional capacitors store energy by electrostatic charge separation on metal plates. Because of the double-layer process and the permeable electrodes’ large surface area, SC can attain more significant capacitance values than conventional capacitors. Second, there are pseudo capacitors that store energy using Faradaic redox reactions, which are rapid reversible chemical adsorption and desorption processes. Through this method, the energy storage capacity is increased beyond what can be achieved using electrostatic means. Hybrid SCs combine the Faradaic redox reactions with the Helmholtz layer [112,113,114]. Table 4 summarizes more details regarding these three types [115].
Known for their unique features, such as their high-power density, long life span, and sustainability, SCs have found significant uses in many sectors. For instance, specific power fluctuations in renewable energy systems and regenerative braking modes may be mitigated, improving their systems’ stability and dependability [116]. Moreover, SCs are heavily utilized in hybrid and electric buses. In Shanghai, by August 2006, supercapacitor electric city buses had begun offering their commercial services in Shanghaist to LIB electric buses, and SC buses do not require specialized charging facilities. Additionally, when the supercapacitor electric city buses remain at the departure station, they are fully charged in a few minutes [112]. Regarding passenger cars, SCs are utilized for various purposes, including high-voltage energy storage systems and engine start-stop, energy regeneration, voltage stabilization, and backup power supply [116,117]. Moreover, using hybrid energy storage systems, incorporating SCs, for EV applications has witnessed promising progress through the years. This hybridization enhances vehicle performance and extends the life cycle of the storage system [118,119,120].

Fuel Cells (FCs)

As we stated earlier, FCs are using hydrogen as a fuel to generate electricity through reactions. They are typically structured with the electrolyte layer in contact between the anode and cathode electrodes. FCs can be divided into categories based on operating temperatures and chemistries. Solid oxide fuel cells (SOFC), alkaline fuel cells (AFC), phosphoric acid fuel cells (PAFC), molten carbonate fuel cells (MCFC), direct methanol fuel cells (DMFC), and proton exchange membrane fuel cells (PEMFCs) are among the several fuel cell types in the industry. Because different fuel cells produce varied operating temperatures and power levels, choosing the suitable fuel cell is crucial. Efficiency and cost are crucial when choosing the finest fuel cell [121].
A fuel cell’s (FC) efficiency is influenced by the power demand placed upon it; typically, higher power draws result in lower efficiency. This decrease in efficiency is primarily due to voltage drops caused by internal resistances. Moreover, fuel cells have a relatively longer response time compared to batteries and SC. FCs are also costly, priced at five times more than internal combustion engines (ICEs). The main cost drivers are the membrane, electrocatalyst, and bipolar plates. Ongoing research aims to develop hydrocarbon membranes to replace perfluorinated membranes [122]. In automotive applications, low-temperature and low-pressure PEMFCs are the most frequent FCs owing to their exceptionally high-power density, lower working temperature (60–80 °C), and reduced corrosion compared to other FCs [78]. More details regarding the key differences among the commercial FCs are summarized in Table 5 [122,123,124].

3.2. Battery Management Systems

Battery management systems (BMSs) are crucial for the secure operation of EVs. While the BESS incorporates different chemicals with their various characteristics in terms of voltage, current, and temperature levels, regardless of the driving patterns and weather conditions, maintaining these values within safe operating levels is mandatory. The main objectives of the BMS shown in Figure 6 include but are not limited to the following aspects [125,126,127,128].
  • Measuring the status of each component of the EV, such as the battery state of charge (SoC), state of health (SoH), and voltage and current levels.
  • Conducting the necessary investigation on this data to assess the battery capacity and ensure cells and battery balance.
  • Predict any degradation possibilities for the BESS.
  • Running the thermal checkup and regulating the BESS temperature.
  • Observing any fault conditions with detection, diagnosis, and mitigation.
  • Regulating the charging and discharging signals through the proper power and control management approaches.
  • Supporting data exchange with other IoT devices toward enhancing data acquisition, storage, and performance optimization.
The decision-making processes involved in battery management systems (BMS) depend on several critical factors, including the estimation of battery state of charge (SOC) and state of health (SOH). These core tools enable the BMS to implement precise control, protection, and management actions. The initial step in achieving this involves estimating these values through various methods, such as direct estimation techniques, battery equivalent models, and data-driven approaches. Each method presents distinct advantages and disadvantages regarding complexity, computational demands, accuracy, and other relevant aspects [128].
Subsequently, these estimated signals are transmitted using specific communication technologies for visualization and management purposes. Based on these measurements, appropriate control signals are generated, utilizing suitable management strategies to regulate the power and operation of the electric vehicle (EV). Technologies such as cloud computing, artificial intelligence, blockchain, and digital twinning can facilitate these processes [125,128]
Given the BMS’s essential role in executing and managing these tasks efficiently, it must be equipped with advanced technologies to ensure smooth, secure, enhanced, and optimized vehicle operation. With ongoing research and development, the future of battery management systems appears promising, heralding a more sustainable and efficient transportation ecosystem.

3.2.1. SOC/SOH Estimation Approaches

While the concept of SoC appears straightforward, achieving accurate estimation is remarkably complex. This complexity arises from the battery’s inherently non-linear electrochemical properties, combined with variations in cell performance and differing operating conditions within each battery pack. Furthermore, these characteristics are not fixed; they evolve over time due to battery aging, fluctuations in temperature, and the complexities of charge and discharge cycles. This dynamic behavior demands a sophisticated and continuous assessment of the battery’s performance parameters, significantly complicating the accurate determination of its available capacity under a wide range of internal and external conditions.
There are only a few methods for directly measuring the SoC and SoH of a battery through its physical and chemical characteristics, such as electrolyte pH, density measurements, and cathodic galvanostatic pulses. However, these techniques require highly precise instruments, which can be costly and often impractical due to the difficulty of accessing the internal components of the battery [58]. As a result, SoC estimation typically relies on parameters that can be easily measured, including voltage, current, impedance, and temperature. Despite the low computational demands of these traditional approaches, they frequently lack accuracy. This inaccuracy is further exacerbated by the flat OCV-SoC curve characteristic of LIBs, making the SoC-voltage relationship less reliable. To overcome these limitations, indirect SoC estimation methods have been introduced. These approaches leverage directly measurable variables from the battery to achieve a more precise SoC estimation [129,130].
Indirect methods for estimating the SoC can be broadly categorized into two main types. The first category includes model-based methods, which rely on battery models to establish a relationship between measurable parameters and SoC. The accuracy of these methods is highly dependent on the specific battery model employed. The second category comprises data-driven approaches, which necessitate a large and diverse dataset to train models that can effectively approximate the relationship between the input data and SoC. It is worth noting that some of these methodologies require the battery to be charged and discharged according to predefined profiles during the feature extraction phase to generate accurate estimates. This process, known as “offline estimation”, is often impractical for many BMS applications. Consequently, there is increasing interest in developing online SoC estimation techniques that can function under the normal operating conditions of batteries. Figure 7 highlights the main SoC/SoH estimation techniques [131,132,133].
Despite the reduced accuracy of Ampere-Hour (AH) models due to their open-loop nature and the extended relaxation time in OCV methods, model-based techniques remain robust and effectively represent the dynamics of batteries. However, integrating these techniques with adaptive algorithms, such as Kalman filters and observers, is crucial to mitigate errors in model parameter calculations. This integration, though complex, presents an opportunity for Data-Driven (DD) models to serve as valuable tools for SoC/SoH estimation, either independently or by simplifying the complexity of model-based methods through predictive analysis of battery dynamics. The adoption of DT technology further enhances this approach, improving the efficiency of BESS and preventing unexpected degradation. Consequently, this review suggests that combining model-based estimation techniques with data-driven approaches, which are rapidly advancing in their ability to identify dynamic patterns within batteries, can yield precise SoC estimations while reducing model complexity. Table 6 provides a summary of the key features and limitations of the frequently used SoC/SoH estimation approaches discussed [128,132,134].

3.2.2. Battery Thermal Management

A battery’s thermal environment has a direct impact on its lifetime, efficiency, and safety. Excessive heat within the battery can cause catastrophic occurrences, such as explosions. As a result, adopting a battery thermal management system (BTMS) is critical to keeping the battery temperature within a safe range, often between 20 °C and 40 °C, therefore avoiding possible risks. The heat dissipation process within the battery is driven by electrochemical reactions, which can pose risks, especially when multiple cells are packed together. High temperatures can compromise battery safety, reduce cycle life, and diminish capacity, particularly in LiFePO4 batteries. BTMS relies on thermal models to accurately simulate the detailed temperature profiles within a battery or cell during operation, accounting for factors like ohmic losses, chemical reactions, and polarization. However, these models require a thorough understanding of the thermodynamic properties of the battery’s materials and components. The complexity of these approaches ranges from simple one-dimensional models, which assume constant heat generation rates, to advanced three-dimensional models that provide a comprehensive description of the battery’s thermophysical properties under varying conditions. These models are crucial for ensuring that the BTMS can effectively control the charging temperature and maintain optimal conditions under various operational scenarios [135,136].
The thermal management system, which can use air, liquid, or Phase-Change Material (PCM), is critical for regulating heat dispersion within the battery pack [137]. The air-cooling system uses ventilation fans or air extractors to provide a flow of cold air across the surface of the battery pack, dispersing heat produced by chemical processes within the battery [138,139]. Despite its use in automobiles such as the Toyota Prius and Honda Insight, this approach frequently results in severe irregular distribution of temperature, especially during extreme temperatures of operation and rates of discharge [140]. In the PCM system, the battery pack is submerged in the phase-change material, which dissipates heat from the outside throughout the phase-change process [141]. However, PCM’s low heat conductivity limits its usefulness, particularly when quick reaction to thermal swelling is required [142]. The liquid cooling system consists of flow channels designed around the battery surfaces, as well as heat exchangers such as a water jacket. This design allows you to modify the liquid temperature, which aids in the absorption of heat created by the battery pack [143,144]. Cooling mediums can be refrigerants, water, or mineral oil, with heat transmitted from the battery to the cooling fluid by direct or indirect contact. This technology provides more consistent temperature distribution and better thermal management than air and PCM systems.

3.3. Technical Challenges and Future Tre Systems

There are several concerns in the field of the BESS for any applications, particularly for the EV market. In the case of EVs, an extended cycle life means that the battery can withstand more years of everyday usage, which is critical for both cost-effectiveness and consumer confidence. A battery’s capacity to retain performance over several cycles is also critical for decreasing the need for premature battery replacements and lowering total ownership costs. The deterioration mechanisms built into batteries’ electrochemical processes have a substantial impact on their lifetime and efficiency. High charge and discharge rates repeated deep drain cycles and exposure to severe temperatures can all cause deterioration, resulting in reduced battery capacity and overall performance. Furthermore, cycle life is heavily influenced by the type of battery chemistry used. For example, LiFePO4 batteries have a longer cycle life than NCA batteries, while having a lower energy density on average. Using a solid electrolyte rather than a liquid electrolyte is at the forefront of research due to its potential to significantly increase cycle life. This breakthrough is due to a decrease in breakdown processes commonly associated with liquid electrolytes. Concurrently, developments in BMS improve the oversight and management of each cell inside a battery pack. These technologies contribute significantly to improving battery cycle life by enhancing charging cycles and regulating temperatures [145,146,147].
Another critical aspect is the growing demand for effective battery recycling and disposal methods due to the environmental risks posed by improper disposal. Batteries include hazardous elements including lead, cobalt, and organic solvents, which, if not properly handled, can cause substantial environmental damage. Recycling is essential for recovering precious minerals such as lithium, cobalt, and nickel, which may be utilized in the manufacturing of new batteries. This not only decreases the need for ecologically harmful mining, but it also helps to cut carbon emissions. However, recycling EV batteries poses various obstacles. For instance, the technological difficulty of recycling LIBs, as opposed to typical LABs, necessitates the use of unique procedures for each battery chemistry. Furthermore, the dismantling procedure is dangerous and labor-intensive. Furthermore, the economic viability of recycling is currently restricted since the expenses of recycling lithium-ion batteries sometimes outweigh the value of the recovered elements. This has led to insufficient investment in large-scale recycling facilities. To address these difficulties, governments and regulatory agencies are enforcing stronger battery recycling standards, such as extended producer responsibility (EPR) rules that make manufacturers liable for their products’ end-of-life management. Recycling technologies are also being improved technologically, such as hydrometallurgical and direct recycling processes, with the goal of increasing efficiency, lowering costs, and recovering a greater percentage of resources while reducing environmental effects [148,149,150,151]
In the same framework, the surge in electric vehicle (EV) manufacturing has resulted in a considerable increase in demand for critical battery materials such as lithium, cobalt, nickel, and graphite. These minerals are critical to the functioning of lithium-ion batteries, yet they are scarce and concentrated in certain geographical locations. However, the mining and processing of these commodities create significant environmental issues. These operations are energy-intensive and can lead to habitat damage, water pollution, and higher carbon emissions. Furthermore, the geographical concentration of these resources leads to supply chain vulnerabilities and geopolitical conflicts, which can affect material stability and price. Therefore, researchers are focusing on a variety of sustainable alternatives and technologies such as batteries with little or no reliance on limited or problematic materials, like cobalt-free batteries. In addition, the notion of a circular economy, in which old batteries are recycled into new ones, is critical for lowering reliance on newly mined resources while mitigating environmental concerns. Battery technology advancements, like solid-state batteries and alternative chemistries like lithium-sulfur or sodium-ion, also offer promise in terms of lowering the demand for crucial raw materials [152,153,154,155].
Another technical issue is the battery memory effect. This phenomenon is most linked with NiCd batteries and causes a loss of battery capacity when the battery is repeatedly recharged before being fully exhausted. This phenomenon happens because of electrochemical reactions within NiCd batteries, which cause crystalline formations to develop on the cadmium electrodes after repeated partial discharges. These crystals impede the battery’s ability to reach full charge, lowering its effective capacity over time. Symptoms include a considerable decrease in runtime and overall performance, particularly for devices that require a high-power output. To reduce this impact, it is advised to complete full drain and recharge cycles on a regular basis, to use contemporary smart chargers with deep discharge capabilities, and to optimize charging cycles using sophisticated BMS. Recent improvements in battery technology, such as NiMH and LIBs, have virtually removed the memory effect due to their differing electrochemical characteristics and superior design, hence improving both performance and longevity [156,157,158].

4. Charging Technologies and Control

Battery technology has mostly been responsible for the market’s expansion. Commercialization of EVs is dependent on the EV battery’s weight, price, charging time, and lifetime [159,160] The parameters of the battery charger significantly impact battery lifetime and charging time, and battery module performance is influenced by how modules are charged and discharged [161]. The charging time must coincide with refueling a conventional ICE vehicle to achieve ultra-fast charging. The EV supply equipment must also be expanded to accommodate higher power levels. Using high-power converters and smart control strategies can shorten the charging period. As a result, numerous studies have developed converter topologies, designs, and control algorithms for EV charging systems. Therefore, this cutting-edge technology relies heavily on battery chargers.

4.1. EV Charging Levels

The charger’s power level provides data concerning the equipment, cost, location, charging time, rate, and impact on the grid. The power levels of EV chargers are categorized into three main levels, as shown in Table 7. Since Level 1 EV chargers draw their power from a standard power plug, they do not need any special installation. The most common EV chargers installed in homes are level 2 chargers, albeit they need certain installation. Level 3 EV chargers, with their high-power delivery capabilities, are called DC fast-charging chargers for commercial use [162]. Extrema fast charging (XFC) can provide an ICE-like refilling experience. Power electronic components for the XFC stations are designed to emphasize front-end AC-DC converter stages and controllers, isolated DC-DC converters, and solid-state transformers (SST). The XFC requires a separate EVSE to deliver high power, and the installation costs are extremely costly. Furthermore, SST offers benefits over traditional line-frequency transformers in terms of galvanic isolation in XFC stations and converting medium voltage into low voltage levels [163].

4.2. EV Charging Methods

EV chargers can be divided into conductive and inductive categories according to how they physically interact, as shown in Figure 8. While the inductive charging method (also known as the wireless charging method, or WLC) uses an electromagnetic field to transfer power to the vehicles, the conductive charging method transfers electrical power by directly contacting the vehicles. However, compared to the WLC, the conductive charging approach is more effective [164]. Thanks to the Wireless Power Transfer (WPT) technology, power transfer is incredibly simple throughout the charging process. The fundamental principle of WPT technology is to create magnetic fields by first converting AC energy to DC energy and then, at a high frequency, converting it back to AC energy [165].

4.2.1. Conductive Chargers

The EV chargers’ place of installation and power rating can be utilized to classify them. Off-board, on-board, and integrated on-board battery chargers (IOBC) with unidirectional and bi-directional power flow capability are the three basic categories into which EV battery chargers can be divided. Bi-directional charging supports Vehicle-to-Grid mode (V2G), thus assisting in injecting power back into the smart grid, preferably during peak-load hours, while unidirectional charging reduces the amount of hardware needed and enhances battery degradation [166].

Off-Board Chargers

Off-board chargers are DC fast chargers that support (level 3) charging and are installed at designated charging stations, which require substantial infrastructure development. However, they offer increased power transfer capabilities by providing the vehicle with a high DC charging current fed directly to the battery. As a result, the charging time of off-board battery chargers can be limited to a few minutes. Consequently, DC charging is considered fast charging. Moreover, this type of charger does not affect the vehicle’s weight since the charger is placed at already-built stations [164].

On-Board Chargers

On-board chargers ease the charging process as they offer mobility since the EV can be directly connected to the grid terminals, yet they can add extra weight and cost to the vehicle. Additionally, the power transfer capability is significantly reduced since the size of the charger is constrained by the vehicle’s space and weight. Therefore, the charging time is increased compared to off-board chargers. Since battery chargers are usually categorized according to their power transfer capability and charging time, onboard chargers support three-phase charging (fast charging) as well as single-phase charging (slow charging) [167].

Integrated On-Board Battery Chargers (IOBC)

IOBC utilizes the already-existing propulsion components in the charging process through a bi-directional DC/AC converter, which works as a rectifier during charging. At the same time, the motor windings are used as a coupling inductor to smooth the grid line currents [138]. This topology can reduce the weight, cost, and space per vehicle by directly charging the battery without needing a bulky separate charging unit upon the vehicle or an already existing infrastructure [168]. The effectiveness of this technique depends on the type of used motor, the number of phases, and the employed converter to ensure that the produced average torque is nullified and achieves minimum hardware reconfiguration to switch between different modes of operation [169].
Figure 9 shows the key differences between the types of conductive chargers. First, the charging unit is in the charging station outside the vehicle for the off-board battery charger, which requires an already-built infrastructure. Therefore, there are no constraints on the charger’s weight, cost, or space. This results in high power transfer capability due to high-rating components. Consequently, the charging time of this charger is the fastest. However, a separate unit is installed on the vehicle for the on-board battery charger type. This topology offers more mobility as the charging can be carried out from power outlets and does not require an infrastructure. However, it adds extra cost, space, and weight per vehicle. As a result, the power transfer capability is restricted, and the charging time is increased. Finally, the integrated onboard charger utilizes the EV propulsion components in the charging by using a bidirectional converter while the motor windings act as filtering inductances. This concept can support the vehicle-to-grid (V2G) mode. However, the machine design and control can be a bit challenging [170].

4.2.2. Wireless Chargers

The Wireless Power Transfer System (WPTS), which has crucial power transmission capabilities, has been developed for charging EVs. The transmitter and the receiver are the electrically isolated power parts that comprise the WPTS for EV charging. The receiver is positioned in the electric vehicle (EV) and supplies power to the battery system, while the transmitter is in the roadway and powered by the grid. This eliminates the need for physical cables and plugs. A power converter and a coupling system are the two components of every wireless power transfer (WPT) section. The transmitter coupling unit produces an alternating field that can be electromagnetic, electric, or magnetic [171]. Various techniques transmit energy, with force and field form variations. The most common techniques are known as inductive, capacitive, resonant inductive, and magnetic gear power transfer [172].

