Next Article in Journal
A Thermodynamic Analysis of an Air-Cooled Proton Exchange Membrane Fuel Cell Operated in Different Climate Regions
Next Article in Special Issue
Simple Power Quality Compensation with Bidirectional Battery Charger for Electric Vehicles in Single-Phase Three-Wire Distribution Feeders
Previous Article in Journal
Carbon Dioxide Emissions and Their Driving Forces of Land Use Change Based on Economic Contributive Coefficient (ECC) and Ecological Support Coefficient (ESC) in the Lower Yellow River Region (1995–2018)
Previous Article in Special Issue
A FACTS Devices Allocation Procedure Attending to Load Share
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Possibilities and Challenges for the Inclusion of the Electric Vehicle (EV) to Reduce the Carbon Footprint in the Transport Sector: A Review

Environment and Sustainability Institute (ESI), University of Exeter, Penryn Campus TR10 9FE, UK
Energies 2020, 13(10), 2602; https://doi.org/10.3390/en13102602
Submission received: 11 April 2020 / Revised: 15 May 2020 / Accepted: 17 May 2020 / Published: 20 May 2020
(This article belongs to the Special Issue Power Electronics and Power Quality 2019)

Abstract

:
To combat global climate change moving towards sustainable, mobility is one of the most holistic approaches. Hence, decarbonization of the transport sector by employing electric vehicles (EVs) is currently an environmentally benign and efficient solution. The EV includes the hybrid EV (HEV), the plug-in hybrid EV (PHEV), and the battery EV (BEV). A storage system, a charging station, and power electronics are the essential components of EVs. The EV charging station is primarily powered from the grid which can be replaced by a solar photovoltaic system. Wide uptake of EVs is possible by improving the technologies, and also with support from the government. However, greenhouse gas emission (GHG) saving potential of the EV is debatable when the required power to charge the EV comes from traditional fossil fuel sources.

1. Introduction

Unwanted emissions of greenhouse gases (GHG) from the burning of fossil fuel have now reached a threatening level which needs an immediate action of prevention by implementing environment-friendly climate policy. In 2015, the International Energy Agency (IEA) set a targeted future energy system scenario to limit the increment of average global temperature to 2 °C, which was later modified to 1.5 °C by 2050 [1]. In 2050, the world population is expected to be 9.8 billion, therefore, around 2 billion road vehicles are expected to be on the roads. Currently, over 90% of the global transport sector relies on oil, and 49% of oil production is consumed by the transport sector alone. Accounting for one-quarter of energy-related GHG emissions in 2009, the transport sector is the most rapidly growing energy consuming sector in the world. Hence, road vehicle electrification is essential to overcome environmental issues [2,3,4,5,6].
Electric vehicles (EVs) which emits no greenhouse gases are believed to be a promising solution to combat climate change and environmental pollution challenges. Robert Anderson first invented the EV using nonrechargeable primary cells between years 1832–1839 [7,8]. Later, other prototypes were invented which did not do well, as they lacked practical rechargeable battery and an electrically efficient motor. In the year 1900, EVs constituted 28% of road vehicles in New York City due to the enhancement of the rechargeable lead–acid battery and the DC electric motor. Until about 1918, EVs were popular, however, this popularity died out due to the presence of oil (gasoline). By 1933, the number of EVs fell to zero because of their slow speed and expensive internal combustion engine (ICE) [9]. ICE vehicles emit carbon dioxide, carbon monoxide, hydrocarbon, and sulphur oxides which results in global warming through greenhouse gas effects and pollution which are harmful to both the environment and humans. Hence, zero-emission vehicles (ZEV) are required as a pollution-prevention strategy. ZEV includes fuel cells and electric vehicles. However, the EV is further classified by the hybrid EV (HEV), the plug-in EV (PHEV), the battery EV (BEV) and the vehicle integrated EV, but only BEV satisfy the ZEV criteria [10,11].
Governments around the world are trying to implement policies, e.g., EV purchase cost incentives, developing EV-charging infrastructure, and enchaining public awareness for EV uptake. The aim of this paper is to present a comprehensive review of the following: details of the EV, the development of storage technologies for the EV, the powering of the EV by the photovoltaic system, challenges for EV uptake, and the present EV scenario worldwide.

2. Electric Vehicles and the Charging Station

The term “electric vehicle” counts any electrically powered vehicle, encompassing cars, electric bikes, motorbikes and other battery-powered vehicles. Hybrid electric vehicles (HEVs), plug-in hybrid electric vehicles (PHEVs) and battery electric vehicles (BEVs) are common types of EV. For PHEVs and BEVs types, the batteries can be externally recharged [12,13].

2.1. Hybrid EV (HEV) and Plug-In Hybrid

A supplementary electric motor and a gasoline internal-combustion engine power the hybrid electric vehicle (HEV) where the electric motor only participates in starting and accelerating the vehicle. HEVs have limited battery capacity which is charged by deacceleration or braking. HEVs are usually considered twice as fuel-efficient and half as polluting compared to conventional diesel- and petrol-powered vehicles because of their no power consumption from the grid. Series, parallel and series–parallel are the three types of propulsion sources employed for the HEV, as shown in Figure 1. In the series type HEV, an electric motor is the propulsion source and batteries are recharged by regenerative braking, generator and an ICE. The series HEV is appropriate for frequent stops and runs, which makes it suitable for city runs. The parallel HEV electric motor and ICE are both connected mechanically and provides vehicle propulsion by transmitting power. As the ICE and electric motor both complement each other, parallel HEVs are suitable for city driving pattern sand highway driving patterns. The Honda Insight and the Ford Escapes are examples of parallel HEV. The combined series–parallel HEV can be run in series or parallel mode. In this structure, the ICE and electric motor are mechanically coupled to wheels and transmission. This combined structure is costly and complicated. A commercial series–parallel HEV is the Toyota Prius.
A plug-in hybrid electric vehicle (PHEV) is a hybrid electric vehicle (HEV) which has the ability to recharge its electrochemical energy storage with electricity from an electric utility grid (off-board source) [14]. Plug-in hybrid electric vehicles are indicated by a “PHEVx” notation, where “x” typically denotes distance in miles which a fully charged PHEV can drive (all electric range) before it requires to operate its engine. A PHEV20 indicates that it can run 20 miles or 32 km all electrically before its first engine turn on [15]. They are usually equipped with a highly efficient ICE and a large capacity battery pack. The PHEV also has series, parallel and series–parallel power train configuration with an additional on-board battery charger. The PHEV has charge-depleting and charge-sustaining operating modes. The PHEV operates in the charge-depleting mode to start up the vehicle, while a low state of battery charge switches the vehicle into the charge-sustaining mode, consequently causing the ICE to begin operating [16].

2.2. BEV

Battery electric vehicles (BEVs) replace the internal combustion engine (ICEV) and the tank with an electric motor powered by a battery, as shown in Figure 2. The BEV is plugged into a charging spot when it is not in use. The BEV does not have conventional engines, fuel tanks, tailpipes and onboard electricity generation provision. Typically, BEVs have a range which is between 60–100+ mi. The absence of a tailpipe allows the BEV to be a tailpipe emissions-free vehicle. Overnight charging using low-cost electricity produced by any type of power station, including renewable, is possible for the BEV. The BEV also possesses sufficient acceleration. However, time consuming battery charging and expensive electricity storage limit the wide range of availability of these vehicles [17,18].
Absence of tail pipe emissions makes the BEV a potential candidate to help to meet the CO2 reduction targets. However, before implying BEVs as zero-emission vehicles, battery manufacturing and disposal and the carbon intensity of electricity generation must be considered [19,20,21]. Battery reuse is an effective option and it was found that the carbon emission reduction potential of reusing batteries is similar to that of moving from oiled based vehicles to the EV [22]. Before the wide market penetration, a charging spot infrastructure and the corresponding investment must be in place. A user satisfaction survey shows that the BEV is now gaining interest amongst consumers [23,24,25].

2.3. Charging Station (CS)

EV charging station controls the energy transfer to the vehicle’s battery [26]. Charging is possible based on power levels and charging mode. The Society of Automotive Engineers (SAE) and the International Electrotechnical Commission (IEC) have standardised the EV charger for different power levels and different charging modes [27]. In regard to power level, there are three levels of charging stations that are available to charge EVs. In general, battery capacity varies from 20 to 60 kWh. Charging level 1 consist a single-phase AC system with up to 3 kW charging power which will take about 7 h to charge a 20 kWh EV battery. For this type charging, the EV connector should have a ground fault interrupter and over current protection. They are commonly located at domestic households and workplace wall boxes. Charging level 2 has a 3-phase AC with up to 24 kW charging power, which needs 1 h to charge a 20 kWh EV battery. For charging level 2, the connector comes under IEC62196-2 type 1, IEC 62196-2 type 2 and GB/T 20234.2 standards. They are located mainly at public charging poles. Charging level 3 has 50 kW DC charging power which provides a fast charging level and charges an EV battery in 20–30 min. Connectors come under SAE J1772, CHAdeMo, and IEC 62196 type 2 standards [28,29]. Fast charging stations which are directly connected to the utility grid also have transformers and rectifiers to produce DC voltage for charging EV batteries in less than 20 min. Hence, a place for a fast charging station can work as an electric fuel station on the motorway (away from the main city) to supply EV energy needed in a short period of time [30,31]. Fast charging station facilities in the public area (main city) are required to spread the wide use of EVs [32]. However fast charging has stronger currents which produce more power losses during transfer and decrease battery lifetime, subsequently reducing the number of total charging cycles. Also, fast DC charging points are more expensive than the AC charger [33].
The location of the charging station for the EV is very essential criteria. Reduction of electric losses from EV charging stations is possible if they are installed near the electric substation [34]. However, most often electric substations are far from EV users (for fast charging, which is the first preference for the consumer). Thus, travel is required to recharge vehicles, which in turn produce EV energy loss. Hence, charging station capacity and the available places to charge a number of EVs are essential parameters in the analysis of economically feasible options [35]. When an EV is connected to a charging station, the control system operates based on several variables, such as the state of charge (SOC) of the battery, current irradiance levels, and the cost of energy from the grid. As EV-CSs are becoming more widespread, other variables such as time spent at the charging location will become important. Figure 3 shows a typical charging facility for EV.