Inductive Coupling Chargers

The two coils that make up the inductive WPTS coupling units are coupled to one another mutually, as shown in the schematic diagram in Figure 10. The transmitting coil is driven by high alternating current frequencies, typically between 10 kHz and 150 kHz or higher. Inductive WPTSs can operate at a higher power rating and can handle more power than capacitive WPTSs due to the increased energy density of the free space magnetic field. However, they produce a lot of electromagnetic noise, have high electric losses in coil resistance, and cannot operate if a metal object is positioned between the coils [173,174].

Inductive Resonant Chargers

Capacitors are added to the coupling coils in resonant WPTS to reduce the voltage and/or current required to power the WPTS. When the coupling coils are in the resonance state, their decreasing magnetic fields cause an electrical current to flow through the resonant circuit, charging the capacitor. As a result, the discharging capacitor also generates an electrical current, which in turn generates a magnetic field in the coupling units, allowing equivalent power to be transmitted between the inductor and the capacitor, as illustrated in Figure 11. The theory of resonance between the coils is the foundation of resonant coupling [175,176].

Capacitive Wireless Chargers

Capacitive wireless charging operates on the principle of creating an electric field between two plates: one in the ground (charging pad) and one on the vehicle. When the vehicle is positioned over the charging pad, capacitive coupling occurs between these plates, allowing for the transfer of electrical energy without direct physical contact [177] as shown in Figure 12. This method typically involves high-frequency alternating current (AC) to create the electric field. The main advantages of capacitive WPTSs are their low losses, low electromagnetic radiation, and the ability to transfer power via metal shields without eddy currents [178]. However, because of the extremely low energy density that can be held between plates in free space, capacitive WPTSs can only be used in low-power applications.

Magnetic Gear Wireless Chargers

The mechanical forces transfer power in the magnetic gear charger system. Certain applications, such as wind power generators and EV motors, use magnetic gear in place of traditional contacting gear. Unlike other frequent applications in EV and wind generators, this technique uses two synchronized permanent magnets as the principal coupling mechanism, and they are positioned side by side instead of coaxially. A current source that feeds into the transmitter winding causes an electro-mechanical torque to be generated on the primary side. At the same speed as the transmitter’s permanent magnet, the primary side permanent magnet rotates, causing the secondary side permanent magnet to rotate as well [179], as shown in Figure 13. Finally, a detailed comparison is presented in Table 8, discussing different charging types concerning performance, efficiency, employed control topologies, and many other aspects.

4.3. Control of Chargers

4.3.1. Control of Conductive Chargers

Control strategies for conductive chargers of electric vehicles (EVs) are crucial for ensuring efficient and effective charging processes. While EVs primarily operate in Grid-to-Vehicle (G2V) mode, future advancements in EV infrastructure are expected to enable Vehicle-to-Grid (V2G), Vehicle-to-Building (V2B), and Vehicle-to-Vehicle (V2V) modes. V2G technology allows EVs to support the grid by discharging power, with incentives for EV owners managed by aggregators [168]. V2B applications can integrate EVs into building energy management systems, leveraging excess energy from renewable sources [198]. V2V technology, using short-range communication, enables cars to exchange information on their status, enhancing traffic safety and efficiency.
Researchers have identified potential negative impacts of EV integration on smart grids, which can be mitigated through optimal management. Two primary control models are used to manage EV charging: centralized and decentralized [199]. In the centralized model, a central controller, such as an aggregator, collects data on EV states and system information to manage charging centrally. This model optimizes resource use but requires a highly reliable communication infrastructure [200]. In contrast, the decentralized model encourages EVs to manage their charging and discharging based on signals from the system operator, preserving private information and allowing EVs to act according to their preferences [201].
Battery chargers for EVs typically include single-phase or three-phase power factor correction (PFC) converters and DC-DC battery interface stages. PFC converters use dual-loop controllers with a fast inner current loop and a slower outer voltage loop to draw a sinusoidal input current and regulate the DC link voltage. DC-DC converters maintain the desired voltage or current at the battery terminals [202]. These control loops are generally managed with linear proportional-integral (PI) controllers in the synchronous DQ-reference frame, although tuning these controllers can be complex. To address this, model predictive control (MPC) methods are increasingly adopted for their fast dynamic response and ability to handle multiple objectives [203].
Implementing MPC-based control architectures for two-stage battery chargers involves using a converter predictive model and cost function minimization to generate switching pulses in the inner loops. In contrast, the outer loops remain the same. This approach offers advantages in dynamic response and objective handling but poses challenges such as variable switching frequency, high computational burden, and dependence on weighting factor selection. Despite these challenges, advanced control strategies are essential for maintaining efficient and reliable EV charging processes, particularly as EV integration into smart grids and broader energy systems continues to evolve [204,205].

4.3.2. Control of Wireless Chargers

Wireless Power Transfer (WPT) systems can be controlled by primary-side, secondary-side, or dual-side controls since both sides have power electronics converters for transmitting and receiving techniques [206,207,208]. Primary-side control involves regulation at the main winding, using a full primary bridge controlled by a regulated transmitter current. In some cases, power factor correction (PFC) is also utilized for regulation [209]. Secondary-side control employs an active rectifier and converter to charge the battery. While this method increases the vehicle’s weight and cost due to additional electronics on the secondary side, it ensures precise control over the charging process [210]. Dual-side control requires control on both the primary and secondary sides, necessitating a communication link between them. This method can involve either independent or coordinated control of both sides, addressing dynamic charging challenges. Dual-side control is ideal for dynamic wireless charging applications, though it requires careful consideration of stability issues when utilizing independent controllers on both sides [211]. The following are the fundamental control methods, as shown in Figure 14.

Primary Side Control

The primary-side control in wireless charging systems involves using a DC/AC converter to regulate the transmitter current, sometimes with power factor correction. It relies on minimal communication with the receiving side, utilizing information such as the state of charge, battery voltage, and state of health to adjust to the load, thereby reducing onboard electronics and lowering vehicle weight and cost [209]. The primary side also controls the working frequency, which must match the receiving side, and regulates the output by adjusting the switching cycles per second [212]. To achieve zero voltage switching, the operating frequency must exceed the resonance frequency of the inductive impedance, although frequency splitting and multiple peaks can occur [213].

Secondary Side Control

This control uses an active AC/DC converter and a DC/DC regulator to manage battery charging, enhancing infrastructure while adding weight and cost due to increased electrical components. Researchers in [214] proposed a model that maximizes efficiency by directly controlling the battery charging current and stabilizing the DC terminal voltage. Advances include simultaneous wireless data and power transmission, which monitors and adjusts charging based on battery health, though it is less suitable for high-frequency communications in limited-power systems [210]. A 3 kW WPT device for supercapacitor recharging uses a Buck converter with a PI controller to handle varying loads, despite performance impacts from resistance fluctuations [215]. Electronically tuning mismatched coils using saturable reactors and decoupled control systems for impedance matching further improved efficiency. Comparative studies highlight that Fuzzy control strategies stabilize parameters more quickly than PI controllers despite the latter’s smooth outputs [216].

Dual Side Control

The Rx-side control regulates the output to track the reference, while the Tx-side control minimizes inverter output power, often using phase shift adjustments. This approach reduces reactive power and operates similarly to a dual-active bridge system [208,217]. In dynamic wireless charging, individual control of both sides is preferable, although stability must be managed using separate controllers. Various voltage control techniques, such as DC voltage regulation, phase shift control, and pulse density modulation, can be applied on either side of the WPT system [207]. Adjustments in the primary side inverter input or secondary side rectifier output voltage help maintain efficiency despite misalignments [218].
Frequency control can mitigate issues by tuning out of resonance coils, allowing for stable operation across a wide coupling range. Advanced control schemes, including constant-current (CC) and constant-voltage (CV) modes, help manage variable loads, although they may face stability challenges due to frequency bifurcation [214,216]. Effective control in WPT systems requires both open-loop and closed-loop strategies to handle power flow, segmentation, and compensation. Closed-loop controls, while complex, offer improved reliability and response consistency [219]. Communication between the Tx and Rx sides is crucial, with stringent requirements for stability and bandwidth. Phase shift control, employed on both sides, adjusts the phase angles to regulate power transfer, supporting bidirectional power flow [220]. Recent advancements and high-power wireless chargers demonstrate the practical applications of these control techniques, as shown in Table 9.

5. Electric Motors

Electric motors are the heart of modern electric vehicles (EVs), being crucial in their performance, efficiency, and overall success. This chapter explores the diverse landscape of electric motor technologies shaping the future of transportation electrification. We delve into five key motor types that have emerged as significant players in the EV industry: Induction Motors (IMs), Permanent Magnet Synchronous Motors (PMSMs), Synchronous Reluctance Motors (SynRMs), Switched Reluctance Motors (SRMs), Axial Flux Motors, and Sub-Harmonic Synchronous Machines (SHSMs).
These motor technologies bring unique advantages and challenges to EV propulsion systems. Induction motors, known for their robustness and cost-effectiveness, have been a mainstay in industrial applications and continue to be used in EVs. Due to their high efficiency and power density, PMSMs have become the dominant choice for many EV manufacturers. SynRMs and SRMs offer simpler construction and the potential for high-speed operation without relying on permanent magnets. Axial flux motors are innovative with their compact design and high torque density. Finally, the emerging SHSM technology aims to combine the benefits of wound rotor machines with brushless operation, potentially overcoming some limitations of traditional PMSMs.
This chapter will examine the working principles, advantages, challenges, and recent developments in these motor technologies. By understanding the strengths and weaknesses of each motor type, we can better appreciate the complexities involved in selecting and optimizing electric motors for the diverse requirements of modern electric vehicles. As the EV market continues to evolve, these motor technologies will play a pivotal role in shaping the future of sustainable transportation.

5.1. Induction Motor (IM)

Due to their robustness, reliability, and cost-effectiveness, induction motors (IMs) have been a significant player in electric vehicle (EV) propulsion systems. These motors, also known as asynchronous motors, operate on the principle of electromagnetic induction and have been successfully implemented in various EV models [221]. The induction motor consists of a stator with poly-phase windings and a rotor with short-circuited conductors. When AC current flows through the stator windings, it creates a rotating magnetic field. This field induces currents in the rotor, producing a magnetic field that interacts with the stator field, resulting in torque production [222].
Induction motors offer several advantages for EV propulsion. They are robust, reliable, and require minimal maintenance due to the absence of brushes or permanent magnets. IMs can operate efficiently over a wide speed range, which is crucial for vehicular applications [223]. The technology is mature, well-understood, and does not rely on rare-earth materials, making it an attractive option for mass production [224]. Another significant advantage is the inherent field-weakening capability of IMs, which allows for extended constant-power operation at high speeds. This characteristic is particularly beneficial for highway driving conditions [225]. Despite their advantages, induction motors face some challenges in EV applications. They generally have lower efficiency at low speeds and partial loads than permanent magnet motors. The rotor losses in IMs can be significant, especially at high speeds, leading to thermal management issues [226]. The power factor of induction motors is typically lower than that of permanent magnet motors, which can result in higher current requirements and potentially larger inverter ratings [227].
Advanced control techniques have been developed to improve the performance of induction motors in EVs. Field-oriented control (FOC) and direct torque control (DTC) are widely used strategies that enable precise torque and flux control, improving dynamic performance and efficiency [228]. Recent developments in sensor-less control techniques have further enhanced the attractiveness of IMs for EV applications by eliminating the need for speed or position sensors, thereby increasing reliability and reducing costs [229]. Research efforts have focused on addressing the limitations of induction motors. Advanced rotor designs, such as copper and optimized slot geometries, have improved efficiency and power density [230]. An example of an induction motor used in an Audi Q8 e-tron is shown in Figure 15 [231]. Integrating wide-bandgap semiconductor devices in motor drives has enabled higher switching frequencies, improved motor performance, and reduced losses [232]. Hybrid designs combining the characteristics of induction and synchronous reluctance motors have also been explored, aiming to leverage the advantages of both motor types [233].

5.2. Permanent Magnet Synchronous Motor (PMSM)

Due to their superior performance characteristics, Permanent Magnet Synchronous Motors (PMSMs), shown in Figure 16 [234], have become the dominant choice for electric vehicle (EV) propulsion systems. These motors combine high efficiency, excellent power density, and a wide constant power speed range, making them ideal for the demanding requirements of modern EVs [235]. PMSMs operate on the principle of electromagnetic interaction between a rotating magnetic field in the stator and permanent magnets in the rotor. The stator typically consists of a three-phase winding arrangement. At the same time, the rotor contains high-strength permanent magnets, usually made from rare-earth materials such as neodymium-iron-boron (NdFeB) or samarium-cobalt (SmCo) [236]. Two main types of PMSMs are used in EVs: surface-mounted PMSMs (SPMSMs) and interior PMSMs (IPMSMs). In SPMSMs, the magnets are attached to the surface of the rotor, while in IPMSMs, the magnets are embedded within the rotor structure. IPMSMs are more common in EV applications due to their ability to achieve higher power density and better field-weakening capabilities [237,238].
PMSMs offer several advantages that make them attractive for EV applications. They can achieve efficiencies over 95% across a wide operating range, contributing to increased vehicle range [239]. Using strong permanent magnets allows for compact designs with high torque-to-weight ratios. PMSMs can operate efficiently at low and high speeds, which is crucial for EV performance. The absence of brushes and slip rings in PMSMs reduces maintenance requirements and improves reliability. Additionally, PMSMs allow for accurate torque and speed control, enhancing vehicle dynamics and drivability [240]. Despite their advantages, PMSMs face some challenges in EV applications. Relying on rare-earth materials for permanent magnets can lead to high and volatile costs. The limited global supply of rare-earth elements has raised geopolitical and sustainability concerns [241]. At high temperatures or under fault conditions, permanent magnets can lose their magnetic properties, potentially leading to motor failure. Compared to induction motors, PMSMs have a more limited constant power speed range, which may require additional gearing for high-speed operation.
To address these challenges, researchers and manufacturers are exploring several avenues. PMSMs are currently being developed using ferrite magnets or other alternative materials to reduce dependence on rare-earth elements [242]. Some researchers are investigating hybrid excitation, combining permanent magnets with electromagnetic excitation to improve field-weakening capabilities and efficiency [243]. Advanced thermal management techniques are implemented to prevent demagnetization and improve power density [244]. Additionally, novel rotor and stator designs are explored to enhance performance and reduce material usage [245]. PMSMs remain at the forefront of electric vehicle propulsion technology due to their excellent performance characteristics. As research continues to address current limitations and improve designs, PMSMs will likely maintain their dominant position in the EV market. The ongoing developments in materials, design optimization, and control strategies promise to enhance the capabilities of PMSMs further, contributing to the advancement of electric vehicle technology.

5.3. Synchronous Reluctance Motor (SynRM) and Switched Reluctance Motor (SRM)

SynRM and SRM have gained increasing attention in the electric vehicle (EV) industry due to their simple and robust construction, absence of permanent magnets, and potential for high-speed operation. These characteristics make SRMs an attractive alternative to permanent magnet and induction motors in certain EV applications [246]. SRMs operate on the principle of magnetic reluctance. The motor consists of a stator with concentrated windings and a rotor made of soft magnetic material with a salient pole structure. When the stator windings are energized, the rotor moves to align itself with the stator’s magnetic field, minimizing the magnetic reluctance path. Continuous rotational motion is achieved by sequentially energizing the stator poles [247].
SynRM and SRM offer several advantages for EV propulsion systems. They have a simple and robust construction, with no windings or permanent magnets on the rotor. This simplicity leads to lower manufacturing costs and improved reliability [248]. SynRM and SRM can operate efficiently at high speeds, which benefits EV applications requiring a wide speed range [249]. The absence of permanent magnets eliminates concerns about rare-earth material supply and demagnetization at high temperatures. This characteristic makes SynRM and SRM particularly suitable for harsh operating environments [250]. Additionally, SynRM and SRM have inherent fault tolerance due to their independent phase structure, enhancing system reliability [251]. However, some hybrid PM-assisted SRMs are used in EV applications to incorporate the advantages of both PM and SRM technologies. One example is the Internal Permanent Magnet Synchronous Reluctance Motor (IPM-SynRM) used in the Tesla Model 3, shown in Figure 17 [252].
Despite their advantages, SynRM and SRM face challenges in widespread EV adoption. One of the primary issues is acoustic noise and vibration, which result from the pulsed nature of torque production and radial forces acting on the stator [253]. Torque ripple is another significant challenge, as it can affect vehicle drivability and passenger comfort [254]. The nonlinear magnetic characteristics of SynRM and SRM make precise torque control more complex than other motor types. This complexity often necessitates sophisticated control strategies and power electronics [255]. Advanced control techniques have been developed to address the challenges associated with SynRM and SRM. Current profiling methods aim to minimize torque ripple by shaping the phase current waveforms [256]. Direct Instantaneous Torque Control (DITC) has shown promise in reducing torque ripple and improving dynamic performance [257]. Model Predictive Control (MPC) has also been applied to SynRM and SRM, demonstrating improvements in efficiency and torque ripple reduction [258]. These advanced control strategies, combined with high-performance power electronics, have significantly enhanced the viability of SynRM and SRM for EV applications.
Recent research has focused on addressing the acoustic noise and torque ripple issues of SynRM and SRM. Novel rotor and stator designs, such as segmented rotor structures and asymmetric stator poles, have shown the potential to reduce acoustic noise and improve torque characteristics [259]. The integration of wide-bandgap semiconductors, such as Silicon Carbide (SiC) devices, in SynRM and SRM drives has enabled higher switching frequencies and more precise current control, improving overall system performance [260]. Multiphase SynRM and SRM configurations, with more than three phases, have been explored to increase power density and reduce torque ripple [261]. These designs offer potential fault tolerance and power segmentation advantages for EV powertrains.

5.4. Axial Flux Motors

Axial flux motors, or axial gap, as shown in Figure 18, have gained increasing attention in the electric vehicle (EV) industry due to their unique geometry and performance characteristics. Unlike traditional radial flux motors, axial flux motors have a disc-shaped structure where the magnetic flux flows parallel to the motor’s rotational axis [262]. The stator and rotor are arranged face-to-face in an axial flux motor. The stator typically contains windings or permanent magnets, while the rotor consists of permanent magnets or a ferromagnetic core, depending on the specific design. The interaction between the magnetic fields of the stator and rotor produces torque, causing rotation [263]. There are several configurations of axial flux motors, including single-sided, double-sided, and multi-stack designs. The single-sided configuration has one stator and one rotor, while the double-sided design features a rotor sandwiched between two stators or vice versa. Multi-stack designs involve multiple rotor-stator pairs arranged in series to increase power output [264].
Axial flux motors offer several advantages that make them attractive for EV applications. They provide high torque density, as the disc-shaped design allows for a larger air gap diameter, resulting in higher torque production per unit volume compared to radial flux motors [266]. The compact axial length of these motors can be advantageous in vehicles where space is at a premium, allowing for more flexible powertrain layouts [267]. Efficient cooling is another benefit, as the disc-shaped structure provides a large surface area for heat dissipation, which can lead to improved thermal management and potentially higher power density [268]. Additionally, axial flux motors can be easily scaled by adding more stator-rotor pairs, allowing for modular designs adapted to different power requirements [269]. Despite their advantages, axial flux motors face some challenges in widespread adoption. The unique geometry of axial flux motors can make mass production more challenging and potentially more expensive than traditional radial flux motors [270]. Strong axial magnetic forces between the stator and rotor can increase mechanical stress on the motor structure, requiring careful design and robust bearings [271]. Some axial flux motor designs can exhibit significant cogging torque, affecting smooth operation at low speeds [272].
Recent research has focused on addressing the challenges associated with axial flux motors. Technological advancements, such as 3D printing of soft magnetic composites, have shown promise in simplifying production and reducing costs [273]. Novel winding configurations and magnet arrangements are being explored to minimize cogging torque and improve efficiency [274]. Recent studies have also focused on integrating axial flux motors with advanced power electronics and control systems. These developments aim to optimize motor performance across a wide operating range, making axial flux motors more competitive with traditional radial flux designs [275].

5.5. Sub-Harmonic Synchronous Machine (SHSM)

The transportation electrification sector has witnessed the rise of Permanent Magnet Synchronous Motors (PMSMs) due to their impressive torque density and efficiency despite the continued prevalence of induction machines in industrial settings [276]. Notwithstanding these merits, the high production costs and fixed flux characteristics of PMSMs have deterred manufacturers from widespread adoption [277]. This challenge has spurred research into PM-less and PM-assisted alternatives, such as wound rotor synchronous and switching reluctance machines, which offer a blend of design simplicity, robustness, efficiency, and power density [278]. Conventional wound rotor synchronous machines, hampered by the drawbacks of brushes and slip rings for DC excitation, have historically been unsuitable for traction applications. The brushless wound rotor synchronous machine (BL-WRSM) topology has been established, which is a promising solution that enables rotor excitation adjustment while maintaining high performance in traction scenarios. Among the various brushless excitation methods, harmonic excitation is elegant, utilizing the synchronous machine’s inherent properties without requiring additional bulky components [279].
Harmonic excitation techniques encompass higher- and lower-order (sub-harmonic) magnetomotive forces (MMF). Studies have revealed the superiority of the sub-harmonic MMF (SH-MMF) approach, which boasts enhanced performance and reduced losses compared to its higher-order counterpart [280]. Consequently, BL-WRSMs employing SH-MMF excitation have emerged as strong contenders for traction applications, offering a trifecta of tunable excitation, high power density, and efficient operation. The evolution of sub-harmonic synchronous machines (SHSMs) began with their initial introduction [281], followed by the development of a single inverter SHSM [282] and its subsequent design optimization [283]. Innovation continued with the creation of novel stator winding configurations, including a 2-layer design for a new SHSM topology [284] and a 3-layer configuration for another SHSM variant [285], as shown in Figure 19. Further advancements saw 2L-SHSM and 3L-SHSM designs adapted for dual inverter configurations [286], enhancing their controllability.
In the quest for even greater torque generation and power density, researchers introduced PM-assisted hybrid SHSM topologies [287,288]. The success of these hybrid designs, particularly the 2L-SHSM, paved the way for exploration into hybrid 3L-SHSM configurations. Recent studies have proposed promising PM-assisted 3L-SHSM designs [289,290], further expanding the horizons of electric machine performance. This ongoing research in SHSM technology exemplifies the relentless pursuit of overcoming PMSM limitations while preserving their advantages, potentially revolutionizing electric motor applications across various industries.
This section has examined various electric motor technologies used in electric vehicles, including Induction Motors, Permanent Magnet Synchronous Motors, Synchronous and Switched Reluctance Motors, Axial Flux Motors, and Sub-Harmonic Synchronous Machines. The comparison Table 10 provided summarizes each motor type’s key characteristics, advantages, and challenges.
This overview demonstrates that no single motor technology is universally superior for all EV applications. Each type offers unique benefits and trade-offs regarding performance, cost, manufacturability, and operational characteristics. The choice of motor depends on specific vehicle requirements, manufacturing considerations, and technological advancements. As the EV industry evolves, we expect continued innovation in these motor technologies, addressing current limitations and potentially leading to hybrid design types. The ongoing development of electric motors will play a crucial role in enhancing future electric vehicles’ performance, efficiency, and affordability.

5.6. Requirements of Motor Performance for Electric Vehicles

Manufacturers must consider various performance requirements when designing or selecting electric motors for EVs to ensure optimal vehicle operation. These requirements are influenced by factors such as vehicle type, intended use, and market segment [291,292]. Table 11 summarizes key performance requirements for electric motors in EV applications.
Power and torque density are crucial for minimizing the motor’s weight and size, directly impacting vehicle performance and range [227,235]. High efficiency across a wide operating range is essential for maximizing the vehicle’s energy utilization and range [236]. A wide and constant power speed range is important for vehicle drivability and performance in various driving conditions [227,235]. Low torque ripple and noise levels improve passenger comfort and overall vehicle refinement [293]. Overload capability is necessary for handling short-term high-power demands, such as during acceleration or hill climbing. Effective thermal management is critical for maintaining performance and ensuring longevity [294].
Reliability is paramount in automotive applications, with motors expected to last the vehicle’s lifetime. Finally, cost considerations are vital for mass-market adoption of EVs [291,293]. These requirements can vary based on the specific vehicle application (e.g., passenger car, commercial vehicle, and high-performance sports car) and may evolve as EV technology advances [291]. Motor designers and vehicle manufacturers must balance these often-competing requirements to create optimal solutions for their target markets [292,294]. Ongoing research and development in electric motor technology continue to push the boundaries of performance, aiming to meet and exceed these requirements [292]. As the EV market matures, we can expect further refinement and potential shifts in these performance criteria to meet evolving consumer expectations and regulatory standards.

6. Standards and Regulations

The standards and regulations for EV charging infrastructure, batteries, and motors are crucial in ensuring safety, compatibility, and efficiency in the electric vehicle ecosystem. Standard organizations like IEC, IEEE, ISO, SAE, ANSI, BIS, and BSI provide comprehensive guidelines covering design, construction, communication, and safety requirements for EV charging stations [295]. These standards are essential for both conductive and inductive charging methods, emphasizing safety and reliability, especially with the increasing use of high-voltage lithium-ion batteries in electric vehicles vehicles [296]. Additionally, the development of international standards for electric vehicles focuses on the importance of efficient energy storage devices like traction batteries and the necessity of universal access to charging infrastructures to promote the widespread adoption of electric vehicles [297]. Furthermore, wireless power transfer (WPT) regulations for EV charging are being established to enhance charging speed and safety, particularly in adverse weather conditions, highlighting the significance of electromagnetic compatibility and human exposure limits [298]. Table 12, Table 13, Table 14 and Table 15 show the related standards included in electric vehicle charging [19,299,300,301].

7. Conclusions

The increasing global reliance on electric vehicles (EVs) is a response to the excessive use of fossil fuels and rising CO2 emissions. EVs enable the integration of alternative energy sources, such as energy storage systems (ESSs) and renewable energy sources (RESs), fostering sustainable mobility and reducing fossil fuel dependency. Despite their benefits, EVs face numerous challenges related to traction systems, storage capacity, charging infrastructure, and energy management technologies. Moreover, the expansion of the EV market requires comprehensive standards and regulations. This review paper, titled “State-of-the-art of Electric Vehicles: Architectures, Control, and Regulations”, provides an in-depth analysis of current advancements in EV components, including energy storage systems, electric motors, and charging technologies. It also explores various control techniques and the prevailing standards and regulations that ensure market harmonization. By synthesizing the latest developments and regulatory frameworks, the paper aims to advance knowledge in the field and support EV technologies’ continued evolution and implementation, ultimately contributing to a more sustainable and efficient transportation future.

Author Contributions

Conceptualization, H.M.H. and S.M.S.H.R.; methodology, H.M.H., A.M.I., R.A.T. and S.M.S.H.R.; validation, H.M.H., A.M.I., R.A.T. and S.M.S.H.R.; investigation, H.M.H., A.M.I., R.A.T., M.S.A. and I.K.; writing—original draft preparation, H.M.H., A.M.I., R.A.T., S.M.S.H.R., M.S.A. and I.K.; writing—review and editing, H.M.H., R.A.T. and O.A.M.; visualization, H.M.H.; supervision, O.A.M.; project administration, O.A.M.; funding acquisition, O.A.M. All authors have read and agreed to the published version of the manuscript.