2.4. EV Battery

The battery is one of the prime components of the EV, as they work as a source for the propulsion of the HEV and the PHEV, while as the major propulsion component for the BEV. Presently available battery technologies for the EV are lead–acid, nickel-based, and lithium-ion. Allowable temperature for a battery to work properly is about +15 °C to +30 °C, however, ambient temperature can vary between −35 °C to +50 °C due to different climates and regions. Furthermore, in a battery heat generation, heat transport, and heat dissipation are three processes which dominate the battery temperature [36]. Efficient thermal management is essential for an efficient EV battery [37].
The lead–acid battery is the oldest technology, discovered in 1859. In a lead–acid battery, negative terminal lead and positive terminal lead oxide both transform into lead sulphate during charging, and diluted sulfuric acid works as a liquid electrolyte. Lead–acid battery technology is the most mature one and has a low cost (100 USD/kWh). However, they also have low specific energy which varies between 20–40 Wh/kg, and they are not suitable when they discharge over 20% of their rated capacity. When operated at a deep rate of the state of charge (SOC), the battery would have a limited life cycle. The heavyweight lead collector lowers the energy and power density of the battery [38,39].
Nickel-based batteries include Ni-Fe, Ni-Cd, Ni-Zn, Ni-MH, and Ni-H2, where nickel hydroxide works as positive electrode and negative electrode materials. Recently Ni-Cd showed potential for EV application, having life cycles of over 2000 or more and energy density. However, Ni-Cd has a high cost which is ten times that of the price of a lead–acid battery. Environmentally benign nickel-metal hydride (NiMH) batteries are also applicable for BEVs and BHEVs due to their safe operation at a high voltage, flexible size. In addition, they are not very expensive and their energy density varies between 60–80 Wh/kg [40,41,42].
Li-ion technology possesses high energy density, high efficiency and a long lifespan [43,44]. Due to the rapid employment of this technology, battery price reduced to 85% compared to 2010 level and now achievable between 130 USD/kWh [45] and 176 USD/kWh [46]. Lithium-based batteries can be categorised as lithium-ion (Li-ion), lithium-ion polymer (LiPo) and lithium-ion phosphate (LiFeP04). The principle operation of lithium-based batteries is common across each category; positively charged lithium ions move between an anode and cathode through an electrolyte. However, they have a risk of fire and explosion during a malfunction [47,48,49]. More recently, developments in silicon-air and lithium-air batteries have produced an eco-friendly, nontoxic and, importantly for the EV industry, they are lightweight and low cost, making them substitute for conventional lithium technologies. This technology has the ability to be commercially available within four years [50]. At subzero temperatures (cold climate countries e.g., Canada, Russia and the Scandinavian Peninsula), Li-ion battery performance degrades. Energy and power density both drastically fall at this condition [51]. At low temperatures, charge transfer kinetics and lithium-ion diffusion become very slow, while the electrolyte conductivity also weakens [52,53,54,55]. It is also reported that for the same current, a Li-ion cell’s available energy at −20 °C is 60% of the room-temperature value [56].
At low temperature, starting the EV is impossible, unless batteries are preheated [57]. The required time to preheat a battery pack can be up to 10–15 min and this preheating process consumes energy from the batteries themselves [58]. The battery management system (BMS) protects a battery from overcharging, overuse and short circuiting. Thus, BMS regulates all engaging activities between the battery and the required load. Overcharging and overuse of the battery can cause excessive heat and can even cause an explosion or flame. BMS for Li-ion batteries are essential as they fail if they are overcharged, causing them to operate outside their safe temperature or completely discharge. Typically, BMS works with multiple responses, where it will initially try to bring back the battery pack voltage by changing power flow in/out from the battery pack. Failure of this step gives authority to BMS to open the battery pack and stop all power flow in/out of it. This step often disables a vehicle, which is an unsafe condition for the vehicle occupants [59,60,61,62]. Table 1 listed the details of the different available energy storage facilities for EVs.

3. Hybrid and Alternative Energy Storage

Highly efficient energy storage technologies which can provide a higher range, weight and cost performance are highly required for the widespread use of the EV. Currently available batteries consume a fraction of stored energy, their life span is not high, and they degrade after charging and discharging a number of times. To meet the peak power demand of the EV, high power density batteries are required, which typically have a higher price than their lower power density counterpart. Increasing the battery size can solve the power density issue, but cost still is a problem. Thermal management, e.g., warming up the battery in cold temperature and cooling it down in hot temperature, is also a key challenge for battery storage.
On the other hand, the supercapacitor has a longer life span and a higher power density which makes it a rapid and effective power supplier. The hybrid battery supercapacitor can be a solution due to its better life span [66,67]. In February 2019, Tesla Inc bought Maxwell, the world’s largest manufacturer of supercapacitors, to overcome the shortage of batteries.

3.1. Battery/Supercapacitor

Supercapacitors (SC) are 95% energy efficient, have specific power ranges between 1000 and 2000 W/kg, possess the longest durability of up to 40 years, require no maintenance, and are insensitive to temperature. Its high-power density is suitable for rapid acceleration and electric braking [68,69,70]. Supercapacitors’ excellent life cycle can be expected to last as long as the car. Supercapacitors store power as static electricity, hence, power delivery is instantaneous from them [71]. The addition of the battery and supercapacitor storage device improves the EV efficiency. Low rated battery capability can be compensated using SC energy storage suitable for electric vehicle application [72].
The primary challenge for the battery and supercapacitor combination is the connection of the SC and battery to the DC bus. Direct connection of the battery and SC to the DC bus is the simplest method, however, fast charging and discharging currents during regenerative acceleration and braking can degrade the battery life span. Also, the supercapacitor’s voltage variation becomes limited as it shares the same terminal with the battery. The battery or supercapacitor connection directly to the DC bus through a DC–DC converter can stop the battery from fluctuation during charging and discharging. Power demand during acceleration and braking is managed by a supercapacitor. High losses due to the inclusion of a converter is the main drawback of this option. Another option is to fully decouple the supercapacitor and battery from the DC bus, making connections exclusively through a DC–DC converter. Additional power losses and costs due to converters are the drawbacks of this system [73]. Connecting SCs directly to the DC bus, without using a power converter, allows one to utilise the SC’s energy-usage capability, which is limited by constant DC bus control [74]. The SC and battery combination has the potential to enhance battery life cycle by up to 50% when compared to the battery-only energy storage system [75]. Employing a smaller rated dc/dc converter can control the voltage of the supercapacitor more so than battery voltage, making it useful in maintaining EVs’ driving conditions [76].

3.2. Fuel Cell/Battery/Supercapacitor

Fuel cells (which are highly efficient and have a high-power density), without combustion, convert the chemical energy of diverse fuels into electricity. They have no tailpipe emissions when driving (as hydrogen to electricity conversion only produces water), silent operation, high reliability and low maintenance, and thus, are considered to be one of the most promising power generation technologies [77,78]. Based on the type of electrolyte, fuel cells include solid oxide fuel cells (SOFCs), polymer electrolyte membrane fuel cells (PEMFCs), and alkaline membrane fuel cells (AMFCs). However, PEMFCs are considered as the most popular energy source for EVs’ engines due to the higher efficiency of the system and the easy accessibility of hydrogen fuel [78,79,80,81]. In 2016, Hyundai and Toyota introduced their fuel cell electric vehicles (FCEV) into the market. At the end of 2016, Honda also launched new environmentally benign hydrogen-powered FCEVs. Nissan, General Motor and Daimler are reportedly set to begin commercialization by 2020 [82]. The present FCEV can drive up to 300–500 km. Refuelling is possible within 10 min from a gas station that has a pressurized hydrogen facility [83,84].
The fuel cell work as the main energy source in hybrid electrical vehicles which are equipped with a fuel cell, battery and supercapacitor, while the supercapacitor and battery served as storage and energy support systems [85,86,87]. During power fluctuations, high power density and dynamic response of the supercapacitor work to relieve stress from the fuel cell and battery. Also, presence of the battery reduces the expense of hydrogen fuel [88]. A fuzzy logic-based control system is beneficial for this combined technology where the logic system will control the charging and discharging of the capacitor bank. This control system has the ability to reduce 14% energy waste [74].

4. Integration of Photovoltaic (PV) in EV

Green power generation from PV technology is extremely promising. In 2017, 500 TWh benign electricity was produced globally from a PV system. In 2019, 500 GW PV was installed, 90 times higher than in 2006 [89]. The use of renewable solar energy is accessible to a wider audience due to the falling cost of the PV systems. Hence, the PV has potential to be source for the EV. A PV powered EV is only suitable for the BEV which can be considered a complete zero emission technology, as nonrenewable and renewable are both still considered as a source.