Funding

The work in this article was partially supported by a grant from the National Science Foundation.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Oladunni, O.J.; Mpofu, K.; Olanrewaju, O.A. Greenhouse gas emissions and its driving forces in the transport sector of South Africa. Energy Rep. 2022, 8, 2052–2061. [Google Scholar] [CrossRef]
  2. Bleviss, D.L. Transportation is critical to reducing greenhouse gas emissions in the United States. WIREs Energy Environ. 2021, 10, e390. [Google Scholar] [CrossRef]
  3. Feng, T.-T.; Yang, Y.-S.; Xie, S.-Y.; Dong, J.; Ding, L. Economic drivers of greenhouse gas emissions in China. Renew. Sustain. Energy Rev. 2017, 78, 996–1006. [Google Scholar] [CrossRef]
  4. Alcántara, V.; Padilla, E. Key sectors in greenhouse gas emissions in Spain: An alternative input–output analysis. J. Ind. Ecol. 2020, 24, 577–588. [Google Scholar] [CrossRef]
  5. Inventory of, U.S. Greenhouse Gas Emissions and Sinks: 1990–2022|US EPA. Available online: https://www.epa.gov/ghgemissions/inventory-us-greenhouse-gas-emissions-and-sinks-1990-2022 (accessed on 29 July 2024).
  6. Rajper, S.Z.; Albrecht, J. Prospects of Electric Vehicles in the Developing Countries: A Literature Review. Sustainability 2020, 12, 1906. [Google Scholar] [CrossRef]
  7. Langbroek, J.H.M.; Franklin, J.P.; Susilo, Y.O. The Effect of Policy Incentives on Electric Vehicle Adoption. Energy Policy 2016, 94, 94–103. [Google Scholar] [CrossRef]
  8. Agrawal, M.; Rajapatel, M.S. Global perspective on electric vehicle 2020. Int. J. Eng. Res. Technol. 2020, 9, 8–11. [Google Scholar] [CrossRef]
  9. Albatayneh, A.; Assaf, M.N.; Alterman, D.; Jaradat, M. Comparison of the Overall Energy Efficiency for Internal Combustion Engine Vehicles and Electric Vehicles. Sci. J. Riga Tech. Univ. Environ. Clim. Technol. 2020, 24, 669–680. [Google Scholar] [CrossRef]
  10. Muratori, M.; Alexander, M.; Arent, D.; Bazilian, M.; Cazzola, P.; Dede, E.M.; Farrell, J.; Gearhart, C.; Greene, D.; Jenn, A.; et al. The rise of electric vehicles—2020 status and future expectations. Prog. Energy 2021, 3, 022002. [Google Scholar] [CrossRef]
  11. Global EV Outlook 2019—Analysis—IEA. Available online: https://www.iea.org/reports/global-ev-outlook-2019 (accessed on 29 July 2024).
  12. Global EV Outlook 2017—Analysis—IEA. Available online: https://www.iea.org/reports/global-ev-outlook-2017 (accessed on 29 July 2024).
  13. Quarterly Electric Car Sales by Region, 2021–2024—Charts—Data & Statistics—IEA. Available online: https://www.iea.org/data-and-statistics/charts/quarterly-electric-car-sales-by-region-2021-2024 (accessed on 29 July 2024).
  14. Hasan, M.K.; Mahmud; Habib, A.A.; Motakabber, S.; Islam, S. Review of electric vehicle energy storage and management system: Standards, issues, and challenges. J. Energy Storage 2021, 41, 102940. [Google Scholar] [CrossRef]
  15. Ntombela, M.; Musasa, K.; Moloi, K. A Comprehensive Review for Battery Electric Vehicles (BEV) Drive Circuits Technology, Operations, and Challenges. World Electr. Veh. J. 2023, 14, 195. [Google Scholar] [CrossRef]
  16. Ehsani, M.; Singh, K.V.; Bansal, H.O.; Mehrjardi, R.T. State of the Art and Trends in Electric and Hybrid Electric Vehicles. Proc. IEEE 2021, 109, 967–984. [Google Scholar] [CrossRef]
  17. Sanguesa, J.A.; Torres-Sanz, V.; Garrido, P.; Martinez, F.J.; Marquez-Barja, J.M. A Review on Electric Vehicles: Technologies and Challenges. Smart Cities 2021, 4, 372–404. [Google Scholar] [CrossRef]
  18. Singh, K.V.; Bansal, H.O.; Singh, D. A comprehensive review on hybrid electric vehicles: Architectures and components. J. Mod. Transp. 2019, 27, 77–107. [Google Scholar] [CrossRef]
  19. Acharige, S.S.G.; Haque, E.; Arif, M.T.; Hosseinzadeh, N.; Hasan, K.N.; Oo, A.M.T. Review of Electric Vehicle Charging Technologies, Standards, Architectures, and Converter Configurations. IEEE Access 2023, 11, 41218–41255. [Google Scholar] [CrossRef]
  20. Ghosh, A. Possibilities and Challenges for the Inclusion of the Electric Vehicle (EV) to Reduce the Carbon Footprint in the Transport Sector: A Review. Energies 2020, 13, 2602. [Google Scholar] [CrossRef]
  21. Thangavel, S.; Mohanraj, D.; Girijaprasanna, T.; Raju, S.; Dhanamjayulu, C.; Muyeen, S.M. A Comprehensive Review on Electric Vehicle: Battery Management System, Charging Station, Traction Motors. IEEE Access 2023, 11, 20994–21019. [Google Scholar] [CrossRef]
  22. Sathiyan, S.P.; Pratap, C.B.; Stonier, A.A.; Peter, G.; Sherine, A.; Praghash, K.; Ganji, V. Comprehensive Assessment of Electric Vehicle Development, Deployment, and Policy Initiatives to Reduce GHG Emissions: Opportunities and Challenges. IEEE Access 2022, 10, 53614–53639. [Google Scholar] [CrossRef]
  23. Stajić, D.; Pfeifer, A.; Herc, L.; Logonder, M. Early Adoption of Battery Electric Vehicles and Owners’ Motivation. Clean. Eng. Technol. 2023, 15, 100658. [Google Scholar] [CrossRef]
  24. Koroma, M.S.; Costa, D.; Philippot, M.; Cardellini, G.; Hosen, M.S.; Coosemans, T.; Messagie, M. Life Cycle Assessment of Battery Electric Vehicles: Implications of Future Electricity Mix and Different Battery End-of-Life Management. Sci. Total Environ. 2022, 831, 154859. [Google Scholar] [CrossRef]
  25. Liu, X.; Zhao, F.; Geng, J.; Hao, H.; Liu, Z. Comprehensive Assessment for Different Ranges of Battery Electric Vehicles: Is It Necessary to Develop an Ultra-Long Range Battery Electric Vehicle? iScience 2023, 26, 106654. [Google Scholar] [CrossRef] [PubMed]
  26. Björnsson, L.H.; Karlsson, S. Electrification of the Two-Car Household: PHEV or BEV? Transp. Res. Part C Emerg. Technol. 2017, 85, 363–376. [Google Scholar] [CrossRef]
  27. Wang, Y.; Biswas, A.; Rodriguez, R.; Keshavarz-Motamed, Z.; Emadi, A. Hybrid Electric Vehicle Specific Engines: State-of-the-Art Review. Energy Rep. 2022, 8, 832–851. [Google Scholar] [CrossRef]
  28. Tran, D.D.; Vafaeipour, M.; El Baghdadi, M.; Barrero, R.; Van Mierlo, J.; Hegazy, O. Thorough State-of-the-Art Analysis of Electric and Hybrid Vehicle Powertrains: Topologies and Integrated Energy Management Strategies. Renew. Sustain. Energy Rev. 2020, 119, 109596. [Google Scholar] [CrossRef]
  29. Yong, J.Y.; Ramachandaramurthy, V.K.; Tan, K.M.; Mithulananthan, N. A Review on the State-of-the-Art Technologies of Electric Vehicle, Its Impacts and Prospects. Renew. Sustain. Energy Rev. 2015, 49, 365–385. [Google Scholar] [CrossRef]
  30. Fantin Irudaya Raj, E.; Appadurai, M. The Hybrid Electric Vehicle (HEV)—An Overview. In Emerging Solutions for e-Mobility and Smart Grids; Springer: Singapore, 2021; pp. 25–36. [Google Scholar] [CrossRef]
  31. Shen, C.; Shan, P.; Gao, T. A Comprehensive Overview of Hybrid Electric Vehicles. Int. J. Veh. Technol. 2011, 2011, 571683. [Google Scholar] [CrossRef]
  32. Yang, C.; Zha, M.; Wang, W.; Liu, K.; Xiang, C. Efficient Energy Management Strategy for Hybrid Electric Vehicles/Plug-in Hybrid Electric Vehicles: Review and Recent Advances under Intelligent Transportation System. IET Intell. Transp. Syst. 2020, 14, 702–711. [Google Scholar] [CrossRef]
  33. Sabri, M.; Danapalasingam, K.A.; Rahmat, M.F. A review on hybrid electric vehicles architecture and energy management strategies. Renew. Sustain. Energy Rev. 2016, 53, 1433–1442. [Google Scholar] [CrossRef]
  34. León, R.; Montaleza, C.; Maldonado, J.L.; Tostado-Véliz, M.; Jurado, F. Hybrid Electric Vehicles: A Review of Existing Configurations and Thermodynamic Cycles. Thermo 2021, 1, 134–150. [Google Scholar] [CrossRef]
  35. Selmi, T.; Khadhraoui, A.; Cherif, A. Fuel Cell–Based Electric Vehicles Technologies and Challenges. Environ. Sci. Pollut. Res. 2022, 29, 78121–78131. [Google Scholar] [CrossRef]
  36. Muthukumar, M.; Rengarajan, N.; Velliyangiri, B.; Omprakas, M.A.; Rohit, C.B.; Raja, U.K. The Development of Fuel Cell Electric Vehicles—A Review. Mater Today Proc. 2021, 45, 1181–1187. [Google Scholar] [CrossRef]
  37. Waseem, M.; Amir, M.; Lakshmi, G.S.; Harivardhagini, S.; Ahmad, M. Fuel Cell-Based Hybrid Electric Vehicles: An Integrated Review of Current Status, Key Challenges, Recommended Policies, and Future Prospects. Green Energy Intell. Transp. 2023, 2, 100121. [Google Scholar] [CrossRef]
  38. Oldenbroek, V.; Wijtzes, S.; Blok, K.; van Wijk, A.J.M. Fuel Cell Electric Vehicles and Hydrogen Balancing 100 Percent Renewable and Integrated National Transportation and Energy Systems. Energy Convers. Manag. X 2021, 9, 100077. [Google Scholar] [CrossRef]
  39. Wu, D.; Ren, J.; Davies, H.; Shang, J.; Haas, O. Intelligent Hydrogen Fuel Cell Range Extender for Battery Electric Vehicles. World Electr. Veh. J. 2019, 10, 29. [Google Scholar] [CrossRef]
  40. Tsang, F.; Potoglou, D.; Wooding, S.; Pedersen, J.S. Bringing the Electric Vehicle to the Mass Market: A Review of Barriers, Facilitators and Policy Interventions; RAND Corporation: Santa Monica, CA, USA, 2012. [Google Scholar]
  41. Sufyan Arshad Ayesha Ashraf, M. Hybrid Electric Vehicles: A General Overview. Int. J. Eng. Appl. Sci. Technol. 2020, 5, 21–26. [Google Scholar]
  42. Veza, I.; Asy’ari, M.Z.; Idris, M.; Epin, V.; Rizwanul Fattah, I.M.; Spraggon, M. Electric Vehicle (EV) and Driving towards Sustainability: Comparison between EV, HEV, PHEV, and ICE Vehicles to Achieve Net Zero Emissions by 2050 from EV. Alex. Eng. J. 2023, 82, 459–467. [Google Scholar] [CrossRef]
  43. Çaǧatay Bayindir, K.; Gözüküçük, M.A.; Teke, A. A Comprehensive Overview of Hybrid Electric Vehicle: Powertrain Configurations, Powertrain Control Techniques and Electronic Control Units. Energy Convers. Manag. 2011, 52, 1305–1313. [Google Scholar] [CrossRef]
  44. Ma, Z.; Luan, Y.; Zhang, F.; Xie, S.; Coskun, S. A Data-Driven Energy Management Strategy for Plug-in Hybrid Electric Buses Considering Vehicle Mass Uncertainty. J. Energy Storage 2024, 77, 109963. [Google Scholar] [CrossRef]
  45. Gong, W.; Liu, C.; Zhao, X.; Xu, S. A Model Review for Controller-Hardware-in-the- Loop Simulation in EV Powertrain Application. IEEE Trans. Transp. Electrif. 2024, 10, 925–937. [Google Scholar] [CrossRef]
  46. Xu, C.; Behrens, P.; Gasper, P.; Smith, K.; Hu, M.; Tukker, A.; Steubing, B. Electric vehicle batteries alone could satisfy short-term grid storage demand by as early as 2030. Nat. Commun. 2023, 14, 119. [Google Scholar] [CrossRef]
  47. Koohi-Fayegh, S.; Rosen, M.A. A Review of Energy Storage Types, Applications and Recent Developments. J. Energy Storage 2020, 27, 101047. [Google Scholar] [CrossRef]
  48. Hussein, H.M.; Abdelrahman, M.S.; Rafin, S.M.S.H.; Kharchouf, I.; Mohammed, O. Enhancing Stability of Hybrid AC/DC Microgrid-Based Battery/Ultracapacitor Energy Storage System. In Proceedings of the 2024 IEEE International Conference on Environment and Electrical Engineering and 2024 IEEE Industrial and Commercial Power Systems Europe (EEEIC/I&CPS Europe), Rome, Italy, 17–20 June 2024. [Google Scholar]
  49. Kumar, D.; Nema, R.K.; Gupta, S. A Comparative Review on Power Conversion Topologies and Energy Storage System for Electric Vehicles. Int. J. Energy Res. 2020, 44, 7863–7885. [Google Scholar] [CrossRef]
  50. Gao, Z.; Xie, H.; Yang, X.; Niu, W.; Li, S.; Chen, S. The Dilemma of C-Rate and Cycle Life for Lithium-Ion Batteries under Low Temperature Fast Charging. Batteries 2022, 8, 234. [Google Scholar] [CrossRef]
  51. Ralls, A.M.; Leong, K.; Clayton, J.; Fuelling, P.; Mercer, C.; Navarro, V.; Menezes, P.L. The Role of Lithium-Ion Batteries in the Growing Trend of Electric Vehicles. Materials 2023, 16, 6063. [Google Scholar] [CrossRef]
  52. Chen, W.; Liang, J.; Yang, Z.; Li, G. A Review of Lithium-Ion Battery for Electric Vehicle Applications and Beyond. Energy Procedia 2019, 158, 4363–4368. [Google Scholar] [CrossRef]
  53. Notter, D.A.; Gauch, M.; Widmer, R.; Wäger, P.; Stamp, A.; Zah, R.; Althaus, H.J. Contribution of Li-Ion Batteries to the Environmental Impact of Electric Vehicles. Environ. Sci. Technol. 2010, 44, 6550–6556. [Google Scholar] [CrossRef]
  54. Bisschop, R.; Willstrand, O.; Amon, F.; Rosenggren, M. Fire Safety of Lithium-Ion Batteries in Road Vehicles; RISE Report 2019; 2019; Volume 50, p. 105. Available online: https://urn.kb.se/resolve?urn=urn:nbn:se:ri:diva-38873 (accessed on 30 July 2024).
  55. Chen, Y.; Kang, Y.; Zhao, Y.; Wang, L.; Liu, J.; Li, Y.; Liang, Z.; He, X.; Li, X.; Tavajohi, N.; et al. A Review of Lithium-Ion Battery Safety Concerns: The Issues, Strategies, and Testing Standards. J. Energy Chem. 2021, 59, 83–99. [Google Scholar] [CrossRef]
  56. Diaz, L.B.; He, X.; Hu, Z.; Restuccia, F.; Marinescu, M.; Barreras, J.V.; Patel, Y.; Offer, G.J.; Rein, G. Review—Meta-Review of Fire Safety of Lithium-Ion Batteries: Industry Challenges and Research Contributions. J. Electrochem. Soc. 2020, 167, 090559. [Google Scholar] [CrossRef]
  57. Chen, Y.; Wang, T.; Tian, H.; Su, D.; Zhang, Q.; Wang, G.; Chen, Y.; Wang, T.; Tian, H.; Su, D.; et al. Advances in Lithium–Sulfur Batteries: From Academic Research to Commercial Viability. Adv. Mater. 2021, 33, 2003666. [Google Scholar] [CrossRef]
  58. Manthiram, A. An Outlook on Lithium Ion Battery Technology. ACS Cent Sci. 2017, 3, 1063–1069. [Google Scholar] [CrossRef]
  59. Ding, Y.; Cano, Z.P.; Yu, A.; Lu, J.; Chen, Z. Automotive Li-Ion Batteries: Current Status and Future Perspectives. Electrochem. Energy Rev. 2019, 2, 1–28. [Google Scholar] [CrossRef]
  60. Hochschulschriften/Increasing Electric Vehicle Range with a Recommendation App Providing ContextSpecific Trip Rankings. Available online: https://epub.jku.at/urn:nbn:at:at-ubl:1-12684 (accessed on 30 July 2024).
  61. Leone, C.; Sturaro, L.; Geroli, G.; Longo, M.; Yaici, W. Design and Implementation of an Electric Skibus Line in North Italy. Energies 2021, 14, 7925. [Google Scholar] [CrossRef]
  62. Sutopo, W.; Maryanie, D.I.; Purwanto, A.; Nizam, M. A Comparative Value Chains Analysis of Battery Technologies for Electric Vehicles. In Proceedings of the 2013 Joint International Conference on Rural Information and Communication Technology and Electric-Vehicle Technology, rICT and ICEV-T 2013, Bandung, Indonesia, 26–28 November 2013. [Google Scholar] [CrossRef]
  63. Windisch-Kern, S.; Gerold, E.; Nigl, T.; Jandric, A.; Altendorfer, M.; Rutrecht, B.; Scherhaufer, S.; Raupenstrauch, H.; Pomberger, R.; Antrekowitsch, H.; et al. Recycling Chains for Lithium-Ion Batteries: A Critical Examination of Current Challenges, Opportunities and Process Dependencies. Waste Manag. 2022, 138, 125–139. [Google Scholar] [CrossRef]
  64. Wu, J.; Dathar, G.K.P.; Sun, C.; Theivanayagam, M.G.; Applestone, D.; Dylla, A.G.; Manthiram, A.; Henkelman, G.; Goodenough, J.B.; Stevenson, K.J. In Situ Raman Spectroscopy of LiFePO4: Size and Morphology Dependence during Charge and Self-Discharge. Nanotechnology 2013, 24, 424009. [Google Scholar] [CrossRef] [PubMed]
  65. Andre, D.; Kim, S.J.; Lamp, P.; Lux, S.F.; Maglia, F.; Paschos, O.; Stiaszny, B. Future Generations of Cathode Materials: An Automotive Industry Perspective. J. Mater. Chem. A Mater. 2015, 3, 6709–6732. [Google Scholar] [CrossRef]
  66. Baczyńska, A.; Niewiadomski, W.; Gonçalves, A.; Almeida, P.; Luís, R. Li-NMC Batteries Model Evaluation with Experimental Data for Electric Vehicle Application. Batteries 2018, 4, 11. [Google Scholar] [CrossRef]
  67. Hammou, A.; Petrone, R.; Diallo, D.; Gualous, H. Estimating the Health Status of Li-Ion NMC Batteries from Energy Characteristics for EV Applications. IEEE Trans. Energy Convers. 2023, 38, 2160–2168. [Google Scholar] [CrossRef]
  68. Liu, X.; Li, K.; Li, X. The Electrochemical Performance and Applications of Several Popular Lithium-Ion Batteries for Electric Vehicles—A Review. Commun. Comput. Inf. Sci. 2018, 925, 201–213. [Google Scholar]
  69. de Castro, R.H.; Romano Espinosa, D.C.; Gobo, L.A.; Kumoto, E.A.; Botelho Junior, A.B.; Tenorio, J.A.S. Design of Recycling Processes for NCA-Type Li-Ion Batteries from Electric Vehicles toward the Circular Economy. Energy Fuels 2024, 38, 5545–5557. [Google Scholar] [CrossRef]
  70. Tran, M.K.; Dacosta, A.; Mevawalla, A.; Panchal, S.; Fowler, M. Comparative Study of Equivalent Circuit Models Performance in Four Common Lithium-Ion Batteries: LFP, NMC, LMO, NCA. Batteries 2021, 7, 51. [Google Scholar] [CrossRef]
  71. Han, X.; Ouyang, M.; Lu, L.; Li, J. Cycle Life of Commercial Lithium-Ion Batteries with Lithium Titanium Oxide Anodes in Electric Vehicles. Energies 2014, 7, 4895–4909. [Google Scholar] [CrossRef]
  72. Al-saadi, M.; Olmos, J.; Saez-de-ibarra, A.; Van Mierlo, J.; Berecibar, M. Fast Charging Impact on the Lithium-Ion Batteries’ Lifetime and Cost-Effective Battery Sizing in Heavy-Duty Electric Vehicles Applications. Energies 2022, 15, 1278. [Google Scholar] [CrossRef]
  73. Guittet, D.; Gasper, P.; Shirk, M.; Mitchell, M.; Gilleran, M.; Bonnema, E.; Smith, K.; Mishra, P.; Mann, M. Levelized Cost of Charging of Extreme Fast Charging with Stationary LMO/LTO Batteries. J. Energy Storage 2024, 82, 110568. [Google Scholar] [CrossRef]
  74. Arya, S.; Verma, S. Nickel-Metal Hydride (Ni-MH) Batteries. Recharg. Batter. 2020, 131–175. [Google Scholar] [CrossRef]
  75. Naim, A.; Vashist, D. Technological Advancements for Reduced Charging Time of Electric Vehicle Batteries: A Review. Smart Struct. Energy Infrastruct. Proc. ICRTE 2022, 2, 99–112. [Google Scholar] [CrossRef]
  76. Arun, V.; Kannan, R.; Ramesh, S.; Vijayakumar, M.; Raghavendran, P.S.; Siva Ramkumar, M.; Anbarasu, P.; Sundramurthy, V.P. Review on Li-Ion Battery vs Nickel Metal Hydride Battery in EV. Adv. Mater. Sci. Eng. 2022, 2022, 7910072. [Google Scholar] [CrossRef]
  77. Fetcenko, M.A.; Ovshinsky, S.R.; Reichman, B.; Young, K.; Fierro, C.; Koch, J.; Zallen, A.; Mays, W.; Ouchi, T. Recent Advances in NiMH Battery Technology. J. Power Sources 2007, 165, 544–551. [Google Scholar] [CrossRef]
  78. Kang, J.; Yan, F.; Zhang, P.; Du, C. Comparison of Comprehensive Properties of Ni-MH (Nickel-Metal Hydride) and Li-Ion (Lithium-Ion) Batteries in Terms of Energy Efficiency. Energy 2014, 70, 618–625. [Google Scholar] [CrossRef]
  79. Aditya, J.P.; Ferdowsi, M. Comparison of NiMH and Li-Ion Batteries in Automotive Applications. In Proceedings of the 2008 IEEE Vehicle Power and Propulsion Conference, VPPC 2008, Harbin, China, 3–5 September 2008. [Google Scholar] [CrossRef]
  80. Chang, S.; Young, K.H.; Lien, Y.L. Reviews of European Patents on Nickel/Metal Hydride Batteries. Batteries 2017, 3, 25. [Google Scholar] [CrossRef]
  81. Zhu, W.H.; Zhu, Y.; Tatarchuk, B.J. Self-Discharge Characteristics and Performance Degradation of Ni-MH Batteries for Storage Applications. Int. J. Hydrogen Energy 2014, 39, 19789–19798. [Google Scholar] [CrossRef]
  82. Young, K.H.; Yasuoka, S. Capacity Degradation Mechanisms in Nickel/Metal Hydride Batteries. Batteries 2016, 2, 3. [Google Scholar] [CrossRef]
  83. Ouyang, L.; Huang, J.; Wang, H.; Liu, J.; Zhu, M. Progress of Hydrogen Storage Alloys for Ni-MH Rechargeable Power Batteries in Electric Vehicles: A Review. Mater. Chem. Phys. 2017, 200, 164–178. [Google Scholar] [CrossRef]
  84. Mohammadi, F.; Saif, M. A Comprehensive Overview of Electric Vehicle Batteries Market. e-Prime Adv. Electr. Eng. Electron. Energy 2023, 3, 100127. [Google Scholar] [CrossRef]
  85. Li, G.; Lu, X.; Kim, J.Y.; Viswanathan, V.V.; Meinhardt, K.D.; Engelhard, M.H.; Sprenkle, V.L. An Advanced Na–FeCl2 ZEBRA Battery for Stationary Energy Storage Application. Adv. Energy Mater. 2015, 5, 1500357. [Google Scholar] [CrossRef]
  86. Sudworth, J.L. Zebra Batteries. J. Power Sources 1994, 51, 105–114. [Google Scholar] [CrossRef]
  87. Dustmann, C.H. Advances in ZEBRA Batteries. J. Power Sources 2004, 127, 85–92. [Google Scholar] [CrossRef]
  88. Sudworth, J.L. The Sodium/Nickel Chloride (ZEBRA) Battery. J. Power Sources 2001, 100, 149–163. [Google Scholar] [CrossRef]
  89. Shamim, N.; Thomsen, E.C.; Viswanathan, V.V.; Reed, D.M.; Sprenkle, V.L.; Li, G. Evaluating ZEBRA Battery Module under the Peak-Shaving Duty Cycles. Materials 2021, 14, 2280. [Google Scholar] [CrossRef]
  90. Leonardi, S.G.; Samperi, M.; Frusteri, L.; Antonucci, V.; D’Urso, C. A Review of Sodium-Metal Chloride Batteries: Materials and Cell Design. Batteries 2023, 9, 524. [Google Scholar] [CrossRef]
  91. Hu, Y.; Wu, X.; Wen, Z.; Hou, M.; Yi, B. Challenges and Thoughts on the Development of Sodium Battery Technology for Energy Storage. Strateg. Study Chin. Acad. Eng. 2021, 23, 94–102. [Google Scholar] [CrossRef]
  92. Nikolic, M.; Schelte, N.; Velenderic, M.; Adjei, F.; Severengiz, S. Life Cycle Assessment of Sodium-Nickel-Chloride Batteries. In Proceedings of the International Renewable Energy Storage Conference (IRES 2022), Düsseldorf, Germany, 20–22 September 2023; pp. 336–362. [Google Scholar]
  93. Dixon, J.; Nakashima, I.; Arcos, E.F.; Ortúzar, M. Electric Vehicle Using a Combination of Ultracapacitors and ZEBRA Battery. IEEE Trans. Ind. Electron. 2010, 57, 943–949. [Google Scholar] [CrossRef]
  94. Capasso, C.; Veneri, O. Integration between Super-Capacitors and ZEBRA Batteries as High Performance Hybrid Storage System for Electric Vehicles. Energy Procedia 2017, 105, 2539–2544. [Google Scholar] [CrossRef]
  95. Moradi, Z.; Lanjan, A.; Tyagi, R.; Srinivasan, S. Review on Current State, Challenges, and Potential Solutions in Solid-State Batteries Research. J. Energy Storage 2023, 73, 109048. [Google Scholar] [CrossRef]
  96. Chen, M.; Winiarz, P.; Luo, J.; Qi, K.; Thomas, F.; Mahdi, L.; Lemaire, J.; Santos, D.M.F. Technological Advances and Market Developments of Solid-State Batteries: A Review. Materials 2024, 17, 239. [Google Scholar] [CrossRef] [PubMed]
  97. Li, S.; Jiang, M.; Xie, Y.; Xu, H.; Jia, J.; Li, J.; Li, S.; Jiang, M.; Xie, Y.; Xu, H.; et al. Developing High-Performance Lithium Metal Anode in Liquid Electrolytes: Challenges and Progress. Adv. Mater. 2018, 30, 1706375. [Google Scholar] [CrossRef]
  98. Machín, A.; Morant, C.; Márquez, F. Advancements and Challenges in Solid-State Battery Technology: An In-Depth Review of Solid Electrolytes and Anode Innovations. Batteries 2024, 10, 29. [Google Scholar] [CrossRef]
  99. Zheng, Y.; Yao, Y.; Ou, J.; Li, M.; Luo, D.; Dou, H.; Li, Z.; Amine, K.; Yu, A.; Chen, Z. A Review of Composite Solid-State Electrolytes for Lithium Batteries: Fundamentals, Key Materials and Advanced Structures. Chem. Soc. Rev. 2020, 49, 8790–8839. [Google Scholar] [CrossRef]
  100. Yang, T.; Wang, C.; Zhang, W.; Xia, Y.; Huang, H.; Gan, Y.; He, X.; Xia, X.; Tao, X.; Zhang, J. A Critical Review on Composite Solid Electrolytes for Lithium Batteries: Design Strategies and Interface Engineering. J. Energy Chem. 2023, 84, 189–209. [Google Scholar] [CrossRef]
  101. Pavlov, D. Lead-Acid Batteries: Science and Technology; Elsevier: Amsterdam, The Netherlands, 2011; pp. 1–643. [Google Scholar] [CrossRef]
  102. Garche, J.; Moseley, P.T.; Karden, E. Lead–acid batteries for hybrid electric vehicles and battery electric vehicles. In Advances in Battery Technologies for Electric Vehicles; Woodhead Publishing: Sawston, UK, 2015; pp. 75–101. [Google Scholar] [CrossRef]
  103. May, G.; Davidson, A.; Monahov, B. Lead batteries for utility energy storage: A review. J. Energy Storage 2018, 15, 145–157. Available online: https://www.sciencedirect.com/science/article/pii/S2352152X17304437 (accessed on 30 July 2024). [CrossRef]
  104. Lopes, P.P.; Stamenkovic, V.R. Past, Present, and Future of Lead–Acid Batteries. Science 2020, 369, 923–924. [Google Scholar] [CrossRef]
  105. Chen, H. Electrolytes of lead-acid batteries. In Lead-Acid Battery Technologies: Fundamentals, Materials, and Applications; CRC Press: Boca Raton, FL, USA, 2015; pp. 137–162. [Google Scholar] [CrossRef]
  106. Gelbke, M.; Mondoloni, C. Flooded starting-lighting-ignition (SLI) and enhanced flooded batteries (EFBs): State-of-the-art’. In Lead-Acid Batteries for Future Automobiles; Elsevier: Amsterdam, The Netherlands, 2017; pp. 149–184. [Google Scholar] [CrossRef]
  107. McKeon, B.B.; Furukawa, J.; Fenstermacher, S. Advanced Lead-Acid Batteries and the Development of Grid-Scale Energy Storage Systems. Proc. IEEE 2014, 102, 951–963. [Google Scholar] [CrossRef]
  108. Vangapally, N.; Penki, T.R.; Elias, Y.; Muduli, S.; Maddukuri, S.; Luski, S.; Aurbach, D.; Martha, S.K. Lead-Acid Batteries and Lead–Carbon Hybrid Systems: A Review. J. Power Sources 2023, 579, 233312. [Google Scholar] [CrossRef]
  109. Noga, M.; Juda, Z. The Application of Nimh Batteries in a Light-Duty Electric Vehicle. Czas. Tech. 2019, 1, 197–222. [Google Scholar] [CrossRef]
  110. Gerssen-Gondelach, S.J.; Faaij, A.P.C. Performance of Batteries for Electric Vehicles on Short and Longer Term. J. Power Sources 2012, 212, 111–129. [Google Scholar] [CrossRef]
  111. Poonam; Sharma, K.; Arora, A.; Tripathi, S.K. Review of Supercapacitors: Materials and Devices. J. Energy Storage 2019, 21, 801–825. [Google Scholar] [CrossRef]
  112. Guo, L.; Hu, P.; Wei, H. Development of Supercapacitor Hybrid Electric Vehicle. J. Energy Storage 2023, 65, 107269. [Google Scholar] [CrossRef]
  113. Gao, D.; Luo, Z.; Liu, C.; Fan, S. A Survey of Hybrid Energy Devices Based on Supercapacitors. Green Energy Environ. 2023, 8, 972–988. [Google Scholar] [CrossRef]
  114. Sharma, P.; Kumar, V. Current Technology of Supercapacitors: A Review. J. Electron. Mater. 2020, 49, 3520–3532. [Google Scholar] [CrossRef]
  115. González, A.; Goikolea, E.; Barrena, J.A.; Mysyk, R. Review on Supercapacitors: Technologies and Materials. Renew. Sustain. Energy Rev. 2016, 58, 1189–1206. [Google Scholar] [CrossRef]
  116. Burke, A.; Liu, Z.; Zhao, H. Present and Future Applications of Supercapacitors in Electric and Hybrid Vehicles. In Proceedings of the 2014 IEEE International Electric Vehicle Conference, IEVC 2014, Florence, Italy, 17–19 December 2014. [Google Scholar] [CrossRef]
  117. Horn, M.; MacLeod, J.; Liu, M.; Webb, J.; Motta, N. Supercapacitors: A New Source of Power for Electric Cars? Econ. Anal. Policy 2019, 61, 93–103. [Google Scholar] [CrossRef]
  118. Xu, H.; Shen, M. The Control of Lithium-Ion Batteries and Supercapacitors in Hybrid Energy Storage Systems for Electric Vehicles: A Review. Int. J. Energy Res. 2021, 45, 20524–20544. [Google Scholar] [CrossRef]
  119. Mounica, V.; Obulesu, Y.P. Hybrid Power Management Strategy with Fuel Cell, Battery, and Supercapacitor for Fuel Economy in Hybrid Electric Vehicle Application. Energies 2022, 15, 4185. [Google Scholar] [CrossRef]
  120. Battery-Supercapacitor Energy, F.; Astaneh, M.; Andric, J.; Lemian, D.; Bode, F. Battery-Supercapacitor Energy Storage Systems for Electrical Vehicles: A Review. Energies 2022, 15, 5683. [Google Scholar] [CrossRef]
  121. Sulaiman, N.; Hannan, M.A.; Mohamed, A.; Majlan, E.H.; Wan Daud, W.R. A Review on Energy Management System for Fuel Cell Hybrid Electric Vehicle: Issues and Challenges. Renew. Sustain. Energy Rev. 2015, 52, 802–814. [Google Scholar] [CrossRef]
  122. Khaligh, A.; Li, Z. Battery, Ultracapacitor, Fuel Cell, and Hybrid Energy Storage Systems for Electric, Hybrid Electric, Fuel Cell, and Plug-in Hybrid Electric Vehicles: State of the Art. IEEE Trans. Veh. Technol. 2010, 59, 2806–2814. [Google Scholar] [CrossRef]
  123. Cunanan, C.; Tran, M.K.; Lee, Y.; Kwok, S.; Leung, V.; Fowler, M. A Review of Heavy-Duty Vehicle Powertrain Technologies: Diesel Engine Vehicles, Battery Electric Vehicles, and Hydrogen Fuel Cell Electric Vehicles. Clean Technol. 2021, 3, 474–489. [Google Scholar] [CrossRef]
  124. İnci, M.; Büyük, M.; Demir, M.H.; İlbey, G. A Review and Research on Fuel Cell Electric Vehicles: Topologies, Power Electronic Converters, Energy Management Methods, Technical Challenges, Marketing and Future Aspects. Renew. Sustain. Energy Rev. 2021, 137, 110648. [Google Scholar] [CrossRef]
  125. Liu, W.; Placke, T.; Chau, K.T. Overview of Batteries and Battery Management for Electric Vehicles. Energy Rep. 2022, 8, 4058–4084. [Google Scholar] [CrossRef]
  126. Waseem, M.; Ahmad, M.; Parveen, A.; Suhaib, M. Battery Technologies and Functionality of Battery Management System for EVs: Current Status, Key Challenges, and Future Prospectives. J. Power Sources 2023, 580, 233349. [Google Scholar] [CrossRef]
  127. Abdelrahman, M.S.; Hussein, H.; Mohammed, O.A. Rule-Based Power and Energy Management System For Shipboard Microgrid With HESS To Mitigate Propulsion and Pulsed Load Fluctuations. In Proceedings of the 2023 IEEE Green Technologies Conference (GreenTech), Denver, CO, USA, 19–21 April 2023; pp. 224–228. [Google Scholar] [CrossRef]
  128. Hussein, H.M.; Aghmadi, A.; Abdelrahman, M.S.; Rafin, S.M.S.H.; Mohammed, O. A Review of Battery State of Charge Estimation and Management Systems: Models and Future Prospective. Wiley Interdiscip Rev. Energy Environ. 2024, 13, e507. [Google Scholar] [CrossRef]
  129. Hussein, H.; Donekal, A.; Aghmadi, A.; Rafin, S.M.S.H.; Mohammed, O.A. State of Charge Estimation Using Data-Driven Models for Inverter-Based Systems. In Proceedings of the 2023 IEEE Design Methodologies Conference, DMC 2023, Miami, FL, USA, 24–26 September 2023. [Google Scholar] [CrossRef]
  130. Rivera-Barrera, J.P.; Muñoz-Galeano, N.; Sarmiento-Maldonado, H.O. SoC Estimation for Lithium-Ion Batteries: Review and Future Challenges. Electronics 2017, 6, 102. [Google Scholar] [CrossRef]
  131. Zhang, R.; Xia, B.; Li, B.; Cao, L.; Lai, Y.; Zheng, W.; Wang, H.; Wang, W. State of the Art of Lithium-Ion Battery SOC Estimation for Electrical Vehicles. Energies 2018, 11, 1820. [Google Scholar] [CrossRef]
  132. Hussein, H.M.; Esoofally, M.; Donekal, A.; Rafin, S.M.S.H.; Mohammed, O. Comparative Study-Based Data-Driven Models for Lithium-Ion Battery State-of-Charge Estimation. Batteries 2024, 10, 89. [Google Scholar] [CrossRef]
  133. Wu, B.; Widanage, W.D.; Yang, S.; Liu, X. Battery Digital Twins: Perspectives on the Fusion of Models, Data and Artificial Intelligence for Smart Battery Management Systems. Energy AI 2020, 1, 100016. [Google Scholar] [CrossRef]
  134. Hussein, H.M.; Rafin, S.M.S.H.; Abdelrahman, M.S.; Mohammed, O.A. Hardware Implementation of a Resilient Energy Management System for Networked Microgrids. World Electr. Veh. J. 2024, 15, 209. [Google Scholar] [CrossRef]
  135. Hu, Y.; Yurkovich, S.; Guezennec, Y.; Yurkovich, B.J. Electro-Thermal Battery Model Identification for Automotive Applications. J. Power Sources 2011, 196, 449–457. [Google Scholar] [CrossRef]
  136. Karimi, G.; Azizi, M.; Babapoor, A. Experimental Study of a Cylindrical Lithium Ion Battery Thermal Management Using Phase Change Material Composites. J. Energy Storage 2016, 8, 168–174. [Google Scholar] [CrossRef]
  137. Garcia-Valle, R.; Lopes, J.A.P. Electric Vehicle Integration into Modern Power Networks; Springer: New York, NY, USA, 2013; pp. 1–325. [Google Scholar]
  138. Zou, C.; Hu, X.; Wei, Z.; Wik, T.; Egardt, B. Electrochemical Estimation and Control for Lithium-Ion Battery Health-Aware Fast Charging. IEEE Trans. Ind. Electron. 2018, 65, 6635–6645. [Google Scholar] [CrossRef]
  139. Tremblay, O.; Dessaint, L.A. Experimental Validation of a Battery Dynamic Model for EV Applications. World Electr. Veh. J. 2009, 3, 289–298. [Google Scholar] [CrossRef]
  140. Chen, D.; Jiang, J.; Kim, G.H.; Yang, C.; Pesaran, A. Comparison of Different Cooling Methods for Lithium Ion Battery Cells. Appl. Therm. Eng. 2016, 94, 846–854. [Google Scholar] [CrossRef]
  141. Rahman, A.; Nur, F.; Hawlader, M.N.A.; Afroz, R. Fuzzy Controlled Evaporative Battery Thermal Management System for EV/HEV. Int. J. Electr. Hybrid Veh. 2015, 7, 22–39. [Google Scholar] [CrossRef]
  142. Liu, H.; Wei, Z.; He, W.; Zhao, J. Thermal Issues about Li-Ion Batteries and Recent Progress in Battery Thermal Management Systems: A Review. Energy Convers. Manag. 2017, 150, 304–330. [Google Scholar] [CrossRef]
  143. Lu, L.; Han, X.; Li, J.; Hua, J.; Ouyang, M. A Review on the Key Issues for Lithium-Ion Battery Management in Electric Vehicles. J. Power Sources 2013, 226, 272–288. [Google Scholar] [CrossRef]
  144. Rao, Z.; Wang, S. A Review of Power Battery Thermal Energy Management. Renew. Sustain. Energy Rev. 2011, 15, 4554–4571. [Google Scholar] [CrossRef]
  145. Naseri, F.; Barbu, C.; Sarikurt, T. Optimal Sizing of Hybrid High-Energy/High-Power Battery Energy Storage Systems to Improve Battery Cycle Life and Charging Power in Electric Vehicle Applications. J. Energy Storage 2022, 55, 105768. [Google Scholar] [CrossRef]
  146. Xia, X.; Li, P. A Review of the Life Cycle Assessment of Electric Vehicles: Considering the Influence of Batteries. Sci. Total Environ. 2022, 814, 152870. [Google Scholar] [CrossRef] [PubMed]
  147. Liu, H.; Chen, F.; Tong, Y.; Wang, Z.; Yu, X.; Huang, R. Impacts of Driving Conditions on EV Battery Pack Life Cycle. World Electr. Veh. J. 2020, 11, 17. [Google Scholar] [CrossRef]
  148. Beaudet, A.; Larouche, F.; Amouzegar, K.; Bouchard, P.; Zaghib, K. Key Challenges and Opportunities for Recycling Electric Vehicle Battery Materials. Sustainability 2020, 12, 5837. [Google Scholar] [CrossRef]
  149. Yu, X.; Li, W.; Gupta, V.; Gao, H.; Tran, D.; Sarwar, S.; Chen, Z.; Yu, X.; Gupta, V.; Gao, H.; et al. Current Challenges in Efficient Lithium-Ion Batteries’ Recycling: A Perspective. Glob. Chall. 2022, 6, 2200099. [Google Scholar] [CrossRef]
  150. Rajaeifar, M.A.; Ghadimi, P.; Raugei, M.; Wu, Y.; Heidrich, O. Challenges and Recent Developments in Supply and Value Chains of Electric Vehicle Batteries: A Sustainability Perspective. Resour. Conserv. Recycl. 2022, 180, 106144. [Google Scholar] [CrossRef]
  151. Costa, C.M.; Barbosa, J.C.; Gonçalves, R.; Castro, H.; Campo, F.J.D.; Lanceros-Méndez, S. Recycling and Environmental Issues of Lithium-Ion Batteries: Advances, Challenges and Opportunities. Energy Storage Mater. 2021, 37, 433–465. [Google Scholar] [CrossRef]
  152. Li, P.; Xia, X.; Guo, J. A Review of the Life Cycle Carbon Footprint of Electric Vehicle Batteries. Sep. Purif. Technol. 2022, 296, 121389. [Google Scholar] [CrossRef]
  153. Wu, H.; Hu, Y.; Yu, Y.; Huang, K.; Wang, L. The Environmental Footprint of Electric Vehicle Battery Packs during the Production and Use Phases with Different Functional Units. Int. J. Life Cycle Assess. 2021, 26, 97–113. [Google Scholar] [CrossRef]
  154. Faisal, S.; Soni, B.P.; Goyal, G.R.; Bakhsh, F.I.; Husain, D.; Ahmad, A. Reducing the Ecological Footprint and Charging Cost of Electric Vehicle Charging Station Using Renewable Energy Based Power System. e-Prime Adv. Electr. Eng. Electron. Energy 2024, 7, 100398. [Google Scholar] [CrossRef]
  155. Lai, X.; Chen, Q.; Tang, X.; Zhou, Y.; Gao, F.; Guo, Y.; Bhagat, R.; Zheng, Y. Critical Review of Life Cycle Assessment of Lithium-Ion Batteries for Electric Vehicles: A Lifespan Perspective. eTransportation 2022, 12, 100169. [Google Scholar] [CrossRef]
  156. Jabbari, V.; Yurkiv, V.; Rasul, G.; Cheng, M.; Griffin, P.; Mashayek, F.; Shahbazian-Yassar, R.; Jabbari, V.; Yurkiv, V.; Rasul, M.G.; et al. A Smart Lithium Battery with Shape Memory Function. Small 2022, 18, 2102666. [Google Scholar] [CrossRef]
  157. Sasaki, T.; Ukyo, Y.; Novák, P. Memory Effect in a Lithium-Ion Battery. Nat. Mater. 2013, 12, 569–575. [Google Scholar] [CrossRef]
  158. Jain, R.; Lakhnot, A.S.; Bhimani, K.; Sharma, S.; Mahajani, V.; Panchal, R.A.; Kamble, M.; Han, F.; Wang, C.; Koratkar, N. Nanostructuring versus Microstructuring in Battery Electrodes. Nat. Rev. Mater. 2022, 7, 736–746. [Google Scholar] [CrossRef]
  159. Roshandel, E.; Mahmoudi, A.; Kahourzade, S.; Tahir, A.; Fernando, N. Propulsion System of Electric Vehicles: Review. In Proceedings of the 2021 31st Australasian Universities Power Engineering Conference, AUPEC 2021, Perth, Australia, 26–30 September 2021. [Google Scholar] [CrossRef]
  160. Elbeshbeshy, A.M.; Matar, O.; Shawky, A.; Alaaeldin, A.; Faiod, H.S.; Gomaa, M.A.; Ahmed, H.A.; Shaheen, A.G.; Kaddah, S.S.; Badr, B.M. Design and Development of a Low-Cost Battery System for Electric Motorcycles. In Proceedings of the 2023 24th International Middle East Power System Conference, MEPCON 2023, Mansoura, Egypt, 19–21 December 2023. [Google Scholar] [CrossRef]
  161. Bala, S.; Tengner, T.; Rosenfeld, P.; Delince, F. The Effect of Low Frequency Current Ripple on the Performance of a Lithium Iron Phosphate (LFP) Battery Energy Storage System. In Proceedings of the 2012 IEEE Energy Conversion Congress and Exposition, ECCE 2012, Raleigh, NC, USA, 15–20 September 2012; pp. 3485–3492. [Google Scholar] [CrossRef]
  162. Yilmaz, M.; Krein, P.T. Review of Battery Charger Topologies, Charging Power Levels, and Infrastructure for Plug-in Electric and Hybrid Vehicles. IEEE Trans. Power Electron. 2013, 28, 2151–2169. [Google Scholar] [CrossRef]
  163. Tu, H.; Feng, H.; Srdic, S.; Lukic, S. Extreme Fast Charging of Electric Vehicles: A Technology Overview. IEEE Trans. Transp. Electrif. 2019, 5, 861–878. [Google Scholar] [CrossRef]
  164. Monteiro, V.; Ferreira, J.C.; Nogueiras Melendez, A.A.; Couto, C.; Afonso, J.L. Experimental Validation of a Novel Architecture Based on a Dual-Stage Converter for Off-Board Fast Battery Chargers of Electric Vehicles. IEEE Trans. Veh. Technol. 2018, 67, 1000–1011. [Google Scholar] [CrossRef]
  165. Na, T.; Yuan, X.; Tang, J.; Zhang, Q. A Review of On-Board Integrated Charger for Electric Vehicles and a New Solution. In Proceedings of the PEDG 2019—2019 IEEE 10th International Symposium on Power Electronics for Distributed Generation Systems, Xi’an, China, 3–6 June 2019; pp. 693–699. [Google Scholar] [CrossRef]