4.1. PV for EV Charging

The inclusion of the EV into the grid-powered charging station enhances the grid instability [90]. A grid-connected charging station that can supply more than 100 kW to fully charge a 36 kWh battery in 20 min can impose energy losses in the grid if it charges ten vehicles simultaneously (imposing 1000 kW load) with the same capacity [91]. The PV system can supply charging power to EVs as a standalone system. The BEV is a particularly well-suited potential candidate which can be powered directly by the PV. The PV generation system is a complete set where components such as the PV generator, battery, charge controller, inverter, and system load are interconnected and directly convert solar irradiance into electricity. These PV power generation plants are not connected to any utility grid. This system possesses several advantages, for example, grid dependency will be reduced, which in turn reduces the risk of grid failure due to EV penetration; and the EV battery can increase the storage facility and can supply everything in the form of vehicle-to-everything (V2E/V2X) which includes vehicle-to-grid (V2G), vehicle-to-home (V2H), vehicle-to-building (V2B), vehicle-to-load (V2L), and vehicle-to-vehicle (V2V) [92,93,94,95,96,97,98].
A standalone PV system can be installed at office buildings, domestic house rooftops, factories, industrial areas, universities, and car parking areas for EV charging (as shown in Figure 4) [99,100]. Generally, employees park their vehicle for a minimum of 8 h during daytime when solar radiations are available and grid electricity demand is also high. Without the need for battery storage, this charging is possible and can mitigate the negative impact of excessive PV-generated power [98]. This “charging while parking” is a very popular concept nowadays [101]. It is also possible to install a PV-powered charging station at a remote location where a large grid is not available [96,97]. A 10.5 kW AC PV array and a 9.6 kWh lithium-ion battery was employed to power lightweight EVs in university campuses [102].
Variability of PV power generation can be an obstacle for a PV–EV combination. Due to the diurnal nature of PV power generation, load charging is often performed during daytime at universities, offices, etc. [103,104]. Using an open source model, it was predicted that nonPV generation may require EV charging to serve the afternoon hours, as only a small portion of transportation demand can be achieved from the high capacity of photovoltaic power generation [105]. Variability can be predicted by employing a probabilistic- or a deterministic-enabled prediction a day ahead of real-time PV output, or a combined state of charge (SOC)-based fair EV charging strategy incorporating a noniterative PV output based on the historical PV ramp data and the real-time measurement [104]. Voltage fluctuation from the PV system was controlled using a large capacity connected in parallel with the PV module [106].
Vehicle to grid (V2G) concept perceives an EV not as load but as a source from where its battery power can be supplied to grid [10,108,109]. Hence, the EV can provide the grid support by regulating voltage and frequency, peak power shaving and spinning reserve [110]. Parked vehicles during grid-connected charging or idle mode can be employed to let active and reactive power flow from the car to the to the grid and power lines. Large-scale deployment of EVs enhances the uncontrolled charging and discharging, significantly influening the power system which can be controlled using a V2G system [111]. For V2H, EV batteries supply energy to a building when the main source of building power cannot meet the building energy demand [112].
A bidirectional converter is an essential power electronics component which allows an EV to be a source for the grid, load, other vehicles and homes. The use of unidirectional converters allows EVs to charge using the charging station (which is primarily powered from the grid), whereas the bidirectional converter allows vehicle to be a source of power and supply to others (grid, load, homes, etc.), hence electricity can flow in both directions. A new topology was investigated where voltage source converters converted solar farm generated power (200 kWh) and further voltage level was modified using a buck boost converter and harmonics, and transients were removed using a low pass filter. A bidirectional converter between the battery storage and DC microgrid maintained the power flow by charging and discharging the battery power [101,113]. The investigation was also performed using PV system battery storage which was directly connected into a medium voltage direct current bus, and with the grid for EV charging. The medium voltage direct current bus voltage played a key role in controlling the system [114]. For Irish climate, the performance of a 6.65 kW PV to charge four EVs was simulated, and the results indicated that in summer, an EV with 90% SOC using home charging facility can run 100 km daily. The performance was also compared for AC and DC distribution systems which showed that the AC system efficiency was 4.67% lower than the DC system over a year [115]. A highly efficient and power-dense three-port converter was developed to integrate the EV, PV and grid to meet the standard of the combination of the CHAdeMO and the Combined Charging System (CCS) [116].

4.2. PV Integrated in the EV (VIPV)

A variety of charging methods utilising PV have been explored suggesting that, subject to local energy tariffs, solar workplace roofs are a favourable solution. However, with the advancement of thin film PV technology, a concept described by Bhatti et al. as a vehicle integrated PV (VIPV) has been suggested as another elegant solution to charging EVs. However, VIPV as the sole power source of a BEV has been shown to be limited due to several factors, including the low power density available. Similarly, [117] deduced that in the study location of Newark which has an array with a peak power of 300 WP and an inverter efficiency of 90%, the VIPV could account for just over 12% of the yearly miles of a Chevy Volt. The study concludes that due to the limited space available on commercial passenger vehicles and subsequent low power density, VIPV may only be considered as a supplement to plug-in charging techniques. The feasibility of VIPV, particularly for large commercial vehicles with big roof surfaces, was conducted using diesel energy equivalent, payback time, potential savings of costs, and CO2 parameters. It was shown that a 1 m2 monocrystalline-based VIPV system integrated into a truck can save up to 1100 litters of diesel in a vehicle with a lifetime of ten years [118]. It was predicted that VIPV is a solution but low light condition and rainy season EV still needs power from external charging station which can be grid powered or SPV system [119].

5. Reduction of Greenhouse Gas Emission and Particulate Matter

Traditional diesel and petrol-based cars or ICE vehicles generate pollutant GHG gas which includes carbon dioxide (CO2), sulphur hexafluoride (SF6), carbon monoxide (CO), hydrocarbons and nitrogen oxides (N2O) and soot or particulate matter. The prime advantage of the EV is that there is no pollutant emission from the tailpipe. In the EU, the EV can save an average of 50–60% of GHG emissions when compared with ICE-based vehicles [120]. In general, the GHG saving potential of the EV can vary from 10% to 60% depends on the type of EV and geographical location [121]. Hence, it is considered that an EV charged by a battery which takes power from grid can bring environment cleanliness. However, while the EV reduces GHG emission, it also increases the demand for grid electricity. This increased grid power generates GHG, as it still uses fossil fuel as its primary energy source. Usually, GHG measurement metrics of EVs only count direct emission-saving from direct burning of fossil fuels and do not include the indirect emission which is associated during the transmission and generation of the electricity production for grid [122,123,124]. Thus, electricity generation from alternative sources such as solar, wind, biomass and nuclear are highly recommended [125]. Indeed, it should rightly be pointed out that a definitive or unique comparison of GHG emissions between EVs and ICEs fully depend on the defined system boundaries and the intrinsic assumptions during the calculations process [126]. It was reported that for Macau, the electric public bus cannot save a considerable amount of GHG emissions when compared to the diesel-powered public bus, as power generation comes from traditional sources. Hence, it can be noted that GHG emission benefits of an EV depend on the electric power sources that are being employed to power the EV and its efficiency, range and operating modes. Decarbonization of the transport sector using EVs is not fully certain as they are not the technological solution for all countries [127].
The IEA predicted that in 2040, 16% enhancement of CO2 emission is possible in the power generating sector when compared to the level in 2014. This emission enhancement is primarily due to the projected 87% surge in global electricity demand. If this production happens through cleaner power generation, the GHG emissions of EVs (particularly BEVs) will be substantially low. However, for the BEV, end of battery life evaluation is also essential. The increasing trend of the BEV enhances the usages of batteries. Recycle of battery is paramount to recover Li and Co which are valuable metals. and to limit Pb, Cd, Cu which are hazardous substances [128].
Traditional diesel and petrol-based cars or ICE vehicles not only generate GHG but also particulate matter. Particles with an aerodynamic diameter less than 10 μm are referred to as PM10, while those with less than 2.5 μm are referred to as PM 2.5. Particulate matter (PM) emission analysis from road traffic shows that 80% of particles are actually PM1 type. PM2.5 and PM10 have adverse impacts on lung disease, acute and chronic bronchitis, asthma attacks, respiratory problems and also possess a risk of inducing lung cancer [129]. According to EU directives, the daily and annual mean of PM10 should be under 50 μg/m3 and 40 μg/m3, respectively. The World Health Organisation’s guideline indicates that the PM2.5 level should be 10 μg/m3 [130,131]. Due to the lack of tailpipe emissions, the BEV does not produce PM from exhaust sources. However, PM from nonexhaust sources (road dust resuspension, road wear, brake wear and tyre wear) for EV is still valid. About 50–85% of traffic generated PM10 and PM2.5 comes from nonexhaust emission. The weight of EVs is higher than that of ICE vehicles due to the battery, which subsequently affects tyre wear. Thus, it is also believed that significant reductions of PM from EVs are only possible by further improvement of battery weight [132,133].

6. Challenges in EV Uptake

6.1. Technical

The available range of EVs provided by companies are not true for most cases and often up to 17% lower ranges than predicted mileage also occur [134]. Ultrafast charging facilities are now coming into the scenario. They have 350 kW capacity and can charge a fully drained car within 10 min, providing cars with a 200 km range. However, cars are not ready to receive this high current flow. Long charging time is an obstacle for EV uptake. Long queues in charging stations are an issue which can be diminished by implementing a “battery swap” system where the discharged battery can be exchanged in any of the planned changing stations for this service, which enable the 100% charged battery on board within a few minutes (similar to oil refilling). Thus, EVs’ driving ranges will be extended.
The lack of a universal standard for the connector is also an issue with smooth penetration of EVs. The connectors of EV chargers vary with country, EV manufacturer, power level shape, size and pin-out [30], as shown in Figure 5. However, the charging station and the battery employed in the car need to be the same brand or design [3]. Moreover, revenue losses from large investments slow down the growth of the charging infrastructures [135,136,137]. Figure 4 illustrates the different connectors for EV battery charging [138]. Currently, two global standards are available for fast charging which include SAE International’s Combined Charging System (CCS) and the CHAdeMO protocol accepted by German and U.S. industries and Japanese car manufacturers, respectively. To lessen the production cost of software and EV charging parts, harmonized standard in collaboration with the International Organization for Standardization (ISO) and International Electrotechnical Commission (IEC) are essential [139].

6.2. Consumer Behaviour in the Context of Socio-Technical Factors

Widespread unavailability of EVs depends primarily on the behaviour of consumers [140]. The intention of purchase and use, consumer readiness and willingness to pay and accept are some factors which limit the widespread implementation of EVs. How customers’ gender, age, education level, income and occupation have an impact on EV purchase is still a debatable topic. From literature-based studies, it is evident that well-educated young and middle-aged male consumers have stronger intentions to adopt EVs [141,142,143,144]. Those educated in technical fields or those engaged in technical professions are more likely to consume EVs [145]. Advanced BEV technologies are easy to understand for technophiles [142,146]. Currently, EV users, irrespective of geography, are primarily well-educated males who have a medium–high income. They use their cars for private purposes and charge them at home during the night [147]. This is confirmed for Austria [148], Canada [149], Germany [145,150], Norway [151], United Kingdom [152], United States [144,153] and Sweden [147]. Nonearly adopters have a tendency to prefer the PHEV due to their lack of knowledge of EV technologies. It is evident that the price of the BEV is higher than conventional ICE-based vehicle. However, while it is reported that the consumer thinks less about the price when they consider purchasing an EV in some literature [143,154], other literature finds that they do consider the high price [155,156].
The EV’s driving range issue is also an obstacle for drivers who are not willing to proceed with the BEV option once they experience the same due to its short range [157]. The limited driving range, the limited availability of public charging stations, and the length of time it takes to charge an EV, especially in a hilly area, causes range-anxiety [158]. A residential charging station may solve this issue [159].
Additionally, diversity in the EV model can attract a range of audiences. The current EV market is not able to excite people from different backgrounds due to lack of EV models. Hence, variation in EV range, style, appealing features and functionality can attract more consumers [139]. The number of cars and number of available members in the family possess a considerable amount of influence on EV consumer. However, it is still not clear whether having more than one vehicle eventually impresses the consumer to buy an EV or not [160].

6.3. Government Support and Policy

Lucrative government policy and incentives can promote the EV in the global market. Upfront initiatives such as tax reduction and subsidies are attractive during the purchase of an EV. In spite of this, the EV can only compete if the tax on ICE vehicles are higher than EVs, which is the case for Norway and Denmark. Norway is a global forerunner for the use of BEVs. In Norway, seventy thousand BEVs were registered, which accounted for approximately 18% of new car sales in 2015. Total BEV and PHEV purchases were 10,000 in 2012, which became 50,000 in 2015 [154]. Some consumers prefer toll exemption and the privilege of being allowed to drive in the bus lane [154]. EV uptake in some countries such as Singapore is low due to the heavy government taxes on vehicles, which include both the ICEVs and EVs [161].
Hence, a government policy which includes financial subsidies, free parking, driving privileges and preferential tax demonstrates a positive influence on consumers’ intentions to purchase EVs [162,163,164].

7. EV Status Worldwide

In 2017, more than 3 million EVs were on the road, and in 2018 this figure reached to 5.2 million, while just ten years ago there were only hundreds [87]. Figure 6 shows the list of countries leading the sales of EVs in 2018. Current global deployment of EVs shows that China has the highest figure with 45%, followed by Europe and the United States with 24% and 22%, respectively [165]. By the end of 2018 1.1 million EVs were added to China s transport sector. Table 2 indicates the EV status and targets for different countries. Data were compared to the year of 2010. Table 3 shows the reason for the global growth of EVs.
Currently, vehicles which are electrified are mostly light-commercial vehicles which include medium truck and two- or three-wheelers. Electric bicycles or ‘e-bikes’ are power-assisted/self-powered bicycles. They have physical pedals, with optional assistance from an electric motor, which can typically support speeds of up to 25 km/h and 45 km/h. Light-weight and efficient lithium-ion batteries are now common for e-bikes, replacing their cheaper, heavy-weight lead–acid counterparts. In the EU, e- bikes have shown a steady growth between 2006 and 2014. In 2014, around 1,325,000 e-bikes were sold in the EU which was 14 times higher than in 2006. E-bikes are mostly imported from China (80%). Germany has the largest e-bike consumption, followed by Austria, Belgium and Switzerland in the EU. Consumers who like to do manual work and who enjoy steep hills and long distances use e- bikes often. By the end of 2018, electric two- and three-wheelers exceeded 300 million, and China is the biggest consumer of them.
Electric buses have potential but need high capacity batteries. In 2018, globally 460,000 electric buses occupied the roads, and Austria has the highest number of EV buses, followed by Belgium and the Netherlands [166]. Recently, Hyundai created an all-electric bus which has a 290 km range and a 256 kWh Li-ion battery pack. In 2018, EVs consumed a total of 58 terawatt-hours (TWh) of electricity globally [166].