  166. Mahobiya, S.; Kumar, S.; Gawre, S.K. Investigation of Different Types of Bidirectional Charging Operations of Electric Vehicles. Lect. Notes Electr. Eng. 2023, 960, 229–243. [Google Scholar] [CrossRef]
  167. Pradhan, R.; Keshmiri, N.; Emadi, A. On-Board Chargers for High-Voltage Electric Vehicle Powertrains: Future Trends and Challenges. IEEE Open J. Power Electron. 2023, 4, 189–207. [Google Scholar] [CrossRef]
  168. Semsar, S.; Soong, T.; Lehn, P.W. On-Board Single-Phase Integrated Electric Vehicle Charger with V2G Functionality. IEEE Trans. Power Electron. 2020, 35, 12072–12084. [Google Scholar] [CrossRef]
  169. Subotic, I.; Bodo, N.; Levi, E. Single-Phase on-Board Integrated Battery Chargers for EVs Based on Multiphase Machines. IEEE Trans. Power Electron. 2016, 31, 6511–6523. [Google Scholar] [CrossRef]
  170. Taha, R.A.; Abdel-Azim, W.E.; Shawier, A.; Metwly, M.Y.; Abdel-Khalik, A.S.; Hamad, M.S.; Hamdy, R.A.; Gadoue, S.; Ahmed, S. Single-Phase Charging of Six-Phase Integrated On-Board Battery Charger Using Predictive Current Control. IEEE Trans. Transp. Electrif. 2024, 10, 540–552. [Google Scholar] [CrossRef]
  171. Khalid, M.; Ahmad, F.; Panigrahi, B.K.; Al-Fagih, L. A Comprehensive Review on Advanced Charging Topologies and Methodologies for Electric Vehicle Battery. J. Energy Storage 2022, 53, 105084. [Google Scholar] [CrossRef]
  172. Kurs, A.; Karalis, A.; Moffatt, R.; Joannopoulos, J.D.; Fisher, P.; Soljačić, M. Wireless Power Transfer via Strongly Coupled Magnetic Resonances. Science 2007, 317, 83–86. [Google Scholar] [CrossRef]
  173. Low, Z.N.; Chinga, R.A.; Tseng, R.; Lin, J. Design and Test of a High-Power High-Efficiency Loosely Coupled Planar Wireless Power Transfer System. IEEE Trans. Ind. Electron. 2009, 56, 1801–1812. [Google Scholar] [CrossRef]
  174. Panchal, C.; Stegen, S.; Lu, J. Review of Static and Dynamic Wireless Electric Vehicle Charging System. Eng. Sci. Technol. Int. J. 2018, 21, 922–937. [Google Scholar] [CrossRef]
  175. Jiang, C.; Chau, K.T.; Liu, C.; Lee, C.H.T. An Overview of Resonant Circuits for Wireless Power Transfer. Energies 2017, 10, 894. [Google Scholar] [CrossRef]
  176. Elbeshbeshy, A.; Rizk, M.; Kaddah, S. Analysis of Inductive Characteristics for Various Helical and Spiral Coil Configurations Supported by COMSOL Multiphysics. Mansoura Eng. J. 2021, 46, 1–9. [Google Scholar] [CrossRef]
  177. Lu, F.; Zhang, H.; Mi, C. A Two-Plate Capacitive Wireless Power Transfer System for Electric Vehicle Charging Applications. IEEE Trans. Power Electron. 2018, 33, 964–969. [Google Scholar] [CrossRef]
  178. Xia, C.Y.; Li, C.W.; Zhang, J. Analysis of Power Transfer Characteristic of Capacitive Power Transfer System and Inductively Coupled Power Transfer System. In Proceedings of the 2011 International Conference on Mechatronic Science, Electric Engineering and Computer, MEC 2011, Jilin, China, 19–22 August 2011; pp. 1281–1285. [Google Scholar] [CrossRef]
  179. Mohamed, A.A.S.; Shaier, A.A.; Metwally, H.; Selem, S.I. Wireless charging technologies for electric vehicles: Inductive, capacitive, and magnetic gear. IET Power Electron. 2023. [Google Scholar] [CrossRef]
  180. Ali, A.; Mousa, H.H.H.; Shaaban, M.F.; Azzouz, M.A.; Awad, A.S.A. A Comprehensive Review on Charging Topologies and Power Electronic Converter Solutions for Electric Vehicles. J. Mod. Power Syst. Clean Energy 2024, 12, 675–694. [Google Scholar] [CrossRef]
  181. Ronanki, D.; Karneddi, H. Electric Vehicle Charging Infrastructure: Review, Cyber Security Considerations, Potential Impacts, Countermeasures, and Future Trends. IEEE J. Emerg. Sel. Top. Power Electron. 2024, 12, 242–256. [Google Scholar] [CrossRef]
  182. Sagar, A.; Kashyap, A.; Nasab, M.A.; Padmanaban, S.; Bertoluzzo, M.; Kumar, A.; Blaabjerg, F. A Comprehensive Review of the Recent Development of Wireless Power Transfer Technologies for Electric Vehicle Charging Systems. IEEE Access 2023, 11, 83703–83751. [Google Scholar] [CrossRef]
  183. Cesiel, D.; Zhu, C. A Closer Look at the On-Board Charger: The Development of the Second-Generation Module for the Chevrolet Volt. IEEE Electrif. Mag. 2017, 5, 36–42. [Google Scholar] [CrossRef]
  184. Nissan Leaf EV Charging Guide—Zapmap. Available online: https://www.zap-map.com/ev-guides/model-charging/nissan-leaf (accessed on 1 July 2024).
  185. Gautam, D.S.; Musavi, F.; Edington, M.; Eberle, W.; Dunford, W.G. An Automotive Onboard 3.3-KW Battery Charger for PHEV Application. IEEE Trans. Veh. Technol. 2012, 61, 3466–3474. [Google Scholar] [CrossRef]
  186. EV Charger Inverter|Car Charging Inverter|Valeo. Available online: https://www.valeo.com/en/charger-inverter-for-electric-vehicles/ (accessed on 1 July 2024).
  187. All There Is to Know about Renault Zoe—Renault Group. Available online: https://www.renaultgroup.com/en/news-on-air/news/all-there-is-to-know-about-charging-renault-zoe/ (accessed on 1 July 2024).
  188. Customer and Product Support|Tesla Support. Available online: https://www.tesla.com/support/home-charging-installation/wall-connector-features (accessed on 1 July 2024).
  189. High Power Charging|High Power Fast Chargers|ABB. Available online: https://new.abb.com/ev-charging/high-power-charging (accessed on 1 July 2024).
  190. Electric Vehicle Electric Vehicle Infrastructure—Enercon. Available online: https://www.enercon.com/markets/power-delivery-grid-infrastructure/ev-electric-vehicle-infrastructure/ (accessed on 1 July 2024).
  191. Excellent Performance Starts with Charging. First Porsche Fast Charging Ppark Featuring 800-Volt Technology Connected to the Grid—Porsche Newsroom. Available online: https://newsroom.porsche.com/en/company/porsche-fast-charging-park-berlin-adlershof-electric-cars-electro-mobility-plug-in-technology-turbo-infrastructure-15814.html (accessed on 1 July 2024).
  192. HEVO: Wireless Charging for Electric Vehicles. Available online: https://hevo.com/ (accessed on 1 July 2024).
  193. WiTricity • Wireless EV Charging. Available online: https://witricity.com/ (accessed on 1 July 2024).
  194. Kim, J.H.; Lee, B.S.; Lee, J.H.; Lee, S.H.; Park, C.B.; Jung, S.M.; Lee, S.G.; Yi, K.P.; Baek, J. Development of 1-MW Inductive Power Transfer System for a High-Speed Train. IEEE Trans. Ind. Electron. 2015, 62, 6242–6250. [Google Scholar] [CrossRef]
  195. Villar, I.; Garcia-Bediaga, A.; Iruretagoyena, U.; Arregi, R.; Estevez, P. Design and Experimental Validation of a 50kW IPT for Railway Traction Applications. In Proceedings of the 2018 IEEE Energy Conversion Congress and Exposition, ECCE 2018, Portland, OR, USA, 23–27 September 2018; pp. 1177–1183. [Google Scholar] [CrossRef]
  196. Bombardier’s PRIMOVE Technology Enters Service on Scandinavia’s First Inductively Charged Bus Line. Available online: https://masstransit.network/mass-transit-news/bombardier-transportation-gmbh/primove-inductively-charged-bus (accessed on 1 July 2024).
  197. Home—WAVE. Available online: https://wavecharging.com/ (accessed on 1 July 2024).
  198. Elbeshbeshy, A.M.; Ghanem, A.; Abulanwar, S.; Deng, F.; Kaddah, S.S.; Rizk, M.E.M. Enhanced Stability in Hybrid AC/DC Microgrids with Controlled Magnetic Energy Router. In Proceedings of the 2023 24th International Middle East Power System Conference, MEPCON 2023, Mansoura, Egypt, 19–21 December 2023. [Google Scholar] [CrossRef]
  199. Zhai, H.; Sands, T. Comparison of Deep Learning and Deterministic Algorithms for Control Modeling. Sensors 2022, 22, 6362. [Google Scholar] [CrossRef] [PubMed]
  200. Zhong, J.; He, L.; Li, C.; Cao, Y.; Wang, J.; Fang, B.; Zeng, L.; Xiao, G. Coordinated Control for Large-Scale EV Charging Facilities and Energy Storage Devices Participating in Frequency Regulation. Appl. Energy 2014, 123, 253–262. [Google Scholar] [CrossRef]
  201. Faddel, S.; Mohammed, O.A. Automated Distributed Electric Vehicle Controller for Residential Demand Side Management. IEEE Trans. Ind. Appl. 2019, 55, 16–25. [Google Scholar] [CrossRef]
  202. Rivera, S.; Kouro, S.; Vazquez, S.; Goetz, S.M.; Lizana, R.; Romero-Cadaval, E. Electric Vehicle Charging Infrastructure: From Grid to Battery. IEEE Ind. Electron. Mag. 2021, 15, 37–51. [Google Scholar] [CrossRef]
  203. Karamanakos, P.; Liegmann, E.; Geyer, T.; Kennel, R. Model Predictive Control of Power Electronic Systems: Methods, Results, and Challenges. IEEE Open J. Ind. Appl. 2020, 1, 95–114. [Google Scholar] [CrossRef]
  204. Hussein, H.M.; Rafin, S.M.S.H.; Kharchouf, I.; Abdelrahman, M.S.; Mohammed, O. Electric Vehicle Performance Enhancement Utilizing Hybrid Energy Storage Systems. In Proceedings of the 2024 IEEE Vehicle Power and Propulsion Conference (VPPC), Washington, DC, USA, 7–10 October 2024. [Google Scholar]
  205. Kouro, S.; Perez, M.A.; Rodriguez, J.; Llor, A.M.; Young, H.A. Model Predictive Control: MPC’s Role in the Evolution of Power Electronics. IEEE Ind. Electron. Mag. 2015, 9, 8–21. [Google Scholar] [CrossRef]
  206. Vishnuram, P.; Suresh, P.; Narayanamoorthi, R.; Vijayakumar, K.; Nastasi, B. Wireless Chargers for Electric Vehicle: A Systematic Review on Converter Topologies, Environmental Assessment, and Review Policy. Energies 2023, 16, 1731. [Google Scholar] [CrossRef]
  207. Patil, D.; McDonough, M.K.; Miller, J.M.; Fahimi, B.; Balsara, P.T. Wireless Power Transfer for Vehicular Applications: Overview and Challenges. IEEE Trans. Transp. Electrif. 2017, 4, 3–37. [Google Scholar] [CrossRef]
  208. Ahmed, A.M.; Khalifa, O.O. Wireless power transfer for electric vehicle charging. AIP Conf. Proc. 2020, 2306, 020026. [Google Scholar] [CrossRef]
  209. Mi, C.C.; Buja, G.; Choi, S.Y.; Rim, C.T. Modern Advances in Wireless Power Transfer Systems for Roadway Powered Electric Vehicles. IEEE Trans. Ind. Electron. 2016, 63, 6533–6545. [Google Scholar] [CrossRef]
  210. Hata, K.; Imura, T.; Hori, Y. Dynamic Wireless Power Transfer System for Electric Vehicles to Simplify Ground Facilities—Power Control and Efficiency Maximization on the Secondary Side. In Proceedings of the IEEE Applied Power Electronics Conference and Exposition—APEC 2016, Long Beach, CA, USA, 20–24 March 2016; pp. 1731–1736. [Google Scholar] [CrossRef]
  211. Diekhans, T.; De Doncker, R.W. A Dual-Side Controlled Inductive Power Transfer System Optimized for Large Coupling Factor Variations and Partial Load. IEEE Trans. Power Electron. 2015, 30, 6320–6328. [Google Scholar] [CrossRef]
  212. Zhang, Y.; Kan, T.; Yan, Z.; Mi, C.C. Frequency and Voltage Tuning of Series-Series Compensated Wireless Power Transfer System to Sustain Rated Power under Various Conditions. IEEE J. Emerg. Sel. Top. Power Electron. 2019, 7, 1311–1317. [Google Scholar] [CrossRef]
  213. Arteaga, J.M.; Aldhaher, S.; Kkelis, G.; Yates, D.C.; Mitcheson, P.D. Multi-MHz IPT Systems for Variable Coupling. IEEE Trans. Power Electron. 2018, 33, 7744–7758. [Google Scholar] [CrossRef]
  214. Zhang, Y.; Zhao, Z. Frequency Splitting Analysis of Two-Coil Resonant Wireless Power Transfer. IEEE Antennas Wirel. Propag. Lett. 2014, 13, 400–402. [Google Scholar] [CrossRef]
  215. Li, Z.; Zhu, C.; Jiang, J.; Song, K.; Wei, G. A 3-KW Wireless Power Transfer System for Sightseeing Car Supercapacitor Charge. IEEE Trans. Power Electron. 2017, 32, 3301–3316. [Google Scholar] [CrossRef]
  216. Huang, Z.; Wong, S.C.; Tse, C.K. Control Design for Optimizing Efficiency in Inductive Power Transfer Systems. IEEE Trans. Power Electron. 2018, 33, 4523–4534. [Google Scholar] [CrossRef]
  217. Li, H.; Li, J.; Wang, K.; Chen, W.; Yang, X. A Maximum Efficiency Point Tracking Control Scheme for Wireless Power Transfer Systems Using Magnetic Resonant Coupling. IEEE Trans. Power Electron. 2015, 30, 3998–4008. [Google Scholar] [CrossRef]
  218. Guidi, G.; Suul, J.A. Minimizing Converter Requirements of Inductive Power Transfer Systems with Constant Voltage Load and Variable Coupling Conditions. IEEE Trans. Ind. Electron. 2016, 63, 6835–6844. [Google Scholar] [CrossRef]
  219. Azad, A.N.; Echols, A.; Kulyukin, V.A.; Zane, R.; Pantic, Z. Analysis, Optimization, and Demonstration of a Vehicular Detection System Intended for Dynamic Wireless Charging Applications. IEEE Trans. Transp. Electrif. 2019, 5, 147–161. [Google Scholar] [CrossRef]
  220. Kumar, A.; Bertoluzzo, M.; Jha, R.K.; Sagar, A. Analysis of Losses in Two Different Control Approaches for S-S Wireless Power Transfer Systems for Electric Vehicle. Energies 2023, 16, 1795. [Google Scholar] [CrossRef]
  221. Cao, Z.; Mahmoudi, A.; Kahourzade, S.; Soong, W.L. An Overview of Electric Motors for Electric Vehicles. In Proceedings of the 2021 31st Australasian Universities Power Engineering Conference (AUPEC), Perth, Australia, 26–30 September 2021; pp. 1–6. [Google Scholar] [CrossRef]
  222. Schlabbach, J.; Rofalski, K.H. Power System Engineering: Planning, Design, and Operation of Power Systems and Equipment; Vch Verlagsgesellschaft Mbh: Weinheim, Germany, 2014; Volume 9783527412600, pp. 1–379. [Google Scholar] [CrossRef]
  223. Yang, Z.; Shang, F.; Brown, I.P.; Krishnamurthy, M. Comparative Study of Interior Permanent Magnet, Induction, and Switched Reluctance Motor Drives for EV and HEV Applications. IEEE Trans. Transp. Electrific. 2015, 1, 245–254. [Google Scholar] [CrossRef]
  224. Bazzi, A.M.; Krein, P.T. Review of Methods for Real-Time Loss Minimization in Induction Machines. IEEE Trans. Ind. Appl. 2010, 46, 2319–2328. [Google Scholar] [CrossRef]
  225. Morimoto, S.; Takeda, Y.; Hirasa, T.; Taniguchi, K. Expansion of Operating Limits for Permanent Magnet Motor by Current Vector Control Considering Inverter Capacity. IEEE Trans. Ind. Appl. 1990, 26, 866–871. [Google Scholar] [CrossRef]
  226. Dorrell, D.G.; Knight, A.M.; Popescu, M.; Evans, L.; Staton, D.A. Comparison of Different Motor Design Drives for Hybrid Electric Vehicles. In Proceedings of the 2010 IEEE Energy Conversion Congress and Exposition, ECCE 2010—Proceedings 2010, Atlanta, GA, USA, 12–16 September 2010; pp. 3352–3359. [Google Scholar] [CrossRef]
  227. Liao, G.; Zhang, W.; Cai, C. Research on a PMSM Control Strategy for Electric Vehicles. Adv. Mech. Eng. 2021, 13, 16878140211051462. [Google Scholar] [CrossRef]
  228. Casadei, D.; Profumo, F.; Serra, G.; Tani, A. FOC and DTC: Two Viable Schemes for Induction Motors Torque Control. IEEE Trans. Power Electron. 2002, 17, 779–787. [Google Scholar] [CrossRef]
  229. Holtz, J. Sensorless Control of Induction Motor Drives. Proc. IEEE 2002, 90, 1359–1394. [Google Scholar] [CrossRef]
  230. Peters, D.T.; Cowie, J.G.; Brush, E.F.; Van Son, D.J. Copper in the Squirrel Cage for Improved Motor Performance. In Proceedings of the IEMDC 2003—IEEE International Electric Machines and Drives Conference, Madison, WI, USA, 1–4 June 2003; Volume 2, pp. 1265–1271. [Google Scholar] [CrossRef]
  231. 2024 Audi Q8 E-Tron®|Premium SUV EV|Audi USA. Available online: https://www.audiusa.com/us/web/en/models/q8-e-tron/q8-e-tron/2024/overview.html (accessed on 30 July 2024).
  232. Millan, J.; Godignon, P.; Perpina, X.; Perez-Tomas, A.; Rebollo, J. A Survey of Wide Bandgap Power Semiconductor Devices. IEEE Trans. Power Electron. 2014, 29, 2155–2163. [Google Scholar] [CrossRef]
  233. Vagati, A.; Pellegrino, G.; Guglielmi, P. Comparison between SPM and IPM Motor Drives for EV Application. In Proceedings of the 19th International Conference on Electrical Machines, ICEM 2010, Rome, Italy, 6–8 September 2010. [Google Scholar] [CrossRef]
  234. Permanent Magnet Synchronous Motor. Available online: https://about-motors.com/motorcontrol/pmsm/ (accessed on 30 July 2024).
  235. Zhu, Z.Q.; Howe, D. Electrical Machines and Drives for Electric, Hybrid, and Fuel Cell Vehicles. Proc. IEEE 2007, 95, 746–765. [Google Scholar] [CrossRef]
  236. Chau, K.T.; Chan, C.C.; Liu, C. Overview of Permanent-Magnet Brushless Drives for Electric and Hybrid Electric Vehicles. IEEE Trans. Ind. Electron. 2008, 55, 2246–2257. [Google Scholar] [CrossRef]
  237. Moghaddam, R.R.; Gyllensten, F. Novel High-Performance SynRM Design Method: An Easy Approach for a Complicated Rotor Topology. IEEE Trans. Ind. Electron. 2014, 61, 5058–5065. [Google Scholar] [CrossRef]
  238. Yang, Y.; He, Q.; Fu, C.; Liao, S.; Tan, P. Efficiency Improvement of Permanent Magnet Synchronous Motor for Electric Vehicles. Energy 2020, 213, 118859. [Google Scholar] [CrossRef]
  239. Pellegrino, G.; Vagati, A.; Boazzo, B.; Guglielmi, P. Comparison of Induction and PM Synchronous Motor Drives for EV Application Including Design Examples. IEEE Trans. Ind. Appl. 2012, 48, 2322–2332. [Google Scholar] [CrossRef]
  240. Briz, F.; Degner, M.; Lorenz, R. Analysis and design of current regulators using complex vectors. IEEE Trans. Ind. Appl. 2000, 36, 817–825. [Google Scholar] [CrossRef]
  241. Hoenderdaal, S.; Espinoza, L.T.; Marscheider-Weidemann, F.; Graus, W. Can a dysprosium shortage threaten green energy technologies? Energy 2013, 49, 344–355. [Google Scholar] [CrossRef]
  242. Ibrahim, M.; Masisi, L.; Pillay, P. Design of variable flux permanent magnet machines using alnico magnets. IEEE Trans. Ind. Appl. 2015, 51, 4482–4491. [Google Scholar] [CrossRef]
  243. Amara, Y.; Vido, L.; Gabsi, M.; Hoang, E.; Ben Ahmed, A.H.; Lecrivain, M. Hybrid excitation synchronous machines: Energy-efficient solution for vehicles propulsion. IEEE Trans. Veh. Technol. 2009, 58, 2137–2149. [Google Scholar] [CrossRef]
  244. El-Refaie, A.M.; Alexander, J.P.; Galioto, S.; Reddy, P.B.; Huh, K.-K.; de Bock, P.; Shen, X. Advanced high-power-density interior permanent magnet motor for traction applications. IEEE Trans. Ind. Appl. 2014, 50, 3235–3248. [Google Scholar] [CrossRef]
  245. Sizov, G.Y.; Ionel, D.M.; Demerdash, N.A.O. Modeling and parametric design of permanent-magnet AC machines using computationally efficient finite-element analysis. IEEE Trans. Ind. Electron. 2012, 59, 2403–2413. [Google Scholar] [CrossRef]
  246. Bilgin, B.; Howey, B.; Callegaro, A.D.; Liang, J.; Kordic, M.; Taylor, J.; Emadi, A. Making the Case for Switched Reluctance Motors for Propulsion Applications. IEEE Trans. Veh. Technol. 2020, 69, 7172–7186. [Google Scholar] [CrossRef]
  247. Krishnan, R. Switched Reluctance Motor Drives: Modeling, Simulation, Analysis, Design, and Applications; CRC Press: Boca Raton, FL, USA, 2017; pp. 1–398. [Google Scholar] [CrossRef]
  248. Miller, T. Optimal design of switched reluctance motors. IEEE Trans. Ind. Electron. 2002, 49, 15–27. [Google Scholar] [CrossRef]