8. Conclusions

Decarbonization of the transport sector could be possible by using EVs as a sustainable and efficient alternative to traditional diesel- and petrol-based vehicles. In this review work, brief details of different EVs, storage facilities, charging EVs through PVs, different socio-technical challenges for EV uptake and global status of EVs have been highlighted. Battery electric vehicle (BEV) is considered to be a true zero-emissions vehicle because of the lack of tailpipe emissions when compared to other types of EV. However, the saving of particulate matter generation from EVs is not significant because of the enhanced weight of BEVs due to the presence of battery storage. Fundamental challenges with EVs are the lack of a suitable energy storage system and an efficient battery management system that could support the competitive mileage when compared to traditional fuel-based vehicles, as well as the lack of a high-performance fast charging facility. An increase in EV uptake depends on government policies that provide lucrative incentives and benefits. Consumer intentions to adopt EVs include, willingness to pay and socio-economic background. To escalate the sale of EV, EV manufactures must pay attention to diversity and creating an appealing model which will attract a large number of consumers.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Allen, M.R.; Dube, O.P.; Solecki, W.; Aragon-Durand, F.; Cramer, W.; Humphreys, S.; Kainuma, M.; Kala, J.; Mahowald, N.; Mulugetta, Y.; et al. Framing and Context. In Global Warming of 1.5 °C. An IPCC Special Report on the Impacts of Global Warming of 1.5 °C above Pre-industrial Levels and Related Global Greenhouse Gas Emission Pathways, in the Context of Strengthening the Global Response to the threat of climate change, sustainable development, and efforts to eradicate poverty; IPCC: Paris, France, 2018. [Google Scholar]
  2. López, I.; Ibarra, E.; Matallana, A.; Andreu, J.; Kortabarria, I. Next generation electric drives for HEV/EV propulsion systems: Technology, trends and challenges. Renew. Sustain. Energy Rev. 2019, 114, 109336. [Google Scholar] [CrossRef]
  3. Martínez-Lao, J.; Montoya, F.G.; Montoya, M.G.; Manzano-Agugliaro, F. Electric vehicles in Spain: An overview of charging systems. Renew. Sustain. Energy Rev. 2017, 77, 970–983. [Google Scholar] [CrossRef]
  4. Van Vliet, O.; Brouwer, A.S.; Kuramochi, T.; Van Den Broek, M.; Faaij, A. Energy use, cost and CO2 emissions of electric cars. J. Power Sources 2011, 196, 2298–2310. [Google Scholar] [CrossRef] [Green Version]
  5. Smith, W.J. Can EV (electric vehicles) address Ireland’s CO2 emissions from transport? Energy 2010, 35, 4514–4521. [Google Scholar] [CrossRef]
  6. Singh, N.; Mishra, T.; Banerjee, R. Greenhouse Gas Emissions in India’s Road Transport Sector. Clim. Chang. Signals Response 2019, 197–209. [Google Scholar] [CrossRef]
  7. Guarnieri, M. When cars went electric, Part 2. IEEE Ind. Electron. Mag. 2011, 5, 61–62. [Google Scholar] [CrossRef]
  8. Thiel, C.; Tsakalidis, A.; Jager-Waldau, A. Will Electric Vehicles Be Killed (again) or Are They the Next Mobility Killer App? Energies 2020, 13, 1828. [Google Scholar] [CrossRef] [Green Version]
  9. Chan, C.C. An Overview of Electric Vehicle Technology. Proc. IEEE 1993, 81, 1202–1213. [Google Scholar] [CrossRef]
  10. Kempton, W.; Letendre, S.E. Electric vehicles as a new power source for electric utilities. Transp. Res. Part D Transp. Environ. 1997, 2, 157–175. [Google Scholar] [CrossRef]
  11. Browne, D.; O’Mahony, M.; Caulfield, B. How should barriers to alternative fuels and vehicles be classified and potential policies to promote innovative technologies be evaluated? J. Clean. Prod. 2012, 35, 140–151. [Google Scholar] [CrossRef]
  12. Bhatti, A.R.; Salam, Z.; Aziz, M.J.B.A.; Yee, K.P.; Ashique, R.H. Electric vehicles charging using photovoltaic: Status and technological review. Renew. Sustain. Energy Rev. 2016, 54, 34–47. [Google Scholar] [CrossRef]
  13. Kumar, M.S.; Revankar, S.T. Development scheme and key technology of an electric vehicle: An overview. Renew. Sustain. Energy Rev. 2017, 70, 1266–1285. [Google Scholar] [CrossRef]
  14. Shamshirband, M.; Salehi, J.; Gazijahani, F.S. Decentralized trading of plug-in electric vehicle aggregation agents for optimal energy management of smart renewable penetrated microgrids with the aim of CO2 emission reduction. J. Clean. Prod. 2018, 200, 622–640. [Google Scholar] [CrossRef]
  15. Markel, T.; Simpson, A. Cost-benefit analysis of plug-in hybrid electric vehicle technology. World Electr. Veh. J. 2007, 1, 294–301. [Google Scholar] [CrossRef] [Green Version]
  16. 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]
  17. Wang, F.; Deng, Y.; Yuan, C. Life cycle assessment of lithium oxygen battery for electric vehicles. J. Clean. Prod. 2020, 264, 121339. [Google Scholar] [CrossRef]
  18. Zhou, Y.; Wen, R.; Wang, H.; Cai, H. Optimal battery electric vehicles range: A study considering heterogeneous travel patterns, charging behaviors, and access to charging infrastructure. Energy 2020, 197, 116945. [Google Scholar] [CrossRef]
  19. Álvarez Fernández, R. A more realistic approach to electric vehicle contribution to greenhouse gas emissions in the city. J. Clean. Prod. 2018, 172, 949–959. [Google Scholar] [CrossRef]
  20. Andwari, A.M.; Pesiridis, A.; Rajoo, S.; Martinez-Botas, R.; Esfahanian, V. A review of Battery Electric Vehicle technology and readiness levels. Renew. Sustain. Energy Rev. 2017, 78, 414–430. [Google Scholar] [CrossRef]
  21. Hannan, M.A.; Hoque, M.M.; Hussain, A.; Yusof, Y.; Ker, P.J. State-of-the-Art and Energy Management System of Lithium-Ion Batteries in Electric Vehicle Applications: Issues and Recommendations. IEEE Access 2018, 6, 19362–19378. [Google Scholar] [CrossRef]
  22. Ahmadi, L.; Yip, A.; Fowler, M.; Young, S.B.; Fraser, R.A. Environmental feasibility of re-use of electric vehicle batteries. Sustain. Energy Technol. Assess. 2014, 6, 64–74. [Google Scholar] [CrossRef]
  23. Kwon, Y.; Son, S.; Jang, K. User satisfaction with battery electric vehicles in South Korea. Transp. Res. Part D Transp. Environ. 2020, 82, 102306. [Google Scholar] [CrossRef]
  24. Li, L.; Wang, Z.; Chen, L.; Wang, Z. Consumer preferences for battery electric vehicles: A choice experimental survey in China. Transp. Res. Part D Transp. Environ. 2020, 78, 102185. [Google Scholar] [CrossRef]
  25. Abotalebi, E.; Scott, D.M.; Ferguson, M.R. Can Canadian households benefit economically from purchasing battery electric vehicles? Transp. Res. Part D Transp. Environ. 2019, 77, 292–302. [Google Scholar] [CrossRef]
  26. Ma, C.T. System planning of grid-connected electric vehicle charging stations and key technologies: A review. Energies 2019, 12, 4201. [Google Scholar] [CrossRef] [Green Version]
  27. 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. [Google Scholar] [CrossRef]
  28. Sheikhi, A.; Bahrami, S.; Ranjbar, A.M.; Oraee, H. Strategic charging method for plugged in hybrid electric vehicles in smart grids; A game theoretic approach. Int. J. Electr. Power Energy Syst. 2013, 53, 499–506. [Google Scholar] [CrossRef]
  29. González, L.G.; Siavichay, E.; Espinoza, J.L. Impact of EV fast charging stations on the power distribution network of a Latin American intermediate city. Renew. Sustain. Energy Rev. 2019, 107, 309–318. [Google Scholar] [CrossRef] [Green Version]
  30. Foley, A.M.; Winning, I.J.; Gallachóir, B.P. State-of-the-art in electric vehicle charging infrastructure. In Proceedings of the 2010 IEEE Vehicle Power and Propulsion Conference, Lille, France, 1–3 September 2010. [Google Scholar]
  31. Knez, M.; Zevnik, G.K.; Obrecht, M. A review of available chargers for electric vehicles: United States of America, European Union, and Asia. Renew. Sustain. Energy Rev. 2019, 109, 284–293. [Google Scholar] [CrossRef]
  32. Fox, G.H. Electric vehicle charging stations: Are we prepared? IEEE Ind. Appl. Mag. 2013, 19, 32–38. [Google Scholar] [CrossRef]
  33. Ruiz, V.; Pfrang, A.; Kriston, A.; Omar, N.; Van den Bossche, P.; Boon-Brett, L. A review of international abuse testing standards and regulations for lithium ion batteries in electric and hybrid electric vehicles. Renew. Sustain. Energy Rev. 2018, 81, 1427–1452. [Google Scholar] [CrossRef]
  34. Kim, J.D. Insights into residential EV charging behavior using energy meter data. Energy Policy 2019, 129, 610–618. [Google Scholar] [CrossRef]
  35. San Román, T.G.; Momber, I.; Abbad, M.R.; Sánchez Miralles, Á. Regulatory framework and business models for charging plug-in electric vehicles: Infrastructure, agents, and commercial relationships. Energy Policy 2011, 39, 6360–6375. [Google Scholar] [CrossRef]
  36. Xia, G.; Cao, L.; Bi, G. A review on battery thermal management in electric vehicle application. J. Power Sources 2017, 367, 90–105. [Google Scholar] [CrossRef]
  37. Hong, S.H.; Jang, D.S.; Park, S.; Yun, S.; Kim, Y. Thermal performance of direct two-phase refrigerant cooling for lithium-ion batteries in electric vehicles. Appl. Therm. Eng. 2020, 173, 115213. [Google Scholar] [CrossRef]
  38. Luo, X.; Wang, J.; Dooner, M.; Clarke, J. Overview of current development in electrical energy storage technologies and the application potential in power system operation. Appl. Energy 2015, 137, 511–536. [Google Scholar] [CrossRef] [Green Version]
  39. Palizban, O.; Kauhaniemi, K. Energy storage systems in modern grids—Matrix of technologies and applications. J. Energy Storage 2016, 6, 248–259. [Google Scholar] [CrossRef]
  40. Taniguchi, A.; Fujioka, N.; Ikoma, M.; Ohta, A. Development of nickel/metal-hydride batteries for EVs and HEVs. J. Power Sources 2001, 100, 117–124. [Google Scholar] [CrossRef]
  41. 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]
  42. Ovshinsky, S.R.; Fetcenko, M.A.; Ross, J. A Nickel Metal Hydride battery for electric vehicles. World Sci. 2008, 169–262. [Google Scholar] [CrossRef]
  43. 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]
  44. Gandoman, F.H.; Jaguemont, J.; Goutam, S.; Gopalakrishnan, R.; Firouz, Y.; Kalogiannis, T.; Omar, N.; Van Mierlo, J. Concept of reliability and safety assessment of lithium-ion batteries in electric vehicles: Basics, progress, and challenges. Appl. Energy 2019, 251, 113343. [Google Scholar] [CrossRef]
  45. Duffner, F.; Wentker, M.; Greenwood, M.; Leker, J. Battery cost modelling: A review and directions for future research. Renew. Sustain. Energy Rev. 2020, 127. Status: Accetped. [Google Scholar] [CrossRef]
  46. Gonzalez-Castellanos, A.; Pozo, D.; Bischi, A. Detailed Li-ion battery characterization model for economic operation. Int. J. Electr. Power Energy Syst. 2020, 116, 105561. [Google Scholar] [CrossRef]
  47. Schuster, S.F.; Brand, M.J.; Berg, P.; Gleissenberger, M.; Jossen, A. Lithium-ion cell-to-cell variation during battery electric vehicle operation. J. Power Sources 2015, 297, 242–251. [Google Scholar] [CrossRef]
  48. Scrosati, B.; Garche, J. Lithium batteries: Status, prospects and future. J. Power Sources 2010, 195, 2419–2430. [Google Scholar] [CrossRef]
  49. Zhang, W.J. Structure and performance of LiFePO4 cathode materials: A review. J. Power Sources 2011, 196, 2962–2970. [Google Scholar] [CrossRef]
  50. Tan, P.; Jiang, H.R.; Zhu, X.B.; An, L.; Jung, C.Y.; Wu, M.C.; Shi, L.; Shyy, W.; Zhao, T.S. Advances and challenges in lithium-air batteries. Appl. Energy 2017, 204, 780–806. [Google Scholar] [CrossRef]
  51. Jaguemont, J.; Boulon, L.; Dubé, Y. A comprehensive review of lithium-ion batteries used in hybrid and electric vehicles at cold temperatures. Appl. Energy 2016, 164, 99–114. [Google Scholar] [CrossRef]
  52. Huang, C.; Sakamoto, J.S.; Wolfenstine, J.; Surampudi, S. The Limits of Low-Temperature Performance of Li-Ion Cells. J. Electrochem. Soc. 2000, 147, 2893–2896. [Google Scholar] [CrossRef] [Green Version]
  53. Smart, M.C.; Ratnakumar, B.V.; Surampudi, S. Electrolytes for Low-Temperature Lithium Batteries Based on Ternary Mixtures of Aliphatic Carbonates. J. Electrochem. Soc. 1999, 146, 486–492. [Google Scholar] [CrossRef]
  54. Zhang, S.S.; Xu, K.; Jow, T.R. Low temperature performance of graphite electrode in Li-ion cells. Electrochim. Acta 2002, 48, 241–246. [Google Scholar] [CrossRef]
  55. Ji, Y.; Zhang, Y.; Wang, C.Y. Li-ion cell operation at low temperatures. J. Electrochem. Soc. 2013, 160, 636–649. [Google Scholar] [CrossRef]
  56. Bugga, R.; Smart, M.; Whitacre, J.; West, W. Lithium ion batteries for space applications. In Proceedings of the 2007 IEEE Aerospace Conference, Big Sky, MT, USA, 3–10 March 2007; pp. 1–7. [Google Scholar]
  57. Guo, S.; Xiong, R.; Wang, K.; Sun, F. A novel echelon internal heating strategy of cold batteries for all-climate electric vehicles application. Appl. Energy 2018, 219, 256–263. [Google Scholar] [CrossRef]
  58. Lv, Y.; Yang, X.; Zhang, G.; Li, X. Experimental research on the effective heating strategies for a phase change material based power battery module. Int. J. Heat Mass Transf. 2019, 128, 392–400. [Google Scholar] [CrossRef]
  59. Rahimi-Eichi, H.; Ojha, U.; Baronti, F.; Chow, M.Y. Battery management system: An overview of its application in the smart grid and electric vehicles. IEEE Ind. Electron. Mag. 2013, 7, 4–16. [Google Scholar] [CrossRef]
  60. Nizam, M.; Maghfiroh, H.; Rosadi, R.A.; Kirana, D. Battery management system design (BMS) for lithium ion batteries Battery Management System Design (BMS) for Lithium Ion Batteries. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2020; p. 030157. [Google Scholar]
  61. 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]
  62. Liu, K.; Li, K.; Peng, Q.; Zhang, C. A brief review on key technologies in the battery management system of electric vehicles. Front. Mech. Eng. 2019, 14, 47–64. [Google Scholar] [CrossRef] [Green Version]
  63. Sbordone, D.; Bertini, I.; Di Pietra, B.; Falvo, M.C.; Genovese, A.; Martirano, L. EV fast charging stations and energy storage technologies: A real implementation in the smart micro grid paradigm. Electr. Power Syst. Res. 2015, 120, 96–108. [Google Scholar] [CrossRef]
  64. Tian, H.; Qin, P.; Li, K.; Zhao, Z. A review of the state of health for lithium-ion batteries: Research status and suggestions. J. Clean. Prod. 2020, 261, 120813. [Google Scholar] [CrossRef]
  65. Varshney, K.; Varshney, P.K.; Gautam, K.; Tanwar, M.; Chaudhary, M. Current trends and future perspectives in the recycling of spent lead acid batteries in India. Mater. Today Proc. 2020. [Google Scholar] [CrossRef]
  66. Kouchachvili, L.; Yaïci, W.; Entchev, E. Hybrid battery/supercapacitor energy storage system for the electric vehicles. J. Power Sources 2018, 374, 237–248. [Google Scholar] [CrossRef]
  67. Song, Z.; Li, J.; Hou, J.; Hofmann, H.; Ouyang, M.; Du, J. The battery-supercapacitor hybrid energy storage system in electric vehicle applications: A case study. Energy 2018, 154, 433–441. [Google Scholar] [CrossRef]
  68. Hannan, M.A.; Hoque, M.M.; Mohamed, A.; Ayob, A. Review of energy storage systems for electric vehicle applications: Issues and challenges. Renew. Sustain. Energy Rev. 2017, 69, 771–789. [Google Scholar] [CrossRef]
  69. Forse, A.C.; Merlet, C.; Griffin, J.M.; Grey, C.P. New perspectives on the charging mechanisms of supercapacitors. J. Am. Chem. Soc. 2016, 138, 5731–5744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Tie, S.F.; Tan, C.W. A review of energy sources and energy management system in electric vehicles. Renew. Sustain. Energy Rev. 2013, 20, 82–102. [Google Scholar] [CrossRef]
  71. Jinrui, N.; Zhifu, W.; Qinglian, R. Simulation and analysis of performance of a pure electric vehicle with a super-capacitor. In Proceedings of the 2006 IEEE Vehicle Power and Propulsion Conference, Windsor, UK, 6–8 September 2006; pp. 2–7. [Google Scholar] [CrossRef]
  72. Soni, S.R.; Upadhyay, C.D.; Chandwani, H. Analysis of battery-super capacitor based storage for electrical vehicle. In Proceedings of the 2015 International Conference on Energy Economics and Environment, Noida, India, 27–28 March 2015. [Google Scholar]
  73. Ostadi, A.; Kazerani, M.; Chen, S.K. Hybrid Energy Storage System (HESS) in vehicular applications: A review on interfacing battery and ultra-capacitor units. In Proceedings of the 2013 IEEE Transportation Electrification Conference and Expo (ITEC), Detroit, MI, USA, 16–19 June 2013. [Google Scholar]
  74. Valdez-Resendiz, J.E.; Rosas-Caro, J.C.; Mayo-Maldonado, J.C.; Claudio-Sanchez, A.; Ruiz-Martinez, O.; Sanchez, V.M. Improvement of ultracapacitors-energy usage in fuel cell based hybrid electric vehicle. Int. J. Hydrog. Energy 2020. [CrossRef]
  75. Shen, J.; Dusmez, S.; Khaligh, A. Optimization of sizing and battery cycle life in battery/ultracapacitor hybrid energy storage systems for electric vehicle applications. IEEE Trans. Ind. Inform. 2014, 10, 2112–2121. [Google Scholar] [CrossRef]
  76. Cao, J.; Emadi, A. A new battery/ultracapacitor hybrid energy storage system for electric, hybrid, and plug-in hybrid electric vehicles. IEEE Trans. Power Electron. 2012, 27, 122–132. [Google Scholar] [CrossRef]
  77. Debe, M.K. Electrocatalyst approaches and challenges for automotive fuel cells. Nature 2012, 486, 43–51. [Google Scholar] [CrossRef]
  78. Stephens, I.E.L.; Rossmeisl, J.; Chorkendorff, I. Toward sustainable fuel cells. Science 2016, 354, 1378–1379. [Google Scholar] [CrossRef]
  79. Priya, K.; Sathishkumar, K.; Rajasekar, N. A comprehensive review on parameter estimation techniques for Proton Exchange Membrane fuel cell modelling. Renew. Sustain. Energy Rev. 2018, 93, 121–144. [Google Scholar] [CrossRef]
  80. Zhang, G.; Jiao, K. Multi-phase models for water and thermal management of proton exchange membrane fuel cell: A review. J. Power Sources 2018, 391, 120–133. [Google Scholar] [CrossRef]
  81. Chen, H.; Song, Z.; Zhao, X.; Zhang, T.; Pei, P.; Liang, C. A review of durability test protocols of the proton exchange membrane fuel cells for vehicle. Appl. Energy 2018, 224, 289–299. [Google Scholar] [CrossRef]
  82. Wang, G.; Yu, Y.; Liu, H.; Gong, C.; Wen, S.; Wang, X.; Tu, Z. Progress on design and development of polymer electrolyte membrane fuel cell systems for vehicle applications: A review. Fuel Process. Technol. 2018, 179, 203–228. [Google Scholar] [CrossRef]
  83. McNicol, B.D.; Rand, D.A.J.; Williams, K.R. Fuel cells for road transportation purposes—Yes or no? J. Power Sources 2001, 100, 47–59. [Google Scholar] [CrossRef]
  84. Li, W.; Long, R.; Chen, H.; Chen, F.; Zheng, X.; He, Z.; Zhang, L. Willingness to pay for hydrogen fuel cell electric vehicles in China: A choice experiment analysis. Int. J. Hydrog. Energy 2020, in press. [Google Scholar] [CrossRef]
  85. 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]
  86. 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]
  87. Lü, X.; Wu, Y.; Lian, J.; Zhang, Y.; Chen, C.; Wang, P.; Meng, L. Energy management of hybrid electric vehicles: A review of energy optimization of fuel cell hybrid power system based on genetic algorithm. Energy Convers. Manag. 2020, 205, 112474. [Google Scholar] [CrossRef]
  88. Fu, Z.; Li, Z.; Si, P.; Tao, F. A hierarchical energy management strategy for fuel cell/battery/supercapacitor hybrid electric vehicles. Int. J. Hydrog. Energy 2019, 44, 22146–22159. [Google Scholar] [CrossRef]
  89. Haegel, N.M.; Atwater, H.; Barnes, T.; Breyer, C.; Burrell, A.; Chiang, Y.-M.; De Wolf, S.; Dimmler, B.; Feldman, D.; Glunz, S.; et al. Terawatt-scale photovoltaics: Transform global energy. Science 2019, 364, 836–838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Richardson, D.B. Electric vehicles and the electric grid: A review of modeling approaches, Impacts, and renewable energy integration. Renew. Sustain. Energy Rev. 2013, 19, 247–254. [Google Scholar] [CrossRef]
  91. Schey, S.; Scoffield, D.; Smart, J. A first look at the impact of electric vehicle charging on the electric grid in the EV project. World Electr. Veh. J. 2012, 5, 667–678. [Google Scholar] [CrossRef] [Green Version]
  92. Corchero, C.; Sanmarti, M. Vehicle-to-Everything (V2X): Benefits and barriers. In Proceedings of the 2018 15th International Conference on the European Energy Market (EEM), Lodz, Poland, 27–29 June 2018; pp. 1–4. [Google Scholar]
  93. Services, V. Fault Tolerant Boost Converter with Multiple Serial. Energies 2020, 13, 1694. [Google Scholar]
  94. Ye, B.; Jiang, J.; Miao, L.; Yang, P.; Li, J.; Shen, B. Feasibility study of a solar-powered electric vehicle charging station model. Energies 2015, 8, 13265–13283. [Google Scholar] [CrossRef] [Green Version]
  95. Tulpule, P.J.; Marano, V.; Yurkovich, S.; Rizzoni, G. Economic and environmental impacts of a PV powered workplace parking garage charging station. Appl. Energy 2013, 108, 323–332. [Google Scholar] [CrossRef]
  96. Ma, T.; Yang, H.; Lu, L. A feasibility study of a stand-alone hybrid solar-wind-battery system for a remote island. Appl. Energy 2014, 121, 149–158. [Google Scholar] [CrossRef]