  249. Rahman, K.; Fahimi, B.; Suresh, G.; Rajarathnam, A.; Ehsani, M. Advantages of switched reluctance motor applications to EV and HEV: Design and control issues. IEEE Trans. Ind. Appl. 2000, 36, 111–121. [Google Scholar] [CrossRef]
  250. Boldea, I.; Tutelea, L.N.; Parsa, L.; Dorrell, D. Automotive electric propulsion systems with reduced or no permanent magnets: An overview. IEEE Trans. Ind. Electron. 2014, 61, 5696–5711. [Google Scholar] [CrossRef]
  251. Ruba, M.; Viorel, I.; Szabó, L. Modular stator switched reluctance motor for fault tolerant drive systems. IET Electr. Power Appl. 2013, 7, 159–169. [Google Scholar] [CrossRef]
  252. Model 3|Tesla’. Available online: https://www.tesla.com/model3 (accessed on 30 July 2024).
  253. Wu, C.-Y.; Pollock, C. Analysis and reduction of vibration and acoustic noise in the switched reluctance drive. IEEE Trans. Ind. Appl. 1995, 31, 91–98. [Google Scholar] [CrossRef]
  254. Li, H.; Bilgin, B.; Emadi, A. An Improved Torque Sharing Function for Torque Ripple Reduction in Switched Reluctance Machines. IEEE Trans. Power Electron. 2019, 34, 1635–1644. [Google Scholar] [CrossRef]
  255. Inderka, R.; De Doncker, R.; Krehenbrink, M. On-line estimation of instantaneous torque in switched reluctance machine control. In Proceedings of the 2000 International Symposium on Industrial Electronics (ISIE 2000), Cholula, Puebla, Mexico, 4–8 December 2000; pp. 385–389. [Google Scholar] [CrossRef]
  256. Fantin, E.; Raj, I.; Appadurai, M. Minimization of Torque Ripple and Incremental of Power Factor in Switched Reluctance Motor Drive. In Recent Trends in Communication and Intelligent Systems; Springer: Singapore, 2021; pp. 125–133. [Google Scholar] [CrossRef]
  257. Fuengwarodsakul, N.; Menne, M.; Inderka, R.; De Doncker, R. High-dynamic four-quadrant switched reluctance drive based on DITC. IEEE Trans. Ind. Appl. 2005, 41, 1232–1242. [Google Scholar] [CrossRef]
  258. Morimoto, S. Trend of permanent magnet synchronous machines. IEEJ Trans. Electr. Electron. Eng. 2007, 2, 101–108. [Google Scholar] [CrossRef]
  259. Widmer, J.D.; Martin, R.; Mecrow, B.C. Optimization of an 80-kW Segmental Rotor Switched Reluctance Machine for Automotive Traction. IEEE Trans. Ind. Appl. 2015, 51, 2990–2999. [Google Scholar] [CrossRef]
  260. Chiba, A.; Kiyota, K.; Hoshi, N.; Takemoto, M.; Ogasawara, S. Development of a rare-earth-free SR motor with high torque density for hybrid vehicles. IEEE Trans. Energy Convers. 2015, 30, 175–182. [Google Scholar] [CrossRef]
  261. Mecrow, B.; Jack, A.; Atkinson, D.; Green, S.; Atkinson, G.; King, A.; Green, B. Design and testing of a four-phase fault-tolerant permanent-magnet machine for an engine fuel pump. IEEE Trans. Energy Convers. 2004, 19, 671–678. [Google Scholar] [CrossRef]
  262. Woolmer, T.J.; McCulloch, M.D. Analysis of the yokeless and segmented armature machine. In Proceedings of the IEEE International Electric Machines and Drives Conference, IEMDC 2007, Antalya, Turkey, 3–5 May 2007; Volume 1, pp. 704–708. [Google Scholar] [CrossRef]
  263. Capponi, F.G.; De Donato, G.; Caricchi, F. Recent advances in axial-flux permanent-magnet machine technology. IEEE Trans. Ind. Appl. 2012, 48, 2190–2205. [Google Scholar] [CrossRef]
  264. Huang, S.; Lipo, T.; Aydin, M.; Lipo, T.A. Axial Flux Permanent Magnet Disc Machines: A Review. Conf. Rec. SPEEDAM 2004, 8, 61–71. [Google Scholar]
  265. Axial Flux Motors|Performance Automotive E-Motors|YASA Ltd. Available online: https://yasa.com/technology/ (accessed on 30 July 2024).
  266. Sergeant, P.; Vansompel, H.; Dupre, L. Performance and implementation issues considering the use of thin laminated steel sheets in segmented armature axial-flux PM machines. In Proceedings of the 2014 XXI International Conference on Electrical Machines (ICEM), Berlin, Germany, 2–5 September 2014; pp. 1363–1369. [Google Scholar] [CrossRef]
  267. Cavagnino, A.; Lazzari, M.; Profumo, F.; Tenconi, A. A comparison between the axial flux and the radial flux structures for PM synchronous motors. IEEE Trans. Ind. Appl. 2002, 38, 1517–1524. [Google Scholar] [CrossRef]
  268. Howey, D.A.; Childs, P.R.N.; Holmes, A.S. Air-gap convection in rotating electrical machines. IEEE Trans. Ind. Electron. 2012, 59, 1367–1375. [Google Scholar] [CrossRef]
  269. Vansompel, H.; Sergeant, P.; Dupre, L.; Bossche, A. A combined wye-delta connection to increase the performance of axial-flux PM machines with concentrated windings. IEEE Trans. Energy Convers. 2012, 27, 403–410. [Google Scholar] [CrossRef]
  270. Bianchi, N.; Bolognani, S.; Luise, F. Potentials and limits of high-speed PM motors. IEEE Trans. Ind. Appl. 2004, 40, 1570–1578. [Google Scholar] [CrossRef]
  271. Parviainen, A.; Niemela, M.; Pyrhonen, J. Modeling of axial flux permanent-magnet machines. IEEE Trans. Ind. Appl. 2004, 40, 1333–1340. [Google Scholar] [CrossRef]
  272. Aydin, M.; Qu, R.; Lipo, T. Cogging Torque Minimization Technique for Multiple-Rotor, Axial-Flux, Surface-Mounted-PM Motors: Alternating Magnet Pole-Arcs in Facing Rotors. In Proceedings of the Conference Record of the 2003 IEEE Industry Applications Conference, 38th IAS Annual Meeting, Salt Lake City, UT, USA, 12–16 October 2003; pp. 555–561. [Google Scholar] [CrossRef]
  273. Kumar, S.N.; Karthikeyan, N.; Mathew, M.; Hossain, J. Development of permanent magnet axial flux generator for small wind turbines. In Proceedings of the 2017 IEEE International Conference on Circuits and Systems (ICCS), Thiruvananthapuram, India, 20–21 December 2017; pp. 43–48. [Google Scholar] [CrossRef]
  274. De Donato, G.; Capponi, F.G.; Caricchi, F. Fractional-slot concentrated-winding axial-flux permanent-magnet machine with core-wound coils. IEEE Trans. Ind. Appl. 2012, 48, 630–641. [Google Scholar] [CrossRef]
  275. Ferraris, L.; Ferraris, P.; Poskovic, E.; Tenconi, A. Theoretic and experimental approach to the adoption of bonded magnets in fractional machines for automotive applications. IEEE Trans. Ind. Electron. 2012, 59, 2309–2318. [Google Scholar] [CrossRef]
  276. Shi, Z.; Sun, X.; Cai, Y.; Yang, Z.; Lei, G.; Guo, Y.; Zhu, J. Torque Analysis and Dynamic Performance Improvement of a PMSM for EVs by Skew Angle Optimization. IEEE Trans. Appl. Supercond. 2019, 29, 1–5. [Google Scholar] [CrossRef]
  277. Lipo, T.A.; Du, Z.S. Synchronous motor drives-a forgotten option. In Proceedings of the 2015 Intl Aegean Conference on Electrical Machines & Power Electronics (ACEMP), 2015 Intl Conference on Optimization of Electrical & Electronic Equipment (OPTIM) & 2015 Intl Symposium on Advanced Electromechanical Motion Systems (ELECTROMOTION), Side, Turkey, 2–4 September 2015. [Google Scholar]
  278. Zhu, Z.Q.; Chu, W.Q.; Guan, Y. Quantitative comparison of electromagnetic performance of electrical machines for HEVs/EVs. CES Trans. Electr. Mach. Syst. 2020, 1, 37–47. [Google Scholar] [CrossRef]
  279. Dai, J.; Hagen, S.; Ludois, D.C.; Brown, I.P. Synchronous Generator Brushless Field Excitation and Voltage Regulation via Capacitive Coupling Through Journal Bearings. IEEE Trans. Ind. Appl. 2017, 53, 3317–3326. [Google Scholar] [CrossRef]
  280. Ayub, M.; Jawad, G.; Kwon, B.-I. Consequent-pole hybrid excitation brushless wound field synchronous machine with fractional slot concentrated winding. IEEE Trans. Magn. 2019, 55, 1–5. [Google Scholar] [CrossRef]
  281. Ali, Q.; Lipo, T.A.; Kwon, B.-I. Design and Analysis of a Novel Brushless Wound Rotor Synchronous Machine. IEEE Trans. Magn. 2015, 51, 1–4. [Google Scholar] [CrossRef]
  282. Hussain, A.; Kwon, B.-I. A new brushless wound rotor synchronous machine using a special stator winding arrangement. Electr. Eng. 2018, 100, 1797–1804. [Google Scholar] [CrossRef]
  283. Hussain, A.; Atiq, S.; Kwon, B.-I. Optimal Design and Experimental Verification of Wound Rotor Synchronous Machine Using Subharmonic Excitation for Brushless Operation. Energies 2018, 11, 554. [Google Scholar] [CrossRef]
  284. Rafin, S.M.S.H.; Ali, Q.; Khan, S.; Lipo, T.A. A novel two-layer winding topology for sub-harmonic synchronous machines. Electr. Eng. 2022, 104, 3027–3035. [Google Scholar] [CrossRef]
  285. Rafin, S.M.S.H.; Ali, Q.; Lipo, T.A. A Novel Sub-Harmonic Synchronous Machine Using Three-Layer Winding Topology. World Electr. Veh. J. 2022, 13, 16. [Google Scholar] [CrossRef]
  286. Rafin, S.M.S.H.; Mohammed, O.A. Novel Dual Inverter Sub-Harmonic Synchronous Machines. IEEE Trans. Magn. 2023, 59, 1–5. [Google Scholar] [CrossRef]
  287. Rafin, S.M.S.H.; Ali, Q.; Mohammed, O.A. Novel PM-Assisted Model of the Two-Layer Sub-Harmonic Synchronous Machines. In Proceedings of the 2023 International Applied Computational Electromagnetics Society Symposium (ACES), Monterey/Seaside, CA, USA, 26–30 March 2023. [Google Scholar] [CrossRef]
  288. Rafin, S.M.S.H.; Ali, Q.; Mohammed, O.A. Hybrid Sub-Harmonic Synchronous Machines Using Series and Parallel Consequent Permanent Magnet. In Proceedings of the 2023 IEEE Energy Conversion Congress and Exposition (ECCE), Nashville, TN, USA, 29 October–2 November 2023; pp. 4163–4167. [Google Scholar] [CrossRef]
  289. Rafin, S.M.S.H.; Ali, Q.; Mohammed, O.A. Dual Inverter Parallel Consequent Pole PM-Assisted Two-Layer Sub-Harmonic Synchronous Machine. In Proceedings of the 2024 IEEE 21st Biennial Conference on Electromagnetic Field Computation (CEFC), Jeju, Republic of Korea, 2–5 June 2024; pp. 1–2. [Google Scholar] [CrossRef]
  290. Rafin, S.M.S.H.; Ali, Q.; Zhao, F.; Mohammed, O.A. Hybrid Three-Layer Sub-Harmonic Synchronous Machine using Consequent Pole Permanent Magnets. In Proceedings of the 2024 IEEE International Magnetic Conference—Short Papers (INTERMAG Short Papers), Rio de Janeiro, Brazil, 5–10 May 2024. [Google Scholar] [CrossRef]
  291. Burress, T.; Campbell, S. Benchmarking EV and HEV Power Electronics and Electric Machines. In Proceedings of the 2013 IEEE Transportation Electrification Conference and Expo: Components, Systems, and Power Electronics—From Technology to Business and Public Policy, ITEC 2013, Detroit, MI, USA, 16–19 June 2013. [Google Scholar] [CrossRef]
  292. Ramesh, P.; Lenin, N.C. High Power Density Electrical Machines for Electric Vehicles-Comprehensive Review Based on Material Technology. IEEE Trans. Magn. 2019, 55, 0900121. [Google Scholar] [CrossRef]
  293. Miller, J.M. Oak Ridge National Laboratory Annual Progress Report for the Power Electronics and Electric Motors Program; The National Academies of Sciences, Engineering, and Medicine: Washington, DC, USA, 2013. [Google Scholar]
  294. Sarlioglu, B.; Morris, C.T.; Han, D.; Li, S. Driving Toward Accessibility: A Review of Technological Improvements for Electric Machines, Power Electronics, and Batteries for Electric and Hybrid Vehicles. IEEE Ind. Appl. Mag. 2017, 23, 14–25. [Google Scholar] [CrossRef]
  295. Sutopo, W.; Nizam, M.; Rahmawatie, B.; Fahma, F. A Review of Electric Vehicles Charging Standard Development: Study Case in Indonesia. In Proceedings of the 2018 5th International Conference on Electric Vehicular Technology (ICEVT), Surakarta, Indonesia, 30–31 October 2018; pp. 152–157. [Google Scholar] [CrossRef]
  296. Wang, B.; Dehghanian, P.; Wang, S.; Mitolo, M. Electrical Safety Considerations in Large-Scale Electric Vehicle Charging Stations. IEEE Trans. Ind. Appl. 2019, 55, 6603–6612. [Google Scholar] [CrossRef]
  297. Kumar, S.; Usman, A.; Rajpurohit, B.S. Battery charging topology, infrastructure, and standards for electric vehicle applications: A comprehensive review. IET Energy Syst. Integr. 2021, 3, 381–396. [Google Scholar] [CrossRef]
  298. Grazian, F.; Shi, W.; Dong, J.; van Duijsen, P.; Soeiro, T.B.; Bauer, P. Survey on standards and regulations for wireless charging of electric vehicles. In Proceedings of the 2019 AEIT International Conference of Electrical and Electronic Technologies for Automotive (AEIT AUTOMOTIVE), Turin, Italy, 2–4 July 2019. [Google Scholar] [CrossRef]
  299. Khalid, M.R.; Khan, I.A.; Hameed, S.; Asghar, M.S.J.; Ro, J.-S. A Comprehensive Review on Structural Topologies, Power Levels, Energy Storage Systems, and Standards for Electric Vehicle Charging Stations and Their Impacts on Grid. IEEE Access 2021, 9, 128069–128094. [Google Scholar] [CrossRef]
  300. Das, H.S.; Rahman, M.M.; Li, S.; Tan, C.W. Electric vehicles standards, charging infrastructure, and impact on grid integration: A technological review. Renew. Sustain. Energy Rev. 2020, 120, 109618. [Google Scholar] [CrossRef]
  301. Rajendran, G.; Vaithilingam, C.A.; Misron, N.; Naidu, K.; Ahmed, R. A comprehensive review on system architecture and international standards for electric vehicle charging stations. J. Energy Storage 2021, 42, 103099. [Google Scholar] [CrossRef]
Figure 1. CO2 emissions by sectors [5].
Figure 1. CO2 emissions by sectors [5].
Electronics 13 03578 g001
Figure 2. Electric car sales by region, 2021–2023 [13].
Figure 2. Electric car sales by region, 2021–2023 [13].
Electronics 13 03578 g002
Figure 3. Main parts of the EV.
Figure 3. Main parts of the EV.
Electronics 13 03578 g003
Figure 4. Ragone plot shows the energy vs. power density comparison of multiple ESS.
Figure 4. Ragone plot shows the energy vs. power density comparison of multiple ESS.
Electronics 13 03578 g004
Figure 5. Comparison of various commercial LIBs.
Figure 5. Comparison of various commercial LIBs.
Electronics 13 03578 g005
Figure 6. Main objectives of the BMS.
Figure 6. Main objectives of the BMS.
Electronics 13 03578 g006
Figure 7. Classifications of LIBs SOC estimation methods.
Figure 7. Classifications of LIBs SOC estimation methods.
Electronics 13 03578 g007
Figure 8. EV charging method classifications.
Figure 8. EV charging method classifications.
Electronics 13 03578 g008
Figure 9. Conductive charger types: (a) off-board charger, (b) on-board charger, and (c) integrated on-board charger.
Figure 9. Conductive charger types: (a) off-board charger, (b) on-board charger, and (c) integrated on-board charger.
Electronics 13 03578 g009
Figure 10. Schematic diagram of inductive coupling chargers.
Figure 10. Schematic diagram of inductive coupling chargers.
Electronics 13 03578 g010
Figure 11. Schematic diagram of inductive resonant chargers.
Figure 11. Schematic diagram of inductive resonant chargers.
Electronics 13 03578 g011
Figure 12. Schematic diagram of capacitive chargers.
Figure 12. Schematic diagram of capacitive chargers.
Electronics 13 03578 g012
Figure 13. Schematic diagram of a magnetic gear wireless charger.
Figure 13. Schematic diagram of a magnetic gear wireless charger.
Electronics 13 03578 g013
Figure 14. Control methods for WPTS.
Figure 14. Control methods for WPTS.
Electronics 13 03578 g014
Figure 15. An induction motor used in the Audi Q8 e-tron [203].
Figure 15. An induction motor used in the Audi Q8 e-tron [203].
Electronics 13 03578 g015
Figure 16. Permanent Magnet Synchronous Motor [234].
Figure 16. Permanent Magnet Synchronous Motor [234].
Electronics 13 03578 g016
Figure 17. Internal permanent magnet synchronous reluctance motor (IPM-SynRM) used in Tesla Model 3 [223].
Figure 17. Internal permanent magnet synchronous reluctance motor (IPM-SynRM) used in Tesla Model 3 [223].
Electronics 13 03578 g017
Figure 18. Axial flux motor for electric vehicle application [265].
Figure 18. Axial flux motor for electric vehicle application [265].
Electronics 13 03578 g018
Figure 19. Sub-harmonic synchronous machine layout: (a) 2-layer subharmonic synchronous machine and (b) 3-layer subharmonic synchronous machine [284,285].
Figure 19. Sub-harmonic synchronous machine layout: (a) 2-layer subharmonic synchronous machine and (b) 3-layer subharmonic synchronous machine [284,285].
Electronics 13 03578 g019
Table 1. Comparison between the common available EV types in the market.
Table 1. Comparison between the common available EV types in the market.
ParametersBEVHEVPHEVFCEV
Source of EnergyESSFossil Fuel/ESSFossil Fuel/ESSHydrogen/ESS
PowertrainElectric MotorIC engine/Electric MotorIC engine/Electric MotorElectric Motor
EmissionNoYesYesNo
NoiseLowMedianMedianLow
All-electric range100–250 milesshort5–50 miles100–300+ miles
Grid connectedYesNoYes
PriceHighMedianMedianHighest
Average top speed94120120100
Charging Time (min)60–4505–1205–1205
Major Issues
  • Battery sizing and management
  • Charging Facilities
  • Battery life cycle
  • Cost
  • Battery sizing and management
  • Control and optimization management of multiple energy resources
  • Fuel cell cost, life cycle, and reliability
  • Hydrogen production
  • Cost
Table 2. Comparison between the commercial chemistries of LIBs.
Table 2. Comparison between the commercial chemistries of LIBs.
ParametersLFPLMONMCNCALTO
Year19931996200819992008
AnodeGraphite carbonGraphite carbonGraphite carbonGraphite carbonLithium titanate
CathodeLithium iron phosphateLithium manganese oxideLithium nickel manganese cobalt oxideLithium nickel cobalt aluminum oxideLithium manganese oxide
SeparatorPolyolefinPolyolefinPolyolefinPolyolefinPolyolefin
Specific energy (Wh/kg)90–150100–150150–220200–26050–80
Thermal
Runaway (°C)
270250210150Highest
Charge (C-rate)10.7–10.7–10.71
Discharge
(C-rate)
111–2110
Nominal Voltage (V)3.2–3.33.7–4.03.8–4.03.6–3.652.3–2.5
WeightLightHeavyLightLightLight
Self-discharge (monthly %)5110-5
Application
  • Industrial
  • Medical
  • EVs
  • Medical
  • Powertrains
  • Power tools
  • Industrial
  • Medical
  • EVs
  • E-bikes
  • Industrial
  • Medical
  • EVs
  • UPS
  • EVs
  • Solar lights
Table 3. Summary of the commonly available batteries in the EV market.
Table 3. Summary of the commonly available batteries in the EV market.
ParametersLead-AcidNiMHLi-Ion
SafetyThermally stableThermally stable with a protective fuseProtection circuit and management are mandated
Specific energy (Wh/kg)30–5060–12090–250
Specific power (w/kg)180Up to 13002500
Cell voltage (V)2.01.23.2–3.7
Life cycle (y)200–300300–500500–2000
Efficiency (%)9090.399
CostLowModeratehigh
Charge time (h)8–162–41–4
Table 4. Comparison among the different supercapacitors’ chemistries.
Table 4. Comparison among the different supercapacitors’ chemistries.
ParametersElectric Double-LayerPseudo-CapacitorHybrid Capacitor
Electrode materialActivated carbonMetal oxidesCarbon/metal oxide
Charge storage mechanismsCharge separationRedox charge transferDouble layer/charge transfer
Specific energy (Wh/kg)5–710–1510–15
Specific power (w/kg)1000–30001000–20001000–2000
Cell voltage (V)2.5–32–3.52–3.3
Life cycle (y)404040
Efficiency (%)>95>95>95
Table 5. Key differences between the different fuel cells.
Table 5. Key differences between the different fuel cells.
ParametersPEMFCPAFCAFCSOFCMCFCDMFC
Temp (°C)<100150–20090–100500–1000600–70060–200
Cell
Voltage (V)
1.11.110.8–10.7–10.2–0.4
Stack power (kW)<1–25050–1001–100<1–3000300–30000.001–100