  97. Vermaak, H.J.; Kusakana, K. Design of a photovoltaic-wind charging station for small electric Tuk-tuk in DR Congo. Renew. Energy 2014, 67, 40–45. [Google Scholar] [CrossRef]
  98. Nunes, P.; Farias, T.; Brito, M.C. Day charging electric vehicles with excess solar electricity for a sustainable energy system. Energy 2015, 80, 263–274. [Google Scholar] [CrossRef]
  99. Coffman, M.; Bernstein, P.; Wee, S. Integrating electric vehicles and residential solar PV. Transp. Policy 2017, 53, 30–38. [Google Scholar] [CrossRef]
  100. Ghotge, R.; Snow, Y.; Farahani, S.; Lukszo, Z.; van Wijk, A. Optimized scheduling of EV charging in solar parking lots for local peak reduction under EV demand uncertainty. Energies 2020, 13, 1275. [Google Scholar] [CrossRef] [Green Version]
  101. Khan, S.; Ahmad, A.; Ahmad, F.; Shafaati Shemami, M.; Saad Alam, M.; Khateeb, S. A Comprehensive Review on Solar Powered Electric Vehicle Charging System. Smart Sci. 2018, 6, 54–79. [Google Scholar] [CrossRef]
  102. Esfandyari, A.; Norton, B.; Conlon, M.; McCormack, S.J. Performance of a campus photovoltaic electric vehicle charging station in a temperate climate. Sol. Energy 2019, 177, 762–771. [Google Scholar] [CrossRef]
  103. Van Roy, J.; Leemput, N.; Geth, F.; Salebien, R.; Buscher, J.; Driesen, J. Operational Electric Vehicle Charging Strategies. IEEE Trans. Sustain. Energy 2014, 5, 264–272. [Google Scholar] [CrossRef]
  104. Islam, M.S.; Mithulananthan, N. PV based EV charging at universities using supplied historical PV output ramp. Renew. Energy 2018, 118, 306–327. [Google Scholar] [CrossRef]
  105. Fattori, F.; Anglani, N.; Muliere, G. Combining photovoltaic energy with electric vehicles, smart charging and vehicle-to-grid. Sol. Energy 2014, 110, 438–451. [Google Scholar] [CrossRef]
  106. Yukita, K.; Kobayashi, Y.; Duy-Dinh, N.; Matsumura, T.; Goto, Y. Suppression of PV output fluctuation using EV in a electric power system. IFAC Pap. 2019, 52, 93–98. [Google Scholar] [CrossRef]
  107. Chandra Mouli, G.R.; Bauer, P.; Zeman, M. System design for a solar powered electric vehicle charging station for workplaces. Appl. Energy 2016, 168, 434–443. [Google Scholar] [CrossRef] [Green Version]
  108. Igualada, L.; Corchero, C.; Cruz-Zambrano, M.; Heredia, F.J. Optimal energy management for a residential microgrid including a vehicle-to-grid system. IEEE Trans. Smart Grid 2014, 5, 2163–2172. [Google Scholar] [CrossRef] [Green Version]
  109. Kramer, B.; Chakraborty, S.; Kroposki, B. A Review of Plug-in Vehicles and Vehicle-to-Grid Capability. In Proceedings of the 2008 34th Annual Conference of IEEE Industrial Electronics, Orlando, FL, USA, 10–13 November 2008; pp. 2278–2283. [Google Scholar]
  110. Ehsani, M.; Falahi, M.; Lotfifard, S. Vehicle to grid services: Potential and applications. Energies 2012, 5, 4076–4090. [Google Scholar] [CrossRef] [Green Version]
  111. Sharma, A.; Sharma, S. Review of power electronics in vehicle-to-grid systems. J. Energy Storage 2019, 21, 337–361. [Google Scholar] [CrossRef]
  112. Barone, G.; Buonomano, A.; Calise, F.; Forzano, C.; Palombo, A. Building to vehicle to building concept toward a novel zero energy paradigm: Modelling and case studies. Renew. Sustain. Energy Rev. 2019, 101, 625–648. [Google Scholar] [CrossRef]
  113. Khan, A.; Memon, S.; Sattar, T. Integration and management of solar energy for electric vehicle charging station. In Solar World Congress 2017-Innovation for the 100% Renewable Energy Transformation; London South Bank University: London, UK, 2017; pp. 943–953. [Google Scholar] [CrossRef]
  114. Torreglosa, J.P.; García-Triviño, P.; Fernández-Ramirez, L.M.; Jurado, F. Decentralized energy management strategy based on predictive controllers for a medium voltage direct current photovoltaic electric vehicle charging station. Energy Convers. Manag. 2016, 108, 1–13. [Google Scholar] [CrossRef]
  115. Kineavy, F.; Duffy, M. Modelling and design of electric vehicle charging systems that include on-site renewable energy sources. In Proceedings of the 2014 IEEE 5th International Symposium on Power Electronics for Distributed Generation Systems (PEDG), Galway, Ireland, 24–27 June 2014; pp. 1–8. [Google Scholar] [CrossRef]
  116. Chandra Mouli, G.R.; Schijffelen, J.; Van Den Heuvel, M.; Kardolus, M.; Bauer, P. A 10 kW Solar-Powered Bidirectional EV Charger Compatible with Chademo and COMBO. IEEE Trans. Power Electron. 2019, 34, 1082–1098. [Google Scholar] [CrossRef] [Green Version]
  117. Birnie, D.P. Solar-to-vehicle (S2V) systems for powering commuters of the future. J. Power Sources 2009, 186, 539–542. [Google Scholar] [CrossRef]
  118. Kronthaler, L.; Maturi, L.; Moser, D.; Alberti, L. Vehicle-integrated Photovoltaic (ViPV) systems: Energy production, Diesel Equivalent, Payback Time; An assessment screening for trucks and busses. In Proceedings of the 2014 Ninth International Conference on Ecological Vehicles and Renewable Energies (EVER), Monte-Carlo, Monaco, 25–27 March 2014. [Google Scholar] [CrossRef]
  119. Manivannan, S.; Kaleeswaran, E. Solar powered electric vehicle. In Proceedings of the First International Conference on Sustainbale Green buildings and Communities, Chennai, India, 18–20 December 2016; pp. 2–5. [Google Scholar]
  120. Moro, A.; Lonza, L. Electricity carbon intensity in European Member States: Impacts on GHG emissions of electric vehicles. Transp. Res. Part D Transp. Environ. 2018, 64, 5–14. [Google Scholar] [CrossRef]
  121. Gómez Vilchez, J.J.; Jochem, P. Powertrain technologies and their impact on greenhouse gas emissions in key car markets. Transp. Res. Part D Transp. Environ. 2020, 80, 102214. [Google Scholar] [CrossRef]
  122. Manjunath, A.; Gross, G. Towards a meaningful metric for the quantification of GHG emissions of electric vehicles (EVs). Energy Policy 2017, 102, 423–429. [Google Scholar] [CrossRef]
  123. Kim, I.; Kim, J.; Lee, J. Dynamic analysis of well-to-wheel electric and hydrogen vehicles greenhouse gas emissions: Focusing on consumer preferences and power mix changes in South Korea. Appl. Energy 2020, 260, 114281. [Google Scholar] [CrossRef]
  124. Gupta, P.; Tong, D.; Wang, J.; Zhuge, W.; Yan, C.; Wu, Y.; Luo, S.; He, X.; Ma, F. Well-to-wheels total energy and GHG emissions of HCNG heavy-duty vehicles in China: Case of EEV qualified EURO 5 emissions scenario. Int. J. Hydrog. Energy 2020, 45, 8002–8014. [Google Scholar] [CrossRef]
  125. Winyuchakrit, P.; Sukamongkol, Y.; Limmeechokchai, B. Do Electric Vehicles Really Reduce GHG Emissions in Thailand? Energy Procedia 2017, 138, 348–353. [Google Scholar] [CrossRef]
  126. Ma, H.; Balthasar, F.; Tait, N.; Riera-Palou, X.; Harrison, A. A new comparison between the life cycle greenhouse gas emissions of battery electric vehicles and internal combustion vehicles. Energy Policy 2012, 44, 160–173. [Google Scholar] [CrossRef]
  127. Abdul-Manan, A.F.N. Uncertainty and differences in GHG emissions between electric and conventional gasoline vehicles with implications for transport policy making. Energy Policy 2015, 87, 1–7. [Google Scholar] [CrossRef]
  128. Choi, Y.; Rhee, S.-W. Current status and perspectives on recycling of end-of-life battery of electric vehicle in Korea (Republic of). Waste Manag. 2020, 106, 261–270. [Google Scholar] [CrossRef]
  129. Zhang, Z.; Zhu, D.; Cui, B.; Ding, R.; Shi, X.; He, P. Association between particulate matter air pollution and lung cancer. Thorax 2020, 75, 85–87. [Google Scholar] [CrossRef]
  130. Jereb, B.; Batkovič, T.; Herman, L.; Šipek, G.; Kovše, Š.; Gregorič, A.; Močnik, G. Exposure to black carbon during bicycle commuting-Alternative route selection. Atmosphere 2018, 9, 21. [Google Scholar] [CrossRef] [Green Version]
  131. Uherek, E.; Halenka, T.; Borken-Kleefeld, J.; Balkanski, Y.; Berntsen, T.; Borrego, C.; Gauss, M.; Hoor, P.; Juda-Rezler, K.; Lelieveld, J.; et al. Transport impacts on atmosphere and climate: Land transport. Atmos. Environ. 2010, 44, 4772–4816. [Google Scholar] [CrossRef]
  132. Amato, F.; Cassee, F.R.; Denier van der Gon, H.A.C.; Gehrig, R.; Gustafsson, M.; Hafner, W.; Harrison, R.M.; Jozwicka, M.; Kelly, F.J.; Moreno, T.; et al. Urban air quality: The challenge of traffic non-exhaust emissions. J. Hazard. Mater. 2014, 275, 31–36. [Google Scholar] [CrossRef]
  133. Timmers, V.R.J.H.; Achten, P.A.J. Non-Exhaust PM Emissions from Battery Electric Vehicles; Elsevier: Amsterdam, The Netherlands, 2018; ISBN 9780128117705. [Google Scholar]
  134. Milligan, R.; Etxebarria, S.; Muneer, T.; Gago, E.J. Driven performance of electric vehicles in Edinburgh and its environs. Energies 2019, 12, 74. [Google Scholar] [CrossRef] [Green Version]
  135. Greene, D.L.; Park, S.; Liu, C. Analyzing the transition to electric drive vehicles in the U.S. Futures 2014, 58, 34–52. [Google Scholar] [CrossRef]
  136. Godina, R.; Rodrigues, E.M.G.; Paterakis, N.G.; Erdinc, O.; Catalão, J.P.S. Innovative impact assessment of electric vehicles charging loads on distribution transformers using real data. Energy Convers. Manag. 2016, 120, 206–216. [Google Scholar] [CrossRef]
  137. Guo, C.; Chan, C.C. Analysis method and utilization mechanism of the overall value of EV charging. Energy Convers. Manag. 2015, 89, 420–426. [Google Scholar] [CrossRef] [Green Version]
  138. Ronanki, D.; Kelkar, A.; Williamson, S.S. Extreme fast charging technology—Prospects to enhance sustainable electric transportation. Energies 2019, 12, 3721. [Google Scholar] [CrossRef] [Green Version]
  139. Haddadian, G.; Khodayar, M.; Shahidehpour, M. Accelerating the Global Adoption of Electric Vehicles: Barriers and Drivers. Electr. J. 2015, 28, 53–68. [Google Scholar] [CrossRef]
  140. Zarazua de Rubens, G. Who will buy electric vehicles after early adopters? Using machine learning to identify the electric vehicle mainstream market. Energy 2019, 172, 243–254. [Google Scholar] [CrossRef]
  141. Prakash, N.; Kapoor, R.; Kapoor, A.; Malik, Y. Gender Preferences for Alternative Energy Transport with Focus on Electric Vehicle. J. Soc. Sci. 2014, 10, 114–122. [Google Scholar] [CrossRef] [Green Version]
  142. Hackbarth, A.; Madlener, R. Consumer preferences for alternative fuel vehicles: A discrete choice analysis. Transp. Res. Part D Transp. Environ. 2013, 25, 5–17. [Google Scholar] [CrossRef] [Green Version]
  143. Hidrue, M.K.; Parsons, G.R.; Kempton, W.; Gardner, M.P. Willingness to pay for electric vehicles and their attributes. Resour. Energy Econ. 2011, 33, 686–705. [Google Scholar] [CrossRef] [Green Version]
  144. Carley, S.; Krause, R.M.; Lane, B.W.; Graham, J.D. Intent to purchase a plug-in electric vehicle: A survey of early impressions in large US cites. Transp. Res. Part D Transp. Environ. 2013, 18, 39–45. [Google Scholar] [CrossRef]
  145. Plötz, P.; Schneider, U.; Globisch, J.; Dütschke, E. Who will buy electric vehicles? Identifying early adopters in Germany. Transp. Res. Part A Policy Pract. 2014, 67, 96–109. [Google Scholar] [CrossRef]
  146. Egbue, O.; Long, S. Barriers to widespread adoption of electric vehicles: An analysis of consumer attitudes and perceptions. Energy Policy 2012, 48, 717–729. [Google Scholar] [CrossRef]
  147. Vassileva, I.; Campillo, J. Adoption barriers for electric vehicles: Experiences from early adopters in Sweden. Energy 2017, 120, 632–641. [Google Scholar] [CrossRef]
  148. Wolf, A.; Seebauer, S. Technology adoption of electric bicycles: A survey among early adopters. Transp. Res. Part A Policy Pract. 2014, 69, 196–211. [Google Scholar] [CrossRef]
  149. Axsen, J.; Goldberg, S.; Bailey, J. How might potential future plug-in electric vehicle buyers differ from current “Pioneer” owners? Transp. Res. Part D Transp. Environ. 2016, 47, 357–370. [Google Scholar] [CrossRef]
  150. Lieven, T.; Mühlmeier, S.; Henkel, S.; Waller, J.F. Who will buy electric cars? An empirical study in Germany. Transp. Res. Part D Transp. Environ. 2011, 16, 236–243. [Google Scholar] [CrossRef]
  151. Graabak, I.; Wu, Q.; Warland, L.; Liu, Z. Optimal planning of the Nordic transmission system with 100% electric vehicle penetration of passenger cars by 2050. Energy 2016, 107, 648–660. [Google Scholar] [CrossRef]
  152. Campbell, A.R.; Ryley, T.; Thring, R. Identifying the early adopters of alternative fuel vehicles: A case study of Birmingham, United Kingdom. Transp. Res. Part A Policy Pract. 2012, 46, 1318–1327. [Google Scholar] [CrossRef] [Green Version]
  153. Hardman, S.; Shiu, E.; Steinberger-Wilckens, R. Comparing high-end and low-end early adopters of battery electric vehicles. Transp. Res. Part A Policy Pract. 2016, 88, 40–57. [Google Scholar] [CrossRef] [Green Version]
  154. Bjerkan, K.Y.; Nørbech, T.E.; Nordtømme, M.E. Incentives for promoting Battery Electric Vehicle (BEV) adoption in Norway. Transp. Res. Part D Transp. Environ. 2016, 43, 169–180. [Google Scholar] [CrossRef] [Green Version]
  155. Erdem, C.; Şentürk, I.; Şimşek, T. Identifying the factors affecting the willingness to pay for fuel-efficient vehicles in Turkey: A case of hybrids. Energy Policy 2010, 38, 3038–3043. [Google Scholar] [CrossRef]
  156. Li, X.; Clark, C.D.; Jensen, K.L.; Yen, S.T.; English, B.C. Consumer purchase intentions for flexible-fuel and hybrid-electric vehicles. Transp. Res. Part D Transp. Environ. 2013, 18, 9–15. [Google Scholar] [CrossRef]
  157. Skippon, S.M.; Kinnear, N.; Lloyd, L.; Stannard, J. How experience of use influences mass-market drivers’ willingness to consider a battery electric vehicle: A randomised controlled trial. Transp. Res. Part A Policy Pract. 2016, 92, 26–42. [Google Scholar] [CrossRef]
  158. Travesset-Baro, O.; Rosas-Casals, M.; Jover, E. Transport energy consumption in mountainous roads. A comparative case study for internal combustion engines and electric vehicles in Andorra. Transp. Res. Part D Transp. Environ. 2015, 34, 16–26. [Google Scholar] [CrossRef] [Green Version]
  159. Melliger, M.A.; van Vliet, O.P.R.; Liimatainen, H. Anxiety vs reality—Sufficiency of battery electric vehicle range in Switzerland and Finland. Transp. Res. Part D Transp. Environ. 2018, 65, 101–115. [Google Scholar] [CrossRef]
  160. Li, W.; Long, R.; Chen, H.; Geng, J. A review of factors influencing consumer intentions to adopt battery electric vehicles. Renew. Sustain. Energy Rev. 2017, 78, 318–328. [Google Scholar] [CrossRef]
  161. Nian, V.; Hari, M.P.; Yuan, J. A new business model for encouraging the adoption of electric vehicles in the absence of policy support. Appl. Energy 2019, 235, 1106–1117. [Google Scholar] [CrossRef]
  162. Zhang, Y.; Yu, Y.; Zou, B. Analyzing public awareness and acceptance of alternative fuel vehicles in China: The case of EV. Energy Policy 2011, 39, 7015–7024. [Google Scholar] [CrossRef]
  163. Sang, Y.N.; Bekhet, H.A. Modelling electric vehicle usage intentions: An empirical study in Malaysia. J. Clean. Prod. 2015, 92, 75–83. [Google Scholar] [CrossRef]
  164. Helveston, J.P.; Liu, Y.; Feit, E.M.D.; Fuchs, E.; Klampfl, E.; Michalek, J.J. Will subsidies drive electric vehicle adoption? Measuring consumer preferences in the U.S. and China. Transp. Res. Part A Policy Pract. 2015, 73, 96–112. [Google Scholar] [CrossRef] [Green Version]
  165. Naumanen, M.; Uusitalo, T.; Huttunen-Saarivirta, E.; van der Have, R. Development strategies for heavy duty electric battery vehicles: Comparison between China, EU, Japan and USA. Resour. Conserv. Recycl. 2019, 151, 104413. [Google Scholar] [CrossRef]
  166. IEA. Global EV Outlook 2019; IEA: Paris, France, 2019. [Google Scholar]
Figure 1. Schematic of typical power train configurations: (a) Series hybrid electric vehicle (HEV), (b) parallel HEV, (c) series–parallel HEV, (d) series plug-in hybrid electric vehicle (PHEV), (e) parallel PHEV, (f) series–parallel PHEV [16].
Figure 1. Schematic of typical power train configurations: (a) Series hybrid electric vehicle (HEV), (b) parallel HEV, (c) series–parallel HEV, (d) series plug-in hybrid electric vehicle (PHEV), (e) parallel PHEV, (f) series–parallel PHEV [16].
Energies 13 02602 g001
Figure 2. Typical power train configuration of the BEV [16].
Figure 2. Typical power train configuration of the BEV [16].
Energies 13 02602 g002
Figure 3. (a) Complete scheme for an EV charging facility ([27]); (b) typical charging stations for EVs in the UK (image courtesy: ROLEC, UK).
Figure 3. (a) Complete scheme for an EV charging facility ([27]); (b) typical charging stations for EVs in the UK (image courtesy: ROLEC, UK).
Energies 13 02602 g003
Figure 4. Concept of a solar powered EV charging station [107].
Figure 4. Concept of a solar powered EV charging station [107].
Energies 13 02602 g004
Figure 5. Global male and female battery charger connectors (Ronanki et al., 2019).
Figure 5. Global male and female battery charger connectors (Ronanki et al., 2019).
Energies 13 02602 g005
Figure 6. The growing sales of EVs since 2009 [87].
Figure 6. The growing sales of EVs since 2009 [87].
Energies 13 02602 g006
Table 1. Details of different available energy storage facilities for EVs [45,63,64,65].
Table 1. Details of different available energy storage facilities for EVs [45,63,64,65].
Storage SystemEnergy Density (Wh/kg)Power Density (W/kg)Energy Efficiency (%)
Lead–acid20–352570–80
NiCd40–6014060
NiMh60–8022050–80
Li-ion100–270300–200085–95
Li-polymer100–200300–100070
Super capacitor25–755000–20,00090+
Table 2. EV targets for different countries by 2050 and comparison with past commitment.
Table 2. EV targets for different countries by 2050 and comparison with past commitment.
CountryTarget (Published in 2010 [30])2020 Present StatusTarget for 2050
Austria2020: 100,000 EVs deployed2018: ~25,000
2020: in Feb 6.7% EV sales
Australia 2012: first cars on road, 2018: mass deployment, 2050: up to 65% of car stock2019:1277 EVs sold
2018: 670 EVs sold
2030: 50% of new cars to be EVs
Canada2018: 500,000 EVs deployed2019: 93,091 EVs on the road, EV sales grew by 125% compared to 2017New light-duty zero-emission vehicle sales by 2025: 10%; 2030: 30%; 2040: 100%
China2011: 500,000 annual production of EVs2011: 8159 EVs and in 2015, 331,092 EVs were sold.
2020: 15,000 EV charging stations to accommodate 5 million EVs
2400: 40% global EV sales
Denmark2020: 200,000 EVs 2019: 4618 BEVs and 3623 PHEVs are sold.2050: Transport sector will be independent of fossil fuel
France2020: 2,000,000 EVs2018: 2% sold card was either a PHEV or a BEV2040: ban on fossil fuel car and Paris will follow this from 2030.
Germany2020: 1,000,000 EVs deployed2019: 24,000 public charging stations2030: 1 million charging stations
Ireland 2020: 10% EV market share2019: 4825 EVs on the road and another 4054 were registered2030: 1 million EVs
Israel 2011: 40,000 EVs, 2012: 40,000 to 100,000 EVs2025: 177,000 EVs on road2030: full switch to EV
Japan2020: 50% market share of next generation vehicles2017: EVs account for 0.4% of market share 2030: 20-30% BEV and PHEV market share
New
Zealand
2020: 5% market share, 2040: 60% market share2021: 64,000 electric vehicles 2030: EVS constituting100% of new vehicles
2050: 100% all lightweight vehicles are EV
Spain2014: 1,000,000 EVs deployed2018: 8000 EVs2040: Banning of diesel, gasoline, hybrid vehicles sale
2050: Permanent ban
Sweden2020: 600,000 EVs deployed2019: PHEVs account for 11% of market share 2030: Halt fossil fuel car
2045: carbon neutral
UK No target figures, but policy to support EVsTransport sector accounts 27%GHG emissionZero emission
USA2015: 1,000,000 PHEV stock2011: 9750 EVs were sold
2015: 71,044 EVs were sold
Los Angeles targets 100% EVs
Table 3. Subsidy for EVs in different countries.
Table 3. Subsidy for EVs in different countries.
CountrySubsidy
Austria
  • EUR 3000 for BEV price of up to 50,000 euro for private use, EUR 60,000 for commercial use
  • EUR 400 for cargo bike
Australia
  • No stamp duty for full EVs
  • 20% discount on annual registration of EVs
Canada
  • CAD 5000 for EV price with a limit of up to CAD 45,000 (for a six-seater or smaller)
China
  • USD 3500 for BEVs
Denmark
  • No tax on BEVs until 2015
France
  • Applicable for vehicles that emit less than 20 g of CO2 per kilometre
  • Up to EUR 6000 for private individuals for a list price of EUR 45,000
  • EUR 900 for a two- or three-wheeler
Germany
  • EUR 6000 (about USD 6700) for electric car price EUR 40,000 (USD 44,500)
Ireland
  • EUR 5000 for BEVs and PHEVs
Israel
Japan
  • USD 7770 for EVs
New Zealand
  • NZD 8000 (USD 4880) for EVs
Spain
  • EUR 5500 for list price EUR 48,400
Sweden
  • EUR 5700 for BEVs
South Korea
  • USD 6600 for BEVs
  • USD 18,600 for FCEVs
UK
  • GBP 3500 for BEVs
USA
  • USD 2500 for BEVs
  • USD 1500 for PHEVs

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/en13102602

AMA Style

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(10):2602. https://doi.org/10.3390/en13102602

Chicago/Turabian Style

Ghosh, Aritra. 2020. "Possibilities and Challenges for the Inclusion of the Electric Vehicle (EV) to Reduce the Carbon Footprint in the Transport Sector: A Review" Energies 13, no. 10: 2602. https://doi.org/10.3390/en13102602

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