Efficiency (%)40–604060605040
Advantages
  • Low operating temperature
  • Fast starting
  • Less corrosion and electrolyte management problem
  • Higher efficiency with CHP
  • Increased tolerance in fuel impurities
  • Higher performance
  • Faster cathode reaction and start
  • Lower material cost
  • Low operating temperature
  • Suitable for CHP
  • Higher efficiency
  • Fuel flexibility
  • Hybrid/gas turbine cycle
  • Suitable for CHP
  • Higher efficiency
  • Fuel flexibility
  • Low cost due to lack of fuel reformer
Disadvantages
  • Expensive catalyst cost
  • Sensitive to hydrogen impurity
  • Expensive catalyst cost
  • Long start time
  • Sensitive to CO2 levels in oxygen and hydrogen
  • High operation temperature, corrosion, and breakdown
  • Long start time
  • High operation temperature
  • Long start
  • Power density
  • Intermediates adhesion to the catalyst surface
Table 6. Comparison between common estimation techniques [128].
Table 6. Comparison between common estimation techniques [128].
TechniqueAdvantagesDisadvantages
Ampere Hour
  • Simple
  • Low computational burden
  • Independent of battery models
  • Aging state is considered
  • An accurate initial value of SOC is needed
  • Error accumulation
  • Battery lifespan is affected by the maximum capacity-determining process
Look-up table-basedOCV
  • Low computational burden
  • Cost-effective
  • Independent of battery models
  • Aging state is considered
  • Suitable only at the high and low SOC levels.
  • Needs a long relaxation time
  • The estimated SOC value is largely affected by the battery hysteresis
EIS
  • Low cost
  • High estimation accuracy
  • Reflects the dynamic characteristics
  • Affected by aging and temperature
  • Influenced by battery hysteresis
  • Varies with battery type and test conditions
Model-Based
  • Insensitive to initial SOC
  • Simulate the capacity fading
  • High correlation between model parameter and estimated accuracy
  • Rely on modeling accuracy
Data Driven
  • Ability to work in battery nonlinear conditions
  • Independent of battery model or characteristics
  • Real-time detection of SOC status
  • High precision
  • Required massive data for model training
  • Large computational efforts
  • Model cannot be generalized
  • Adaption to aging state is not possible
Table 7. EV charging levels.
Table 7. EV charging levels.
Power Level TypesCharger LocationTypical UsePower LevelCharging Time
Level 1 (Opportunity/Slow)
120 V (US) or 230V (EU)
On-board
1-phase
Charging at
home or office
  • 1.4 kW (12 A)
  • 1.9 kW (20 A)
  • 4–11 h
  • 11–36 h
Level 2 (Primary)
240 V (US)
300V (EU)
On/off-board
1- or 3-phase
Charging at private or public outlets
  • 4 kW (17 A)
  • 8 kW (32 A)
  • 19.2 kW (80 A)
  • 1–4 h
  • 2–6 h
  • 2–3 h
Level 3 (Fast)
(208–600 Vac or Vdc)
Off/on-board
3-phase
Commercial, analogous to a filling station
  • 50 kW
  • 100 kW
  • 0.4–1 h
  • 0.2–0.5 h
Extreme Fast Charging (XFC)
1000 Vdc and above
Off-boardCommercial>350 kW~5 min
Table 8. Comparison between charger types [19,180,181,182].
Table 8. Comparison between charger types [19,180,181,182].
FeatureOnboard ChargersOffboard ChargersWireless Chargers
Control Topologies
  • Constant Voltage (CV)
  • Constant Curent (CC)
  • Combination of CV and CC
  • Fast Charging Techniques
  • High Power DC Charging (CHAdeMO, CCS)
  • Resonant Inductive Coupling
  • Magnetic Resonance Coupling
  • Frequency Control
  • Phase Control
  • Dynamic Load Matching
Power Conversion Stages
  • AC-DC: Rectification Techniques
  • DC-DC: Buck, Boost, Buck-Boost Converters
  • AC-DC: Advanced Rectification Techniques
  • DC-DC: High Power Converters, Inverter Technologies
  • Power Electronics: Rectifiers and Inverters
Communication Protocols
  • CAN (Controller Area Network)
  • LIN (Local Interconnect Network)
  • PLC (Power Line Communication)
  • OCPP (Open Charge Point Protocol)
  • Nearfield Communication (NFC)
  • Standard Communication Protocols
Efficiency
  • Moderate Efficiency
  • Dependent on control strategy and power electronics
  • High Efficiency
  • Advanced control strategies optimize performance
  • Lower Efficiency compared to wired methods
  • Efficiency decreases with misalignment and air gap size
Performance
  • Good for regular overnight charging
  • Limited by onboard power conversion capabilities
  • Excellent for rapid, high-power charging
  • High power delivery, up to hundreds of kW
  • Suitable for convenience and ease of use
  • Challenges in maintaining
Cost
  • Integrated into the vehicle, increasing vehicle cost
  • Lower infrastructure cost
  • High infrastructure cost
  • Requires substantial installation and maintenance
  • High implementation cost
  • Cost of coils and alignment systems
Practical Applications
  • Home Charging
  • Work Charging
  • Public Fast Charging Stations
  • Fleet Operations
  • Home and Public Wireless Charging
  • Dynamic Charging (In Motion Charging)
Advantages
  • Convenient for daily use
  • Integrated into the vehicle
  • Rapidly reduces charging time
  • Supports long-distance travel
  • Ultimate convenience with no physical connection
  • Reduced wear on connectors and ports
Challenges
  • Limited by the vehicle’s onboard systems
  • Complex control algorithms
  • Requires precise monitoring and switching
  • High infrastructure and installation costs
  • Potential thermal management issues
  • Impact on battery longevity
  • Lower efficiency, alignment, and air gap challenges
  • More complex control strategies
  • Higher implementation costs
Charger ManufacturerConventional On-board
  • Chevy Volt [183]
  • Nissan Leaf [184]
  • Delta Q [185]
Integrated on-board
  • Valeo Dual inverter [186]
  • Renault Chameleon [187]
  • Tesla Supercharger V2 [188]
  • ABB Tera HP 150-kW [189]
  • Enercon E-charger [190]
  • Porsche Modular Fast Charging Park A [191]
Static
Dynamic
Table 9. Comparison of control methods for WPT [207].
Table 9. Comparison of control methods for WPT [207].
Control StrategyAdvantagesDisadvantages
Primary Side Control
  • Maintains constant current for the transmitter.
  • Load-independent control
  • Eliminates the need for communication links between primary and secondary
  • Lower system efficiency due to load independence
Secondary Side Control
  • Provides constant current control
  • Enables implementation of maximum efficiency control
  • Necessitates controlled power electronics on the secondary side
Dual Side Control with Communication
  • Achieves desired power control and maximum efficiency
  • Supports bidirectional power flow
  • Requires secure and minimal delay wireless communication between primary and secondary
Dual Side Control without Communication
  • Allows independent control of primary and secondary
  • Facilitates desired power with maximum efficiency
  • Potential stability issues due to conflicts between primary and secondary control
Table 10. Comparison between electric motors used in EV applications.
Table 10. Comparison between electric motors used in EV applications.
Motor TypeKey CharacteristicsAdvantagesChallenges
IM
  • Asynchronous operation
  • Rotor with short-circuited conductors
  • Robust and reliable
  • Cost-effective
  • Wide speed range
  • No rare-earth materials
  • Good field-weakening capability
  • Lower efficiency at low speeds
  • Significant rotor losses at high speeds
  • Lower power factor
PMSM
  • Permanent magnets in the rotor
  • Synchronous operation
  • High efficiency (>95%)
  • Excellent power density
  • Wide speed range
  • Accurate torque and speed control
  • Reliance on rare-earth materials
  • Risk of demagnetization
  • Limited constant power speed range
SynRM and
SRM
  • No permanent magnets
  • Salient pole rotor structure
  • Simple and robust construction
  • Lower manufacturing costs
  • High-speed operation capability
  • Fault tolerance
  • Acoustic noise and vibration
  • Torque ripple
  • Complex control requirements
Axial Flux
  • Disc-shaped structure
  • Axial magnetic flux path
  • High torque density
  • Compact axial length
  • Efficient cooling
  • Scalable design
  • Manufacturing challenges
  • Strong axial forces
  • Potential for high cogging torque
SHSM
  • Brushless wound rotor design
  • Sub-harmonic MMF excitation
  • Tunable excitation
  • High power density
  • Efficient operation
  • No need for rare-earth materials
  • Relatively new technology
  • Complex winding configurations
Table 11. Key performance requirements for electric motors in EV applications.
Table 11. Key performance requirements for electric motors in EV applications.
Performance MetricTypical RequirementImportance
Power Density>5 kW/kgHigh
Torque Density>20 Nm/kgHigh
Peak Efficiency>95%High
Wide Speed Range0–15,000 RPMMedium to High
Constant Power Speed Range>3:1Medium
Torque Ripple<5%Medium
Overload Capability150–200% for short durationsMedium
Noise and Vibration<80 dBAMedium
Thermal ManagementOperating temp < 150 °CHigh
Reliability>150,000 miles/15 yearsHigh
Cost< USD 10/kWHigh
Table 12. Battery standards.
Table 12. Battery standards.
StandardDescription
IEC 62133Developed by the International Electrotechnical Commission (IEC), this standard specifies safety requirements for portable sealed secondary lithium cells and batteries, covering electrical, mechanical, thermal criteria, and testing procedures.
UL 1642Underwriters Laboratories (UL) created this safety standard for lithium batteries. It includes various tests to evaluate battery safety, often required for lithium-ion batteries in consumer electronics.
UN 38.3Established by the United Nations, these safety standards are crucial for transporting lithium-ion batteries. They involve multiple tests assessing resistance to environmental factors such as mechanical stress, vibration, and temperature changes, and compliance is essential for international shipping.
ISO 12405This standard pertains to the testing and characterizing lithium-ion cells and modules used in electric vehicles, addressing electrical properties, mechanical integrity, thermal behavior, and abuse testing, along with performance, safety, and environmental considerations.
IEEE 1625Issued by the Institute of Electrical and Electronics Engineers (IEEE), this standard covers rechargeable lithium-ion battery packs in portable computing devices, focusing on safety issues like overcharging, over-discharging, and short-circuit prevention to minimize the risk of battery-related incidents.
NFPA 70Published by the National Fire Protection Association (NFPA) as part of the National Electrical Code (NEC), this standard governs the installation of energy storage systems, including lithium-ion batteries, specifying fire safety, ventilation, electrical, and other safety measures.
Table 13. Communication and safety standards.
Table 13. Communication and safety standards.
StandardDescription
SAE J2293Details the communication and network requirements necessary for efficient EV charging infrastructure. This standard ensures interoperability and seamless data exchange within the charging network.
SAE J2344Establishes comprehensive safety guidelines for electric vehicles, focusing on minimizing risks during operation and charging. It covers aspects such as electrical safety, crashworthiness, and thermal management.
SAE J2847Defines the specific communication protocols required between various components of the EV charging system. This includes interactions between the EV, charging station, and the grid to ensure effective and safe power transfer.
IEC 63110Sets the management protocol for EV charging and discharging infrastructure, ensuring standardized communication for energy management, billing, and user authentication across different systems and platforms.
IEC 62752Outlines standards for cable control and protection devices used in EV charging, including requirements for detecting faults, disconnecting power, and ensuring user safety.
IEC 61851Specifies comprehensive safety regulations for EV charging stations, covering electrical safety, operational requirements, and compatibility with different EV models to ensure safe and efficient charging.
ISO 15118Defines the communication protocols between an EV and a charging station, facilitating features like Plug and Charge. This standard also includes protocols for vehicle-to-grid (V2G) interactions, enabling bi-directional energy flow.
ISO 17409Specifies safety requirements for conductive charging systems in electric vehicles, focusing on ensuring safe operation during connection, charging, and disconnection.
ISO 18246Sets safety standards for charging electric motorcycles and mopeds, addressing risks associated with electrical faults, overheating, and improper connections.
GB/T 27930-2015Establishing the communication protocol between off-board conductive chargers and the EV battery management system ensures coordinated and efficient charging.
NEC 625, NEC 626Specifies safety measures for off-board EV charging, including requirements for electrical installation, fire safety, and grounding.
UL2231, UL2251, UL2202Defines protection device requirements for EV charging equipment, covering aspects such as overcurrent protection, isolation, and durability to enhance safety.
CHAdeMOSpecifies the communication requirements for DC fast charging, ensuring that EVs and charging stations can effectively exchange data to manage high-power charging sessions safely.
Table 14. Connectors standards.
Table 14. Connectors standards.
StandardDescription
SAE J1772Defines the general physical requirements for conductive charging connectors used in electric vehicles. This includes specifications for plug design, durability, and safety features to ensure reliable and secure connections.
IEC 62196Specifies the requirements and testing procedures for plugs, socket outlets, vehicle connectors, and inlets for conductive charging. This standard ensures compatibility and safety across different EV models and charging stations.
GB/T
20234.1-2015
Details the physical requirements for connectors and interfaces used in EV charging, aligning with international standards such as IEC 62196 and SAE J1772 to promote global interoperability.
CHAdeMOSets the physical requirements for DC fast charging connectors, ensuring they can handle high currents and voltages while maintaining safety and reliability during fast charging sessions.
Table 15. System requirements standards.
Table 15. System requirements standards.
StandardDescription
SAE J1772Specifies the general electrical and performance requirements for conductive charging systems in electric vehicles. It covers aspects such as voltage levels, current capacity, and system architecture to ensure efficient and safe charging.
SAE J1773Provides standards for inductively coupled charging systems, detailing the requirements for wireless power transfer between the charging pad and the EV, including alignment, efficiency, and safety measures.
SAE J2293Establishes standards for both on-board and off-board charging equipment, covering conductive AC and DC charging and inductive charging. It also defines power requirements and system architecture for these charging methods.
SAE L2894/2Specifies requirements for power quality in EV charging systems, ensuring that the power supplied meets specific standards for voltage stability, harmonics, and overall quality to protect both the vehicle and the grid.
SAE J2953Sets standards for the interoperability of electric vehicles with the grid, ensuring that EVs can effectively communicate and interact with different types of charging infrastructure for seamless charging experiences.
SAE J2954Provides guidelines for wireless power transfer (WPT) for light-duty plug-in/electric vehicles, including alignment methodology and system requirements to ensure efficient and safe wireless charging.
SAE J3068Specifies the requirements for three-phase AC charging of electric vehicles, detailing the necessary electrical parameters and system configurations to support high-power AC charging.
IEC 60038Specifies voltage standards for EV charging applications, ensuring that the voltage levels used in charging systems are compatible with both the EVs and the grid infrastructure.
IEC 60664-1Provides standards for installation coordination of EV charging equipment in low voltage supply systems, covering insulation requirements and protection against electrical surges.
IEC 61851Specifies the general requirements for conductive charging systems for electric vehicles, including safety, compatibility, and performance standards to ensure reliable and efficient charging.
IEC 61980Defines wireless power transfer (WPT) systems standards, specifying requirements for systems operating up to 1000 V AC and 1500 V DC to ensure safe and effective wireless charging.
GB/T
18487.1-2015
Sets general requirements for conductive charging systems, aligning with IEC 61851 to ensure compatibility and safety across different EV charging infrastructures.
GB/T
20234.2-2015
Specifies the parameters for AC charging interfaces, including rated voltage, current, and maximum power to ensure reliable and efficient AC charging.
GB/T
20234.3-2015
Defines the parameters for DC charging interfaces, specifying rated voltage and current to support high-power DC fast charging.
UL2594, UL1741Specifies requirements for EV supply equipment concerning converters and controllers, ensuring that these components meet safety and performance standards.
CHAdeMODetails the electrical requirements for DC fast charging, ensuring EVs can safely handle high-power charging sessions.
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.

Share and Cite

MDPI and ACS Style

Hussein, H.M.; Ibrahim, A.M.; Taha, R.A.; Rafin, S.M.S.H.; Abdelrahman, M.S.; Kharchouf, I.; Mohammed, O.A. State-of-the-Art Electric Vehicle Modeling: Architectures, Control, and Regulations. Electronics 2024, 13, 3578. https://doi.org/10.3390/electronics13173578

AMA Style

Hussein HM, Ibrahim AM, Taha RA, Rafin SMSH, Abdelrahman MS, Kharchouf I, Mohammed OA. State-of-the-Art Electric Vehicle Modeling: Architectures, Control, and Regulations. Electronics. 2024; 13(17):3578. https://doi.org/10.3390/electronics13173578

Chicago/Turabian Style

Hussein, Hossam M., Ahmed M. Ibrahim, Rawan A. Taha, S. M. Sajjad Hossain Rafin, Mahmoud S. Abdelrahman, Ibtissam Kharchouf, and Osama A. Mohammed. 2024. "State-of-the-Art Electric Vehicle Modeling: Architectures, Control, and Regulations" Electronics 13, no. 17: 3578. https://doi.org/10.3390/electronics13173578

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop