Accessibility Links

  • Skip to content
  • Skip to search IOPscience
  • Skip to Journals list
  • Accessibility help
  • Accessibility Help

Click here to close this panel.

Purpose-led Publishing is a coalition of three not-for-profit publishers in the field of physical sciences: AIP Publishing, the American Physical Society and IOP Publishing.

Together, as publishers that will always put purpose above profit, we have defined a set of industry standards that underpin high-quality, ethical scholarly communications.

We are proudly declaring that science is our only shareholder.

The rise of electric vehicles—2020 status and future expectations

Matteo Muratori 13,1 , Marcus Alexander 2 , Doug Arent 1 , Morgan Bazilian 3 , Pierpaolo Cazzola 4 , Ercan M Dede 5 , John Farrell 1 , Chris Gearhart 1 , David Greene 6 , Alan Jenn 7 , Matthew Keyser 1 , Timothy Lipman 8 , Sreekant Narumanchi 1 , Ahmad Pesaran 1 , Ramteen Sioshansi 9 , Emilia Suomalainen 10 , Gil Tal 7 , Kevin Walkowicz 11 and Jacob Ward 12

Published 25 March 2021 • © 2021 IOP Publishing Ltd Progress in Energy , Volume 3 , Number 2 Focus on Transport Electrification Citation Matteo Muratori et al 2021 Prog. Energy 3 022002 DOI 10.1088/2516-1083/abe0ad

Article metrics

70178 Total downloads

Permissions

Get permission to re-use this article

Share this article

Author e-mails.

[email protected]

Author affiliations

1 National Renewable Energy Laboratory, Golden, CO, United States of America

2 Electric Power Research Institute, Palo Alto, CA, United States of America

3 Colorado School of Mines, Golden, CO, United States of America

4 International Transport Forum in Paris, France

5 Toyota Research Institute of North America, Ann Arbour, MI, United States of America

6 University of Tennessee, Knoxville, TN, United States of America

7 University of California, Davis, CA, United States of America

8 University of California, Berkeley, CA, United States of America

9 The Ohio State University, Columbus, OH, United States of America

10 Institut VEDECOM, Versailles, France

11 Calstart, Pasadena, CA, United States of America

12 Carnegie Mellon University, Pittsburgh, PA, United States of America

Author notes

13 Author to whom any correspondence should be addressed.

Matteo Muratori https://orcid.org/0000-0003-1688-6742

Doug Arent https://orcid.org/0000-0002-4219-3950

Morgan Bazilian https://orcid.org/0000-0003-1650-8071

Ahmad Pesaran https://orcid.org/0000-0003-0666-1021

Emilia Suomalainen https://orcid.org/0000-0002-6339-2932

Gil Tal https://orcid.org/0000-0001-7843-3664

Jacob Ward https://orcid.org/0000-0002-8278-8940

  • Received 3 August 2020
  • Accepted 27 January 2021
  • Published 25 March 2021

Peer review information

Method : Single-anonymous Revisions: 3 Screened for originality? Yes

Buy this article in print

Electric vehicles (EVs) are experiencing a rise in popularity over the past few years as the technology has matured and costs have declined, and support for clean transportation has promoted awareness, increased charging opportunities, and facilitated EV adoption. Suitably, a vast body of literature has been produced exploring various facets of EVs and their role in transportation and energy systems. This paper provides a timely and comprehensive review of scientific studies looking at various aspects of EVs, including: (a) an overview of the status of the light-duty-EV market and current projections for future adoption; (b) insights on market opportunities beyond light-duty EVs; (c) a review of cost and performance evolution for batteries, power electronics, and electric machines that are key components of EV success; (d) charging-infrastructure status with a focus on modeling and studies that are used to project charging-infrastructure requirements and the economics of public charging; (e) an overview of the impact of EV charging on power systems at multiple scales, ranging from bulk power systems to distribution networks; (f) insights into life-cycle cost and emissions studies focusing on EVs; and (g) future expectations and synergies between EVs and other emerging trends and technologies. The goal of this paper is to provide readers with a snapshot of the current state of the art and help navigate this vast literature by comparing studies critically and comprehensively and synthesizing general insights. This detailed review paints a positive picture for the future of EVs for on-road transportation, and the authors remain hopeful that remaining technology, regulatory, societal, behavioral, and business-model barriers can be addressed over time to support a transition toward cleaner, more efficient, and affordable transportation solutions for all.

Export citation and abstract BibTeX RIS

This article was updated on 29 April 2021 to add the name of the fifth author and to correct the name of the eighth author.

1. Introduction

First introduced at the end of the 1800s, electric vehicles (EVs) 12 have been experiencing a rise in popularity over the past few years as the technology has matured and costs (especially of batteries) have declined substantially. Worldwide support for clean transportation options (i.e. low emissions of greenhouse gasses [GHG] to mitigate climate change and criteria pollutants) has promoted awareness, increased charging opportunities, and facilitated adoption of EVs. EVs present numerous advantages compared to fossil-fueled internal-combustion-engine vehicles (ICEVs), inter alia: zero tailpipe emissions, no reliance on petroleum, improved fuel economy, lower maintenance, and improved driving experience (e.g. acceleration, noise reduction, and convenient home and opportunity recharging). Further, when charged with clean electricity, EVs provide a viable pathway to reduce overall GHG emissions and decarbonize on-road transportation. This decarbonization potential is important, given limited alternative options to liquid fossil fuels. The ability of EVs to reduce GHG emissions is dependent, however, upon clean electricity. Therefore, EV success is intertwined closely with the prospect of abundant and affordable renewable electricity (in particular solar and wind electricity) that is poised to transform power systems (Jacobson et al 2015 , Kroposki et al 2017 , Gielen et al 2019 , IEA 2020b ). Coordinated actions can produce beneficial synergies between EVs and power systems and support renewable-energy integration to optimize energy systems of the future to benefit users and offer decarbonization across sectors (CEM 2020 ). A cross-sectoral approach across the entire energy system is required to realise clean future transformation pathways (Hansen et al 2019 ). EVs are expected to play a critical role in the power system of the future (Muratori and Mai).

EV success is increasing rapidly since the mid-2010s. EV sales are breaking previous records every year, especially for light-duty vehicles (LDVs), buses, and smaller vehicles such as three-wheelers, mopeds, kick-scooters, and e-bikes (IEA 2017 , 2018a , 2019 , 2020 ). To date, global automakers are committing more than $140 billion to transportation electrification, and 50 light-duty EV models are available commercially in the U.S. market (Moore and Bullard 2020 ). Approximately 130 EV models are anticipated by 2023 (AFDC 2020 , Moore and Bullard 2020 ). Future projections of the role of EVs in LDV markets vary widely, with estimates ranging from limited success (∼10% of sales in 2050) to full market dominance, with EVs accounting for 100% of LDV sales well before 2050. Many studies project that EVs will become economically competitive with ICEVs in the near future or that they are already cost-competitive for some applications (Weldon et al 2018 , Sioshansi and Webb 2019 , Yale E360 2019 , Kapustin and Grushevenko 2020 ). However, widespread adoption requires more than economic competitiveness, especially for personally owned vehicles. Behavioral and non-financial preferences of individuals on different technologies and mobility options are also important (Lavieri et al 2017 , Li et al 2017 , McCollum et al 2018 , Ramea et al 2018 ). EV adoption beyond LDVs has been focused on buses, with significant adoption in several regions (especially China). Electric trucks also are receiving great attention, and Bloomberg New Energy Finance (BloombergNEF) projects that by 2025, alternative fuels will compete with, or outcompete, diesel in long-haul trucking applications (Moore and Bullard 2020 ). These recent successes are being driven by technological progress, especially in batteries and power electronics, greater availability of charging infrastructure, policy support driven by environmental benefits, and consumer acceptance. EV adoption is engendering a virtuous circle of technology improvements and cost reductions that is enabled and constrained by positive feedbacks arising from scale and learning by doing, research and development, charging-infrastructure coverage and utilization, and consumer experience and familiarity with EVs.

Vehicle electrification is a game-changer for the transportation sector due to major energy and environmental implications driven by high vehicle efficiency (EVs are approximately 3–4 times more efficient than comparable ICEVs), zero tailpipe emissions, and reduced petroleum dependency (great fuel diversity and flexibility exist in electricity production). Far-reaching implications for vehicle-grid integration extend to the electricity sector and to the broader energy system. A revealing example of the role of EVs in broader energy-transformation scenarios is provided by Muratori and Mai, who summarize results from 159 scenarios underpinning the special report on Global Warming of 1.5 °C (SR1.5) by Intergovernmental Panel on Climate Change (IPCC). Muratori and Mai also show that transportation represents only ∼2% of global electricity demand currently (with rail responsible for more than two-thirds of this total). They show that electricity is projected to provide 18% of all transportation-energy needs by 2050 for the median IPCC scenario, which would account for 10% of total electricity demand. Most of this electricity use is targeted toward on-road vehicle electrification. These projections are the result of modeling and simulations that are struggling to keep pace with the EV revolution and its role in energy-transformation scenarios as EV technologies and mobility are evolving rapidly (McCollum et al 2017 , Venturini et al 2019 , Muratori et al 2020 ). Recent studies explore higher transportation-electrification scenarios: for example, Mai et al ( 2018 ) report a scenario in which 75% of on-road miles are powered by electricity, and transportation represents almost a quarter of total electricity use during 2050.

Vehicle electrification is a disruptive element in energy-system evolution that radically changes the roles of different sectors, technologies, and fuels in long-term transformation scenarios. Traditionally, energy-system-transformation studies project minimal end-use changes in transportation-energy use over time (limited mode shifting and adoption of alternative fuels), and the sector is portrayed as a 'roadblock' to decarbonization. In many decarbonization scenarios, transportation is seen traditionally as one of the biggest hurdles to achieve emissions reductions (The White House 2016 ). These scenarios rely on greater changes in the energy supply to reduce emissions and petroleum dependency (e.g. large-scale use of bioenergy, often coupled to carbon capture and sequestration) rather than demand-side transformations (IPCC 2014 , Pietzcker et al 2014 , Creutzig et al 2015 , Muratori et al 2017 , Santos 2017 ). In most of these studies, electrification is limited to some transportation modes (e.g. light-duty), and EVs are not expected to replace ICEVs fully (The White House 2016 ). More recently, however, major mobility disruptions (e.g. use of ride-hailing and vehicle ride-sharing) and massive EV adoption across multiple applications are proposed (Edelenbosch et al 2017 , Van Vuuren et al 2017 , Hill et al 2019 , E3 2020 , Zhang and Fujimori 2020 ). These mobility disruptions allow for more radical changes and increase the decarbonization role of transportation and highlight the integration opportunities between transportation and energy supply, especially within the electricity sector. For example, Zhang and Fujimori ( 2020 ) highlight that for vehicle electrification to contribute to climate-change mitigation, electricity generation needs to transition to clean sources. They note that EVs can reduce mitigation costs, implying a positive impact of transport policies on the economic system (Zhang and Fujimori 2020 ). In-line with these projections, many countries are establishing increasingly stringent and ambitious targets to support transport electrification and in some cases ban conventional fossil fuel vehicles (Wentland 2016 , Dhar et al 2017 , Coren 2018 , CARB 2020 , State of California 2020 ).

EV charging undoubtedly will impact the electricity sector in terms of overall energy use, demand profiles, and synergies with electricity supply. Mai et al ( 2018 ) show that in a high-electrification scenario, transportation might grow from the current 0.2% to 23% of total U.S. electricity demand in 2050 and significantly impact system peak load and related capacity costs if not controlled properly. Widespread vehicle electrification will impact the electricity system across the board, including generation, transmission, and distribution. However, expected changes in U.S. electricity demand as a result of vehicle electrification are not greater than historical growth in load and peak demand. This finding suggests that bulk-generation capacity is expected to be available to support a growing EV fleet as it evolves over time, even with high EV-market growth (U.S. DRIVE 2019 ). At the same time, many studies have shown that 'smart charging' and vehicle-to-grid (V2G) services create opportunities to reduce system costs and facilitate the integration of variable renewable energy (VRE). Charging infrastructure that enables smart charging and alignment with VRE generation, as well as business models and programs to compensate EV owners for providing charging flexibility, are the most pressing required elements for successfully integrating EVs with bulk power systems. At the local level, EV charging could increase and change electricity loads significantly, which could impact distribution networks and power quality and reliability (FleetCarma 2019 ). Distribution-network impacts can be particularly critical for high-power charging and in cases in which many EVs are concentrated in a specific location, such as clusters of residential LDV charging and possibly fleet depots for commercial vehicles (Muratori 2018 ).

This paper provides a timely status of the literature on several aspects of EV markets, technologies, and future projections. The paper focuses on multiple facets that characterize technology status and the role of EVs in transportation decarbonization and broader energy-transformation pathways focusing on the U.S. context. As appropriate, global context is provided as well. Hybrid EVs (for which liquid fuel is the only source of energy) and fuel cell EVs (that power an electric powertrain with a fuel cell and on-board hydrogen storage) have some similarities with EVs and could complement them for many applications. However, these technologies are not reviewed in detail here. The remainder of this paper is structured as follows. Section 2 focuses on the status of the light-duty-EV market and provides a comparison of projections for future adoption. Section 3 provides insights on market opportunities beyond LDVs. Section 4 offers a review of cost and performance evolution for batteries, power electronics, and electric machines that are key components of EV success. Section 5 reviews charging-infrastructure status and focuses on modeling and analysis studies used to project charging-infrastructure requirements, the economics of public charging, and some considerations on cybersecurity and future technologies (e.g. wireless charging). Section 6 provides an overview of the impact of EV charging on power systems at multiple scales, ranging from bulk power systems to distribution networks. Section 7 provides insights into life-cycle cost and emissions studies focusing on EVs. Finally, section 8 touches on future expectations.

1.1. Summary of take-away points

1.1.1. ev adoption.

  • The global rate of adoption of light-duty EVs (passenger cars) has increased rapidly since the mid-2010s, supported by three key pillars: improvements in battery technologies; a wide range of supportive policies to reduce emissions; and regulations and standards to promote energy efficiency and reduce petroleum consumption.
  • Adoption of advanced technologies has been underestimated historically in modeling and analyses; EV adoption is projected to remain limited until 2030, and high uncertainty is shown afterward with widely different projections from different sources. However, EVs hold great promise to replace conventional LDVs affordably.
  • Barriers to EV adoption to date include consumer skepticism toward new technology, high purchase prices, limited range and lack of charging infrastructure, and lack of available models and other supply constraints.
  • A major challenge facing both manufacturers and end-users of medium- and heavy-duty EVs is the diverse set of operational requirements and duty cycles that the vehicles encounter in real-world operation.
  • EVs appear to be well suited for short-haul trucking applications such as regional and local deliveries. The potential for battery-electric models to work well in long-haul on-road applications has yet to be established, with different studies indicating different opportunities.

1.1.2. Batteries and other EV technologies

  • Over the last 10 years, the price of lithium-ion battery packs has dropped by more than 80% (from over $1000 kWh −1 to $156 kWh −1 at the end of 2019, BloombergNEF 2020 ). Further price reduction is needed to achieve EV purchase-price parity with ICEVs.
  • Over the last 10 years, the specific energy of a lithium-ion battery cell has almost doubled, reaching 240 Wh kg −1 (BloombergNEF 2020 ), reducing battery weight significantly.
  • Reducing or eliminating cobalt in lithium-ion batteries is an opportunity to lower costs and reduce reliance on a rare material with controversial supply chains.
  • While batteries are playing a key role in the rise of EVs, power electronics and electric motors are also key components of an EV powertrain. Recent trends toward integration promise to deliver benefits in terms of increased power density, lower losses, and lower costs.

1.1.3. Charging infrastructure

  • With a few million light-duty EVs on the road, currently, there is about one public charge point per ten battery electric vehicles (BEVs) in U.S. (although most vehicles have access to a residential charger).
  • Given the importance of home charging (and the added convenience compared to traditional refueling at public stations), charging solutions in residential areas comprising attached or multi-unit dwellings is likely to be essential for EVs to be adopted at large scale.
  • Although public charging infrastructure is clearly important to EV purchasers, how best to deploy charging infrastructure in terms of numbers, types, locations, and timing remains an active area for research.
  • The economics of public charging vary with location and station configuration and depend critically on equipment and installation costs, incentives, non-fuel revenues, and retail electricity prices, which are heavily dependent on station utilization.
  • The electrification of medium- and heavy-duty commercial trucks and buses might introduce unique charging and infrastructure requirements compared to those of light-duty passenger vehicles.
  • Wireless charging, specifically high-power wireless charging (beyond level-2 power), could play a key role in providing an automated charging solution for tomorrow's automated vehicles.

1.1.4. Power system integration

  • Accommodating EV charging at the bulk power-system level (generation and transmission) is different in each region, but there are no major known technical challenges or risks to support a growing EV fleet, especially in the near term (approximately one decade).
  • At the local level, however, EV charging can increase and change electricity loads significantly, causing possible negative impacts on distribution networks, especially for high-power charging.
  • The integration of EVs into power systems presents opportunities for synergistic improvement of the efficiency and economics of electromobility and electric power systems, and EVs can support grid planning and operations in several ways.
  • There are still many challenges for effective EV-grid integration at large scale, linked not only to the technical aspects of vehicle-grid-integration (VGI) technology but also to societal, economic, business model, security, and regulatory aspects.
  • VGI offers many opportunities that justify the efforts required to overcome these challenges. In addition to its services to the power system, VGI offers interesting perspectives for the full exploitation of synergies between EVs and VRE as both technologies promise large-scale deployment in the future.

1.1.5. Life-cycle cost and emissions

  • Many factors contribute to variability in EV life-cycle emissions, mostly the carbon intensity of electricity, charging patterns, vehicle characteristics, and even local climate. Grid decarbonization is a prerequisite for EVs to provide major GHG-emissions reductions.
  • Existing literature suggests that future EVs can offer 70%–90% lower GHG emissions compared to today's ICEVs, most obviously due to broad expectations for continued grid decarbonization.
  • Operational costs of EVs (fuel and maintenance) are typically lower than those of ICEVs, largely because EVs are more efficient than ICEVs and have fewer moving parts.

1.1.6. Synergies with other technologies and future expectations

  • Vehicle electrification fits in broader electrification and mobility macro-trends, including micro-mobility in urban areas, new mobility business models regarding ride-hailing and car-sharing, and automation that complement well with EVs.
  • While EVs are a relatively new technology and automated vehicles are not readily available to the general public, the implications and potential synergies of these technologies operating in conjunction are significant.
  • The coronavirus pandemic is impacting transportation markets negatively, including those for EVs, but long-term prospects remain undiminished.
  • Several studies project major roles for EVs in the future, which is reflected in massive investment in vehicle development and commercialization, charging infrastructure, and further technology improvement. Consumer adoption and acceptance and technology progress form a virtuous self-reinforcing circle of technology-component improvements and cost reductions.
  • EVs hold great promise to replace ICEVs affordably for a number of on-road applications, eliminating petroleum dependence, improving local air quality and enabling GHG-emissions reductions, and improving driving experiences.
  • Forecasting the future, including technology adoption, remains a daunting task. However, this detailed review paints a positive picture for the future of EVs across a number of on-road applications.

2. Status of electric-LDV market and future projections

This section provides a current snapshot of the electric-LDV market in a global and U.S. context, but focuses on the latter. The global rate of adoption of electric LDVs has increased rapidly since the mid-2010s 13 . By the end of 2019, the global EV fleet reached 7.3 million units—up by more than 40% from 2018—with more than 1.25 million electric LDVs in the U.S. market alone (IEA 2020 ). EV sales totaled more than 2.2 million in 2019, exceeding the record level that was attained in 2018, despite mixed performances in different markets. Electric-LDV sales increased in Europe and stagnated or declined in other major markets, particularly in China (with a significant slowdown due to changes in Chinese subsidy policy in July 2019), Japan, and U.S. U.S. EV adoption varies greatly geographically—nine counties in California currently see EVs accounting for more than 10% of sales (8% on average for California as a whole), but national-level sales remain at less than 3% (Bowermaster 2019 ). BEV sales exceeded those of plug-in hybrid electric vehicles (PHEVs) in all regions.

The rapid increase in EV adoption is underpinned by three key pillars:

  • (a)   Improvements and cost reductions in battery technologies, which were enabled initially by the large-scale application of lithium-ion batteries in consumer electronics and smaller vehicles (e.g. scooters, especially in China, IEA 2017 ). These developments offer clear and growing opportunities for EVs and HEVs to deliver a reduced total cost of ownership (TCO) in comparison with ICEVs.
  • (b)   A wide range of supportive policy instruments for clean transportation solutions in major global markets (Axsen et al 2020 ), which are mirrored by private-sector investment. These developments are driven by environmental goals (IPCC 2014 ), including reduction of local air pollution. These policy instruments support charging-infrastructure deployment (Bedir et al 2018 ) and provide monetary (e.g. rebates and vehicle-registration discounts) and non-monetary (e.g. access to high-occupancy-vehicle lanes and preferred parking) incentives to support EV adoption (IEA 2018a , AFDC 2020 ).
  • (c)   Regulations and standards that support high-efficiency transportation solutions and reduce petroleum consumption (e.g. fuel-economy standards, zero-emission-vehicle mandates, and low-carbon-fuel standards). These regulations are being supported by technology-push measures, consisting primarily of economic instruments (e.g. grants and research funds) that aim to stimulate technological progress (especially batteries), and market-pull measures (e.g. public-procurement programs) that aim to support the deployment of clean-mobility technologies and enable cost reductions due to technology learning, scale, and risk mitigation.

Transport electrification also has started a virtuous self-reinforcing circle. Battery-technology developments and cost reductions triggered by EV adoption provide significant economic-development opportunities for the companies and countries intercepting the battery and EV value chains. Adoption of alternative vehicles both is enabled and constrained by powerful positive feedback arising from scale and learning by doing, research and development, consumer experience and familiarity with technologies (e.g. neighborhood effect), and complementary resources, such as fueling infrastructure (Struben and Sterman 2008 ). In this context, more diversity in make and model market offerings is supporting vehicle adoption. As of April 2020, there are 50 EV models available commercially in U.S. markets (AFDC 2020 ), and ∼130 are anticipated by 2023 (Moore and Bullard 2020 ).

Measures that support transport electrification have been, and increasingly shall be, accompanied by policies that control for the unwanted consequences. Thus, the measures need to be framed in the broader energy and industry contexts.

When looking at the future, EV-adoption forecasts remain highly uncertain. Technology-adoption projections are used by a number of stakeholders to guide investments, inform policy design and requirements (Kavalec et al 2018 ), assess benefits of previous and ongoing efforts (Stephens et al 2016 ), and develop long-term multi-sectoral assessments (Popp et al 2010 , Kriegler et al 2014 ). However, projecting the future, including technology adoption, is a daunting task. Past projections often have turned out to be inaccurate. Still, progress has been made to address projection uncertainty (Morgan and Keith 2008 , Reed et al 2019 ) and contextualize scenarios to explore alternative futures in a useful way. Scenario analysis is used largely in the energy-environment community to explore the possible implications of different judgments and assumptions by considering a series of 'what if' experiments (BP 2019 ).

Adoption of advanced technologies historically has been underestimated in modeling and analysis results (e.g. Creutzig et al 2017 ), which fail to capture the rapid technological progress and its impact on sales. Historical experiences suggest that technology diffusion, while notoriously difficult to predict, can occur rapidly and with an extensive reach (Mai et al 2018 ). Projecting personally owned LDV sales is particularly challenging because decisions are made by billions of independent (not necessarily rational) decision-makers valuing different vehicle attributes based on incomplete information (e.g. misinformation and skepticism toward new technologies) and limited financial flexibility.

Many studies make projections for future EV sales (see figure 1 for a comparison of different projections). Some organizations (e.g. Energy Information Administration [EIA]) historically have been conservative in projecting EV success, mostly due to scenario constraints and assumptions. Still, U.S. EV-sales projections from EIA in recent years are much higher than in the past. Others (e.g. BloombergNEF and Electric Power Research Institute [EPRI]) consistently have been more optimistic in terms of EV sales and continue to adjust sales projections upward. Policy ambition for EV adoption is also optimistic. For example, in September 2020, California passed new legislation that requires that by 2035 all new car and passenger-truck sales be zero-emission vehicles (and that all medium- and heavy-duty vehicles be zero-emission by 2045) (California, 2020). Projected EV sales and outcomes from major energy companies vary widely, ranging from somewhat limited EV adoption (e.g. ExxonMobil) to full market success (e.g. Shell). A survey from Columbia University (Kah 2019 ) considers 17 studies and shows that 'EV share of the global passenger vehicle fleet is not projected to be substantial before 2030 given the long lead time in turning over the global automobile fleet' and that 'the range of EVs in the 2040 fleet is 10% to 70%'. The studies compared in figure 1 show an even greater variability for 2050 projections, ranging from 13% to 100% of U.S. EV adoption for LDVs.

Figure 1.

Figure 1.  Electric LDV (BEV and PHEV) new sales projections from numerous international sources. Unless otherwise noted, data refer to new U.S. sales. AEO2015 = EIA Annual Energy Outlook 2015, Reference Scenario. AEO2017 = EIA Annual Energy Outlook 2017, Reference Scenario. AEO2020 = EIA Annual Energy Outlook 2020, Reference Scenario. AEO2020HO = EIA Annual Energy Outlook 2020, High Oil Scenario. EFS Med = National Renewable Energy Laboratory (NREL) Electrification Futures Study, Medium Scenario. EFS High = NREL Electrification Futures Study, High Scenario. EPRI Med = EPRI Plug-in Electric Vehicle Projections: Scenarios and Impacts, Medium Scenario. EPRI High = EPRI Plug-in Electric Vehicle Projections: Scenarios and Impacts, High Scenario. EPRI NEA = EPRI National Electrification Assessment, Reference Scenario. GEVO NP = IEA Global EV Outlook 2019, New Policies Scenario. GEVO CEM = IEA Global EV Outlook 2019, Clean Energy Ministerial 30@30 Campaign Scenario. BNEF = BloombergNEF EV Outlook 2020. Equinor Riv = Equinor 2019 Energy Perspectives, Rivalry Scenario. Equinor Ren = Equinor 2019 Energy Perspectives, Renewal Scenario. Shell Sky = Shell Sky Scenario. ExxonMobil = 2019 ExxonMobil Outlook for Energy. IEEJ Ref = The Institute of Energy Economics, Japan. 2019 Outlook, Reference Scenario (global sales). IEEJ Adv = The Institute of Energy Economics, Japan. 2019 Outlook, Advanced Technologies Scenario (global sales). CA ZEV Mandate = California zero-emission vehicle (ZEV) Executive Order N-79-20 (September 2020).

Download figure:

The future remains uncertain, but there is a clear trend in projections of light-duty EV sales toward more widespread adoption as the technology improves, consumers become more familiar with the technology, automakers expand their offerings, and policies continue to support the market.

A number of studies analyze the drivers of EV adoption (Vassileva and Campillo 2017 , Priessner et al 2018 ) and highlight several barriers for EVs to achieve widespread success, including consumer skepticism for new technologies (Egbue and Long 2012 ); uncertainty around environmental benefits (consumers wonder whether EVs actually are green; see section 7 for more clarity on the environmental benefits of EVs) and continued policy support; unclear battery aging/resale value; high costs (Haddadian et al 2015 , Rezvani et al 2015 , She et al 2017 ); lack of charging infrastructure (Melaina et al 2017 , Narassimhan and Johnson 2018 ); range anxiety (the fear of being unable to complete a trip) associated with shorter-range EVs; longer refueling times compared to conventional vehicles (Franke and Krems 2013 , Neubauer and Wood 2014 ; Melaina et al 2017 ); dismissive and deceptive car dealerships (De Rubens et al 2018 ); and other EV-supply considerations, such as limited model availability and limited supply chains.

A recent review of 239 articles published in top-tier journals focusing on EV adoption draws attention to 'relatively neglected topics such as dealership experience, charging infrastructure resilience, and marketing strategies as well as identifies much-studied topics such as charging infrastructure development, TCO, and purchase-based incentive policies' (Kumar and Alok 2019 ). Similar reviews published recently focus on different considerations, such as market heterogeneity (Lee et al 2019a ), incentives and policies (Hardman 2019 , Tal et al 2020 ), and TCO (Hamza et al 2020 ). Other than some limited discussions on business models and TCO, the literature is focused on one side of the story, namely demand. However, the availability (makes and models) of EVs is extremely limited compared to ICEVs (AFDC 2020 ). This is justified, in part, by new technologies requiring time to be introduced, and, in part, by the higher manufacturer revenues associated with selling and providing maintenance for ICEVs. Moreover, slow turnover in legacy industry (Morris 2020 ) and other supply constraints can be a major barrier to widespread EV uptake (Wolinetz and Axsen 2017 , De Rubens et al 2018 ). Kurani ( 2020 ) argues that in most cases, 'Results of large sample surveys and small sample workshops mutually reinforce the argument that continued growth of PEV markets faces a barrier in the form of the inattention to plug-in electric vehicles (PEVs) of the vast majority of car-owning and new-car-buying households even in a place widely regarded as a leader. Most car-owning households are not paying attention to PEVs or the idea of a transition to electric-drive.'

3. EVs beyond light-duty applications

While much of the recent focus on vehicle electrification is with LDVs and small two- or three-wheelers (primarily in China), major progress also is being made with the electrification of medium- and heavy-duty vehicles. This includes heavy-duty trucks of various types, urban transit buses, school buses, and medium-duty vocational vehicles. As of the end of 2019, there were about 700 000 medium- and heavy-duty commercial EVs in use around the world (EV Volumes 2020 , IEA 2020 ).

A major challenge facing both manufacturers and end-users of medium- and heavy-duty vehicles is the diverse set of operational requirements and duty cycles that the vehicles encounter in real-world operation. When designing powertrain configurations and on-board energy-storage needs for new technologies, it is of critical importance to represent vehicle behavior accurately for different operations, including possible changes triggered by electrification (Delgado-Neira 2012 ). Medium- and heavy-duty vehicles can require a large number of powertrain and battery configurations, control strategies, and charging solutions. These needs depend on vehicle type (covering the full U.S. gross vehicle weight ratings [GVWR] spectrum from class 3 to class 8, 10 001–80 000 lb [4536–36 287 kg]), commercial operational situations and activities, and diverse drive cycles and charging opportunities (e.g. depot-based operations vs. long-haul). An example of this potential variability and its effect on the required battery capacity across multiple vehicle vocations is shown in figure 2 (Smith et al 2019 ).

Figure 2.

Figure 2.  Battery capacity requirements vs. weight class for medium- and heavy-duty vehicles (Smith et al 2019 ).

Another example of the highly variable use cases for medium- and heavy-duty EVs shows energy efficiencies range between 0.8 kWh mile −1 and 3.2 kWh mile −1 (0.5–2.5 kWh km −1 ) (Gao et al 2018 ). If the on-board energy-storage needs for these vehicles are considered, assuming a daily operational range of between 50 miles and 200 miles (80–322 km), this results in battery-size requirements between 40 kWh and 640 kWh (assuming that the vehicle is recharged once daily). If additional charging strategies are considered (with their variability in expected charge times and associated power ratings), the range of vehicle-hardware and charging-infrastructure possibilities increases further. When adding variability across use cases with respect to temperature effects, battery-capacity degradation, payload, and road grade, it becomes clear that medium- and heavy-duty truck manufacturers face a significant challenge in designing, developing, and manufacturing systems that are able to meet the diverse operational requirements.

There are potential synergies between components of light-duty and medium- and heavy-duty electric vehicles. However, the requirements of medium- and heavy-duty vehicles place much greater burdens on powertrain components. The power and energy needs in heavy-duty applications are much larger than in light-duty applications. Heavy-duty vehicles could demand twice the peak power, four times the torque, and can consume more than five times the energy per mile (or km) driven compared to LDVs. In addition to using more energy per mile (or km) driven, typically, commercial vehicles drive many more miles (or km) per day, requiring much larger batteries and possibly much higher-power charging. Moreover, heavy-duty-vehicle users expect their vehicles to last more than a million miles, pointing to significantly higher durability requirements for heavy-duty-vehicle components (Smith et al 2019 ). Overall, these requirements, in combination with the needs for very high durability and very high-power drivelines and charging, may cause battery chemistries of heavy-duty vehicle batteries to diverge from those that are used in LDVs, hindering economies of scale. Demands for high efficiency, high power, and lower weight will put pressure on commercial vehicles to work at higher voltages than LDVs do. While LDVs are designed typically with powertrains that operate at a few hundred volts, it may be desirable to design large EVs with kilovolt powertrains. This will have a particularly significant impact on power electronics and could drive the development of wide-bandgap power electronics.

Historically, EVs have not been considered capable alternatives to heavy-duty diesel trucks (above 33 000 lb [14 969 kg] GVWR) due to high capital costs, high energy and power requirements, and weight and range-related battery constraints. International Council on Clean Transportation (ICCT), for example, suggests that while conventional EV-charging methods may be sufficient for small urban commercial vehicles, overhead catenary or in-road charging are required for heavier vehicles (Moultak et al 2017 ). Recent studies dispute this, anticipating a much greater opportunity for EVs to replace diesel trucks in the short-term, even for long-haul applications (Mai et al 2018 , McCall and Phadke 2019 , Borlaug et al Forthcoming ), but the potential for battery-electric models to work well in long-haul applications has yet to be established (NACFE 2018 ). Studies show that a significant amount of payload capacity will be consumed by batteries, potentially up to 7 tons or 28% of capacity in a truck with a 500 mile (805 km) range with 1100 kWh battery capacity (Burke and Fulton 2019 ). Thus, batteries would reduce significantly the amount of cargo that can be carried. Other studies suggest this could be much less―on the order of 4% of lost payload capacity for 500 mile range (805 km) trucks and with overall lower TCO than diesel trucks (Phadke et al 2019 ). For short-haul applications, such as port drayage and regional or local deliveries, EVs appear well suited and battery weight may not affect the cargo or payload capacity adversely. Several heavy-duty battery-electric trucks for short- and medium-haul applications have been developed and tested in recent years by Balqon, Daimler Trucks NA, Peterbilt, TransPower, Tesla, US Hybrid, Volvo, and others (AFDC 2020 ).

Urban buses are also a major emerging market for electrification. In California, Innovative Clean Transit rules require transit agencies to transition completely to zero-emission technologies (batteries or fuel cells), with all new bus purchases being zero-emission by 2029 (CARB 2018 ). Eight of the ten largest transit agencies in California already are adopting zero-emission technologies into their fleets (CARB 2018 ). In a comparative study of urban buses running on diesel, compressed natural gas, diesel hybrid, fuel cells, and batteries, the battery buses are estimated to have the lowest CO 2 emissions in both California and Finland bus duty cycles at the time of the study (Lajunen and Lipman 2016 ). This study also shows that battery buses have only slightly higher overall costs per mile (or km) than fossil-fuel-based alternatives. Future projections out to 2030 show that electric buses have the lowest overall life-cycle costs, especially when CO 2 costs are included (Lajunen and Lipman 2016 ).

Medium-duty delivery vehicles (typically 10 000–33 000 lb [4536–14 969 kg] GVWR) are another attractive emerging area for electrification. The goods-delivery market is growing at approximately 9% per year in recent years, with a projected $343 billion global industry value in 2020 (Accenture 2015 ). The 'last mile' delivery vehicles that are needed for this market are undergoing changes and present good opportunities for electrification. Amazon, for example, has announced plans to purchase 100 000 custom-designed Rivian electric delivery vans by 2030, with 10 000 of the vehicles delivered by late 2022 (Davies 2019 ).

A significant challenge with electrifying these heavy- and medium-duty vehicles revolves around the installation of the required charging infrastructure (either at depots or along highways). While LDVs typically charge at power levels of 3 kW–10 kW, and potentially 50 kW–250 kW with DC fast chargers (DCFCs), a heavy-duty vehicle may require higher-power charging, depending on its duty cycle. Fleets of these vehicles charging in one location, such as a truck depot or travel center, may require several megawatts of power. This requires expensive charging infrastructure, potentially including costly and time-consuming distribution-grid upgrades, to provide the higher voltage and current levels that are needed. For example, a single 350 kW DCFC that may be suitable for heavy-duty applications costs almost $150 000 today (Nelder and Rogers 2019 , Nicholas 2019 ). These costs would, in turn, impact the business case for vehicle electrification. Potential costs of grid upgrades to support these new electrical loads would be additional expenses that may or may not be supported by the local utility, depending on the circumstances. To enable reliable, low-cost charging, which is crucial when considering the TCO for a fleet owner, the installation and operational costs of the charging infrastructure must be optimized, requiring engagement with power-supply stakeholders.

4. Batteries, power electronics, and electric machines

Electrification is a key aspect of modern life, and electric motors and machines are prevalent in manufacturing, consumer electronics, robotics, and EVs (Zhu and Howe 2007 ). One reason for the recent success and rise in adoption of EVs is the use of advanced lithium-ion batteries with improved performance, life, and lower cost. Improved energy and power performance, increased cycle and calendar life, and lower costs are leading to EVs with longer electric range and better acceleration at lower cost premia that are attracting consumers. This section summarizes the state-of-the-art for batteries and for power electronics, electric machines, and electric traction drives in terms of cost, performance, power and energy density, and reliability, and highlights some research challenges, pathways, and targets for the future.

4.1. Batteries

Over the last 10 years, the price of a lithium-ion battery pack has dropped by almost 90% from over $1000 kWh −1 in 2010 to $156 kWh −1 at the end of 2019 (BloombergNEF 2020 ). Meanwhile, the specific energy of a lithium-ion battery cell has almost doubled from 140 Wh kg −1 to 240 Wh kg −1 during that same window of time (BloombergNEF 2020 ). The improvement in performance and cost comes mainly from engineering improvements, use of materials with higher capacities and voltages, and development of methods to increase stability for longer life and improved safety. Improvements in cell, module, and pack design also help to improve performance and lower costs. Increases in manufacturing volume due to EV sales contribute significantly to cost reductions (Nykvist and Nilsson 2015 , Nykvist et al 2019 ). However, further reductions in battery costs, along with a reduction in the cost of electric machines and power electronics, are needed for EVs to achieve purchase-price parity with ICEVs. This parity is estimated by U.S. Department of Energy (DOE) to be achieved at battery costs of ∼$100 kWh −1 (preferably less than $80 kWh −1 ) (VTO, 2020 ). At that point, EVs should have both a purchase- and a lifetime-operating-cost benefit over ICEVs. Such cost benefits are likely to trigger drastic increases in EV sales. Figure 3 shows the observed price of lithium-ion battery packs from 2010 to 2018, as well as estimated prices through 2030. BloombergNEF projects that by 2024 the price for original equipment manufacturers (OEMs) to acquire battery packs will go below $100 kWh −1 and reach ∼$60 kWh −1 by 2030 if high levels of investments continue in the future (BloombergNEF 2020 ).

Figure 3.

Figure 3.  Evolution of battery prices over the last 10 years and future projections (Goldie-Scot 2019 ). BloombergNEF 2019.

The typical anode material that is used in most lithium-ion EV batteries is graphite (Ahmed et al 2017 ). Research is underway to utilize silicon, in addition to graphite, due to its higher specific-energy capacity. For cathodes, there is more variety (Lee et al 2019 , Manthiram 2020 ). Consumer electronics such as mobile phones and computers almost exclusively have used lithium cobalt oxide, LiCoO 2 , due to its high specific-energy density (Keyser et al 2017 ). Most EV manufacturers (except Tesla) have avoided using LiCoO 2 in EVs due to its high cost and safety concerns. Lithium iron phosphate also has been used for electric cars and buses because of its long life and better safety and power capabilities. However, due to its low specific-energy density (110 Wh kg −1 ) when paired with a graphite anode, lithium iron phosphate is not used commonly for light-duty EVs in U.S. In recent years, battery makers and vehicle OEMs have moved to lithium nickel manganese cobalt oxides (NMC) with varying ratios of the three transition metals. Initially, OEMs used NMC111 (the numbers represent the molar fractions of nickel, manganese, and cobalt, which are equal in this case), but they have transitioned to NMC532 and utilize NMC622 now while working to stabilize the NMC811 cathode structure. The goal is eventually to reduce the amount of cobalt in the cathode to less than 5% and perhaps even eliminate the use of cobalt. The use of these cathodes with higher specific-energy density and less cobalt leads to lower battery cost per unit energy ($ kWh −1 ). Table 1 shows the specific energy and estimated (bottom-up) cost from Argonne National Laboratory's BatPaC Battery Performance and Cost model (Ahmed et al 2016 ) based on large-volume material processing, cell manufacturing, and pack manufacturing.

Table 1.  Calculated specific energy and cost of advanced lithium-ion batteries with different cathode/anode chemistries. Numbers are from BatPaC (Ahmed et al 2016 ) and are intended for relative comparison only. Final values can change depending on the components used and production volume, and costs reported do not reflect what a negotiated price could be between a battery and EV maker.

The cost of batteries is expected to decline in the future due to improved capacity of materials (such as Si anodes), increased percentage of active material components, use of lower-cost elements (no cobalt), improved packaging, and continued automation to increase yield while leading to a longer electric range. However, price increases for certain metals such as Ni and Li could prevent achieving those lower-battery-cost projections. Moreover, different battery chemistries can lead to very different costs and specific energies. For example, table 1 shows results obtained from bottom-up calculations with Argonne National Laboratory's BatPaC Battery Performance and Cost Model (Ahmed et al 2016 ), for a 100 kWh battery pack showing great variability in battery cost and performance for different chemistries.

Opportunities to improve performance and reduce costs further are being pursued in a number of major research areas. The battery community is investigating a number of materials, with the aim of reducing the cost and increasing the energy density of battery systems (Deign and Pyper 2018 ). Future work will involve utilizing silicon (Salah et al 2019 ) or lithium metal (Zhang et al 2020 ) as the anode while utilizing high-energy cathodes, such as NMC811 or lithium sulfur (Zhu et al 2019 ). Reducing the amount of critical material in lithium-ion batteries, especially cobalt, is an opportunity to lower the cost of batteries and improve supply-chain resilience. The private and public sectors are working toward developing new cathode materials along these lines (Li et al 2009 , 2017b ). Research and development (R&D) projects are underway to develop infrastructure and recycling technologies to collect batteries and recover the key battery materials economically and environmentally (Harper et al 2019 ). Reuse of end-of-life batteries from EVs would delay the need for additional battery materials, which should have positive environmental benefits (Neubauer et al 2012 ). Different battery technologies also are being explored. To increase energy density, reduce cost, and improve safety, the battery community is pursuing development of solid-state batteries with solid-state electrolytes (Randau et al 2020 ) that have ionic conductivities approaching those of today's liquid electrolyte systems. Solid-state lithium batteries enable the use of metallic lithium anodes, together with solid electrolytes and high-energy cathodes (such as high-nickel NMC or sulfur). Lithium-metal batteries based on solid electrolytes can, in principle, alleviate the safety concerns with current lithium-ion batteries with a flammable organic electrolyte. The main challenges facing lithium-metal anodes are dendritic growth, especially at low temperatures and higher current rates. Dendritic growth could lead to short circuit and thermal runaway and low Coulombic efficiency leading to poor cycle life (Xia et al 2019 ). Slow ion transport through the solid-state electrolyte leading to low power densities and manufacturing challenges, including poor mechanical integrity, pose additional challenges. Significant R&D activities are focused on developing solid-state electrolytes that prevent dendrite growth, have high ionic conductivity, good voltage-stability windows, and low impedance at the electrode–electrolyte interface. Recent cathode formulations in Li-S cells overcome the polysulfide problem, which could lead to lower efficiency and cycle life. Nevertheless, the deployment of cells with lower electrolyte-to-sulfur ratios for scale-up to large sizes is a remaining challenge. It may take another 5 to 10 years to mass-produce solid-state lithium batteries for EV applications.

As is discussed in section 5 , a network of fast chargers and batteries that can handle high charging-power rates is needed to address any potential barriers to widespread EV adoption. Research is focusing on developing batteries that can be charged very quickly (e.g. 80% of capacity in less than 15 min). A number of challenges to high-power charging, such as lithium plating, thermal management, and poor cycle life, need to be addressed (Ahmed et al 2017 ; DOE 2017 , Michelbacher et al 2017 ). Significant efforts also have focused on developing electrochemical and thermal modeling of batteries for EV applications (Kim et al 2011 , Chen et al 2016 , Keyser et al 2017 , Zhang et al 2017 ) to improve battery lifetime and efficiency in real-world applications. These efforts include lifetime-estimation and degradation modeling under different real-world climate and driving conditions (Hoke et al 2014 , Neubauer and Wood 2014 , Liu et al 2017b , Harlow et al 2019 , Li et al 2019 ); simplified models for control and diagnostics (e.g. state-of-charge estimation) (Muratori et al 2010 , Fan et al 2013 , Cordoba-Arenas et al 2015 , Bartlett et al 2016 ); and developing effective thermal management and control strategies (Pesaran 2001 , Serrao et al 2011 ).

Besides EV applications, batteries can offer energy-storage solutions for hybrid- or distributed-energy systems. These solutions include the use of batteries in integrated configurations with wind or solar photovoltaic (PV) systems or with EV fast-charging stations (Bernal-Agustín and Dufo-Lopez 2009 , Badwawi et al 2015 , Muratori et al 2019a ). Batteries also can provide stabilization and flexibility and can improve resilience and efficiency for power systems in general, especially for critical services or when a high share of variable power generation (e.g. from solar or wind) is expected (Divya and Østergaard 2009 , Denholm et al 2013 ; De Sisternes et al 2016 ). Lithium-ion batteries that have been developed for EV applications have found their way into stationary applications (Pellow et al 2020 ) because of their lower cost and modularity compared to other energy-storage technologies (Chen et al 2020 ). Moreover, EV batteries can be reused or repurposed at the end of their 'vehicle life' (usually considered when energy storage capacity drops below 70%–80% of the original nominal value, (Podias et al 2018 )) for stationary applications, improving their economic and environmental performance (Assuncao et al 2016 , Ahmadi et al 2017 , Martinez-Laserna et al 2018 , Olsson et al 2018 , Kamath et al 2020 ).

4.2. Power electronics, electric machines, and electric-traction-drive systems

While batteries are playing a key role in the rise of EVs, power electronics and electric motors and machines are also key components of an EV powertrain. Traditionally, the motor and power electronics drive were separate components in an EV. However, recent trends toward integration promise to deliver benefits in terms of increased power density, lower losses, and lower costs compared with separate motor and motor-drive solutions (Reimers et al 2019 ). Figure 4 shows the 2020 power density for power electronics, electric machines, and electric-traction-drive system from some example commercial vehicles as well as the 2025 DOE and U.S. DRIVE Partnership targets for near-term improvements (U.S. DRIVE 2017 , Chowdhury et al 2019 ). Commercially available vehicles exceed the 2020 power-density target. However, the 2025 target is at least a factor of six to eight higher than current commercial baselines. U.S. DRIVE Partnership also proposes electric-traction-drive-cost targets for 2020 and 2025: $8 kW −1 and $6 kW −1 , respectively, both of which are challenging targets (U.S. DRIVE 2017 , Chowdhury et al 2019 ). The authors are not aware of commercial systems meeting the 2020 target, and the 2025 target represents a further 33% reduction.

Figure 4.

Figure 4.  Integrated electric-drive system and inverter power density for several commercial light-duty vehicles and DOE targets (data from U.S. DRIVE 2017 , Chowdhury et al 2019 ).

Improvements in compact power electronics and electric machines are applicable to novel emerging wheel-integrated solutions as well (Iizuka and Akatsu 2017 , Fukuda and Akatsu 2019 ). The development of advanced electric traction drive with improved efficiency is a strategy for increasing the range of electric-drive vehicles. In addition to this, chassis light-weighting is another strategy that is being pursued by the industry and the research community for increasing EV driving ranges. There are several technical challenges in meeting the DOE power-density targets (shown in figure 4 ). Challenges in meeting related DOE cost targets remain as well. A range of integration approaches are proposed in the literature, including surface mounting the power electronics on the motor housing (Nakada, Ishikawa, and Oki 2014 ), mounting on the motor stator iron (Wheeler et al 2005 ), and piecewise integration. Piecewise integration involves modularizing both power modules and machine stators into smaller units (Brown et al 2007 ). In all cases, the close physical positioning of the power electronics relative to the machine and the associated harsh thermal environment necessitate new concepts related to the active cooling of both components. A first strategy may be to isolate the power electronics from the machine thermally using parallel cooling mechanisms (Wheeler et al 2005 ). Another approach may be to use a fully integrated, series-connected, active-cooling loop (Tenconi et al 2008 , Gurpinar et al 2018 ). In either case, cost benefits may be realized through the possible elimination or combination of cooling loops. Significant research also has been focused on reducing rare-earth and heavy-rare-earth materials within the electric machines because that is an additional important pathway to reduce costs (U.S. DRIVE 2017 ).

Higher levels of integration go hand-in-hand with the utilization of wide-bandgap (WBG) semiconductor devices, which may be used at higher operational temperatures (e.g. >200 °C versus 150 °C for silicon) with reduced switching loss (Millán et al 2014 ). However, the adoption of WBG devices requires new packaging technologies to support the end goals of high temperature, high frequency, higher voltages, and more compact footprints. High-performance electrical interconnects (Cheng et al 2013 ), die-attach (Liu et al 2020 ), encapsulation (Cao et al 2010 ), and power-module-substrate technologies (Stockmeier et al 2011 ), along with thermal management and reliability of these technologies (Moreno et al 2014 , Paret et al 2016 , 2019 ), are critical aspects to consider. The new materials, devices, and components must be cost-effective and high-temperature-capable to be compatible with WBG devices. The downsizing of passive electrical components is another added benefit of adopting WBG devices and a further necessity for integrated machine-drive packaging solutions. Fortunately, the higher switching frequencies that are supported by WBG devices enable the downsizing of both the inductors and capacitors found in a traditional power-control unit (Hamada et al 2015 ). The development of economically viable and high-temperature-capable passives, capacitors in particular (Caliari et al 2013 ), is an area of great interest.

Besides EV applications, power electronics and electric machines with low cost, high performance, and high reliability are important for numerous energy-efficiency and renewable-energy applications, such as solar inverters, generators and electric drives for wind, grid-tied medium-voltage power electronics, and sensors and electronics for high-temperature geothermal applications (PowerAmerica 2020 ).

5. Charging infrastructure

Infrastructure planning and deploying an ecosystem of cost-effective and convenient public and private chargers is central to supporting EV adoption (CEM 2020 ). The lack of a sufficient refueling infrastructure has hampered many past efforts to promote alternatives to petroleum fuels (McNutt and Rodgers 2004 ). Extensive research is being done to address the diverse challenges that are posed by a transition from fossil-fuelled ICEVs to EVs and the special role of charging infrastructure in this transition (Muratori et al 2020b ).

At the end of 2019, there were an estimated 7.3 million EV chargers (or plugs) worldwide, of which almost 0.9 million were public, including approximately 264 000 public DCFCs (81% in China) (IEA 2020 ). Significant government support and private investments are helping to expand the network of public charging stations worldwide. With about 7.2 million light-duty BEVs on the road, there is about one public charger per 10 light-duty BEVs, and most vehicles have access to a residential charger. However, the number of public chargers per BEV varies widely among the 10 countries with the most BEVs (figure 5 ) because of different strategies for deploying fast versus slow public chargers. In addition to these LDV chargers, IEA estimated there are 184 000 fast chargers dedicated to electric buses (95% in China).

Figure 5.

Figure 5.  Public charging availability by country in 2019, measured as Level-1 and Level-2 chargers per BEV and DCFC per 10 BEVs (Data from IEA 2020 ).

Studies show consistently that today's EVs do the majority (50%–80%) of their charging at home, followed by at work (15%–25% when workers use their vehicles to commute), and using public chargers (only about 5% of charging) (Hardman et al 2018 ). PHEVs conduct more charging at home than BEVs do, and they rely more on level-1 charging (Tal et al 2019 ). While single-household detached residences readily can accommodate level-1 or -2 charging, multi-unit dwellings require curbside public charging or installations in shared parking facilities (Hall and Lutsey 2017 ). Historical data on the charging behavior of California BEV owners reveals that 11% of their charging sessions were at level 1, 72% were at level 2, and 17% used DCFCs (Tal et al 2019 ). Use of DCFCs is lowest for BEVs with less than 100 miles (161 km) of range, highest for medium-range BEVs, and lower again for BEVs with ranges of 300 miles (483 km) or more.

5.1. Charging-siting modeling

Public charging infrastructure is clearly important to EV purchasers and supports EV sales by adding value (Narassimhan and Johnson 2018 , Greene et al 2020 ). However, how best to deploy charging infrastructure, in terms of numbers, types, locations, and timing remains an active area for research (Ko et al 2017 , Funke et al 2019 provide reviews). The literature includes many examples of geographically and temporally detailed models to optimize the location, number, and types of charging stations (e.g. Wood et al 2017 , Wu and Sioshansi 2017 , Zhao et al 2019 ). Geographically and temporally detailed data recording the movements of PEVs and their charging behavior are scarce. With few exceptions (e.g. Gnann et al 2018 ), simulation analyses rely on conventional ICEV databases (e.g. Dong et al 2014 , Wood et al 2015 , 2018 ), which do not reflect the changes PEV owners will make to maximize the utility of PEVs.

Given the importance of home charging, access to chargers for on-street parking in residential areas comprising attached or multi-unit dwellings is likely to be essential for PEVs to be adopted at large scale. Grote et al ( 2019 ) employ heuristic methods with geographical-information systems to locate curbside chargers in urban areas using a combination of census and parking data. The works of Nie and Ghamami ( 2013 ), Ghamami et al ( 2016 ), and Wang et al ( 2019 ) are examples of the variety of optimization methods that are applied to design DCFC networks to support intercity travel. Despite these examples, applied research is hindered by the scarcity of data on long-distance vehicle travel by PEVs (Eisenmann and Plötz 2019 ). Jochem et al ( 2019 ) estimate that 314 DCFC stations could provide minimum coverage of EU intercity routes with approximately 0.7 charging points per 1000 BEVs. Using a database of simulated U.S. intercity travel, He et al ( 2019 ) employ a mixed-integer model to optimize the location and number of DCFCs. They conclude that 250 stations could serve 98% of the long-distance miles of BEVs with ranges of 150 miles (241 km) or greater but only 73% of the long-distance miles of 100 mile range (161 km range) BEVs. Similarly, Wood et al ( 2017 ) estimate that 400 DCFC stations are required to cover the U.S. interstate-highway network with a 40 mile (64 km) spacing between stations. Others consider the optimal location of dynamic, wireless charging in combination with stationary charging (Liu and Wang 2017 ).

Optimization models for locating chargers to support commercial PEV fleets also appear in the literature (Jung et al 2014 , Shahraki et al 2015 ). In the future, if vehicle sharing becomes much more common, the downtime for charging could be an important disadvantage for PEVs. Using an integer model to optimize station allocation and PEV assignment, Roni et al ( 2019 ) find that charging time represents 72%–75% of vehicle downtime but that charging time could be reduced by almost 50% by optimal deployment of charging stations.

5.2. Beyond LDV charging

The electrification of medium- and heavy-duty commercial trucks and buses introduces unique charging and infrastructure requirements compared to those of LDVs. These requirements stem from the significantly higher battery capacities required on-board the vehicles, potentially shorter charging-dwell times (due to the in-service time requirements of the vehicles), and the potential of large facility charging loads (due to multiple trucks or buses charging in one location). One challenge is to understand the costs associated with the multitude of charging scenarios for commercial vehicles for current operations as well as future operations. It is expected that on-road freight vehicle miles (or km) traveled will increase by 75% from 2012 to 2045 (McCall and Phadke 2019 ). This increase may bring about new business models and potentially new charging-infrastructure approaches to meet this demand with electrified trucks. California's Innovative Clean Transit regulation, which will require California transit agencies to adopt zero-emission buses by 2040, is likely to drive large charging-infrastructure investments for buses (CARB 2018 ).

Today's commercial diesel-powered trucks in small fleets typically are fueled at publicly available on-road fueling stations, while nearly half of trucks in fleets of 10+ vehicles use company-owned facilities (Davis and Boundy 2020 ). Likewise, commercial EVs are charged primarily in fleet-owned facilities as their daily schedule allows (most often overnight). This depot-charging approach, which enables seamless integration of EVs into fleet logistics, might limit the electrification of some vehicle segments in the long term due to the battery capacity that is needed to satisfy their daily-range requirements (the need to complete their full-day function) and return to the facility to recharge fully 14 . Some studies suggest that long-haul battery-electric trucks are technically feasible and economically compelling (Phadke et al 2019 ) while others are more skeptical (Held et al 2018 ). Publicly available, high-power charging or en-route charging infrastructure for commercial vehicles could enable electrification for longer-distance vehicles (by enabling smaller on-board battery-capacity needs), but this scenario has cost challenges. En-route, high-power charging of over 1 MW might be needed to enable 500 miles (805 km) or more of daily driving. Installation of a 20 MW truck-charging station in California (capable of multiple 1.5 MW charge events for heavy-duty freight vehicles) is estimated to cost as much as 15 million USD. McCall and Phadke ( 2019 ) estimate that as many as 750 of these stations are needed to electrify the fleet of California Class-8 combination trucks. Charging commercial vehicles at depots requires additional infrastructure costs to install lower-power EV-supply equipment networks (e.g. 50 kW–100 kW) capable of charging multiple vehicles at these lower rates. These depot charging systems also will challenge existing facility electrical systems by adding a significant load that was not planned previously at the facility (Borlaug et al Forthcoming ).

5.3. Economics of public charging

PEV-charging economics vary with location and station configuration and depend critically on equipment and installation costs and retail electricity prices, which are dependent on utilization (Muratori et al 2019b , Borlaug et al 2020 ). In the early stages of market development, when there are relatively few vehicles, future demand is uncertain, and most charging is done at an EV's home base (Nigro and Frades 2015 , Madina et al 2016 ). Public charging stations tend to be lightly used during these initial stages (e.g. INL 2015 ), which poses a difficult challenge for private investment. Understanding and quantifying the value of public charging is hindered by lack of experience with PEVs on the part of consumers (Ito et al 2013 , Greene et al 2020 , Miele et al 2020 ) and the complexity of network effects in the evolution of alternative-fuel-vehicle markets (Li et al 2017a ). Nevertheless, it is likely that DCFCs will be profitable with sufficient demand. Considering vehicle ranges of between 100 km and 300 km and charging-power levels of between 50 kW and 150 kW, Gnann et al ( 2018 ) conclude that charger-usage fees could be between 0.05 € kWh −1 and 0.15 € kWh −1 in addition to the cost of electricity. The estimates were based on simulations with average daily occupancy of charging points of 10%–25% and peak-hour utilization of 20%–70%. In their simulations, utilization rates increase with increasing charger power and decrease with increased EV range. For intercity travel along European Union highways, Jochem et al ( 2019 ) estimate that a surcharge of 0.05 € kWh −1 of DCFC would make a minimal coverage of 314 stations (with 20 charge plugs each) profitable, even for station capital costs of one million EUR. He et al ( 2019 ) optimize DCFC locations along U.S. intercity routes and conclude that providing an adequate nationwide charging network for long-distance travel by 100 mile (161 km) range BEVs is more economical than increasing vehicle range and reducing the number of charging stations. Muratori et al ( 2019a ) consider a set of charging scenarios from real-world data and thousands of U.S. electricity retail rates. They conclude that batteries can be highly effective at mitigating electricity costs associated with demand charges and low station utilization, thereby reducing overall DCFC costs.

Early estimates show that the cost of public DCFC in U.S. can vary widely based on the station characteristics and level of use (Muratori et al 2019a ). Numerous new technology options are being explored to provide lower-cost electricity for light-duty passenger and medium- and heavy-duty commercial BEVs. Increasing the range of EVs through higher-power public charging stations as well as accommodating new potential BEV business models, such as transportation-network companies or automated vehicles, are driving new charging-technology solutions. Managed charging solutions that are available today can provide increased value to the BEV owner (lower electricity costs), charging station owner (lower operating costs), or grid operator (lower infrastructure-investment costs). For example, a managed-charging solution has been adopted and is currently in operation at a Santa Clara Valley Transportation Authority depot to charge a fleet of Proterra electric buses optimally to ensure minimal stress on the grid (Ross 2018 ).

5.4. Emerging charging technologies

Wireless charging, specifically high-power wireless charging (beyond level-2 power levels), could play a key role in providing an automated charging solution for tomorrow's automated vehicles (Lukic and Pantic 2013 , Qiu et al 2013 , Miller et al 2015 , Feng et al 2020 ). Wireless charging also can enable significant electric range for BEVs by providing en-route opportunity charging (static or dynamic charging opportunities). If a network of wireless charging options is available to provide convenient and fast en-route charging, it could help reduce the amount of battery that is needed on-board a vehicle and reduce the cost of ownership for a BEV owner. Wireless charging is being developed for power levels of up to 300 kW for LDVs, 500 kW for medium-duty vehicles, and 1000 kW for heavy-duty vehicles. Bidirectional functionality, improved efficiency, interoperability of different systems, improved cybersecurity, and increased human-safety factors continue to be developed (Ozpineci et al 2019 ).

Connectivity and communication advances will enable new BEV-charging infrastructure and managed charging solutions. However, emerging cybersecurity threats also are being identified and should be addressed. There are concerns associated with data exchange, communications network, infrastructure, and firmware/software elements of the EV infrastructure (Chaudhry and Bohn 2012 ), and new charging-system security requirements and protocols are being developed to address these concerns (ElaadNL 2017 ). New emulation and simulation platforms also are being developed to address these threats and help understand the consequences and value of mitigating cyberattacks that could affect BEVs, electric-vehicle-supply equipment, or the electric grid (Sanghvi et al 2020 ).

6. Vehicle-grid integration (VGI)

Connecting millions of EVs to the power system, as may occur in the coming decades in major cities, regions, and countries around the world, introduces two fundamental themes: (a) challenges to meet reliably overall energy and power requirements, considering temporal load variations, and (b) VGI opportunities that leverage flexible vehicle charging ('smart charging') or V2G services to provide power-system services from connected vehicles. Multiple studies, which are reviewed in detail below, investigate the potential load growth, impact on load shapes, and infrastructure implications of increased EV adoption. These works focus especially on impact on distribution systems and opportunities for flexible charging to reshape aggregate power loads. Mai et al ( 2018 ), for example, shows that in a high-electrification scenario, transportation might grow from the current 0.2% to 23% of total U.S. electricity demand by 2050. This growth would impact system peak load and related capacity costs significantly if not controlled properly. In-depth analytics indicate a complex decision framework that requires critical understanding of potential future mobility demands and business models (e.g. ride-hailing, vehicle sharing, and mobility as a service), technology evolution, electricity-market and retail-tariff design, infrastructure planning (including charging), and policy and regulatory design (Codani et al 2016 , Eid et al 2016 , Knezovic et al 2017 , Borne et al 2018 , Hoarau and Perez 2019 , Gomes et al 2020 , Muratori and Mai 2020 , Thompson and Perez 2020 ).

While accommodating EV charging at the bulk-power (generation and transmission) level will be different in each region, no major technical challenges or risks have been identified to support a growing EV fleet, especially in the near term (FleetCarma 2019 , U.S. DRIVE 2019 , Doluweera et al 2020 ). At the same time, many studies show that smart charging and V2G create opportunities to reduce system costs and facilitate VRE integration (Sioshansi and Denholm 2010 , Weiller and Sioshansi 2014 , IRENA 2019 , Zhang et al 2019 ). Therefore, charging infrastructure that enables smart charging (e.g. widespread residential and workplace charging) and alignment with VRE generation and business models and programs to compensate EV owners for providing charging flexibility are critical elements for successful integration of EVs with bulk power systems.

6.1. Impact of EV loads on distribution systems

At the local level, EV charging can increase and change electricity loads significantly, having possible negative impacts on distribution networks (e.g. cables and distribution transformers) and power quality or reliability (Khalid et al 2019 ). Residential EV charging represents a significant increase in household electricity consumption that can require upgrades of the household electrical system which, unless managed properly, may exceed the maximum power that can be supported by distribution systems, especially for legacy infrastructure and during times of high electricity utilization (e.g. peak hours and extreme days) (IEA 2018b ). The impact of EVs on distribution systems also is influenced by the simultaneous adoption of other distributed energy resources, e.g. rooftop PV panels. While this interdependency complicates assessing the impact of EV charging, Fachrizal et al ( 2020 ) show that the two technologies support one other. Similarly, Vopava et al ( 2020 ) show that line overloads caused by rooftop PV panels can be reduced (but not avoided) by increasing EV adoption and vice versa.

The impact of EV charging on distribution systems is particularly critical for high-power charging and in cases in which many EVs are concentrated in specific locations, such as clusters of residential LDV charging and possibly fleet depots for commercial vehicles (Saarenpää et al 2013 , Liu et al 2017a , Muratori 2018 ). Smart charging, by which EV charging is timed based on signals from the grid and electricity prices that vary over time, or other forms of control, can help to minimize the impact of EV charging on distribution networks. However, smart charging requires both appropriate business models and signals (with related communication and distributed-control challenges). The market for distribution-system operators to provide such services is not mature yet (Everoze 2018 , Crozier, Morstyn, and McCulloch 2020 ). Time-varying pricing schemes, which are effective at influencing the timing of EV charging (PG&E 2017 ), typically do not include any distribution-level considerations. Thus, while consumers are responsive to such signals, the business models to include distribution-level metrics still are lacking. Moreover, price signals are offered usually to a large consumer base with the intent of reshaping the overall system load. At the local level, however, multiple consumers responding to the same signal might cause 'rebound peaks' (Li et al 2012 , Muratori and Rizzoni 2016 ) that can overstress distribution systems, calling for coordination among consumers connected to the same distribution network (e.g. direct EV-charging control from an intermediate aggregator).

Charging of larger commercial vehicles and highway fast-charging stations typically involves higher power levels: DCFC is typically at 50 kW/plug today, but power levels are increasing rapidly. Commercial charging locations with multiple plugs co-located at a specific location may lead to possible MW-level loads, which is roughly equivalent to the peak load of a large hotel. Commercial DCFC may require costly upgrades to distribution systems that can impact the cost-effectiveness of public fast charging heavily, especially if stations experience low utilization (Garrett and Nelder 2016 , Muratori et al 2019b ). While charging timing and speed at commercial stations is less flexible (consumers want to charge and leave or commercial fleets must meet business requirements), business models are often already in place to incentivize curbing maximum peak power from commercial installations. For example, demand charges (a fixed monthly payment that is proportional to the peak power that is drawn during a given month) are fairly common in U.S. retail tariffs and provide a reason to limit peak power. Furthermore, Muratori et al ( 2019a ) show that distributed batteries can be effective at mitigating the cost associated with demand charges by up to 50%, especially for 'peaky' or low-utilization EV-charging loads. Batteries also can facilitate coupling EV-charging stations with local solar electricity production or can provide grid services (Megel, Mathieu, and Andersson 2015 ), generating additional revenue.

6.2. Value of managed ('smart') EV charging for power systems

The integration of EVs into power systems presents several opportunities for synergistic improvement of the efficiency and economics of electromobility and electric power systems. These synergies stem from two inherent characteristics of EVs and power systems. Demand response and other forms of demand-side flexibility can be of value for power-system planning and operations (Albadi and El-Saadany 2007 , 2008 , Su and Kirschen 2009 , Muratori et al 2014 ). Contemporaneously, most personal-vehicle driving patterns entail vehicle-use for mobility purposes a relatively small proportion of the time (Kempton and Letendre 1997 ). If EVs are grid-connected for extended periods of time, they can provide demand-side flexibility in the form of smart charging or V2G services. Such use of an EV can improve its economics by leveraging cheaper electricity at little incremental cost (e.g. the costs of monitoring, communication, and control equipment that are needed to manage smart charging). EVs can support grid planning and operations in a number of ways. Figure 6 summarizes the key support services that EVs can provide. These services include reducing peak load and generation-, transmission-, and distribution-capacity requirements, deferring system upgrades, providing load response, supporting power-system dispatch (including VRE integration and real-time energy and operating reserves), providing energy arbitrage, and supporting power quality and end retail consumers.

Figure 6.

Figure 6.  Summary of opportunities for EVs to provide demand-side flexibility to support power system planning and operations across multiple timescales.

Habib et al ( 2015 ) and Thompson and Perez ( 2020 ) provide detailed surveys of different potential uses of EVs for smart charging and V2G services. This includes active- and reactive-power services, load balancing, power-quality-related services (e.g. managing flicker and harmonics), retail-bill management, resource adequacy, and network deferral. In addition, Habib et al ( 2015 ) discuss different standards and technology needs relating to V2G services.

Kempton and Letendre ( 1997 ) provide the first description of the concept of EVs providing grid services, either in the form of smart charging or bidirectional V2G services (which can involve discharging EV batteries). Denholm and Short ( 2006 ) study the benefits of controlled overnight charging of PHEVs for valley-filling purposes. They demonstrate that with proper control of vehicle charging, up to 50% of the vehicle fleet could be electrified without needing new generation capacity to be built and at substantial savings compared to using liquid fuels for transportation. They show also that under conservative utility-planning practices, PHEVs could replace a significant portion of low-capacity-factor generating capacity by providing peaking V2G services. Tomić and Kempton ( 2007 ) examine the economics of using EVs for the provision of frequency reserves and demonstrate that such services can yield substantive revenues to vehicle owners in a variety of wholesale markets. Thompson and Perez ( 2020 ) conduct a meta-analysis of V2G services and value streams and find that power-focused services are of greater value than energy-focused services. They distinguish the two types of services based on the extent to which EV batteries must be discharged and degraded. Sioshansi and Denholm ( 2010 ) come to a similar conclusion in comparing the value of using PHEV batteries for energy arbitrage and operating reserves.

Another important synergy between EVs and power systems is using the flexibility of EV charging to manage the integration of VRE into power systems (Mwasilu et al 2014 , Weiller and Sioshansi 2014 ). Hoarau and Perez ( 2018 ) develop a framework for examining the synergies between EV charging and the integration of photovoltaic-solar resources into power systems. They find that the spatial footprint across which solar resources and EVs are deployed and the regulatory, policy, and market barriers to cooperation between solar resources and EVs are critical to realizing these synergies. Szinai et al ( 2020 ) find that controlled EV charging in California under its 2025 renewable-portfolio standards can reduce operational costs and renewable curtailment compared to unmanaged charging. They find that properly designed time-of-use retail tariffs can achieve some, but not all, of the benefits of controlled EV charging. They show also that these two approaches to managing EV charging (controlling EV charging directly and time-of-use tariffs) reduce the cost of infrastructure that is necessary to accommodate EV charging relative to a case of uncontrolled EV charging. Chandrashekar et al ( 2017 ) conduct an analysis of the Texas power system and find similar benefits of controlled EV charging in reducing wind-integration costs. Coignard et al ( 2018 ) show that under California's 2020 renewable-portfolio standards, controlled EV charging can deliver the same renewable-integration benefits that California's energy-storage mandate does but at substantially lower costs. They show that bidirectional V2G services deliver up to triple the value of controlled EV charging. Kempton and Tomić ( 2005 ) show that high penetrations of wind energy in U.S. could be accommodated at relatively low costs if 3% of the vehicle fleet provides frequency reserves and 8%–38% of the fleet provides operating reserves and energy-storage services to avoid wind curtailment. Loisel et al ( 2014 ) and Zhang et al ( 2019 ) conduct more forward-looking analyses of the synergies between EVs and renewables. The former examines German systems, and the latter examines California systems under potential renewable-deployment scenarios in the year 2030.

An important assumption underlying these works is that EV owners (or aggregations of EVs) are exposed to prices that signal the value of these services and that there are regulatory and business models that allow such services to be exploited (i.e. consumer are willing to engage in these programs and are compensated properly for providing flexible charging). Several pilot studies suggest that EV owners have interest in participating in utility-run controlled-charging programs and that a set of different compensation strategies beyond time-varying electricity pricing might maximize engagement (Geske and Schumann 2018 , Hanvey 2019 , Küfeoğlu et al 2019 , Delmonte et al 2020 ).

Niesten and Alkemade ( 2016 ) survey the literature on these topics and numerous European and U.S. pilot programs in terms of value generation for V2G services. They find that the ability of an aggregator to scale is related to its ability to develop a financially viable business model for V2G services. Another important consideration is the availability of control and communication technologies to manage EV charging based on power-system conditions. Key considerations in the design of control strategies are robustness in the face of uncertainty (e.g. renewable availability, EV-arrival times and charge levels upon arrivals, and EV-departure times), data privacy, and robustness to communication or other failures. Le Floch et al ( 2015a ), Le Floch et al ( 2015b ), ( 2016 ), develop a variety of distributed and partial-differential-equation-based algorithms for controlling EV charging. Rotering and Ilic ( 2011 ) develop a dynamic-optimization-based approach to control EV charging and bidirectional V2G services (with a focus on the provision of ancillary services). Donadee and Ilic ( 2014 ) develop a Markov decision process to optimize the offering behavior of EVs that participate in wholesale electricity markets to provide frequency reserves.

6.3. Remaining challenges for effective vehicle-grid integration at scale

There are still many challenges to tackle before smart charging and V2G can be deployed effectively at large scale. These challenges are linked to the technical aspects of VGI technology but also to societal, economic, security, resilience, and regulatory questions (Noel et al 2019a ). With regard to the technical challenges of VGI, existing barriers notably include battery degradation, charger availability and efficiency, communication standards, cybersecurity, and aggregation issues (Eiza and Ni 2017 , Sovacool et al 2017 , Noel et al 2019a ).

While the technical aspects of VGI are studied widely, this is much less the case for its key societal aspects. Societal issues include the environmental performance of VGI, its impact on natural resources, consumer acceptance and awareness, financing and business models, and social justice and equity (Sovacool et al 2018 ). There are also various regulatory and political challenges linked to clarifying the regulatory frameworks applicable to VGI as well as market-design issues, such as the proper valuation of VGI services and double taxation (Noel et al 2019a ) and the trade-offs between bulk power and distribution-system needs. Regulatory changes may be required to enable distribution-network operators and EV owners (or aggregators) to take a more active role in electricity markets. The Parker project, an experimental project on balancing services from an EV fleet, underlines some of the barriers to providing ancillary services, such as metering requirements (Andersen et al 2019 ). It is argued that insufficient regulatory action might keep us from attaining the full economic and environmental benefits of V2G (Thompson and Perez 2020 ) and that regulations are lagging behind technological developments (Freitas Gomes et al 2020 ). The lack of defined business models is seen by many experts as a key impediment (Noel et al 2019b ).

Major challenges that are linked to data-related aspects of VGI, including who has the right to access data from EVs (e.g. the state of EV batteries and charging) and how these data can be exploited, remain. Privacy concerns are one of the major obstacles to user acceptance (as is fear of loss of control over charging) (Bailey and Axsen 2015 ). In addition, there are also questions linked to cybersecurity (Noel et al 2019a ).

Nevertheless, VGI offers many opportunities that justify the efforts required to overcome these challenges. In addition to its services to the power system, VGI offers interesting perspectives for the full exploitation of synergies between EVs and renewable energy sources as both technologies promise large-scale deployment in the future (Kempton and Tomić 2005 , Lund and Kempton 2008 ). Exploiting EV batteries for VGI also is appealing from a life-cycle perspective, as the manufacturing of EV batteries has a non-trivial environmental footprint (Hall and Lutsey 2018 ). However, there are a few future developments that might compromise the potential of VGI, most notably cheaper batteries (including second-life EV batteries) that might compete with EVs for many potential services (Noel et al 2019b ). In addition, the impacts of new mobility business models, such as the rise of vehicle- and ride-sharing, on grid services remain unclear. Although smart charging will come first in the path toward grid integration, V2G services have the potential to provide additional value (Thingvad et al 2016 ).

7. Life-cycle cost and emissions

EVs differ from conventional ICEVs on an emissions basis. While the operation of gasoline- or diesel-powered ICEVs produces GHG and pollutant emissions that are discharged from the vehicle tailpipe, EVs have no tailpipe emissions. In a broader context, EVs still can be associated with so-called 'upstream' emissions from the processes that generate, transmit, and distribute the electricity that is used for their charging. Fueling an ICEV also involves upstream 'fuel-cycle' emissions from the raw-material extraction and transportation, refining, and final-product-delivery processes that make gasoline or diesel fuel available at a retail pump. These fuel-cycle emissions give rise to the colloquial jargon 'well-to-pump' emissions. Accordingly, a 'well-to-wheels' (WTW) life-cycle analysis (LCA) is an appropriate framework for comparing EV and ICEV emissions. WTW considers both upstream emissions from the fuel cycle ('well-to-pump') and direct emissions from vehicle operation ('pump-to-wheels') for a standardized functional unit and temporal period. WTW studies have a history of over three decades of use to evaluate direct and indirect emissions related to fuel production and vehicle operations (Wang 1996 ). WTW emissions are expressed typically on a per-mile or per-kilometer basis over a vehicle's assumed lifetime.

WTW analyses typically focus only on fuel production and vehicle operation. Some studies consider broader system boundaries that include vehicle production and decommissioning (i.e. recycling and scrappage) in an LCA framework. This broader system boundary considers what is commonly called the 'vehicle cycle' and provides a so-called 'cradle-to-grave' (or 'C2G') analysis. Vehicle-cycle emissions typically account for 5%–20% of today's ICEV C2G emissions and can be as low as 15% or as high as 80% of today's BEV emissions, depending on the underlying electricity-generation mix. Lower-carbon mixes result in vehicle-cycle emissions accounting for a greater portion of total emissions. As an extreme illustrative example, the case of zero-carbon electricity implies that vehicle-cycle emissions account for 100% of C2G emissions. In general, BEV vehicle-cycle emissions are 25% to 100% higher than their ICEV counterpart (Samaras and Meisterling 2008 , Ambrose and Kendall 2016 , Elgowainy et al 2016 , Hall and Lutsey 2018 , Ricardo 2020 ). As this section explores, higher initial BEV vehicle-cycle emissions almost always are counterbalanced by lower emissions during vehicle operation (with notable exceptions in cases in which BEVs are charged from especially high-emissions electricity).

Even including upstream emissions, EVs are championed as a critical technology for decarbonizing transportation (in line with anticipated widespread grid decarbonization). National Research Council ( 2013 ) identifies EVs as one of several technologies that could put U.S. on a path to reducing transportation-sector GHG emissions to 80% below 2005 levels in 2050. Furthermore, National Research Council ( 2013 ) estimates that BEVs would reduce emissions by 53%–72% compared to ICEVs in 2030. IEA ( 2019 ) contends, similarly, that EVs can reduce WTW GHG emissions by half versus equivalent ICEVs in 2030. Recently published literature also agrees, even on a C2G basis, estimating that future EV pathways offer 70%–90% lower GHG emissions compared to today's ICEVs (Elgowainy et al 2018 ). As such, the broad view across national, international, and academic-research perspectives is that EVs offer the potential to reduce transportation-related GHG emissions by 53% to 90% in the future.

Several studies find that EVs already reduce WTW GHG emissions today by as little as 10% or as much as 41% on average versus comparable ICEVs based on current electricity-production mixes. Samaras and Meisterling ( 2008 ), who are among the first to relate a range of potential electricity carbon intensities to associated EV-lifecycle emissions explicitly, estimate a 38%–41% GHG emissions benefit for EVs powered by the average 2008 U.S. grid. Hawkins et al ( 2012a ), informed by a meta-study of 51 previous LCAs, highlight great variations based on different electricity generation assumptions and vehicle lifetime. Hawkins et al ( 2012b ) estimate a decline of 10%–24% global warming potential (a measure proportional to GHG emissions) for EVs powered by the average 2012 European electricity mix. Elgowainy et al 2016 , 2018 ) estimate that EVs emit 20%–35% fewer GHG emissions when operating on the average 2014 U.S. grid mix.

Many factors contribute to variability in EV WTW emissions and estimated reduction opportunities compared to ICEVs—electricity-carbon intensity, charging patterns, vehicle characteristics, and even local climate (Noshadravan et al 2015 , Requia et al 2018 ). To illustrate these variabilities, figure 7 compares WTW GHG emissions of EVs versus comparable ICEVs. Relative emissions reductions are generally larger for larger vehicles. Woo et al ( 2017 ) find that electrifying SUVs reduces emissions more than electrifying sub-compact vehicles on a WTW basis versus comparable ICEVs (30%–45% and 10%–20%, respectively, assuming median national grid mixes). Ellingsen et al ( 2016 ) find that large EVs emit proportionally less than small EVs compared to comparable ICEVs on a C2G basis (27% and 19%, respectively).

Figure 7.

Figure 7.  WTW GHG emissions for EVs versus comparable ICEVs on average and with illustrative variability by market segment, electricity generation pathway, grid mix, and ambient temperature.

Low-carbon electricity can lead to greater reductions in EV emissions. Electricity that is produced from coal, which has a high carbon intensity, can increase EV emissions by as much as 40% or decrease EV emissions by as much as 5% compared to an ICEV (depending on other assumptions). Conversely, electricity from hydropower, nuclear, solar, or wind, all of which offer near-zero carbon intensities, can decrease EV emissions by more than 95% compared to an ICEV (Woo et al 2017 ). Such variability in electricity-generation pathways affects the relative benefits of real-world grid mixes. For example, while EVs offer 30%–65% lower emissions versus comparable ICEVs on average in Europe (Woo et al 2017 , Moro and Lonza 2018 ), in individual countries relative emissions can range from as much as 95% lower to 60% higher (Orsi et al 2016 , Moro and Lonza 2018 ). Typically, U.S. EVs provide emissions reductions, but in some regions EV emissions are higher compared to an efficient ICEV (Reichmuth 2020 ). Changes in regional climate and daily weather add further variability: EV emissions can vary between 40% and 50% lower than a comparable ICEV even when charged from the same grid mix (Yuksel et al 2016 ). While outside the scope of a typical WTW comparison, the additional consideration of refueling infrastructure (i.e. gasoline stations for ICEVs and recharging equipment for EVs) is estimated to increase EV emissions by 4%–8% compared to a more modest 0.3%–0.7% increase for ICEV emissions (Lucas et al 2012 ).

When assessing EV emissions, average or marginal grid-emission factors are considered (Anair and Mahmassani 2012 , Traut et al 2013 , EPRI 2015 , Nealer and Hendrickson 2015 , Nealer et al 2015 , Elgowainy et al 2018 ), leading to significantly different results. Average emissions factors consider all electricity loads as equivalent, while marginal emission factors consider EVs as an additional load on top of existing electricity demands and estimate the associated incremental generation emissions. Marginal emissions could be higher or lower than average, depending on the relative emissions of marginal plants compared to the average in different regions. Different questions lead to using average or marginal metrics. Proper assessment of indirect EV emissions associated with electricity generation is complicated by numerous factors, including timescale (short or long term, aggregate or temporally explicit), system boundaries, impact of EV loads on power-system-expansion and -operation decisions, and non-trivial supply-demand synergies and allocation complexities. Yang ( 2013 ) reviews different grid-emissions-allocation methods concluding that there is no ideal approach to the allocation of emissions to specific end-use and stressing how different assumptions make it difficult to determine EV emissions and compare them to other alternatives and across studies. Nealer and Hendrickson ( 2015 ) discuss whether it is more appropriate to use marginal or average grid-emission factors to estimate EV emissions, concluding that 'average emissions may be the most accessible for long-term comparisons given the assumptions that must be made about the future of the electricity grid.'

Just as EVs offer typically a WTW-emissions reduction compared to ICEVs while shifting those emissions from the tailpipe to upstream, EVs shift costs as well. Operational (fuel and maintenance) costs of EVs are typically lower than those of ICEVs, largely because EVs are more efficient than ICEVs and have fewer moving parts. While data are still scarce, a recent Consumer Reports study estimates that maintenance and repair costs for EVs are about half over the life of the vehicle and that a typical EV owner who does most fueling at home can expect to save an average of $800 to $1000 a year on fuel costs over an equivalent ICEV (Harto 2020 ). Insideevs ( 2018 ) estimates a saving of 23% in servicing costs over the first 3 years and 60 000 miles (96 561 km). Borlaug et al ( 2020 ) estimate fuel savings between $3000 and $10 500 compared with gasoline vehicles (over a 15 year time horizon). However, vehicle capital costs for EVs are higher (principally due to the relatively high cost of EV batteries). In general, studies use a TCO metric to combine and compare initial capital costs with operational costs over a vehicle's lifetime. While some studies find that EVs are typically cost-competitive with ICEVs (Weldon et al 2018 ), others find that EVs are still more costly, even on a TCO basis (Breetz and Salon 2018 , Elgowainy et al 2018 ), or that the relative cost depends on other contextual factors, such as vehicle lifetime and use, economic assumptions, and projected fuel prices. Longer travel distance and smaller vehicle sizes favor relatively lower EV TCO (Wu et al 2015 ), as do lower relative electricity-versus-gasoline price differentials (Lévay et al 2017 ). Despite these differences regarding TCO conclusions across studies, there is general agreement that future EV costs will decline (Dumortier et al 2015 , Wu et al 2015 , Elgowainy et al 2018 ).

The existing literature suggests future EV emissions will decline, in large part due to expectations for continued grid decarbonization (Elgowainy et al 2016 , 2018 , Woo et al 2017 , Cox et al 2018 ). For example, Ambrose et al ( 2020 ) anticipate that evolution in vehicle types and designs could accelerate future decreases for EV GHG emissions. Several studies also posit repurposing used EV batteries for stationary applications could accrue additional GHG benefits (Ahmadi et al 2014 , 2017 , Olsson et al 2018 , Kamath et al 2020 ). Cox et al ( 2018 ) suggest future connectivity and automation technologies will enable energy-optimized EV-recharging behavior and associated lower carbon emissions. Similarly, future EV costs also are expected to decline as battery costs continue to decline (cf section 4 ), and new mobility modes such as ride-hailing lead to higher vehicle use that favors the business case for highly efficient EVs compared to ICEVs.

8. Synergies with other technologies, macro trends, and future expectations

Vehicle electrification fits within broader electrification trends, including power-system decarbonization and mobility changes. The latter include micro-mobility in urban areas, new mobility business models revolving around 'shared' services as opposed to vehicle ownership (e.g. ride-hailing and car-sharing), ride pooling, and automation. These trends are driven partially by the larger availability of efficient and cost-effective electrified technologies (Mai et al 2018 ) and the prospect of abundant and affordable renewable electricity and by other technological and behavioral changes (e.g. real-time communication). Abundant and affordable renewable electricity is a conditio sine qua non for EVs to provide a pathway to decarbonize road transportation. Direct use of PV on-board vehicles (i.e. PV-powered or solar vehicles) also is being considered. However, this concept still faces many challenges (Rizzo 2010 , Aghaei et al 2020 ). Yamaguchi et al ( 2020 ) show potential synergies for integration but also highlight that for this technology to be successful, the development of high-efficiency (>30%), low‐cost, and flexible PV modules is essential.

Urban micro-mobility is emerging recently as an alternative to traditional mobility modes providing consumers in most metropolitan areas worldwide with convenient options for last-mile transportation (Clewlow 2019 , Zarif et al 2019 , Tuncer and Brown 2020 ). Virtually all micro-mobility solutions use all-electric powertrains. Shared electric scooters and bikes (often dockless), e.g. those pioneered by Lime and Bird in the U.S., are experiencing rapid success and are 'the fastest-ever U.S. companies to reach billion-USD valuations, with each achieving this milestone within a year of inception' (Ajao 2019 ). Future expectations for micro-mobility remain uncertain due to issues related to sidewalk congestion, safety, and vandalism (heavily impacting the business case for these technologies). However, the nexus with EVs has not been questioned.

Similarly, ride-hailing—matching drivers with passengers at short notice for one-off rides through a smartphone application, which date back to Uber's introducing the concept in 2009—is an attractive alternative to traditional transportation solutions. These mobility-as-a-service solutions cater to the consumer's need for quick, convenient, and cost-effective transportation and may lead to drops in car-ownership and driver-licensure rates (Garikapati et al 2016 , Clewlow and Mishra 2017 , Movmi 2018 , Walmsley 2018 , Henao and Marshall 2019 , Arevalo 2020 ). After just over 10 years, ride-hailing is widely available and extremely successful, with hundreds of millions of consumers worldwide and 36% of U.S. consumers having used ride-hailing services (Mazareanu 2019 ). While most ride-hailing vehicles today are ICEVs (in line with the existing LDV stock), many ride-hailing companies are exploring electrification opportunities (Slowik et al 2019 ). EVs offer a number of potential advantages as high vehicle usage promotes a more favorable business model for recovering the higher EV purchase price by leveraging cheaper fuel costs (Borlaug et al 2020 ). At the same time, long-range vehicles and effective charging solutions are required for ride-hailing companies to transition to EVs (Tu et al 2019 ). Moreover, EVs can mitigate additional fuel use and emissions related to increased travel, mostly due to deadheading, which is estimated to be ∼85% (Henao and Marshall 2019 ). EVs also provide access to restricted areas in some cities (driving some regional goals for ride-hailing electrification). For example, Uber aims for half of its London fleet to be electric by 2021 and 100% electric by 2025 (Slowik et al 2019 ).

Automation trends are also poised to have the potential to disrupt transportation as we know it. The combination of electric and connected automated vehicles (CAVs) is hypothesized to offer natural synergies, including easier integration with CAV sensors and a greater affinity for cheaper fuels aligning with greater travel (Sperling 2018 ). The chief counterargument relates to high power requirements for a heavily instrumented CAV, which would deplete EV batteries quickly and may be accommodated better with PHEV powertrains. Wireless EV charging, both stationary and dynamic, increases the potential synergies enabling autonomous recharging. Also, CAVs may be required to maximize the efficiency of dynamic wireless charging. In fact, without the alignment accuracy enabled by CAVs, in-road dynamic charging may have limited efficacy. The literature on these synergies is relatively sparse, though some studies are beginning to investigate the implications of combining EV and CAV technologies.

Even though the technology is not widely available commercially, several studies are beginning to examine how consumer preferences may be influenced by the combination of connected, automated, and electric vehicles 15 . Thiel et al ( 2020 ), for example, highlight how full EV success may emerge as automated shared vehicles become predominant in a world where the border between public and private transport will cease to exist. Tsouros and Polydoropoulou ( 2020 ) develop a survey combining traditional attributes (e.g. car type and vehicle style) alongside future technology attributes (e.g. fuel type and degree of automation) and estimate preferences using a latent-class structural-regression approach. They find a specific class of consumers, described as technology-savvy, who have a high proclivity for both alternative-fuel vehicle technologies and higher degrees of automation. While the proportion of the population that can be classified as technology-savvy is unclear, Tsouros and Polydoropoulou ( 2020 ) provide early compelling evidence that consumers see explicit value in the combination of EVs and automation. Hardman et al ( 2019 ) provide a complementary perspective of early adopters of automated vehicles based on a survey of existing U.S. EV owners. Similar to the work of Tsouros and Polydoropoulou ( 2020 ), Hardman et al ( 2019 ) find that the type of consumers who would pursue automated vehicles have similar lifestyles, attitudes, and socio-demographic profiles as EV adopters. These include high-income consumers, with high levels of knowledge about technology features, who have positive perceptions of CAV attributes and technology in general, provided that safety concerns are resolved.

Another benefit of the combined technologies is the potential to integrate charging events better with the needs of the electricity grid. Several studies assess the combination of these technologies with new mobility services such as car-sharing systems to optimize VGI. Iacobucci et al ( 2018 ) consider a case study in Tokyo of the ability of connected, automated EVs to be dispatched to respond to both transportation demand and charging to meet demands and constraints of the electricity system. The authors observe the vehicles can take advantage of a variety of different time-of-day pricing structures—leading to a tradeoff between wait times and cost benefits from lower fuel prices. They find that the vehicles in Tokyo can supply on the order of 3.5 MW of charging flexibility per 1000 vehicles, even during times of high mobility demand. Miao et al ( 2019 ) conduct a similar study in a generic region. The authors develop an algorithm that simulates operational behavior of the connected, automated EV technology that includes trip demand and vehicle usage, vehicle relocation, and vehicle charging. Their results indicate that charging behavior is highly sensitive to different levels of charging due to the length of charging—which can affect service provision of trip demand.

The final topic of study considering synergistic opportunities between connected, automated vehicles and EVs focused on emissions benefits. Taiebat et al 2018 explore the environmental impacts of automated vehicles showing net positive environmental impacts at the local vehicle-urban levels due to improved efficiency, but acknowledge that greater vehicle utilization and shifts in travel patterns might to offset some of these benefits. Of course, EVs provide the significant benefit of eliminating tailpipe criteria-pollutant emissions, yielding significant human-health benefits. Regarding GHG emissions, two of the earliest studies on this topic examine the net effect of automation on reducing transport GHG emissions (Brown et al 2014 , Wadud et al 2016 ). Greenblatt and Saxena ( 2015 ) conduct a case-study application of connected and automated vehicles in taxi fleets and find large emissions benefits associated with electrification. They find a decrease of GHG emissions intensities ranging from 87% to 94% below comparable ICEVs in 2014 and 63% to 82% below hybrid electric vehicles (HEVs) in 2030. The total emissions benefit is augmented relative to privately owned vehicles due to the higher travel intensity of taxi vehicles. Following these earlier works, additional case studies examine the hypothetical application of automated and electric fleets. These include two studies in Austin, Texas. Loeb and Kockelman ( 2019 ) examine a variety of scenarios to simulate the operation of different vehicle fleets replacing current-day transportation network companies and taxis. The primary goal of their work is to estimate costs associated with operation. They find that automated EVs are the most profitable and provide the best service among the vehicle-technology options that they examine. Gawron et al ( 2019 ) also perform a case study in Austin, Texas, but focus on the emissions benefits of electrifying an automated taxi fleet. They find that nearly 60% of emissions and energy in a base case CAV fleet can be reduced by electrifying powertrains. These improvements can be pushed up to 87% when coupled with grid decarbonization, dynamic ride-sharing, and various system- and technology-efficiency improvements. These results are consistent with a more generalized study by Stogios et al ( 2019 ), who, in a similar approach simulating fleet behavior, find that emissions from CAVs are most dramatically improved via electrification.

While EVs are a relatively new technology and automated vehicles are not widely available commercially, the implications and potential synergies of electrification and automation operating in conjunction are significant. The studies mentioned in this section are investigating a broad set of impacts when CAVs are coupled with EVs. Future research is necessary to generalize and refine many of these results. However, the potential for transformative changes to transportation emissions is clear.

8.1. Expectations for the future

EVs hold great promise to replace ICEVs for a number of on-road applications. EVs can provide a number of benefits, including addressing reliance on petroleum, improving local air quality, reducing GHG emissions, and improving driving experience. Vehicle electrification aligns with broader electrification and decarbonization trends and integrates synergistically with mobility changes, including urban micro-mobility, automation, and mobility-as-a-service solutions. The effective integration of EVs into power systems presents numerous opportunities for synergistic improvement of the efficiency and economics of electromobility and electric power systems, with EVs capable of supporting power-system planning and operations in several ways. Full exploitation of the synergies between EVs and VRE sources offers a path toward affordable and clean energy and mobility for all, as both technologies promise large-scale deployment in the future. To enable such a future continued technology progress, investments in charging infrastructure (and related building codes), consumer education, effective and secure VGI programs, and regulatory and business models supporting all aspects of vehicle electrification are all critical elements.

The coronavirus pandemic is impacting LDV sales in most countries negatively, and 2020 EV sales are expected to be lower than 2019, marking the first decline in a decade (BloombergNEF 2020 ). However, sales of ICEVs are set to drop even faster and, despite the crisis, EV sales could reach a record share of the overall LDV market in 2020 (Gul et al 2020 ). Despite these short-term setbacks, long-term prospects for EVs remain undiminished (BloombergNEF 2020 ).

Several studies project major roles for EVs in the future, which is reflected in massive investment in vehicle development and commercialization, charging infrastructure, and further technology improvement, especially in batteries and their supply chains. Consumer adoption and acceptance and technology progress form a virtuous self-reinforcing circle of technology-component improvements and cost reductions that can enable widespread adoption. Forecasting the future, including technology adoption, remains a daunting task. Nevertheless, this detailed review paints a positive picture for the future of EVs for on-road transport. The authors remain hopeful that technology, regulatory, societal, behavioral, and business-model barriers can be addressed over time to support a transition toward cleaner, more efficient, and affordable mobility solutions for all.

Acknowledgments

The authors thank Paul Denholm, Elaine Hale, Trieu Mai, Caitlin, Murphy, Bryan Palmintier, and Dan Steinberg for valuable comments on figure 6 , as well as two anonymous reviewers for helpful comments on the paper. This work was co-authored by National Renewable Energy Laboratory (NREL), which is operated by Alliance for Sustainable Energy, LLC, for U.S. Department of Energy (DOE) under Contract No. DE-AC36-08GO28308. No funding was received to support this work. The views expressed in this article do not necessarily represent the views of DOE or the U.S. Government. The findings and conclusions in this publication are those of the authors alone and should not be construed to represent any official U.S. Government determination or policy, or the views of any of the institutions associated with this study's authors.

 EVs are defined as vehicles that are powered with an on-board battery that can be charged from an external source of electricity. This definition includes plug-in hybrid electric vehicles (PHEVs) and battery electric vehicles (BEVs). EVs often are referred to as plug-in electric vehicles (PEVs).

 Transport electrification is confined not only to electric LDVs. Transport electrification includes a wide range of other vehicles, spanning from small vehicles that are used for urban mobility, such as three-wheelers, mopeds, kick-scooters, and e-bikes, to large urban buses and delivery vehicles. In 2019, the number of electric two-wheelers on the road exceeded 300 million and buses approached 0.6 million (IEA 2019 , Business Wire 2020 ), with new deliveries in 2019 close to 100 thousand units (EV Volumes 2020 ).

 Just over 10% of the U.S. heavy-duty truck (Class 7–8) population requires an operating range of 500 miles (805 km) or more, while nearly 80% operate within a 200 mile (322 km) range and around 70% within 100 miles (161 km). Only ∼25% of heavy truck VMT require an operating range of over 500 miles (805 km) (Borlaug et al Forthcoming ).

 As a counterargument, Tesla states that 'all new Tesla cars come standard with advanced hardware capable of providing Autopilot features today, and full self-driving capabilities in the future—through software updates designed to improve functionality over time'.

  • Search Menu
  • Browse content in A - General Economics and Teaching
  • Browse content in A1 - General Economics
  • A10 - General
  • A11 - Role of Economics; Role of Economists; Market for Economists
  • A12 - Relation of Economics to Other Disciplines
  • A13 - Relation of Economics to Social Values
  • A14 - Sociology of Economics
  • Browse content in A2 - Economic Education and Teaching of Economics
  • A20 - General
  • A29 - Other
  • A3 - Collective Works
  • Browse content in B - History of Economic Thought, Methodology, and Heterodox Approaches
  • Browse content in B0 - General
  • B00 - General
  • Browse content in B1 - History of Economic Thought through 1925
  • B10 - General
  • B11 - Preclassical (Ancient, Medieval, Mercantilist, Physiocratic)
  • B12 - Classical (includes Adam Smith)
  • B13 - Neoclassical through 1925 (Austrian, Marshallian, Walrasian, Stockholm School)
  • B14 - Socialist; Marxist
  • B15 - Historical; Institutional; Evolutionary
  • B16 - History of Economic Thought: Quantitative and Mathematical
  • B17 - International Trade and Finance
  • B19 - Other
  • Browse content in B2 - History of Economic Thought since 1925
  • B20 - General
  • B21 - Microeconomics
  • B22 - Macroeconomics
  • B23 - Econometrics; Quantitative and Mathematical Studies
  • B24 - Socialist; Marxist; Sraffian
  • B25 - Historical; Institutional; Evolutionary; Austrian
  • B26 - Financial Economics
  • B27 - International Trade and Finance
  • B29 - Other
  • Browse content in B3 - History of Economic Thought: Individuals
  • B30 - General
  • B31 - Individuals
  • Browse content in B4 - Economic Methodology
  • B40 - General
  • B41 - Economic Methodology
  • B49 - Other
  • Browse content in B5 - Current Heterodox Approaches
  • B50 - General
  • B51 - Socialist; Marxian; Sraffian
  • B52 - Institutional; Evolutionary
  • B53 - Austrian
  • B54 - Feminist Economics
  • B55 - Social Economics
  • B59 - Other
  • Browse content in C - Mathematical and Quantitative Methods
  • Browse content in C0 - General
  • C00 - General
  • C02 - Mathematical Methods
  • Browse content in C1 - Econometric and Statistical Methods and Methodology: General
  • C10 - General
  • C12 - Hypothesis Testing: General
  • C13 - Estimation: General
  • C14 - Semiparametric and Nonparametric Methods: General
  • C18 - Methodological Issues: General
  • C19 - Other
  • Browse content in C2 - Single Equation Models; Single Variables
  • C20 - General
  • C21 - Cross-Sectional Models; Spatial Models; Treatment Effect Models; Quantile Regressions
  • C22 - Time-Series Models; Dynamic Quantile Regressions; Dynamic Treatment Effect Models; Diffusion Processes
  • C23 - Panel Data Models; Spatio-temporal Models
  • C25 - Discrete Regression and Qualitative Choice Models; Discrete Regressors; Proportions; Probabilities
  • Browse content in C3 - Multiple or Simultaneous Equation Models; Multiple Variables
  • C30 - General
  • C32 - Time-Series Models; Dynamic Quantile Regressions; Dynamic Treatment Effect Models; Diffusion Processes; State Space Models
  • C34 - Truncated and Censored Models; Switching Regression Models
  • C38 - Classification Methods; Cluster Analysis; Principal Components; Factor Models
  • Browse content in C4 - Econometric and Statistical Methods: Special Topics
  • C43 - Index Numbers and Aggregation
  • C44 - Operations Research; Statistical Decision Theory
  • Browse content in C5 - Econometric Modeling
  • C50 - General
  • Browse content in C6 - Mathematical Methods; Programming Models; Mathematical and Simulation Modeling
  • C60 - General
  • C61 - Optimization Techniques; Programming Models; Dynamic Analysis
  • C62 - Existence and Stability Conditions of Equilibrium
  • C63 - Computational Techniques; Simulation Modeling
  • C65 - Miscellaneous Mathematical Tools
  • C67 - Input-Output Models
  • Browse content in C8 - Data Collection and Data Estimation Methodology; Computer Programs
  • C82 - Methodology for Collecting, Estimating, and Organizing Macroeconomic Data; Data Access
  • C89 - Other
  • Browse content in C9 - Design of Experiments
  • C90 - General
  • C91 - Laboratory, Individual Behavior
  • C92 - Laboratory, Group Behavior
  • C93 - Field Experiments
  • Browse content in D - Microeconomics
  • Browse content in D0 - General
  • D01 - Microeconomic Behavior: Underlying Principles
  • D02 - Institutions: Design, Formation, Operations, and Impact
  • D03 - Behavioral Microeconomics: Underlying Principles
  • Browse content in D1 - Household Behavior and Family Economics
  • D10 - General
  • D11 - Consumer Economics: Theory
  • D12 - Consumer Economics: Empirical Analysis
  • D13 - Household Production and Intrahousehold Allocation
  • D14 - Household Saving; Personal Finance
  • Browse content in D2 - Production and Organizations
  • D20 - General
  • D21 - Firm Behavior: Theory
  • D22 - Firm Behavior: Empirical Analysis
  • D23 - Organizational Behavior; Transaction Costs; Property Rights
  • D24 - Production; Cost; Capital; Capital, Total Factor, and Multifactor Productivity; Capacity
  • D25 - Intertemporal Firm Choice: Investment, Capacity, and Financing
  • Browse content in D3 - Distribution
  • D30 - General
  • D31 - Personal Income, Wealth, and Their Distributions
  • D33 - Factor Income Distribution
  • D39 - Other
  • Browse content in D4 - Market Structure, Pricing, and Design
  • D40 - General
  • D41 - Perfect Competition
  • D42 - Monopoly
  • D43 - Oligopoly and Other Forms of Market Imperfection
  • D46 - Value Theory
  • Browse content in D5 - General Equilibrium and Disequilibrium
  • D50 - General
  • D51 - Exchange and Production Economies
  • D57 - Input-Output Tables and Analysis
  • D58 - Computable and Other Applied General Equilibrium Models
  • Browse content in D6 - Welfare Economics
  • D60 - General
  • D61 - Allocative Efficiency; Cost-Benefit Analysis
  • D62 - Externalities
  • D63 - Equity, Justice, Inequality, and Other Normative Criteria and Measurement
  • D64 - Altruism; Philanthropy
  • D69 - Other
  • Browse content in D7 - Analysis of Collective Decision-Making
  • D71 - Social Choice; Clubs; Committees; Associations
  • D72 - Political Processes: Rent-seeking, Lobbying, Elections, Legislatures, and Voting Behavior
  • D73 - Bureaucracy; Administrative Processes in Public Organizations; Corruption
  • D74 - Conflict; Conflict Resolution; Alliances; Revolutions
  • Browse content in D8 - Information, Knowledge, and Uncertainty
  • D80 - General
  • D81 - Criteria for Decision-Making under Risk and Uncertainty
  • D82 - Asymmetric and Private Information; Mechanism Design
  • D83 - Search; Learning; Information and Knowledge; Communication; Belief; Unawareness
  • D84 - Expectations; Speculations
  • D85 - Network Formation and Analysis: Theory
  • D86 - Economics of Contract: Theory
  • D87 - Neuroeconomics
  • Browse content in D9 - Micro-Based Behavioral Economics
  • D91 - Role and Effects of Psychological, Emotional, Social, and Cognitive Factors on Decision Making
  • Browse content in E - Macroeconomics and Monetary Economics
  • Browse content in E0 - General
  • E00 - General
  • E01 - Measurement and Data on National Income and Product Accounts and Wealth; Environmental Accounts
  • E02 - Institutions and the Macroeconomy
  • Browse content in E1 - General Aggregative Models
  • E10 - General
  • E11 - Marxian; Sraffian; Kaleckian
  • E12 - Keynes; Keynesian; Post-Keynesian
  • E13 - Neoclassical
  • E16 - Social Accounting Matrix
  • E17 - Forecasting and Simulation: Models and Applications
  • Browse content in E2 - Consumption, Saving, Production, Investment, Labor Markets, and Informal Economy
  • E20 - General
  • E21 - Consumption; Saving; Wealth
  • E22 - Investment; Capital; Intangible Capital; Capacity
  • E23 - Production
  • E24 - Employment; Unemployment; Wages; Intergenerational Income Distribution; Aggregate Human Capital; Aggregate Labor Productivity
  • E25 - Aggregate Factor Income Distribution
  • E26 - Informal Economy; Underground Economy
  • E27 - Forecasting and Simulation: Models and Applications
  • Browse content in E3 - Prices, Business Fluctuations, and Cycles
  • E30 - General
  • E31 - Price Level; Inflation; Deflation
  • E32 - Business Fluctuations; Cycles
  • E37 - Forecasting and Simulation: Models and Applications
  • Browse content in E4 - Money and Interest Rates
  • E40 - General
  • E41 - Demand for Money
  • E42 - Monetary Systems; Standards; Regimes; Government and the Monetary System; Payment Systems
  • E43 - Interest Rates: Determination, Term Structure, and Effects
  • E44 - Financial Markets and the Macroeconomy
  • E49 - Other
  • Browse content in E5 - Monetary Policy, Central Banking, and the Supply of Money and Credit
  • E50 - General
  • E51 - Money Supply; Credit; Money Multipliers
  • E52 - Monetary Policy
  • E58 - Central Banks and Their Policies
  • Browse content in E6 - Macroeconomic Policy, Macroeconomic Aspects of Public Finance, and General Outlook
  • E60 - General
  • E61 - Policy Objectives; Policy Designs and Consistency; Policy Coordination
  • E62 - Fiscal Policy
  • E63 - Comparative or Joint Analysis of Fiscal and Monetary Policy; Stabilization; Treasury Policy
  • E64 - Incomes Policy; Price Policy
  • E65 - Studies of Particular Policy Episodes
  • Browse content in F - International Economics
  • Browse content in F0 - General
  • F00 - General
  • F01 - Global Outlook
  • F02 - International Economic Order and Integration
  • Browse content in F1 - Trade
  • F10 - General
  • F11 - Neoclassical Models of Trade
  • F12 - Models of Trade with Imperfect Competition and Scale Economies; Fragmentation
  • F13 - Trade Policy; International Trade Organizations
  • F14 - Empirical Studies of Trade
  • F15 - Economic Integration
  • F16 - Trade and Labor Market Interactions
  • F17 - Trade Forecasting and Simulation
  • F18 - Trade and Environment
  • Browse content in F2 - International Factor Movements and International Business
  • F20 - General
  • F21 - International Investment; Long-Term Capital Movements
  • F22 - International Migration
  • F23 - Multinational Firms; International Business
  • Browse content in F3 - International Finance
  • F30 - General
  • F31 - Foreign Exchange
  • F32 - Current Account Adjustment; Short-Term Capital Movements
  • F33 - International Monetary Arrangements and Institutions
  • F34 - International Lending and Debt Problems
  • F35 - Foreign Aid
  • F36 - Financial Aspects of Economic Integration
  • F37 - International Finance Forecasting and Simulation: Models and Applications
  • F39 - Other
  • Browse content in F4 - Macroeconomic Aspects of International Trade and Finance
  • F40 - General
  • F41 - Open Economy Macroeconomics
  • F42 - International Policy Coordination and Transmission
  • F43 - Economic Growth of Open Economies
  • F44 - International Business Cycles
  • F45 - Macroeconomic Issues of Monetary Unions
  • F47 - Forecasting and Simulation: Models and Applications
  • Browse content in F5 - International Relations, National Security, and International Political Economy
  • F50 - General
  • F51 - International Conflicts; Negotiations; Sanctions
  • F53 - International Agreements and Observance; International Organizations
  • F54 - Colonialism; Imperialism; Postcolonialism
  • F55 - International Institutional Arrangements
  • F59 - Other
  • Browse content in F6 - Economic Impacts of Globalization
  • F60 - General
  • F61 - Microeconomic Impacts
  • F62 - Macroeconomic Impacts
  • F63 - Economic Development
  • F64 - Environment
  • F65 - Finance
  • Browse content in G - Financial Economics
  • Browse content in G0 - General
  • G00 - General
  • G01 - Financial Crises
  • Browse content in G1 - General Financial Markets
  • G10 - General
  • G11 - Portfolio Choice; Investment Decisions
  • G12 - Asset Pricing; Trading volume; Bond Interest Rates
  • G13 - Contingent Pricing; Futures Pricing
  • G14 - Information and Market Efficiency; Event Studies; Insider Trading
  • G15 - International Financial Markets
  • G18 - Government Policy and Regulation
  • G19 - Other
  • Browse content in G2 - Financial Institutions and Services
  • G20 - General
  • G21 - Banks; Depository Institutions; Micro Finance Institutions; Mortgages
  • G22 - Insurance; Insurance Companies; Actuarial Studies
  • G23 - Non-bank Financial Institutions; Financial Instruments; Institutional Investors
  • G24 - Investment Banking; Venture Capital; Brokerage; Ratings and Ratings Agencies
  • G28 - Government Policy and Regulation
  • Browse content in G3 - Corporate Finance and Governance
  • G30 - General
  • G32 - Financing Policy; Financial Risk and Risk Management; Capital and Ownership Structure; Value of Firms; Goodwill
  • G33 - Bankruptcy; Liquidation
  • G34 - Mergers; Acquisitions; Restructuring; Corporate Governance
  • G35 - Payout Policy
  • G38 - Government Policy and Regulation
  • Browse content in G5 - Household Finance
  • G51 - Household Saving, Borrowing, Debt, and Wealth
  • Browse content in H - Public Economics
  • Browse content in H1 - Structure and Scope of Government
  • H10 - General
  • H11 - Structure, Scope, and Performance of Government
  • H12 - Crisis Management
  • Browse content in H2 - Taxation, Subsidies, and Revenue
  • H20 - General
  • H22 - Incidence
  • H23 - Externalities; Redistributive Effects; Environmental Taxes and Subsidies
  • H25 - Business Taxes and Subsidies
  • H26 - Tax Evasion and Avoidance
  • Browse content in H3 - Fiscal Policies and Behavior of Economic Agents
  • Browse content in H4 - Publicly Provided Goods
  • H40 - General
  • H41 - Public Goods
  • Browse content in H5 - National Government Expenditures and Related Policies
  • H50 - General
  • H53 - Government Expenditures and Welfare Programs
  • H55 - Social Security and Public Pensions
  • H56 - National Security and War
  • Browse content in H6 - National Budget, Deficit, and Debt
  • H60 - General
  • H62 - Deficit; Surplus
  • H63 - Debt; Debt Management; Sovereign Debt
  • H68 - Forecasts of Budgets, Deficits, and Debt
  • Browse content in H7 - State and Local Government; Intergovernmental Relations
  • H70 - General
  • H74 - State and Local Borrowing
  • H77 - Intergovernmental Relations; Federalism; Secession
  • Browse content in I - Health, Education, and Welfare
  • Browse content in I0 - General
  • I00 - General
  • Browse content in I1 - Health
  • I10 - General
  • I12 - Health Behavior
  • I14 - Health and Inequality
  • I15 - Health and Economic Development
  • Browse content in I2 - Education and Research Institutions
  • I20 - General
  • I21 - Analysis of Education
  • I23 - Higher Education; Research Institutions
  • I24 - Education and Inequality
  • I26 - Returns to Education
  • Browse content in I3 - Welfare, Well-Being, and Poverty
  • I30 - General
  • I31 - General Welfare
  • I32 - Measurement and Analysis of Poverty
  • I38 - Government Policy; Provision and Effects of Welfare Programs
  • Browse content in J - Labor and Demographic Economics
  • Browse content in J0 - General
  • J00 - General
  • J01 - Labor Economics: General
  • J08 - Labor Economics Policies
  • Browse content in J1 - Demographic Economics
  • J10 - General
  • J13 - Fertility; Family Planning; Child Care; Children; Youth
  • J15 - Economics of Minorities, Races, Indigenous Peoples, and Immigrants; Non-labor Discrimination
  • J16 - Economics of Gender; Non-labor Discrimination
  • J18 - Public Policy
  • Browse content in J2 - Demand and Supply of Labor
  • J20 - General
  • J21 - Labor Force and Employment, Size, and Structure
  • J22 - Time Allocation and Labor Supply
  • J23 - Labor Demand
  • J24 - Human Capital; Skills; Occupational Choice; Labor Productivity
  • J26 - Retirement; Retirement Policies
  • J28 - Safety; Job Satisfaction; Related Public Policy
  • J29 - Other
  • Browse content in J3 - Wages, Compensation, and Labor Costs
  • J30 - General
  • J31 - Wage Level and Structure; Wage Differentials
  • J32 - Nonwage Labor Costs and Benefits; Retirement Plans; Private Pensions
  • J33 - Compensation Packages; Payment Methods
  • J38 - Public Policy
  • Browse content in J4 - Particular Labor Markets
  • J40 - General
  • J41 - Labor Contracts
  • J42 - Monopsony; Segmented Labor Markets
  • J44 - Professional Labor Markets; Occupational Licensing
  • J45 - Public Sector Labor Markets
  • J46 - Informal Labor Markets
  • J48 - Public Policy
  • J49 - Other
  • Browse content in J5 - Labor-Management Relations, Trade Unions, and Collective Bargaining
  • J50 - General
  • J51 - Trade Unions: Objectives, Structure, and Effects
  • J52 - Dispute Resolution: Strikes, Arbitration, and Mediation; Collective Bargaining
  • J53 - Labor-Management Relations; Industrial Jurisprudence
  • J54 - Producer Cooperatives; Labor Managed Firms; Employee Ownership
  • J58 - Public Policy
  • Browse content in J6 - Mobility, Unemployment, Vacancies, and Immigrant Workers
  • J60 - General
  • J61 - Geographic Labor Mobility; Immigrant Workers
  • J62 - Job, Occupational, and Intergenerational Mobility
  • J63 - Turnover; Vacancies; Layoffs
  • J64 - Unemployment: Models, Duration, Incidence, and Job Search
  • J65 - Unemployment Insurance; Severance Pay; Plant Closings
  • J68 - Public Policy
  • J69 - Other
  • Browse content in J7 - Labor Discrimination
  • J71 - Discrimination
  • J78 - Public Policy
  • Browse content in J8 - Labor Standards: National and International
  • J80 - General
  • J81 - Working Conditions
  • J83 - Workers' Rights
  • J88 - Public Policy
  • Browse content in K - Law and Economics
  • Browse content in K0 - General
  • K00 - General
  • Browse content in K1 - Basic Areas of Law
  • K11 - Property Law
  • K12 - Contract Law
  • K13 - Tort Law and Product Liability; Forensic Economics
  • Browse content in K2 - Regulation and Business Law
  • K20 - General
  • K21 - Antitrust Law
  • K22 - Business and Securities Law
  • K23 - Regulated Industries and Administrative Law
  • K25 - Real Estate Law
  • Browse content in K3 - Other Substantive Areas of Law
  • K31 - Labor Law
  • K39 - Other
  • Browse content in K4 - Legal Procedure, the Legal System, and Illegal Behavior
  • K40 - General
  • K41 - Litigation Process
  • K42 - Illegal Behavior and the Enforcement of Law
  • Browse content in L - Industrial Organization
  • Browse content in L0 - General
  • L00 - General
  • Browse content in L1 - Market Structure, Firm Strategy, and Market Performance
  • L10 - General
  • L11 - Production, Pricing, and Market Structure; Size Distribution of Firms
  • L12 - Monopoly; Monopolization Strategies
  • L13 - Oligopoly and Other Imperfect Markets
  • L14 - Transactional Relationships; Contracts and Reputation; Networks
  • L16 - Industrial Organization and Macroeconomics: Industrial Structure and Structural Change; Industrial Price Indices
  • Browse content in L2 - Firm Objectives, Organization, and Behavior
  • L20 - General
  • L21 - Business Objectives of the Firm
  • L22 - Firm Organization and Market Structure
  • L23 - Organization of Production
  • L24 - Contracting Out; Joint Ventures; Technology Licensing
  • L25 - Firm Performance: Size, Diversification, and Scope
  • L26 - Entrepreneurship
  • L29 - Other
  • Browse content in L3 - Nonprofit Organizations and Public Enterprise
  • L30 - General
  • L31 - Nonprofit Institutions; NGOs; Social Entrepreneurship
  • L32 - Public Enterprises; Public-Private Enterprises
  • L33 - Comparison of Public and Private Enterprises and Nonprofit Institutions; Privatization; Contracting Out
  • L39 - Other
  • Browse content in L4 - Antitrust Issues and Policies
  • L40 - General
  • L41 - Monopolization; Horizontal Anticompetitive Practices
  • L44 - Antitrust Policy and Public Enterprises, Nonprofit Institutions, and Professional Organizations
  • Browse content in L5 - Regulation and Industrial Policy
  • L50 - General
  • L52 - Industrial Policy; Sectoral Planning Methods
  • Browse content in L6 - Industry Studies: Manufacturing
  • L60 - General
  • L61 - Metals and Metal Products; Cement; Glass; Ceramics
  • L66 - Food; Beverages; Cosmetics; Tobacco; Wine and Spirits
  • L67 - Other Consumer Nondurables: Clothing, Textiles, Shoes, and Leather Goods; Household Goods; Sports Equipment
  • Browse content in L7 - Industry Studies: Primary Products and Construction
  • L78 - Government Policy
  • Browse content in L8 - Industry Studies: Services
  • L80 - General
  • L82 - Entertainment; Media
  • Browse content in L9 - Industry Studies: Transportation and Utilities
  • L97 - Utilities: General
  • L98 - Government Policy
  • Browse content in M - Business Administration and Business Economics; Marketing; Accounting; Personnel Economics
  • Browse content in M0 - General
  • M00 - General
  • Browse content in M1 - Business Administration
  • M10 - General
  • M12 - Personnel Management; Executives; Executive Compensation
  • M13 - New Firms; Startups
  • M16 - International Business Administration
  • Browse content in M2 - Business Economics
  • M21 - Business Economics
  • Browse content in M3 - Marketing and Advertising
  • M37 - Advertising
  • Browse content in M4 - Accounting and Auditing
  • M41 - Accounting
  • M49 - Other
  • Browse content in M5 - Personnel Economics
  • M51 - Firm Employment Decisions; Promotions
  • M52 - Compensation and Compensation Methods and Their Effects
  • M54 - Labor Management
  • M55 - Labor Contracting Devices
  • Browse content in N - Economic History
  • Browse content in N0 - General
  • N00 - General
  • N01 - Development of the Discipline: Historiographical; Sources and Methods
  • Browse content in N1 - Macroeconomics and Monetary Economics; Industrial Structure; Growth; Fluctuations
  • N10 - General, International, or Comparative
  • N11 - U.S.; Canada: Pre-1913
  • N12 - U.S.; Canada: 1913-
  • N13 - Europe: Pre-1913
  • N14 - Europe: 1913-
  • N15 - Asia including Middle East
  • N17 - Africa; Oceania
  • Browse content in N2 - Financial Markets and Institutions
  • N20 - General, International, or Comparative
  • N23 - Europe: Pre-1913
  • N24 - Europe: 1913-
  • N25 - Asia including Middle East
  • N26 - Latin America; Caribbean
  • Browse content in N3 - Labor and Consumers, Demography, Education, Health, Welfare, Income, Wealth, Religion, and Philanthropy
  • N30 - General, International, or Comparative
  • N32 - U.S.; Canada: 1913-
  • N34 - Europe: 1913-
  • Browse content in N4 - Government, War, Law, International Relations, and Regulation
  • N43 - Europe: Pre-1913
  • Browse content in N5 - Agriculture, Natural Resources, Environment, and Extractive Industries
  • N50 - General, International, or Comparative
  • N51 - U.S.; Canada: Pre-1913
  • N52 - U.S.; Canada: 1913-
  • N55 - Asia including Middle East
  • N7 - Transport, Trade, Energy, Technology, and Other Services
  • Browse content in N8 - Micro-Business History
  • N80 - General, International, or Comparative
  • Browse content in O - Economic Development, Innovation, Technological Change, and Growth
  • Browse content in O1 - Economic Development
  • O10 - General
  • O11 - Macroeconomic Analyses of Economic Development
  • O12 - Microeconomic Analyses of Economic Development
  • O13 - Agriculture; Natural Resources; Energy; Environment; Other Primary Products
  • O14 - Industrialization; Manufacturing and Service Industries; Choice of Technology
  • O15 - Human Resources; Human Development; Income Distribution; Migration
  • O16 - Financial Markets; Saving and Capital Investment; Corporate Finance and Governance
  • O17 - Formal and Informal Sectors; Shadow Economy; Institutional Arrangements
  • O18 - Urban, Rural, Regional, and Transportation Analysis; Housing; Infrastructure
  • O19 - International Linkages to Development; Role of International Organizations
  • Browse content in O2 - Development Planning and Policy
  • O20 - General
  • O23 - Fiscal and Monetary Policy in Development
  • O24 - Trade Policy; Factor Movement Policy; Foreign Exchange Policy
  • O25 - Industrial Policy
  • Browse content in O3 - Innovation; Research and Development; Technological Change; Intellectual Property Rights
  • O30 - General
  • O31 - Innovation and Invention: Processes and Incentives
  • O32 - Management of Technological Innovation and R&D
  • O33 - Technological Change: Choices and Consequences; Diffusion Processes
  • O34 - Intellectual Property and Intellectual Capital
  • O35 - Social Innovation
  • O38 - Government Policy
  • O39 - Other
  • Browse content in O4 - Economic Growth and Aggregate Productivity
  • O40 - General
  • O41 - One, Two, and Multisector Growth Models
  • O43 - Institutions and Growth
  • O44 - Environment and Growth
  • O47 - Empirical Studies of Economic Growth; Aggregate Productivity; Cross-Country Output Convergence
  • Browse content in O5 - Economywide Country Studies
  • O50 - General
  • O51 - U.S.; Canada
  • O52 - Europe
  • O53 - Asia including Middle East
  • O54 - Latin America; Caribbean
  • O55 - Africa
  • Browse content in P - Economic Systems
  • Browse content in P0 - General
  • P00 - General
  • Browse content in P1 - Capitalist Systems
  • P10 - General
  • P11 - Planning, Coordination, and Reform
  • P12 - Capitalist Enterprises
  • P13 - Cooperative Enterprises
  • P14 - Property Rights
  • P16 - Political Economy
  • P17 - Performance and Prospects
  • Browse content in P2 - Socialist Systems and Transitional Economies
  • P20 - General
  • P21 - Planning, Coordination, and Reform
  • P25 - Urban, Rural, and Regional Economics
  • Browse content in P3 - Socialist Institutions and Their Transitions
  • P30 - General
  • P31 - Socialist Enterprises and Their Transitions
  • P35 - Public Economics
  • P36 - Consumer Economics; Health; Education and Training; Welfare, Income, Wealth, and Poverty
  • P37 - Legal Institutions; Illegal Behavior
  • Browse content in P4 - Other Economic Systems
  • P40 - General
  • P41 - Planning, Coordination, and Reform
  • P46 - Consumer Economics; Health; Education and Training; Welfare, Income, Wealth, and Poverty
  • P48 - Political Economy; Legal Institutions; Property Rights; Natural Resources; Energy; Environment; Regional Studies
  • Browse content in P5 - Comparative Economic Systems
  • P50 - General
  • P51 - Comparative Analysis of Economic Systems
  • P52 - Comparative Studies of Particular Economies
  • Browse content in Q - Agricultural and Natural Resource Economics; Environmental and Ecological Economics
  • Browse content in Q0 - General
  • Q00 - General
  • Q01 - Sustainable Development
  • Browse content in Q1 - Agriculture
  • Q15 - Land Ownership and Tenure; Land Reform; Land Use; Irrigation; Agriculture and Environment
  • Q18 - Agricultural Policy; Food Policy
  • Browse content in Q3 - Nonrenewable Resources and Conservation
  • Q30 - General
  • Browse content in Q4 - Energy
  • Q41 - Demand and Supply; Prices
  • Q42 - Alternative Energy Sources
  • Q48 - Government Policy
  • Browse content in Q5 - Environmental Economics
  • Q50 - General
  • Q54 - Climate; Natural Disasters; Global Warming
  • Q56 - Environment and Development; Environment and Trade; Sustainability; Environmental Accounts and Accounting; Environmental Equity; Population Growth
  • Q57 - Ecological Economics: Ecosystem Services; Biodiversity Conservation; Bioeconomics; Industrial Ecology
  • Browse content in R - Urban, Rural, Regional, Real Estate, and Transportation Economics
  • Browse content in R0 - General
  • R00 - General
  • Browse content in R1 - General Regional Economics
  • R10 - General
  • R11 - Regional Economic Activity: Growth, Development, Environmental Issues, and Changes
  • R12 - Size and Spatial Distributions of Regional Economic Activity
  • R15 - Econometric and Input-Output Models; Other Models
  • Browse content in R2 - Household Analysis
  • R20 - General
  • Browse content in R3 - Real Estate Markets, Spatial Production Analysis, and Firm Location
  • R30 - General
  • R31 - Housing Supply and Markets
  • R4 - Transportation Economics
  • Browse content in R5 - Regional Government Analysis
  • R51 - Finance in Urban and Rural Economies
  • R58 - Regional Development Planning and Policy
  • Browse content in Y - Miscellaneous Categories
  • Browse content in Y1 - Data: Tables and Charts
  • Y10 - Data: Tables and Charts
  • Browse content in Y3 - Book Reviews (unclassified)
  • Y30 - Book Reviews (unclassified)
  • Browse content in Y8 - Related Disciplines
  • Y80 - Related Disciplines
  • Browse content in Z - Other Special Topics
  • Browse content in Z0 - General
  • Z00 - General
  • Browse content in Z1 - Cultural Economics; Economic Sociology; Economic Anthropology
  • Z10 - General
  • Z11 - Economics of the Arts and Literature
  • Z12 - Religion
  • Z13 - Economic Sociology; Economic Anthropology; Social and Economic Stratification
  • Z18 - Public Policy
  • Browse content in Z2 - Sports Economics
  • Z29 - Other
  • Advance articles
  • Editor's Choice
  • Author Guidelines
  • Submission Site
  • Open Access
  • About Cambridge Journal of Economics
  • About the Cambridge Political Economy Society
  • Editorial Board
  • Advertising and Corporate Services
  • Self-Archiving Policy
  • Dispatch Dates
  • Terms and Conditions
  • Journals on Oxford Academic
  • Books on Oxford Academic

Issue Cover

Article Contents

1. introduction, 2. paris purposes and the future we made, 3. the problem of unmaking, 4. conclusion: unmaking and is paris possible, conflict of interest statement, bibliography.

  • < Previous

Electric vehicles: the future we made and the problem of unmaking it

  • Article contents
  • Figures & tables
  • Supplementary Data

Jamie Morgan, Electric vehicles: the future we made and the problem of unmaking it, Cambridge Journal of Economics , Volume 44, Issue 4, July 2020, Pages 953–977, https://doi.org/10.1093/cje/beaa022

  • Permissions Icon Permissions

The uptake of battery electric vehicles (BEVs), subject to bottlenecks, seems to have reached a tipping point in the UK and this mirrors a general trend globally. BEVs are being positioned as one significant strand in the web of policy intended to translate the good intentions of Article 2 of the Conference of the Parties 21 Paris Agreement into reality. Governments and municipalities are anticipating that a widespread shift to BEVs will significantly reduce transport-related carbon emissions and, therefore, augment their nationally determined contributions to emissions reduction within the Paris Agreement. However, matters are more complicated than they may appear. There is a difference between thinking we can just keep relying on human ingenuity to solve problems after they emerge and engaging in fundamental social redesign to prevent the trajectories of harm. BEVs illustrate this. The contribution to emissions reduction per vehicle unit may be less than the public initially perceive since the important issue here is the lifecycle of the BEV and this is in no sense zero-emission. Furthermore, even though one can make the case that BEVs are a superior alternative to the fossil fuel-powered internal combustion engine, the transition to BEVs may actually facilitate exceeding the carbon budget on which the Paris Agreement ultimately rests. Whether in fact it does depends on the nature of the policy that shapes the transition. If the transition is a form of substitution that conforms to rather than shifts against current global scales and trends in private transportation, then it is highly likely that BEVs will be a successful failure. For this not to be the case, then the transition to BEVs must be coordinated with a transformation of the current scales and trends in private transportation. That is, a significant reduction in dependence on and individual ownership of powered vehicles, a radical reimagining of the nature of private conveyance and of public transportation.

According to the UK Society of Motor Manufacturers and Traders (SMMT), the Tesla Model 3 sold 2,685 units in December 2019, making it the 9th best-selling car in the country in that month (by new registrations; in August, a typically slow month for sales, it had been 3rd with 2,082 units sold; Lea, 2019; SMMT, 2019 ). As of early 2020, battery electric vehicles (BEVs) such as the new Hyundai Electric Kona had a two-year waiting list for delivery and the Kia e-Niro a one-year wait. The uptake of electric vehicles, subject to bottlenecks, seems to have reached a tipping point in the UK and this transcends the popularity of any given model. This possible tipping point mirrors a general trend globally (however, see later for quite what this means). At the regional, national and municipal scale, public health and environmentally informed legislation are encouraging vehicle manufacturers to invest heavily in alternative fuel vehicles and, in particular, BEVs and plug-in hybrid vehicles (PHEVs), which are jointly categorised within ‘ultra-low emission vehicles’ (ULEVs). 1 According to a report by Deloitte, more than 20 major cities worldwide announced plans in 2017–18 to ban petrol and diesel cars by 2030 or sooner ( Deloitte, 2018 , p. 5). All the major manufacturers have or are launching BEV models, and so vehicles are becoming available across the status and income spectrum that has in the past determined market segmentation. According to the consultancy Frost & Sullivan (2019) , there were 207 models (143 BEVs, 64 PHEVs) available globally in 2018 compared with 165 in 2017.

In 2018, the UK government published its Road to Zero policy commitment and introduced the Automated and Electric Vehicles Act 2018 , which empowers future governments to regulate regarding the required infrastructure. Road to Zero announced an ‘expectation’ that between 50% and 70% of new cars and vans will be electric by 2030 and the intention to ‘end the sale of new conventional petrol and diesel cars and vans by 2040’, with the ‘ambition’ that by 2050 almost all vehicles on the road will be ‘zero-emission’ at the point of use ( Department for Transport, 2018 ). Progress towards these goals was to be reviewed 2025. 2 However, on 4 February 2020, Prime Minister Boris Johnson announced that in the run-up to Conference of the Parties (COP)26 in Glasgow (now postponed), Britain would bring forward its 2040 goal to 2035. The UK is a member of the Clean Energy Ministerial Campaign (CEM), which launched the EV30@30 initiative in 2017, and its Road to Zero policy commitments broadly align with those of many European countries. 3 Norway has longstanding generous incentives for BEVs ( Holtsmark and Skonhoft, 2014 ) and 31% of all cars sold in 2018 and just under 50% in the first half of 2019 in Norway were BEVs. According to the International Energy Agency (IEA), Norway is the per capita global leader in electric vehicle uptake ( IEA, 2019A ). 4

BEVs, then, are being positioned as one significant strand in the web of policy intended to translate the good intentions of Article 2 of the COP 21 Paris Agreement into reality (see Morgan, 2016 ; IEA, 2019A , pp. 11–2). Clearly, governments and municipalities are anticipating that a widespread shift to electric vehicles will significantly reduce transport-related carbon emissions and, therefore, augment their nationally determined contributions (NDCs) to emissions reduction within the Paris Agreement. And, since the BEV trend is global, the impacts potentially also apply to countries whose relation to Paris is more problematic, including the USA (for Trump and his context, see Gills et al. , 2019 ). However, matters are more complicated than they may appear. Clearly, innovation and technological change are important components in our response to the challenge of climate change. However, there is a difference between thinking we can just keep relying on human ingenuity to solve problems after they emerge and engaging in fundamental social redesign to prevent the trajectories of harm. BEVs illustrate this. In what follows we explore the issues.

The aim of this paper, then, is to argue that it is a mistake to claim, assert or assume that BEVs are necessarily a panacea for the emissions problem. To do so would be an instance of what ecological economists refer to as ‘technocentrism’, as though simply substituting BEVs for existing internal combustion engine (ICE) vehicles was sufficient. The literature on this is, of course, vast, if one consults specialist journals or recent monographs (e.g. Chapman, 2007 ; Bailey and Wilson, 2009 ; Williamson et al. , 2018 ), but remains relatively under-explored in general political economy circles at a time of ‘Climate Emergency’, and so warrants discussion in introductory and indicative fashion, setting out, however incompletely, the range of issues at stake. To be clear, the very fact that there is a range is itself important. BEVs are technology, technologies have social contexts and social contexts include systemic features and related attitudes and behaviours. Technocentrism distracts from appropriate recognition of this. At its worse, technocentrism fails to address and so works to reproduce a counter-productive ecological modernisation: the technological focus facilitates socio-economic trends, which are part of the broader problem rather than solutions to it. In the case of BEVs, key areas to consider and points to make include:

Transport is now one of, if not, the major source of carbon emissions in the UK and in many other countries. Transport emissions stubbornly resist reduction. The UK, like many other countries, exhibits contradictory trends and policy claims regarding future carbon emissions reductions. As such, it is an error to simply assume prior emissions reduction trends will necessarily continue into the future, and the new net-zero goal highlights the short time line and urgency of the problem.

Whilst BEVs are, from an emissions point of view, a superior technology to ICE vehicles, this is less than an ordinary member of the public might think. ‘Embodied emissions’, ‘energy mix’ and ‘life cycle’ analysis all matter.

There is a difference between ‘superior technology’ and ‘superior choice’, the latter must also take account of the scale of and general trend growth in vehicle ownership and use. It is this that creates a meaningful context for what substitution can be reasonably expected to achieve.

A 1:1 substitution of BEVs for ICE vehicles and general growth in the number of vehicles potentially violates the Precautionary Principle. It creates a problem that did not need to exist, e.g. since there is net growth, it involves ‘emission reductions’ within new emissions sources and this is reckless. Inter alia , a host of fallacies and other risks inherent to the socio-economy of BEVs and resource extraction/dependence also apply.

As such, it makes more sense to resist rather than facilitate techno-political lock-in or path-dependence on private transportation and instead to coordinate any transition to BEVs with a more fundamental social redesign of public transport and transport options.

This systematic statement should be kept in mind whilst reading the following. Cumulatively, the points stated facilitate appropriate consideration of the question: What kind of solution are BEVs to what kind of problem? And we return to this in the conclusion. It is also worth bearing in mind, though it is not core to the explicit argument pursued, that an economy is a complex evolving open system and economics has not only struggled to adequately address this in general, it has particularly done so in terms of ecological issues (for relevant critique, see especially the work of Clive Spash and collected, Fullbrook and Morgan, 2019 ). 5 Since we assume limited prior knowledge on the part of the reader, we begin by briefly setting out the road to the current carbon budget problem.

The United Nations Framework Convention on Climate Change (UNFCCC) was created in 1992. Article 2 of the Convention states its goal as, the ‘stabilization of greenhouse gas concentrations in the atmosphere at a level that would prevent dangerous anthropogenic interference with the climate system’ ( UNFCCC, 1992 , p. 4; Gills and Morgan, 2019 ). Emissions are cumulative because emitted CO 2 can stay in the atmosphere for well over one hundred years (other greenhouse gases [GHGs] tend to be of shorter duration). Our climate future is made now. The Intergovernmental Panel on Climate Change (IPCC) collates existent models to produce a forecast range and has typically used atmospheric CO 2 of 450 ppm as a level likely to trigger a 2°C average warming. This has translated into a ‘carbon budget’ restricting total cumulative emissions to the lower end of 3,000+ Gigatonnes of CO 2 (GtCO 2 ). In the last few years, climate scientists have begun to argue that positive feedback loops with adverse warming and other climatological and ecological effects may be underestimated in prior models (see Hansen et al. , 2017 ; Steffen et al. , 2018 ). Such concerns are one reason why Article 2 of the UNFCCC COP 21 Paris Agreement included a goal of at least trying to do better than the 2°C target—restricting warming to 1.5°C. This further restricts the available carbon budget. However, current Paris Agreement country commitments stated as NDCs look set to exceed the 3,000+ target in a matter of a few short years ( UNFCCC, 2015 ; Morgan, 2016 , 2017 ).

Since the industrial revolution began, we have already produced more than 2,000 GtCO 2 . Total annual emissions have increased rather than decreased over the period in which the problem has been recognised. The United Nations Environment Program (UNEP) publishes periodic ‘emissions gap’ reports. Its recent 10-year summary report notes that emissions grew at an average 1.6% per year from 2008 to 2017 and ‘show no signs of peaking’ ( Christensen and Olhoff, 2019 , p. 3). In 2018, the 9th Report stated that annual emissions in 2017 stood at a record of 53.5 Gigatonnes of CO 2 and equivalents (GtCO 2e ) ( UNEP, 2018 , p. xv). This compares to less than 25 GtCO 2 in 2000 and far exceeds on a global basis the level in the Kyoto Protocol benchmark year of 1990. According to the 9th Emissions Gap Report, 184 parties to the Paris Agreement had so far provided NDCs. If these NDCs are achieved, annual emissions in 2030 are projected to still be 53 GtCO 2e . However, if the current ‘implementation deficit’ continues global annual emissions could increase by about 10% to 59 GtCO 2e . This is because current emissions policy is not sufficient to offset the ‘key drivers’ of ‘economic growth and population growth’ ( Christensen and Olhoff, 2019 , p. 3). By sharp contrast, the IPCC Global Warming of 1.5 ° C report states that annual global emissions must fall by 45% from the 2017 figure by 2030 and become net zero by mid-century in order to achieve the Paris target ( IPCC, 2018 ). According to the subsequent 10th Emissions Gap Report, emissions increased yet again to 55.3 GtCO 2e in 2018 and, as a result of this adverse trend, emissions need to fall by 7.6% per year from 2020 to 2030 to achieve the IPCC goal, and this contrasts with less than 4% had reductions begun in 2010 and 15% if they are delayed until 2025 ( UNEP 2019A ). Current emissions trends mean that we will achieve an additional 500 GtCO 2 quickly and imply an average warming of 3 to 4°C over the rest of the century and into the next. We are thus on track for the ‘dangerous anthropogenic interference with the climate system’ that the COP process is intended to prevent ( UNFCCC, 1992 , p. 4). According to the 10th Emissions Gap Report, 78% of all emissions derive from the G-20 nations, and whilst many countries had recognised the need for net zero, only 5 countries of the G-20 had committed to this and none had yet submitted formal strategies. COP 25, December 2019, meanwhile, resulted in no overall progress other than on measurement and finance (for detailed analysis, see Newell and Taylor, 2020 ). As such, the situation is urgent and becoming more so.

Problems, moreover, have already begun to manifest ( UNEP 2019B , 2019B ; IPCC 2019A , 2019B ). Climate change does not respect borders, some countries may be more adversely affected sooner than others, but there is no reason to assume that cumulative effects will be localised. Moreover, there is no reason to assume that they will be manageable based on our current designs for life. In November 2019, several prominent systems and climate scientists published a survey essay in Nature highlighting nine critical climate tipping points that we are either imminently approaching or may have already exceeded ( Lenton et al. , 2018 ). In that same month, more than 11,250 scientists from 153 countries (the Alliance of World Scientists) signed a letter published in BioScience concurring that we now face a genuine existential ‘Climate Emergency’ and warning of ‘ecocide’ if ‘major transformations’ are not forthcoming ( Ripple et al. , 2019 ). We live in incredibly complex interconnected societies based on long supply chains and just in time delivery–few of us (including nations) are self-sufficient. Global human civilisation is extremely vulnerable and the carbon emission problem is only one of several conjoint problems created by our expansionary industrialised-consumption system. Appropriate and timely policy solutions are, therefore, imperative. Cambridge now has a Centre for the Study of Existential Risk and Oxford a Future of Humanity Institute (see also Servigne and Stevens, 2015 ). This is serious research, not millenarian cultishness. The Covid-19 outbreak only serves to underscore the fragility of our systems. As Michael Marmot, Professor of epidemiology has commented, the outbreak reveals not only how political decisions can make systems more vulnerable, but also how governments can, when sufficiently motivated, take immediate and radical action (Harvey, 2020). To reiterate, however, according to both the IPCC and UNEP, emissions must fall drastically. 6

Policy design and implementation are mainly national (domestic). As such, an initial focus on the UK provides a useful point of departure to contextualise what the transition to BEVs might be expected to achieve.

The UK is a Kyoto and Paris signatory. It is a member of the European Emissions Trading Scheme (ETS). The UK Climate Change Act 2008 was the world’s first long-term legally binding national framework for targeted statutory reductions in emissions. The Act required the UK to reduce its emissions by at least 80% by 2050 (below the 1990 baseline; this has been broadly in line with subsequent EU policy on the subject). 7 The Act put in place a system of five yearly ‘carbon budgets’ to keep the UK on an emissions reduction pathway to 2050. The subsequent carbon budgets have been produced with input from the Committee on Climate Change (CCC), an independent body created by the 2008 Act to advise the government. In November 2015, the CCC recommended a target of 57% below 1990 levels by the early 2030s (the fifth carbon budget). 8 Following the Paris Agreement’s new target of 1.5°C and the IPCC and UNEP reports late 2018, the CCC published the report Net Zero: The UK’s contribution to stopping global warming ( CCC, 2019 ). 9 The CCC report recognises that Paris creates additional responsibility for the UK to augment and accelerate its targets within the new bottom-up Paris NDC procedure. The CCC recommended an enhanced UK net-zero GHG emissions target (formally defined in terms of long-term and short-term GHGs) by 2050. This included emissions from aviation and shipping and with no use of strategies that offset or swap real emissions. In June 2019, Theresa May, then UK Prime Minister, committed to adopt the recommendation using secondary legislation (absorbed into the 2008 Act—but without the offset commitment). So, the UK is one of the few G-20 countries to, so far, provide a formal commitment on net zero, though as the UNEP notes, a commitment is not itself necessarily indicative of a realisable strategy. The CCC responded to the government announcement:

This is just the first step. The target must now be reinforced by credible UK policies, across government, inspiring a strong response from business, industry and society as a whole. The government has not yet moved formally to include international aviation and shipping within the target , but they have acknowledged that these sectors must be part of the whole economy strategy for net zero. We will assist by providing further analysis of how emissions reductions can be delivered in these sectors through domestic and international frameworks. 10

The development of policy is currently in flux during the Covid-19 lockdown and whilst Brexit reaches some kind of resolution. As noted in the Introduction section, however, May’s replacement, Boris Johnson has signalled his government’s commitment to achieving its statutory commitments. However, this has been met with some scepticism, not least because it has not been clear what new powers administrative bodies would have and over and above this many of the Cabinet are from the far right of the Conservative Party, and are on record as climate change sceptics or have a voting record of opposing environmentally focussed investment, taxes, subsidies and prohibitions (including the new Environment Secretary, George Eustice, formerly of UKIP). The policy may and hopefully will change, becoming more concrete, but it is still instructive to assess context and general trends.

The UK has one of the best records in the world on reducing emissions. However, given full context, this is not necessarily a cause for congratulation or confidence. It would be a mistake to think that emissions reduction exhibits a definite rate that can be projected from the past into the future. 11 This applies both nationally and globally. Some sources of relative reduction that are local or national have different significance on a global basis (they are partial transfers) and overall the closer one approaches net zero the more resistant or difficult it is likely to become to achieve reductions. The CCC has already begun to signal that the UK is now failing to meet its existent budgets. This follows periods of successive emissions reductions. According to the CCC, the UK has reduced its GHG emissions by approximately one-third since 1990. ‘Per capita emissions are now close to the global average at 7–8 tCO 2 e/person, having been over 50% above in 2008’ ( CCC, 2019 , p. 46). Other analyses are even more positive. According to Carbon Brief, emissions have fallen in seven consecutive years from 2013 to 2019 and by 40% compared with the 1990 benchmark. Carbon Brief claim that since 2010 the UK has the fastest rate of emissions reduction of any major economy. However, it concurs with the CCC that future likely reductions are less than the UK’s carbon budgets and that the new net-zero commitment requires: amounting to only an additional 10% reduction over the next decade to 2030. 12

Moreover, all analyses agree that the reduction has mainly been achieved by reducing coal output for use in electricity generation (switching to natural gas) and by relative deindustrialisation as the UK economy has continued to grow—manufacturing is a smaller part of a larger service-based economy. 13 And , the data are based on a production focussed accounting system. The accounting system does not include all emissions sources. It does not include those that the UK ‘imports’ based on consumption. UK consumption-based emissions per year are estimated to be about 70% greater than the production measure (for different methods, see DECC, 2015 ). 14 If consumption is included, the main estimates for falling emissions change to around a 10% reduction since 1990. Moreover, much of this has been achieved by relatively invisible historic transitions as the economy has evolved in lock-step with globalisation. That is, reductions have been ones that did not require the population to confront behaviours as they have developed. No onerous interventions have been imposed, as yet . 15 However, it does not follow that this can continue, since future reductions are likely to be more challenging. The UK cannot deindustrialise again (nor can the global economy, as is, simply deindustrialise in aggregate if final consumption remains the primary goal), and the UK has already mainly switched from coal energy production. Emissions from electricity generation may fall but it also matters what the electricity is being used to power. In any case, future emissions reductions, in general, require more effective changes in other sectors, and this necessarily seems to require everyone to question their socio-economic practices. Transport is a key issue.

As a ‘satellite’ of its National Accounts, the UK Office for National Statistics (ONS) publishes Environmental Accounts and these data are used to measure progress. Much of the data refer to the prior year or earlier. In 2017, UK GHG emissions were reported to be 566 million tonnes CO 2 e (2% less than 2016 and, as already noted about one-third of the 1990 level; ONS, 2019 ). The headline accounts break this down into four categories (for which further subdivisions are produced by various sources) and we can usefully contrast 1990 and recent data ( ONS, 2019 , p. 4):

The Environmental Accounts’ figures indicate some shifting in the relative sources of emissions over the last 30 years. As we have intimated, electricity generation and manufacturing have experienced reduced emissions, though they are far from zero; household and transport, meanwhile, have remained stubbornly high. Moreover, the accounts are also slightly misleading for the uninitiated, since transport refers to the industry and not all transport. Domestic car ownership and use are part of the household sector, and it is the continued dependence on car ownership that provides, along with heating and insulation issues, one of the major sources of the persistently high level of household emissions. The UK Department for Business, Energy and Industrial Strategy (DBEIS) provides differently organised statistics and attributes cars to its transport category and uses a subsequent residential category rather than household category. The Department’s statistical release in 2018 thus attributes a higher 140 MtCO 2 e to transport for 2016, whilst the residential category is a correspondingly lower figure of approximately 106 MtCO 2 e. The 140 MtCO 2 e is just slightly less than the equivalent figure for 1990, although transport achieved a peak of about 156 MtCO 2 e in 2005 ( DBEIS, 2018 , pp. 8–9). As of 2016, transport becomes the largest source of emissions based on DBEIS data (exceeding energy supply) whilst households become the largest in the Environmental Accounts. In any case, looking across both sets of accounts, the important point here is that since 1990 transport as a source of emissions has remained stubbornly high. Transport emissions have been rising as an industrial sector in the Environmental Accounts or relatively consistent and recently rising in its total contribution in the DBEIS data. The CCC Net Zero report draws particular attention to this. Drawing on the DBEIS data, it states that ‘Transport is now the largest source of UK GHG emissions (23% of the total) and saw emissions rise from 2013 to 2017’ ( CCC, 2019 , p. 48). More generally, the report states that despite some progress in terms of the UK carbon budgets, ‘policy success and progress in reducing emissions has been far from universal’ ( CCC, 2019 , p. 48). The report recommends ( CCC, 2019 , pp. 23–6, 34):

A fourfold increase by 2050 in low carbon (renewables) electricity

Developing energy storage (to enhance the use of renewables such as wind)

Energy-efficient buildings and a shift from gas central heating and cooking

Halting the accumulation of biodegradable waste in landfills

Developing carbon capture technology

Reducing agricultural emissions (mainly dairy but also fertiliser use)

Encouraging low or no meat diets

Land management to increase carbon retention/absorption

Rapid transition to electric vehicles and public transport

As we noted in the Introduction section, the UK Department for Transport Road To Zero document stated a goal of ending the sale of conventional diesel- and petrol-powered ICE vehicles by 2040. The CCC suggested improving on this:

Electric vehicles. By 2035 at the latest all new cars and vans should be electric (or use a low-carbon alternative such as hydrogen). If possible, an earlier switchover (e.g. 2030) would be desirable, reducing costs for motorists and improving air quality. This could help position the UK to take advantage of shifts in global markets. The Government must continue to support strengthening of the charging infrastructure, including for drivers without access to off-street parking. ( CCC, 2019 , p. 34)

The UK government’s response to these and other similar suggestions has been to bring the target date forward to 2035 and to propose that the prohibition will also apply to hybrids. However, the whole is set to go out to consultation and no detail has so far (early 2020) been forthcoming. In its 11 March 2020 Budget, the government also committed £1 billion to ‘green transport solutions’, including £500 million to support the rollout of the electric vehicle charging infrastructure, whilst extending the current grant/subsidy scheme for new electric vehicles (albeit at a reduced rate of £3000 from £3500 per new registration). It has also signalled that it may tighten the timeline for sales prohibition further to 2030. 16 As a policy, much of this is, ostensibly at least, positive, but there is a range of issues that need to be considered regarding what is being achieved. The context of transition matters and this may transcend the specifics of current policy.

3.1 BEV transition: life cycles?

The CCC is confident that a transition to electric vehicles can be a constructive contribution to achieving net-zero emissions by mid-century. However, the point is not unequivocal. The previously quoted CCC communique following the UK government’s commitment to implement Net Zero uses the phrase ‘credible UK policies, across government, inspiring a strong response from business, industry and society as a whole’, and the CCC report places an emphasis on BEVs and a transition to public transport. The relative dependence between these two matters (and see Conclusion). BEVs are potentially (almost) zero emissions in use. But they are not zero emissions in practice. Given this, then the substitution of BEVs for current carbon-powered ICEs is potentially problematic, depending on trends in ownership of and use of powered vehicles (private transportation). These points will become clearer as we proceed.

BEVs are not zero emission in context and based on the life cycle. This is for two basic reasons. First, a BEV is a powered vehicle and so the source of power can be from carbon-based energy supply sources (and this varies with the ‘energy mix’ of electricity production in different countries; IEA, 2019A , p. 8). Second, each new vehicle is a material product. Each vehicle is made of metals, plastics, rubber and so forth. Just the cabling in a car can be 60 kg of metals. All the materials must be mined and processed, or synthesised, the parts must be manufactured, transported and assembled, transported again for sale and then delivered. For example, according to the SMMT in 2016, only 12% of cars sold in the UK were built in the UK and 80% of those built in the UK were exported in that year. Some components (such as a steering column) enter and exit the UK multiple times whilst being built and modified and before final assembly. Vehicle manufacture is a global business in terms of procuring materials and a mainly regional (in the international sense) business in terms of component manufacture for assembly and final sales. Power is used throughout this process and many miles are travelled. Moreover, each vehicle must be maintained and serviced thereafter, which compounds this utilisation of resources. BEVs are a subcategory of vehicles and production locations are currently more concentrated than for vehicles in general (Tesla being the extreme). 17 In any case, producing a BEV is an economic activity and it is not environmentally costless. As Georgescu-Roegen (1971) noted long ago and ecologically minded economists continue to highlight (see Spash, 2017 ; Holt et al. , 2009 ), production cannot evade thermodynamic consequences. In terms of BEVs, the primary focus of analysis in this second sense of manufacturing as a source of contributory emissions has been the carbon emissions resulting from battery production. Based on current technology, batteries are heavy (a significant proportion of the weight of the final vehicle) and energy intensive to produce.

Comparative estimates regarding the relative life cycle emissions of BEVs with equivalent fossil fuel-powered vehicles are not new. 18 Over the last decade, the number of life cycle studies has steadily risen as the interest in and uptake of BEVs have increased. Clearly, there is great scope for variation in findings, since the energy mix for electricity supply varies by country and the assumptions applied to manufacturing can vary between studies. At the same time, the general trend over the last decade has been for the energy mix in many countries to include more renewables and for manufacturing to become more energy efficient. This is partly reflected in metrics based on emissions per $GDP, which in conjunction with relative expansion in service sectors are used to establish ‘relative decoupling’. So, given that both the energy mix of power production and the emissions derived from production can improve, then one might expect a general trend of improved emissions claims for BEVs in recent years and this seems to be the case.

For example, if we go back to 2010, the UK Royal Academy of Engineering found that technology would likely favour PHEVs over BEVs in the near future because the current energy mix and state of battery technology indicated that emissions deriving from charging were typically higher for BEVs than an average ordinary car’s fuel consumption—providing a reason to persist with ICE vehicles or, more responsibly, choose hybrids over pure electric ( Royal Academy of Engineering, 2010 ). Using data up to 2013, but drawing on the previous decade, Holtsmark and Skonhoft (2014) come to similar conclusions based on the most advanced BEV market—Norway. Focussing mainly on energy mix (with acknowledgement that a full life cycle needs to be assessed) they are deeply sceptical that BEVs are a significant net reduction in carbon emissions ( Holtsmark and Skonhoft, 2014 , pp. 161, 164). Neither the Academy nor Holtsmark and Skonhoft are merely sceptical. The overall point of the latter was that more needed to be done to accelerate the use of low or no carbon renewables for power infrastructure (a point the CCC continues to make). This, of course, has happened in many places, including the UK. That is, acceleration of the use of renewables, though it is by no means the case government can take direct credit for this in the UK (and there is also evidence on a global level that a transition to clean energy from fossil fuel forms is much slower than some data sources indicate; see Smil, 2017A , 2017B ). 19 In terms of BEVs, however, recent analyses are considerably more optimistic regarding emissions potential per BEV (e.g. Hoekstra, 2019 ; Regett et al. , 2019 ). Research by Staffell et al. (2019) at Imperial for the power corporation, Drax, provides some interesting insights and contemporary metrics.

Staffell et al. split BEVs into three categories based on conjoint battery and vehicle size: a 30–45 kWh battery car, equivalent to a mid-range or standard car; a heavier, longer-range, 90–100 kWh battery car, equivalent to a luxury or SUV model; and a 30–40 kWh battery light van. They observe that a 40-litre tank of petrol releases 90–100 kgCO 2 when burnt and the ‘embodied’ emissions represented by the manufacture of a standard lithium-ion battery are estimated at 75–125 kgCO 2 per kWh. They infer that every kWh of power embodied in the manufacture of a battery is, therefore, approximately equivalent to using a full tank of petrol. For example, a 30 kWh battery embodies thirty 40-litre petrol tank’s worth of emissions. The BEV’s are also a source of emissions based on the energy mix used to charge the battery for use. The in-use emissions for the BEV are a consequence of the energy consumed per km and this depends on the weight of car and efficiency of the battery. 20 They estimate 33 gCO 2 per km for standard BEVs, 44–54 gCO 2 for luxury and SUVs and 40 gCO 2 for vans. In all cases, this is significantly less than an equivalent fossil-fuel vehicle.

The insight that the estimates and comparisons are leading towards is that the battery embodies an ‘upfront carbon cost’ which can be gradually ‘repaid’ by the saving on emissions represented by driving a BEV compared with driving an equivalent fossil fuel-powered vehicle. That is, the environmental value of opting for BEVs increases over time. Moreover, if the energy mix is gradually becoming less carbon based, this effect is likely to improve further. Based on these considerations, Staffell et al. estimate that it may take 2–4 years to repay the embodied emissions in the battery for a standard BEV and 5 to 6 for the luxury or SUV models. Fundamentally, assuming 15 years to be typical for the on-the-road life expectancy of a vehicle, they find lifetime emissions for each BEV category are lower than equivalent fossil-fuel vehicles.

Still, the implication is that BEVs are not zero emission. Moreover, the degree to which this is so is likely to be significantly greater than a focus on the battery alone indicates. Romare and Dahlöff (2017) , assess the life-cycle of battery production (not use), and in regard of the stages of battery production find that the manufacturing stages account for about 50% of the emissions and the mining and processing stages about the same. They infer that there is significant scope for further emissions reductions as manufacturing processes improve and the Drax study seems to confirm this. However, whilst the battery may be the major component, as we have already noted, vehicle manufacture is a major process in terms of all components and in terms of distance travelled in production and distribution. It is also worth noting that the weight of batteries creates strong incentives to opt for lighter materials for other parts of the vehicle. Most current vehicles are steel based. An aluminium vehicle is lighter, but the production of aluminium is more carbon intensive than steel, so there are also further hidden trade-offs that the positive narrative for BEVs must consider. 21

The general point worth emphasising here is that there is basic uncertainty built into the complex evolving process of transition and change. There is a basic ontology issue here familiar in economic critique: there is no simple way to model the changes with confidence, and in broader context confidence in modelling may itself be a problem here when translated into policy, since it invites complacency. 22 That said, the likely direction of travel is towards further improvements in the energy mix and improvements in battery technology. Both these may be incremental or transformational depending on future technologies (fusion for energy mix and organics and solid-state technologies for batteries perhaps). 23 But one must still consider time frames and ultimate context. 24 The context is a carbon budget and the need for radical reductions in emissions by 2030 and net zero by mid-century. Consider: if just the battery of a car requires four years to be paid back then there is no significant difference in the contribution to emissions from the vehicle into the mid 2020s. For larger vehicles, this becomes the later 2020s, and each year of delay in transition for the individual owner is another year closer to 2030. Since transport is (stubbornly) the major source of emissions in the UK and a major source in the world, this is not irrelevant. BEVs can readily be a successful failure in Paris terms. This brings us to the issue of trends in vehicle ownership and substitutions. This also matters for what we mean by transition.

3.2 Substitutions and transformations: successful failure?

There are many ways to consider the problem of transition. Consider the ‘Precautionary Principle’. This is Principle 15 of the 1992 Rio Declaration: ‘In order to protect the environment, the precautionary principle shall be widely applied by the States [UN members] according to their capabilities. Where there are threats of serious or irreversible damage, lack of full scientific certainty shall not be used as a reason for postponing cost-effective measures to prevent environmental degradation’ (UNCED). Assuming we can simply depend on unrealised technology potentially violates the Principle. Why is this so? If BEVs are a source of net emissions, then each new vehicle continues to contribute to overall emissions. The current number of vehicles to be replaced, therefore, is a serious consideration, as is any growth trend. Here, social redesign rather than merely adopting new technology is surely more in accordance with the Precautionary Principle. BEVs may be sources of lower emissions than fossil fuel-powered vehicles, but it does not follow that we are constrained to choose between just these two options or that it makes sense to do so in aggregate, given the objective of radical and rapid reduction in emissions. If time is short and numbers of vehicles are large and growing then the implication is that substitution of BEVs should (from a precautionary point of view) occur in a context that is oppositional to this growing trend. That is, the goal should be one of reducing private car ownership and use, and increasing the availability, pervasiveness and use of public transport (and alternatives to private vehicle ownership). This is an issue compounded by the finding that there is an upfront carbon cost from BEVs. Some consideration of current vehicle numbers and trends in the UK and globally serve to reinforce the point.

The UK Department for Transport publishes annual statistics for vehicle licensing. According to the 2019 statistical release for 2018 data, there were 38.2 million licensed vehicles in Britain and 39.4 million including Northern Ireland ( Department for Transport, 2019 ). Vehicles are categorised into cars, light goods vehicles, heavy goods vehicles, motorcycles and buses and coaches. Cars comprised 31.5 million of the total (82%) and the total represented a 1.2% increase in the year 2017. There is, furthermore, a long-term year-on-year trend increase in vehicles since World War II and over the last 20 years that growth (the net change as new vehicles are licensed and old vehicles taken off the road) has averaged 630,000 vehicles per year ( Department for Transport, 2019 , p. 7). This is partly accounted for not only by population growth, and business growth, but also by an increase in the number of vehicles per household. According to the statistical release, 2.9 million new vehicles were registered in 2018, and though this was about 5% fewer than 2017 the figure remained broadly consistent with long-term trends in numbers and still represented growth (contributing to the stated 1.2% increase). 25 Of the total new registrations in 2018, 2.3 million were cars and 360,000 were light goods vehicles. Around 2 million has been typical for cars.

The point to take from these metrics is that numbers are large and context matters. Cars represent 31.5 million emission sources and there are 39.4 million vehicles in the UK. Replacing these 1:1 reproduces an emissions problem. Replacing them in conjunction with an ownership growth trend exacerbates the emissions problem that then has to be resolved. If around 2 million new cars are registered per year then the point at which the BEVs amongst these new registrations can be assumed to begin payback for embodied emissions prior to the point at which they become net sources of reduced (and not zero ) emissions is staggered over future years based on the rate of switching. There are then also net new vehicles. Given there are 31.5 million cars to be replaced over time (plus net growth), there is a high likelihood of significant transport emissions up to and beyond 2030. The problem, of course, is implicit in the Department for Transport policy commitment to end sales of petrol and diesel vehicles by 2035 and ensure all vehicles are zero-emission in use by 2050. Knowingly committing to this ingrained emission problem, given we have already recognised the urgency and challenge of the carbon budget and the ‘stubbornness’ of transport emissions, is not prudent, if alternatives exist . It is producing a problem that need not exist purely because enabling car ownership and use is a line of least resistance in policy terms (it requires the least change in behaviour and thus provokes limited opposition). It is also worth noting that the UK, like most countries, has an ‘integrated’ transport policy. However, the phrasing disguises the relative levels of investment between different modes of transport. Austerity politics may have resulted in declining road quality in the UK but, in general terms, the UK is still committed to heavy investment in and expansion of its road system. 26 This infrastructure investment not only seems ‘economically rational’, but it is also a matter of relative emphasis and ‘lock-in’. The future policy is predicated on the dominance of road use and thus vehicle use.

The crux of the matter here is how we view political expedience. Surely this hinges on the consequences of policy failure. That is, the failure to implement an effective policy given the genuine problem expressed in the goal of 1.5 or 2°C. ‘Alternatives’ may seem unrealistic, but this is a matter of will and policy—of rational social design rather than impossibility. The IPCC and other sources suggest that achieving the Paris goals requires mobilisation of a kind not previously seen outside of wartime. Policy can pivot on this quite quickly, even if perhaps this can seem unlikely in 2020. Climate events may make this necessary and popular pressure and opinion may be transformed. This is currently uncertain. Positions on this may yet move quite quickly.

Lock-in also implies an underlying sociological issue. This is important to consider regarding simply opting for substitution without greater emphasis on reduction. Even if substitution occurs smoothly, it places greater pressure on areas of reduction over which we have less control as societies and involves an orientation that has further potential policy consequences that cannot be readily quantified and which increase the overall uncertainty regarding NDCs. As any modern historian, urban geographer or sociologist will attest, car ownership has been imbricate with the development and design—the configuration—of modern societies, and it has been deeply integrated into identity. Cars are social technologies and philosophers also have much to say about this sociality in general (e.g. Faulkner and Runde, 2013 ; Lawson, 2017 ). Cars are more than merely convenient; they are sources of autonomy and status (e.g. John Urry explored the sociology of ‘automobility’; see, Dennis and Urry, 2009 ). As such, the more that environmental and transport policy validate the car, then the more that the car is normalised through socialisation for the citizen, perhaps leading to citizens being more prepared to countenance locked-in harms (congestion, etc.) prior to change, in turn, making it less likely (sub)urban spaces are redesigned in ways predicated on the absence of (or severe limits to) private transport. The trend in many countries over the car era has been that building roads leads to more car use, which leads to congestion, which leads to more roads (especially in concentrated zones around [sub]urban spaces).

According to the UK Ordnance Survey, Britain has increased its total road surface by 132 square miles over the decade since 2010 (a 9% increase). According to the UK Department for Transport, vehicle traffic increased by 0.8% in 2019 (September to September) to 330.1 billion miles travelled and car travel, as a subset, increased to 258 billion miles (a 1.5% increase). 27 The 11 March 2020 Budget seems to confirm the trend. Whilst it commits around £1 billion to ‘green transport solutions’, this is in the context of a £27 billion announced investment in roads, including upgrading and a proposed 4,000 miles of new road. As the Green Party MP, Caroline Lucas, noted there is a basic disconnect here, since this seems set to increase the UK’s dependence on private transport, when it makes more sense to begin to curtail that dependence, given how significant the UK’s transport emissions are. 28 So, within the various tensions in policy, there seems to be a tendency to facilitate techno-political lock-in or path-dependence on private transportation. As Mattioli et al. (2020) argue, the multiple strands of policy and practice that maintain car dependence contribute to ‘carbon lock-in’. The systemic consequences matter both for the perpetuation of fossil fuel vehicle use in the short term and, given they are not net zero for emissions, powered vehicles in the longer term. Not only does this matter in the UK, but it also matters globally. All the issues stated are reproduced globally. Moreover, in some ways, they are compounded for countries where widespread car ownership is relatively new.

3.3 The fallacy of composition, problems that need not exist and resource risk

Estimates vary for the global total number of vehicles. According to Wards Intelligence, the global total was 1.32 billion in 2016 ( Petit, 2017 ). Extrapolated estimations imply that the total likely increased to more than 1.5 billion in 2019. In 1976, the figure was 342 million and in 1996, 670 million, so the trend implies an approximate doubling every 20 years, which if it continued would imply a figure approaching 3 billion by end of the 2030s. Clearly, it is problematic to simply extrapolate a linear trend, but it is not unreasonable to assume a general trend of growth. Observed experience is that many ‘developed’ country middle-class households have accommodated more than one car per household. This is classically the case in the USA. In 2017, the USA, with a population of 325.7 million in that year, reported a total of 272.5 million registered vehicles compared with 193 million in 1990 ( Statista, 2019A ). In any case, the world population is still growing, incomes are growing and many countries are far from a position of one car per household. China with a population of 1.3 billion overtook the USA in the total number of registered vehicles around 2016 to 2017, with 300.3 million registered vehicles in March of 2017 (Zheng, 2017). Growth is rapid and the China Traffic Bureau of the Ministry of Public Security reported a total of 325 million registered vehicles, December 2018, an increase of 15.56 million in the year ( China Daily , 2018 ). The People’s Republic is now the world’s largest car market and the number of registered cars increased to 240 million in 2018 ( Statista, 2019B ). India too has rapidly growing car ownership and on a lesser scale this is replicated across the developing world.

For our purposes, two well-known concepts and a further resource dependence risk seem to apply here. First, there is patently a ‘fallacy of composition’ issue. That is, the assumption that many can do what few previously did without changing the conditions or producing different (adverse) consequences than arose when only a few adopted that behaviour or activity. Those consequences are climatological and ecological. It remains the case that we are socialised to desire and appreciate cars and it remains a fact that private transport can be extremely convenient. It can also, given the commentary above, appear hypocritical to be suggesting shifting to a far greater reliance on public transport, since this implicitly involves denying to developing country citizens a facet of modernity enjoyed previously by developed country citizens. But this is a distraction from the underlying collective interest in reduced car ownership and use. It denies the basic premise that a Precautionary Principle applies to all and that societies that are not yet car dependent have the opportunity to avoid a problem, rather than have to manage it via either moving straight to private transport BEVs or a transition from fossil fuel-powered ICEs to BEVs with all that entails in terms of ingrained emissions. Policy may be mainly domestic, but climate change is global and aggregate effects do not respect borders, which brings us to a second concept or risk that may be exacerbated.

Second, a ‘quasi-Jevons’ effect’ may apply. Growth of vehicle use is a problem of resource use and this is a thermodynamic and emissions problem. However, it is, as we have noted, also the case that battery technology and energy mix for BEVs are improving. So, this may involve significant declines in relative cost, which in turn may create a tendency for BEV ownership to accelerate which could exacerbate net growth in numbers of vehicles. Net growth could ironically be to the detriment of emissions savings. Whether this is so, depends, in part, on what kind of overall transport policy countries adopt and whether consumers, corporations and markets are allowed to be the arbiter of which area of transport dominates. It also depends, in part, on what materials are required for future batteries. Current technology implies massive increases in costs based on securing sources of lithium and cobalt as battery demand rises. So even if a Jevons’ effect is avoided, a different issue may apply. Resource procurement is a Precautionary Principle issue since effective BEVs at the kind of numbers necessary to substitute for all vehicles seem to require technological transformation—without it, multiple problems apply whilst emissions remain ingrained.

For example, when the UK CCC announced its 2035 recommendation to accelerate the BEV transition, members of the Security of Supply of Mineral Resources (SSMR) project wrote a research note to the CCC (Webster, 2019). They pointed out that the current total European demand for cobalt is 19,800 tonnes and that producing the batteries to replace 2.3 million cars in the UK (in accordance with contemporary statistics for new registrations) would require 15,600 tonnes. The UK would also need 20,000 tonnes of lithium, which is 45% of the current total European demand. If we replicate this ramping up of demand across Europe and the globe for vehicles, recognising that there are other growing demands for the minerals and metals (including batteries for other purposes) then it seems unlikely that supply can respond, unless dependence on lithium and cobalt (and other constituents) falls sharply as technology changes. Clearly, the problem is also contingent on the uptake of BEVs. Over recent years, there has, in fact, been an oversupply of the main materials for battery production because several of the main mining corporations anticipated that battery demand would take off faster than it actually has. For example, global prices of cobalt, nickel and lithium carbonate have increased significantly over the last decade but have fallen in 2018 to the end of 2019. However, industry analysis indicates that current annual global production is the equivalent of about 10 million standard BEVs based on current technology, and as the previous statistics on global vehicle numbers (see also next section) indicate, this is far less than transition via substitution would seem to require in the next decade. 29

Shortages and price rises, therefore, are if not inevitable, at least likely. Currently, about 60% of the cost of a BEV is the battery and 80% of that 60% (about 50% of the vehicle) is the cost of battery materials. It is, therefore, important to achieve secure supply and stable costs. The further context here is the issue of UK domestic battery capacity. In 2013, the government created the Advanced Propulsion Centre (APC) with a 10 year £500 million investment commitment matched by industry. The APC’s remit is to address supply chain issues for electric vehicles. Not unexpectedly, the APC quickly identified lack of domestic battery production capacity as a major impediment. In response in 2016 another government initiative, Innovate UK set up the Faraday Battery Challenge to encourage domestic capacity and innovation. The Battery Industrialisation Centre was then set up in Coventry, to attract manufacturers in the supply chain for BEVs to locate there, focussed around a centre of research excellence. However, the APC has no control over the global supply and prices of battery materials, the investment and location decisions of battery manufacturers or the necessary infrastructure for BEVs to be a feasible technology. 30 For example, according to the APC, if domestic BEV demand were 500,00 per year by 2025, then the UK would need three ‘gigafactories’. Battery manufacture is currently dominated by LG Chem and Samsung in South Korea, CATL in China and Panasonic in Japan. None of these have current plans to build a gigafactory in the UK. In any case, there is a further problem here which raises a whole set of environmental and ethical issues explored in ecological circles under the general heading ‘extractivism’ (see, e.g. Dunlap, 2019 ). As time goes by, the UK and the world may become dependent on high price supplies of materials drawn from unstable or hostile regimes (the Democratic Republic of Congo, etc.), which is a risk in many ways (and a likely source of Dutch disease—the ‘resource curse’—for unstable regimes). So, not placing a relative emphasis on substituting BEVs for ICEs and not endorsing the current vehicle growth trend (which is different as a suggestion than rejecting BEVs entirely) avoid multiple problems and risks.

It is also worth noting that simple market decisions can have a further collective adverse consequence based on individual consumer preference and reasoning, which may also affect BEVs in the short term. Many current BEVs have smaller or low efficiency batteries and thus short ranges. These favour urban use for short journeys, but most people own cars with a view also to range further afield. As such, it seems likely that until the technology is all long range (and the charging infrastructure is pervasive) many consumers, if the choice exists and income allows, will own BEVs as an additional vehicle, not a replacement vehicle. 31 This may be a short-term issue, given the regulatory changes focussed from 2030 to 2040 in many countries. But, again, from a Paris point of view, taking the IPCC 1.5°C and UNEP Emissions Gap reports into consideration, this matters. This brings us to a final issue. What is the actual take-up of BEVs (and ULEVs)? How rapid is the transition? In the Introduction section, I suggested that the UK had reached a tipping point and that this mirrored a general trend globally. This, however, needs context.

3.4 How many electric vehicles?

The data emerging in recent years and stated in the Introduction section are a step-change, but as a possible tipping point it begins from a low base and BEVs (the least emitting of the low emission vehicles) are a subset, albeit a rapidly expanding one, of ULEVs. According to the UK Department for Transport statistical release for 2018, there were 200,000 ULEVs registered in total, of which 63,992 ULEVs were newly registered in that year ( Department for Transport, 2019 , p. 4). 93% of the total registrations were cars and the total constitutes a 39% increase on the year 2017 total and a 20% increase in the rate of registration—there were just 9,500 ULEVs at the beginning of 2010 (so, about 20 times greater in a decade). However, the 2018 data mean that ULEVs accounted for just 0.5% of all licensed vehicles and were still only 2.1% of all new registrations in that year. Preliminary data available early 2020 indicate continued growth with almost 38,000 new BEV registrations in 2019, a 144% year-on-year increase. As a recent UK House of Commons Briefing Paper notes, however, the government prefers to emphasise the percentage changes in take-up rather than the percentages of the absolute numbers or the absolute numbers themselves ( Hirst, 2019 ). The International Energy Agency (IEA) places the UK in its leading countries list by ULEV and BEV market share (measured by the percentage of total annual registration): Norway dominates, followed by Iceland, Sweden, the Netherlands and then a significant drop-off to a trailing group including China, the USA, Germany, the UK, Japan, France, Canada and South Korea. However, the market share in this trailing group is less than 5% in every case (see appended Figure 1 ). China, given its size (and because of the urgency of its urban air quality problems and its capacity for authoritarian implementation), dominates the raw numbers in terms of total ULEVs and BEVs. All this notwithstanding, the IEA confirms the general point that up-take is accelerating, but the base is low and so achieving total ULEV or BEV coverage is some way off:

The global electric car fleet exceeded 5.1 million in 2018, up by 2 million since 2017, almost doubling the unprecedented amount of new registrations in 2017. The People’s Republic of China… remained the world’s largest electric car market with nearly 1.1 million electric cars sold in 2018 and, with 2.3 million units, it accounted for almost half of the global electric car stock. Europe followed with 1.2 million electric cars and the United States with 1.1 million on the road by the end of 2018 and market growth of 385000 and 361000 electric cars from the previous year. Norway remained the global leader in terms of electric car market share at 46% of its new electric car sales in 2018, more than double the second-largest market share in Iceland at 17% and six-times higher than the third-highest Sweden at 8%. In 2018, electric buses continued to witness dynamic developments, with more than 460000 vehicles on the world’s road, almost 100000 more than in 2017…In freight transport, electric vehicles (EVs) were mostly deployed as light-commercial vehicles (LCVs), which reached 250000 units in 2018, up 80000 from 2017. Medium truck sales were in the range of 1000–2000 in 2018, mostly concentrated in China. ( IEA, 2019A , p. 9)

Over the next few years, it seems likely we will see rapid changes in these metrics. There is a great deal of discussion in policy analysis regarding bottlenecks and impediments and these, of course, are also important (consumer uncertainty, ‘range anxiety’, availability of sufficient infrastructure for charging and so on). 32 However, as everything argued so far indicates regarding transition and trends, underlying the whole is the conditionality of success and the potential for failure, involving avoidable ingrained emission and risks. There is a basic difference between a superior technology and a superior choice since the latter is a socio-economic matter of context: of rates of change, scales and substitutions. Ultimately, this creates deep concerns in terms of achieving the Paris goals. The IEA explores two forecast scenarios for the uptake of ULEVs. Both involve a projection of annual ULEV sales and total stock to 2030 ( IEA, 2019A ). First a ‘New Policies’ Scenario. This takes the current policy commitments of individual countries and extrapolates. By 2030, the scenario projects global ULEV sales at 23 million in that year and a total stock of 130 million. This is considerably less than 30% of all vehicles now and in 2030. Second, the EV30@30 Scenario. This assumes an accelerated commitment that adopts the @30 goals (notably 30% annual sales share for BEVs by 2030; IEA, 2019A , pp. 29–30). By 2030, the scenario projects global ULEV sales at 43 million in that year and a total stock of 250 million. Again, this is less than 30% of all vehicles now and in 2030.

The figures, of course, are highly conditional, but the point is clear, even the best-case scenario currently being anticipated has ULEVs and BEVs as a minority of all vehicles in 2030—and 2030 is a key year for achieving Paris, according to the October 2018 IPCC 1.5°C report. Moreover, it is notable that the projections assume continuous growth in the number of vehicles (and so continuous growth in ICE vehicles) and the major areas of numerical growth in BEVs continue to be China, so some significant part of the anticipated total will be new ingrained emissions that arguably did not need to exist. 33 Again, this is highly conditional but it at least creates questions regarding what is being ‘saved’ when the IEA claims that the New Policies Scenario results in 2.5 million barrels a day less demand for oil in 2030 and the EV30@30 Scenario 4.3 million barrels a day ( IEA, 2019A , p. 7). 34 Less of more is not a saving in an objective sense, if this is a preventable future, and it is not a rational way to set about ‘saving’ the planet. It remains the case, of course, that this is better than nothing, but it is deeply questionable whether in policy terms any of this is the ‘best that can be done’. As stated in the Introduction section, technocentrism distracts from appropriate recognition of this. At its worse, technocentrism fails to address and so works to reproduce a counter-productive ecological modernisation: the technological focus facilitates socio-economic trends, which are part of the broader problem rather than solutions to it. The important inference is that there are multiple reasons to think that greater emphasis on social redesign and less private transport avoids successful failure and is more in accordance with the Precautionary Principle.

I ended the introduction to this essay by stating that we would be exploring the foregrounding question: What kind of solution are BEVs to what kind of problem? It should be clearer now what was meant by this. Ultimately, the balance between private and public transport matters if the Paris goals are to be achieved. Equally clearly, this is not news to the UK CCC or to any serious analyst of electric vehicles and the transport issue for our climatological and ecological future (again, e.g. Chapman, 2007 ; Bailey and Wilson, 2009 ; Williamson et al. , 2018 ; Mattioli et al. , 2020 ). At the same time, the context and issues are not widely understood and the problems are often understated, at least in so far as, discursively, most weight is placed on stating progress in achieving a transition to ULEVs and BEVs. This is technocentric. Despite its general concerns and careful critical stance, the CCC is also partly guilty of this. For example, Ewa Kmietowicz, Transport Team Leader of the CCC Secretariat, refers to the UK Road to Zero strategy as a ‘lost opportunity’, and the CCC identifies a number of shortfalls in the strategy. 35 However, the general thrust of the CCC position is to focus on a rapid transition to BEVs and to overcoming bottlenecks. 36 Broader feasibility is subsumed under general assumptions about continued economic expansion and expansion of the transport system. So, there is more of a situation of complementarity (with caveats) between public and private transport, and the whole becomes an exercise in types of investment within expansionary trends, rather than a more radical recognition of the fundamental problems that we ought to think about avoiding. It is also worth noting that many of the major advocates of BEVs are industry organisations. The UK Society of Motor Manufacturers and Traders, for example, are not unconcerned but they are not impartial either; they have a vested interest in the vehicle industry and its growth. For industry, ULEVs and BEVs are an opportunity before they are a solution to a problem. There are, however, recognitions that a rethink is required. These range from direct activism, such as ‘Rocks in the Gearbox’ (along the lines of Extinction Rebellion), to analysis from establishment think tanks, such as the World Economic Forum 37 , and statements from government oversight committees. For example, the UK Commons Science and Technology Committee (CSTC) not only endorses the CCC 2035 accelerated BEV target but also states more explicitly:

In the long-term, widespread personal vehicle ownership does not appear to be compatible with significant decarbonisation. The Government should not aim to achieve emissions reductions simply by replacing existing vehicles with lower-emissions versions. Alongside the Government’s existing targets and policies, it must develop a strategy to stimulate a low-emissions transport system, with the metrics and targets to match. This should aim to reduce the number of vehicles required, for example by: promoting and improving public transport; reducing its cost relative to private transport; encouraging vehicle usership in place of ownership; and encouraging and supporting increased levels of walking and cycling. ( CSTC, 2019 )

This, as Caroline Lucas suggests, speaks to the need to coordinate public and private transport policy more effectively and clearly, and there is a need for broader informed debate here. In political ecological circles, for example, there is a growing critique of the tensions encapsulated in the concept of an ‘environmental state’ (see Koch, 2019 ). That is the coordination and coherence of environmental imperatives with other policy concerns. State-rescaling and degrowth and postgrowth work highlight the profound problems that are now starting to emerge as states come to terms with the basic mechanisms that have been built into our economies and societies (see also Newell and Mulvaney, 2013 ; Newell, 2019 ). 38 New thinking is required and this extends to the social ontology and theory we use to conceptualise economies (see Spash and Ryan, 2012 ; Lawson, 2012 , 2019 ) and political formations (see Bacevic, 2019 ; Patomäki, 2019. Covid-19 does not change this ( Gills, 2020 ).

In transport terms, there are many specific issues to consider. Some solutions are simple but overlooked because we are always thinking in terms of sophisticated innovations and inventions. However, we do not need to conform to the logics of ‘technological fixes’, that we somehow think will enable the impossible, to perhaps see some scope in ‘fourth industrial revolution’ transformations ( Center for Global Policy Solutions, 2017 ; Morgan, 2019B ). For example, public transport may also extend to a future where no individual owns a range extensive powered vehicle (perhaps just local scooters for the young and mobility scooters for the infirm) and instead a system operates of autonomous fleet vehicles that are coordinated by artificial intelligence with logistics implemented through Smartphone calendar access booking systems—and coordination functions could maximise sharing, where vehicles could also be (given no drivers are involved) adaptable connective pods that chain together to minimise congestion and energy use. This seems like science fiction now, and perhaps a little ridiculous, but a few years ago so did the Smartphone. And the technology already exists in infancy. Such a system could be either state-funded and run or private partnership and franchise, but in either case, it radically redraws the transport environment whilst working in conformity with the geography of living spaces we have already developed. Will is what is required and if the outcome of COP24 ( UNFCCC, 2018 ) and COP25 ( Newell and Taylor, 2020 ) with limited progress towards the Paris goals persists, then it seems likely that emissions will accumulate rapidly in the near future and the likelihood of a serious climate event with socio-economic consequences rises. At that stage, more invasive statutory and regulatory intervention may start to occur as the carbon budget becomes a more urgent target. Prohibitions, transport rationing and various other possibilities may then be on the agenda if we are to unmake the future we are currently writing and, to mix metaphors, avoid a road to nowhere.

None declared

Thanks to two anonymous reviewers for extensive and useful comment—particularly regarding the systematic statement of issues in the Introduction section and for additional useful references. Jamie Morganis Professor of Economic Sociology at Leeds Beckett University, UK. He coedits the Real-World Economics Review with Edward Fullbrook. RWER is the world’s largest subscription based open access economics journal. He has published widely in the fields of economics, political economy, philosophy, sociology, and international politics. His recent books include: Modern Monetary Theory and its Critics (ed. with E. Fullbrook, WEA Books, 2020), Economics and the ecosystem (ed. with E. Fullbrook, WEA Books, 2019); Brexit and the political economy of fragmentation: Things fall apart (ed. with H. Patomäki, Routledge, 2018); Realist responses to post-human society (ed. with I. Al-Amoudi, Routledge, 2018); Trumponomics: Causes and consequences (ed. with E. Fullbrook, College Publications, 2017); What is neoclassical economics? (ed., Routledge, 2015); and Piketty’s capital in the twenty-first century (ed. with E. Fullbrook, College Publications, 2014).

Bacevic , J . 2019 . Knowing neoliberalism , Social Epistemology , vol. 33 , no. 4 , 380 – 92

Google Scholar

Bailey , I. and Wilson , G . 2009 . Theorising transitional pathways in response to climate change: technocentrism, ecocentrism, and the carbon economy , Environment and Planning A , vol. 41 , no. 10 , 2324 – 41

CCC. 2019 . Net Zero: The UK’s Contribution to Stopping Global Warming , London , Author

Google Preview

Center for Global Policy Solutions. 2017 . Stick Shift: Autonomous Vehicles, Driving Jobs and the Future of Work , Washington DC , Author

Chapman , L . 2007 . Transportation and climate change: a review , Journal of Transport Geography , vol. 15 , no. 5 , 354 – 67

China Daily. 2018, December 1 . China has 325 million motor vehicles , China Daily

Christensen , J. and Olhoff , A . 2019 . Lessons from a Decade of Emissions Gap Assessments , Nairobi , UNEP

CSTC. 2019 . Clean Growth: Technologies for Meeting the UK’s Emissions Reduction Targets , London , Author

DBEIS. 2018 . Annex: 1990–2016 UK Greenhouse Gas Emissions, Final Figures by end User , London , Author

DECC. 2015 . Different Approaches to Reporting UK Greenhouse Gas Emissions , London , Author

Deloitte. 2018 . Battery Electric Vehicles. New market. New entrants. New challenges , London , Author

Dennis , K. and Urry , J . 2009 . After the Car , Cambridge , Polity

Department for Transport. 2018 . The Road to Zero: Next Steps Towards Cleaner Road Transport and Delivering our Industrial Strategy , London , Author

Department for Transport. 2019 . Vehicle Licensing Statistics: Annual 2018 , London , Author

Dunlap , A . 2019 . Wind, coal and copper: the politics of land grabbing, counterinsurgency and the social engineering of extraction , Globalizations vol. 17 , no. 4 , 661 – 82

Environmental Audit Committee. 2016 . Sustainability in the Department of Transport [Third Report of Sessions 2016–17], London , House of Commons

Faraday Institution. 2019 . UK Electric Vehicle and Battery Production Potential to 2040 , London , Author

Faulkner , P. and Runde , J . 2013 . Technological objects, social positions and the transformational model of social activity , MIS Quarterly , vol. 37 , no. 3 , 803 – 18

Frost & Sullivan. 2019 . Global Electric Vehicle Market Outlook, 2019 , Author , London

Fullbrook , E. and Morgan , J. (eds.). 2019 . Economics and the Ecosystem , London , World Economic Association Books

Georgescu-Roegen , N . 1971 . The Entropy Law and the Economic Process , Cambridge and London , Harvard University Press

Gills , B . 2020 . Deep Restoration: from the Great Implosion to the Great Awakening , Globalizations , vol. 17 , no. 4 , 577 – 9

Gills , B. and Morgan , J . 2019 . Global Climate Emergency: after COP24, climate science, urgency, and the threat to humanity , Globalizations

Gills , B. , Morgan , J. and Patomäki , H . 2019 . President Trump as status dysfunction , Organization , vol. 26 , no. 2 , 291 – 301

Hansen , J. et al.  2017 . Young people’s burden: requirement of negative CO2 emissions , Earth System Dynamics , vol. 8 , 577 – 616

Harvey , F . 2020, March 28 . Tackle climate crisis and poverty with zeal of Covid-19 fight scientists urge , The Guardian

Hirst , D . 2019 . ‘ Electric Vehicles and Infrastructure’ , Briefing Paper no. CBP07480, London , House of Commons Library

Hoekstra , A . 2019 . The underestimated potential of battery electric vehicles to reduce emissions , Joule , vol. 3 , no. 6 , 1412 – 4

Holt , R. , Pressman , S. and Spash , C. (eds.). 2009 . Post Keynesian and Ecological Economics , Cheltenham , Edward Elgar

Holtsmark , B. and Skonhoft , A . 2014 . The Norwegian support and subsidy policy for electric cars. Should it be adopted by other countries? Environmental Science & Policy , vol. 42 , 160 – 8

IEA. 2019A . Global EV Outlook 2019: Scaling up the Transition to Electric Mobility , Paris , Author

IEA. 2019B . World Energy Outlook , Paris , Author

IPCC. 2018 . Global Warming of 1.50C: Summary for Policymakers , Geneva , Author

IPCC. 2019A . IPCC Special Report on Climate Change, Desertification, Land Degradation Sustainable Land Management Food Security and Greenhouse Gas fluxes in Terrestrial Ecosystems , Geneva , Author

IPCC. 2019B . The Ocean and Cryosphere in a Changing Climate , Geneva , Author

Koch , M . 2019 . The state in the transformation to a sustainable postgrowth economy , Environmental Politics , vol. 29 , no. 1 , 115 – 33

Lawson , C . 2012 . Aviation lock-in and emissions trading , Cambridge Journal of Economics , vol. 36 , no. 5 , 1221 – 43

Lawson , C . 2017 . Technology and Isolation , Cambridge , Cambridge University Press

Lawson , T . 2019 . The Nature of Social Reality: Issues in Social Ontology , London , Routledge

Lea , R . 2019, September 6 . Tesla Model 3 enters sales chart at No3 , The Times

Lenton , T. , Rockstrom , J. , Gaffney , O. , Rahmstorf , S. , Richardson , K. , Steffen , W. and Schellnuber , H . 2018 . Climate tipping points too risky to bet against , Nature , vol. 575 , 592 – 5

Manzetti , S. and Mariasiu , F . 2015 . Electric vehicle battery technologies: from present state to future systems , Renewable and Sustainable Energy Reviews , vol. 51 , 1004 – 12

Mattioli , G. , Roberts , C. , Steinberger , J. , and Brown , A . 2020 . The political economy of car dependence: a systems of provision approach , Energy Research & Social Science , vol. 66 , 1 – 18

Morgan , J . 2016 . Paris COP21: power that speaks the truth? Globalizations , vol. 13 , no. 6 , 943 – 51

Morgan , J . 2017 . Piketty and the growth dilemma revisited in the context of ecological economics , Ecological Economics , vol. 136 , 169 – 77

Morgan , J . 2019A . Intervention, policy and responsibility: economics as over-engineered expertise?, pp. 145 – 63 in Dolfsma , W. and Negru , I. (eds.), 2019 The Ethical Formation of Economists , London , Routledge

Morgan , J . 2019B . Will we work in twenty-first century capitalism? A critique of the fourth industrial revolution literature , Economy and Society , vol. 48 , no. 3 , 371 – 98

Morgan , J. and Patomäki , H . 2017 . Contrast explanation in economics: its context, meaning, and potential , Cambridge Journal of Economics , vol. 41 , no. 5 , 1391 – 418

Nasir , A. and Morgan , J . 2018 . The unit root problem: affinities between ergodicity and stationarity, its practical contradictions for central bank policy, and some consideration of alternatives , Journal of Post Keynesian Economics , vol. 41 , no. 3 , 339 – 63

National Audit Office. 2019 . Department of Transport Sustainability Update , London , Author

Newell , P . 2019 . Transformismo or transformation? The global political economy of energy transitions , Review of International Political Economy , vol. 26 , no. 1 , 25 – 48

Newell , P. and Mulvaney , D . 2013 . The political economy of the “just transition” , The Geographical Journal , vol. 179 , no. 2 , 132 – 40

Newell , P. and Taylor , O . 2020 . Fiddling while the planet burns? COP 25 in perspective , Globalizations , vol. 17 , no. 4 , 580 – 92

ONS. 2019 . UK Environmental Accounts: 2019 , London , Author

Patomäki , H . 2019, July 9 . ‘The Climate Movement? What’s Next?’, available at https://patomaki.fi/en/2019/07/the-climate-movement-whats-next/

Petit , S . 2017 . World vehicle population rose 2.6% in 2016 , Ward Intelligence

Regett , A. , Mauch , W. and Wagner , U . 2019 . Carbon Footprint of Electric Vehicles – A Plea for More Objectivity , available at https://www.ffe.de/attachments/article/856/Carbon_footprint_EV_FfE.pdf

Ripple , W. , Wolf , C. , Newsome , T. , Barnbard , P. , Moomaw , W. and 11,258 signatories. 2019 . World Scientists’ warning of a climate emergency , BioScience , vol. 70 , no. 1 , 8 – 12

Romare , M. and Dahlöff , L . 2017 . ‘ The Life Cycle Energy Consumption and Greenhouse gas Emissions From Lithium-Ion Batteries’ , Report No. C243, Stockholm , IVL Swedish Environmental Research Institute

Royal Academy of Engineering. 2010 . Electric Vehicles: Charged with Potential , London , Author

Servigne , P. and Stevens , R . 2015 . Comment tout peut s’efondrer , Paris , Science Humaines , available at https://www.seuil.com/ouvrage/comment-tout-peut-s-effondrer-pablo-servigne/ 9782021223316

Smil , V . 2017A . Energy and Civilization , Boston , MIT Press

Smil , V . 2017B . Energy Transitions , Colorado , Praeger

SMMT. 2019 . UK Electric Car Registrations Surge in August but it’s a Long Road to Zero and Barriers Must be Addressed , Press Release September 5th, London , Author

Spash , C. (ed.). 2017 . Routledge Handbook of Ecological Economics: Nature and Society , New York , Routledge

Spash , C . 2020 . A tale of three paradigms: realising the revolutionary potential of ecological economics , Ecological Economics vol. 169 ,

Spash , C. and Ryan , A . 2012 . Economic schools of thought on the environment: investigating unity and division , Cambridge Journal of Economics , vol. 36 , no. 5 , 1091 – 121

Staffell , I. Green , R. Gross , R. and Green , T . 2019 . How clean is my car? , Electric Insights Quarterly , vol. Q2 , 7 – 10

Statista. 2019A . Number of Motor Vehicles Registered in the United States from 1990 to 2017 (in 1000s) , [data updated June 2019], available at https://www.statista.com/statistics/183505/number-of-vehicles-in-the-united-states-since-1990/

Statista. 2019B . Car “Parc” in China from 2007 to 2018 (millions) , [data updated August 2019], available at https://www.statista.com/statistics/285306/number-of-car-owners-in-china/

Steffen , W. et al.  2018 . Trajectories of the Earth System in the Anthropocene , Proceedings of the National Academy of Sciences of the USA , vol. 115 , 8252 – 9

Taylor , M . 2015 . The Political Ecology of Climate Change Adaptation , London , Routledge/Earthscan

UNEP. 2012 . Global Environmental Outlook Report 5: Environment for the Future We Want , New York , Author

UNEP. 2018 . Emissions Gap Report 2018 , 9th ed., New York , Author

UNEP. 2019A . Emissions Gap Report 2019 , 10th ed., New York , Author .

UNEP. 2019B . Global Environmental Outlook Report 6: Healthy Planet Healthy People , New York , Author

UNFCCC. 1992 . United Nations Framework Convention on Climate Change , New York , Author

UNFCCC. 2015 . Adoption of the Paris Agreement and Annex: Paris Agreement , Paris , Author

UNFCCC. 2018 . Katowice Texts , Katowice , Author

Webster , B . 2019, June 6 . Britain “could be held to ransom” on electric cars , The Times

Williamson , K. , Satre-Meloy , A. , Velasco , K. and Green , K . 2018 . Climate Change Needs Behaviour Change: Making the Case for Behavioural Solutions to Reduce Global Warming , Arlington, VA , Centre for Behaviour and the Environment

Zheng , S . 2017, April 19 . China now has over 300 million vehicles… that’s almost America’s total population , South China Morning Post

Global electric car sales and market share, 2013–18.

Global electric car sales and market share, 2013–18.

Source : IEA (2019, p. 10).

ULEV refers to vehicles that emit less than 75 gCO 2 per km. This essentially means BEVs, PHEVs, range-extended (typically an auxiliary fuel tank) electric vehicles, fuel cell (non-plug-in) electric vehicles and hybrid models (non-plug in vehicles with a main fuel tank but whose battery recharges and which drive short distances in electric mode).

Note, there is little sign of legislative and regulatory detail to plans as of early 2020. Furthermore, there is a difference between acknowledging that the uptake of alternatively fuelled vehicles, including BEVs, is growing and drawing the inference that UK government policy (channelled primarily via the Department for Transport) is as effective as it might be (see Environmental Audit Committee, 2016 ; National Audit Office, 2019 and also later discussions).

CEM is coordinated by the IEA and is an initiative lead by Canada and China (but including a steadily growing number of signatory countries). The EV30@30 initiative aims to achieve a 30% annual sales share for BEVs by 2030.

IEA headline statistics include plug-in hybrids so 2018 becomes 46% for Norway (IEA, 2019A, p. 10).

For example, Spash (2020) and Spash and Ryan (2012) . One might also note the work of John O’Neill at Manchester University. Perhaps the most prominent ‘realist’ working on transport and ecology is Petter Naess, at Norwegian University of Life Sciences.

The UNEP 9th Report calls for a 55% reduction by 2030.

The initial rationale in 2008 was that to achieve a maximum limit of 2°C warming global emissions needed to fall from the levels at that time to 20–24 GtCO 2 e with an implied average of 2.1–2.6 t CO 2 per capita on a global basis in 2050. This translated to a 50–60% reduction to the then global total. Since UK emissions were above average per capita, the UK reduction required was estimated at about 80%. Given that emissions then increased and atmospheric ppm has risen the original calculations are now mainly redundant.

For the work of the CCC, see: https://www.theccc.org.uk/about/ .

The report also provides useful context regarding the UN sustainable development goals ( CCC, 2019 : p. 66) and CCC thinking on growth and economics ( CCC, 2019 : pp. 46–7).

https://www.theccc.org.uk/2019/06/11/response-to-government-plan-to-legislate-for-net-zero-emissions-target/ .

And further methodological issues apply in economics (see; Morgan and Patomäki, 2017 ; Nasir and Morgan, 2018 ; Morgan, 2019A ).

For a full analysis, see https://www.carbonbrief.org/analysis-uks-co2-emissions-have-fallen-29-per-cent-over-the-past-decade . The Carbon Brief analysis omits shipping and aviation. As the campaign group Transport and Environment notes UK shipping was responsible for 14.4 MtCO 2 , which is the third highest in Europe (after the Netherlands and Spain) and shipping is exempt from tax on fossil fuels under EU law. See p. 20: https://www.transportenvironment.org/sites/te/files/publications/Study-EU_shippings_climate_record_20191209_final.pdf .

UK coal use for energy supply reduced by approximately 90% from 1990 to 2017 and in 2019 amounted to just 2% of the energy mix and in 2019 the UK went two weeks without using any coal at all for power production (the first time since 1882); 1990 to 2010 natural gas use steadily increased from a near-zero base but has declined since 2010 as use of renewables has grown. Coal use in manufacturing has decreased by 75% from 1990 to 2017 ( ONS, 2019 ). As noted, some assessments place the reduction in total emissions at around 40% based on other metrics and the tabulated figures I provide indicate yet another percentage— all however are trend decreases indicative of a general direction of travel.

‘Embedded emissions’ or the UK carbon footprint is addressed by the UK Department for Environment Food and Rural Affairs (Defra). To be clear, there is a whole set of further issues that one might address in regard of measurement of emissions—how they are attributed and what this means (where created, where induced through demand, which state, what corporation and so different ‘Cartesian’ claims regarding the significance of location are possible), and this is indicative of the conflict over representation and partition of responsibility (so whilst the climate does not care about borders, they have infected measurement and policy). There is no scientifically neutral way to achieve this, merely different sets of criteria with different consequences (I thank an anonymous referee for extended comment on this, see also Taylor, 2015 ; who argues that adaptation politics produces a focus on governance within existing political and economic structures based on borders, etc.).

Congestion charges in London or a plastic bag tax do not meet this threshold.

This is supported, for example, by The Climate Group’s EV100 initiative: a voluntary scheme where corporations commit to making electric the ‘new normal’ of their vehicle fleets by 2030 (recognising that over half of annual new registrations are owned by businesses) https://www.theclimategroup.org/project/ev100 .

Until recently Tesla had one main production centre in California. However, it now also has a $5 billion factory in Shanghai and plans for a factory in Berlin. Tesla is currently the world’s largest producer of BEVs (368,000 units in 2019), followed by the Chinese company BYD Auto (195,000 units in 2019). Tesla was founded in July 2003 by Martin Eberhard and Mark Tarpenning in response to General Motors scrapping its EV programme (as unprofitable). Elon Musk joined as a HNWI first-round investor in February 2004 (he put in $6.5 m of the total $7.5 m and became chairman of the Tesla board); Eberhard was initially CEO but was removed and replaced by Musk in 2007 and Tarpenning left in 2008. Tesla floated on the Nasdaq in June 2010 at $17 per share and exceeded $500 per share for the first time in January 2020. Tesla is the USA’s most valuable car manufacturer by market capitalisation (worth more than Ford and GM combined).

The European Commission’s collaborative research forum JEC has been producing ‘well-to-wheels’ analyses of energy efficiency of different engine technologies since the beginning of the century. The USA periodically publishes the findings of its GREET model (the Greenhouse gases Regulated Emissions and Energy use in Transportation model). See https://greet.es.anl.gov .

For example, since 1985 according to Carbon Brief global coal use in power production measured in terawatt hours only reduced in 2009 and 2015 (though it seems likely to do so in 2019); China notably continues to build coal-fired power plants though the rate of growth of use has slowed. (According to the IEA Coal report, 2019, China consumed 3,756 million tonnes of coal in 2018 (a 1% increase) and India 986 million tonnes (a 5% increase). Renewables are a growing part of an expanding global energy system.

https://www.carbonbrief.org/analysis-global-coal-power-set-for-record-fall-in-2019 .

Staffell et al . observe that the British electricity grid produces an average 204 gCO 2 per kWh in 2019 and a standard petrol car emits 120–160 gCO 2 per km.

This is a point made by Richard Smith. There are, of course, alternatives to aluminium. One should also note that manufacturers are responding to consumer preference by increasing the average size of models and this is increasing the weight and resource use. In February 2020, for example, Which Magazine analysed 292 popular car models and found that they were on average 3.4% or 67 kg heavier than older models and this was offsetting some of the efficiency gains for emissions.

And the argument this is leading to is that it makes far greater sense to default to greater dependence on prudential social redesign, rather than optimistic technocentrism, behind which is techno-politics.

For discussion of battery technology and scope for improvement, see Manzetti and Mariasiu (2015) and Faraday Institution (2019) . Currently, most BEVs use lithium-ion phosphate, nickel-manganese cobalt oxide or aluminium oxide batteries. Liquid electrolyte constituents require containment and shielding. Specifically, a battery creates a flow of electrons from the positive electrode (the cathode made of a lithium metal oxide, etc. from the previous list) through a conducting electrolyte medium (lithium salt in an organic solution) to a negative electrode (the anode made typically of carbon, since early experiment with metals tended to produce excess heating and fire). This creates a current. Charging flows to the anode and discharge oxidises the anode which must then be recharged. The batteries are relatively low ‘energy density’ and can be a fire hazard when they heat. Given the chemical constituents, battery disposal is also a significant environmental hazard (see IEA, 2019A: pp. 8, 22–3). A ‘solid-state’ battery uses a specially designed (possibly glass or ceramic) solid medium that allows ions to travel through from one electrode to another. The solid-state technology is in principle higher energy density, much lighter and more durable. The implication is higher kWh batteries with greater range, charging capacity and durability and efficiency. Jeremy Dyson has reportedly invested heavily in solid-state technology and though his proposed own brand BEV is not now going ahead, reports indicate the battery technology investment will continue.

One might also consider hydrogen battery technology. Hydrogen fuel cell technology for vehicles is different than BEV. The vehicle has a tank in the rear for compressed cooled gas, which supplies the cell at the front of the car whilst driving. Refuelling is a rapid pumping process rather than a long wait. The gas has two possible origins: natural gas conversion where ‘steam methane reformation’ separates methane into hydrogen and CO 2 or water electrolysis, where grid AC electricity is converted to DC, which is applied to water and using a membrane splits it into hydrogen and waste oxygen. Currently, over 95% of hydrogen is from the former. Major investors in hydrogen technology are Shell (for natural gas conversion), IMT Power (in partnership with Shell) for water conversion and Toyota whose Mirai model is hydrogen powered.

Though fewer new cars were registered than in previous years, this significant metric for the total number of vehicles is the cumulative number of registrations (taking into account cars no longer registered). There are, however, some underlying issues: uncertainty regarding the status of diesel cars and problems of availability, cost and trust in BEVs seems to be causing many people in the UK to delay buying a new car; the expansion of Uber meanwhile has had a generational and urban effect, reducing car ownership as an aspiration amongst the young.

And re aviation, a new runway at Heathrow between 2026 and 2050.

See: https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attachment_data/file/852708/provisional-road-traffic-estimates-gb-october-2018-to-september-2019.pdf .

See: https://greenworld.org.uk/article/budget-deeply-disappointing-says-caroline-lucas

For example, global production of cobalt in 2018 was 120,000 tonnes, and production of about 2 million BEVs currently requires around 25,000 tonnes, so 10 million BEVs would require all of the current output. Cobalt traded at more than US$90,000 per ton 2018 but had fallen to around US$30,000 at the end of 2019.

In the UK, the current daily consumption of petrol and diesel for road transport is about 125 million litres or about 45 billion litres per year. So, BEVs are essentially substituting for this scale of energy use, shifting demand to electricity generation. National Grid attempted to model this in 2017. Their forecast (highly contingent obviously) suggests that if all cars sold by 2040 were BEVs and thus the car market was dominated by BEVs by 2050 and if most vehicles were charged at peak times in 2050 then an additional 30 gigawatts of electricity would be required. This is about 50% greater than the current peak winter demand in 2017. This was widely reported in the press. This best/worst case, of course, does not allow for innovative solutions such as off-peak home charging pioneered by Ovo and other niche suppliers. However, even with such solutions, there will still be a net increase in required capacity from the system. This has been estimated at about 10 new Hinckley power stations.

One possible long-term solution currently in development is toughened solar panel devices that can be laid as a road or car park surfaces, enabling contact recharging of the vehicle (in motion or otherwise). There are, however, multiple problems with the technology so far.

For example, analysis from Capital Economics suggests a three-way charging split is likely to develop: home recharging is likely to dominate, followed by an on-route charging model (substituting for current petrol forecourts at roadside) and destination recharging (given charging is slower than filling a fuel tank it makes sense to transform car parks at destinations into charging centres—supermarkets, etc.). They estimate UK demand at 25 million BEV chargers by 2050 of which all but 2.6 million will be home charging. As of early 2020, there were 8,400 filling stations which might be fully converted. Tesco has a reported commitment to install 2,400 charging points. These are issues frequently reported in the press.

This point can also be made in other ways. Not only does the emissions saving relate to net new sources of cars, but the contrast is also in terms of trend changes in the size of vehicle. According to the recent IEA World Energy Outlook report ( IEA, 2019B ), the number of SUVs is increasing and these consume around 25% more fuel than a mid-range car. If current growth trends continue (SUVs are 42% of new sales in China, 30% in India and about 50% in the USA), the IEA projects that the take-up of ICE SUVs will more than offset any marginal gains in emissions from the transition to BEVs.

It is also the case that the projected ‘savings’ from ULEVs are likely inaccurate. Following the EU, most countries adopted (and manufacturers report using) the Worldwide Harmonised Light Vehicle Test Procedure (WLTP). This became mandatory in the UK from September 2018. The WLTP is the new laboratory defined test for car distance-energy metrics. Vehicles are tested at 23°C, but without associated use of A/C or heating. Though claimed to as realistic than its predecessors, it is still basically unrealistic. Temperature range for ULEVs has significant consequences for battery performance and for use of on-board services, so real distance travelled per unit of energy is liable to be less. For similar reasons, ICEs will also travel less distance per litre of fuel so this is not a comparative gain for ICEs, it is likely a comparative loss to all of us if we rely on the figures.

See https://www.theccc.org.uk/2018/07/10/road-to-zero-a-missed-opportunity/ .

See https://www.theccc.org.uk/2018/07/10/governments-road-to-zero-strategy-falls-short-ccc-says/ .

See https://www.weforum.org/agenda/2019/08/shared-avs-could-save-the-world-private-avs-could-ruin-it/ .

For practical network initiatives, see, for example, https://climatestrategies.org .

Email alerts

Citing articles via.

  • Contact Cambridge Political Economy Society
  • Recommend to your Library

Affiliations

  • Online ISSN 1464-3545
  • Print ISSN 0309-166X
  • Copyright © 2024 Cambridge Political Economy Society
  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Institutional account management
  • Rights and permissions
  • Get help with access
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

The current research on electric vehicle

Ieee account.

  • Change Username/Password
  • Update Address

Purchase Details

  • Payment Options
  • Order History
  • View Purchased Documents

Profile Information

  • Communications Preferences
  • Profession and Education
  • Technical Interests
  • US & Canada: +1 800 678 4333
  • Worldwide: +1 732 981 0060
  • Contact & Support
  • About IEEE Xplore
  • Accessibility
  • Terms of Use
  • Nondiscrimination Policy
  • Privacy & Opting Out of Cookies

A not-for-profit organization, IEEE is the world's largest technical professional organization dedicated to advancing technology for the benefit of humanity. © Copyright 2024 IEEE - All rights reserved. Use of this web site signifies your agreement to the terms and conditions.

Advertisement

Advertisement

Electric Vehicles Charging Technology Review and Optimal Size Estimation

  • Original Article
  • Open access
  • Published: 02 October 2020
  • Volume 15 , pages 2539–2552, ( 2020 )

Cite this article

You have full access to this open access article

  • Morris Brenna 1 ,
  • Federica Foiadelli 1 ,
  • Carola Leone   ORCID: orcid.org/0000-0002-4399-3052 1 &
  • Michela Longo 1  

77k Accesses

135 Citations

3 Altmetric

Explore all metrics

Many different types of electric vehicle (EV) charging technologies are described in literature and implemented in practical applications. This paper presents an overview of the existing and proposed EV charging technologies in terms of converter topologies, power levels, power flow directions and charging control strategies. An overview of the main charging methods is presented as well, particularly the goal is to highlight an effective and fast charging technique for lithium ions batteries concerning prolonging cell cycle life and retaining high charging efficiency. Once presented the main important aspects of charging technologies and strategies, in the last part of this paper, through the use of genetic algorithm, the optimal size of the charging systems is estimated and, on the base of a sensitive analysis, the possible future trends in this field are finally valued.

Similar content being viewed by others

research paper on electric vehicles pdf

A comprehensive review on hybrid electric vehicles: architectures and components

Krishna Veer Singh, Hari Om Bansal & Dheerendra Singh

research paper on electric vehicles pdf

Life cycle assessment of electric vehicles: a systematic review of literature

Pabitra Kumar Das, Mohammad Younus Bhat & Shambhu Sajith

research paper on electric vehicles pdf

Applications of Lithium-Ion Batteries in Grid-Scale Energy Storage Systems

Tianmei Chen, Yi Jin, … Qiang Chen

Avoid common mistakes on your manuscript.

1 Introduction

Growing concern of carbon dioxide emissions, greenhouse effects and rapid depletion of fossil fuels raise the necessity to produce and adopt new eco-friendly sustainable alternatives to the internal combustion engine (ICE) driven vehicles. For this reason, in the last decade, EVs have become in some way widespread, principally because of their negligible flue gas emissions and lesser reliance on oil. It is estimated that by 2022, EVs will be over 35 million in the World. However, a critical problem associated with EVs is that their high penetration causes significant issues on the power distribution grid such as: power quality deterioration, enhanced damaged of line, downturn of distribution transformers, increased distortion and higher fault current [ 1 , 2 ]. One efficient approach to relieve the effect is to integrate local power generation such as renewable energy sources (RESs) into the EV charging infrastructure [ 3 , 4 , 5 ].

Batteries can be charged through conductive or inductive methods [ 6 ]. Inductive charger are wireless charging systems (WCS). WCS can be stationary, which means that they can only be utilized when the car is parked or in stationary modes, such as in car parks, garages, or at traffic signals, or they can be dynamic. This latter method allows battery charging while the vehicle is in motion. In general, WCS can bring some advantages in the form of aesthetic quality, reliability, durability and user friendliness. Anyway, due to some challenges such as electromagnetic compatibility (EMC) issues, limited power transfer, bulky and expensive structures, shorter range, and lower efficiency [ 7 , 8 ], inductive chargers are not largely commercialized and employed as the conductive ones. In this paper only conductive charging strategies are analyzed.

Based on their power ratings EV battery chargers can be divided into level 1, level 2 and level 3. Table 1 summarize the characteristics of the three different power levels [ 9 , 10 ].

A battery charger can allow a unidirectional or bidirectional power flow at all power levels. The bidirectional power flow adds to the grid-to-vehicle interaction (G2V) also the vehicle-to-grid (V2G) mode [ 11 ]. This latter technology can bring significant improvement in the overall reliability of the distribution grid, since in case of system failure, peak load demand or other unexpected scenarios, with a bidirectional power flow, the EVs can be used as back up generation, supplying the energy back to the grid when needed [ 12 ]. With V2G, as all the energy storage systems, EVs battery can be used not only as back up resource but also to improve the power quality, the stability and the operating cost of distribution network. Moreover, in the long run, V2G could reduce investment in new power generation infrastructure [ 13 , 14 , 15 , 16 ]. All the just listed reasons are increasing researchers’ interest in this technology.

Not only the choice of the charging technology, but also the selection of the correct charging method is a feature that has to be considered during the charging procedure. The most popular charging strategies to recharge Li-ion batteries are constant-current/constant-voltage (CC/CV) and pulse current charging methods [ 17 , 18 ]. However these methods do not take into account the several internal process of the battery which influence its charging capability and aging. As a result, some promising charging strategies which are based on more complete models of lithium-ion batteries are now under research [ 19 ].

EV battery and power electronic devices costs are steadily falling [ 20 , 21 ]. This rapid decline is mostly due to a growing manufacturing industry, which is constantly increasing the knowledge, the number of applications and the improvements of both these technologies. As the costs are falling, the trends of EV battery energy density, gross weight and semiconductor devices performances are following exactly the opposite direction; in fact, with equal capacities, batteries are becoming ever smaller and lighter and the power electronic devices ever more performing [ 22 ]. All this impacts the choice and the size of the charging systems. The sizing procedure of a suitable charging system is made even more difficult by the presence of many different technologies of the onboard and off-board chargers and also different cost, dimensions, weights, power rating; and so on.

This the paper is organized as follow. In Sects. 2 and 3 , respectively, on-board and off-board charger most common architectures are presented and their operating principles are explained. In Sect. 4 , the concept of fast charging stations is introduced. The available and most suitable charging methods are listed in Sect. 5 with particular attention to the charging methods most suitable for direct-current (DC) fast charging. Finally, a genetic algorithm is used in Sect. 6 to estimate the optimal charging system size and its possible future trends.

2 Onboard Charger

Battery chargers can be implemented inside (on-board) or outside (off-board) the vehicle. Onboard battery chargers (OBC) are limited by size, weight and volume [ 23 ] for this reason they are usually compatible with level 1 and level 2 chargers. They usually have unidirectional power transfer capability; nevertheless in some case the configuration, a bidirectional power transfer can be achieved. Figure  1 shows the typical architecture of an electric vehicle charging system, in such figure both the on-board charger and the off-board one are represented.

figure 1

Charging system configuration for electric vehicle

2.1 Two Stage

Onboard chargers are typically composed by two stages: a front-end AC–DC stage and a back-end DC-DC stage. Very different topologies are proposed in literature for both the converters.

The front-end rectifier usually contains a boost power factor correction (PFC) converter to achieve high power factor and low harmonic distortion. The rectifier stage can be performed by a half-bridge, full-bridge or multilevel diode bridge. Half-bridge rectifier is less expensive since it contains less number of diodes/switches, full-bridge rectifier is more complex but the components are subjected to lower stresses. If instead higher power ratings should be achieved a good choice for the ac–dc converter is a multilevel configuration. Figure  2 shows a full-bridge diode rectifier with a conventional PFC boost converter [ 24 , 25 ]. By substituting all the diodes with active switches, a bidirectional power flow can be obtained.

figure 2

Full-bridge rectifier with conventional PFC boost converter

In PFC converter the interleaved boost converter is becoming more and more popular. As shown in Fig.  3 , an interleaved boost converter simply consists in two boost converters in parallel, operating 180° out of phase [ 26 ]. The main aim of this interleaving is to increase the output current by reducing the input current ripple and hence by reducing the overall volume of the input ElectroMagnetic Interference (EMI) filter and of the boost inductor [ 23 , 27 , 28 , 29 ]. On the other hand, interleaving means increasing the cost and the complexity of the design.

figure 3

Full-bridge rectifier with interleaved PFC boost converter

The front-end converter is followed by a second dc-dc converter. Resonant power converters are very common to perform this second stage, because of the potential to achieve at the same time both higher switching frequency and lower switching losses [ 29 ]. Among all the resonant converter, LLC configuration, shown in Fig.  4 , is achieving resounding interest, thanks to its several advantages over other resonant topologies, such as: (1) the ability to operate at zero-voltage switching (ZVS) or zero-current switching (ZCS), (2) containing a high frequency transformer it performs the galvanic isolation between the grid and the EV, (3) a wide output voltage regulation is possible, (4) the output filter consists only in a capacitor and not in an LC filter [ 29 , 30 , 31 ].

figure 4

Example of two-stage onboard charger with LLC converter

However in literature other types of dc–dc converter have been proposed. In [ 32 ] a bidirectional buck-boost non-isolated dc/dc converter is proposed whose principal aim is to allow a reduction of the dc-link capacitor. In [ 33 ] is proposed for the dc–dc stage a cascade structure of a high-frequency LLC converter followed by a buck converter. The advantage of this configuration consists in a constant switching frequency operation of the active switches of the LLC converter while the charge control is performed by the buck converter. On the other and by adding an additional stage to the dc–dc converter, the complexity and the cost of the converter increases. Finally, in [ 34 ] a dc–dc non-isolated buck converter is employed. This latter topology is easy to implement but it works only if the dc-link voltage is higher than the battery pack voltage.

2.2 Single Stage

If the ac–dc rectifier is combined with the dc-dc converter, a single stage battery charger is obtained. This topology of battery charger is used if lower cost and size are required [ 23 , 34 ], in fact single stage battery charger allows the elimination of some bulky and expensive components such as inductors and dc-link capacitors [ 35 , 36 ] which instead are required in two-stage charger. However, the drawback is that single stage battery chargers with non-isolated converter suffer from a limited conversion ratio, which limits their application for the wide range of output voltage. If instead a high frequency isolation is present, as in the OCB configuration proposed in [ 35 ], the low frequency component generated by the rectification stage pass through the high frequency transformer leading to large magnetizing current. Moreover to achieve power factor correction, a large number of diodes and active switches could be necessary [ 36 ], increasing in this way the complexity of the configuration and hence decreasing the reliability of the overall charger.

2.3 Integrated

To maximize the reduction of components number and hence to further reduce the size, weight and cost of the battery charger several integrated topologies have been proposed and studied in literature [ 23 , 37 , 38 , 39 , 40 ]. The concept of integration consists of reusing some of the drivetrain components (inverter and motor windings) to implement the onboard charging system. However, some problems may born from this combination: the configurations proposed in [ 37 , 38 ] required access to inaccessible points of the motor windings, in [ 39 ] a rearrangement of the motor windings is necessary during the transition between different operation modes. Finally, by using the charger configuration proposed in [ 40 ], even if neither access to the neutral point of the motor windings nor their rearrangement are required, the control of the active switches becomes more difficult.

2.4 Multifunctional

The last type of proposed OBC are the so called multifunctional OBCs. In this type of battery charger some components are shared to accomplish different aims. In this way higher fuel efficiency can be reached by smaller and lighter design. In [ 41 ], the proposed multifunctional battery charger can charge the auxiliary battery via the propulsion battery when the vehicle is in a driving state, acting in this way as an OBC and as low-voltage dc-to-dc converter (LDC) jointly. In [ 42 ], a similar configuration, shown in Fig.  5 , with the same duties is presented.

figure 5

Multifunctional OBC proposed in [ 42 ]

3 Off-Board Charger

Level 3 charger, because of their rating powers, are usually installed outside the vehicle (off-board). Also for level 3 off-board charger a large amount of different solutions is studied in literature [ 43 , 44 , 45 , 46 , 47 , 48 , 49 , 50 , 51 , 52 , 53 , 54 , 55 , 56 , 57 , 58 , 59 , 60 , 61 , 62 ]. Since it is mandatory to guarantee galvanic isolation between the AC supply circuit and the DC output circuit according to the IEC EN 61,851–23 standard, in this paragraph only isolated off-board charger have been presented.

The off-board charging system is most commonly composed of two stages: a grid-facing AC/DC converter followed by a DC/DC converter providing an interface to EV battery. Based on the converter topology, both these stages can allow unidirectional or bidirectional power flow.

3.1 Bidirectional AC/DC Converter

One of the most widely used bidirectional AC/DC converter is the three-phase LCL active rectifier [ 43 , 44 , 45 ], whose scheme is reported in Fig.  6 .

figure 6

Three-phase LCL active rectifier [ 44 ]

The advantages of this type of converter are low harmonic input currents, bidirectional power flow and power factor (PF) regulation.

In [ 46 ] and [ 47 ] the front end ac–dc conversion is performed by a neutral-point-clamped (NPC) three-phase three-level converter. This converter has been used to increase the power density and to achieve low current harmonics distortion. Another advantage is that it allows the creation of a bipolar dc bus which can be used for the implementation of partial-power converters. However, the NPC causes imbalance of power and, as a result, voltage balancing problem across the DC bus capacitors.

3.2 Unidirectional AC/DC Converter

The most common unidirectional AC/DC converter used in off-board charging system is the Vienna rectifier [ 48 , 49 , 50 ]. It has advantages such as low voltage stress on each switch and high efficiency. However, the main limitations are the restricted reactive power control and the need of a dc-link capacitor voltage balancing. In [ 48 ] the authors propose a 25 kW off-board charger prototype composed by a single-switch Vienna rectifier, as depicted in Fig.  7 , and four three level dc/dc modules parallel connected.

figure 7

Vienna rectifier scheme proposed in [ 48 ]

3.3 Bidirectional DC/DC Converter

The main isolated DC/DC converter used in case of bidirectional power flow is the dual active bridge (DAB) [ 51 , 52 , 53 , 54 ], reported in Fig.  8 , and its variants (resonant DAB [ 53 ], multilevel DAB [ 54 ]). In particular, this topology is gaining interest thanks to the capabilities of the new wide-bandgap semiconductor (Gan/SiC) devices which enabled the converter efficiency and power density improvements [ 55 ].

figure 8

Dual active bridge scheme

3.4 Unidirectional DC/DC Converter

If unidirectional power flow is required, a LLC resonant converter is chosen, in most cases, as power interface between the dc bus of the AC/DC converter and the EV battery because of to its advantages over other resonant topologies [ 56 , 57 ], such as: the ability to operate at zero-voltage switching (ZVS) or zero-current switching (ZCS), it allows a wide output voltage regulation, the output filter consists only of a capacitor and not of an inductor and capacitor (LC) filter [ 58 ].

Another unidirectional DC/DC converter used in case of unidirectional off-board charger is the phase shifted full bridge converter [ 59 , 60 ], whose scheme is reported in Fig.  9 . This type of converter has different advantages such as high power density, low magnetic interference and high efficiency which make it well suitable for the implementation in battery chargers [ 61 ].

figure 9

Phase shift full bridge converter

4 Fast Charging Stations

To reduce the driving range anxiety and hence to support a stronger increase of the penetration of EVs worldwide there is the need of a charging system which is able to replace the current existing oil station. A fast charging station (FCS) can allow the charging of an EV at 80% within a half of hour from its depletion, but to reduce the charging time from 7–8 h to 30 min, FCS requires high power from the grid and for this reason they are usually connected to the MV network [ 63 , 64 , 65 ], even if some FCS connected to the LV grid are proposed too [ 66 ]. The connection of such charging stations requires a huge capital investment and it could easily overload the distribution network. Another critical aspect to be considered consists in the voltage drop that the connection of FCS can cause along the lines of the distribution networks, which according to the standard EN50160 has to remain lower than 10%.

According to [ 67 ] the impact of fast charging stations on distribution MV grid can be mitigated with the use of energy storage systems (ESSs) which can shave peak power demand and provide additional network services. Moreover, ESS can also increase the voltage level in case of too high voltage drop along the lines, this service requires the implementation of a voltage control.

In order to further minimize the impact of the FCS on the grid, renewable energy resources can integrated within the FCS too [ 68 ]. In fact, in normal operation, during daytime, the EV batteries can be charged from the solar PV by reducing in this way the possibility of overloading the MV network. In night-time, instead, when solar energy is not available the EV batteries can be charged from the grid. EVs also can support to the grid at the peak load demand if needed. By this way, the grid will never become unstable with a high pulse power of charging from EVs.

4.1 Architecture

The configuration chosen in [ 63 ] for the ultra-fast charging station is shown in Fig.  10 . The isolation between the AC side and the DC side is performed by the line-frequency transformer. The ac–dc stage is common to all the EV, in fact the output of the Cascade H Bridge (CHB) multilevel converter creates a unipolar common DC voltage bus at which all dc–dc converters are connected. The power interface between the dc bus and the integrated storage or the EV battery is performed by a LLC resonant converter.

figure 10

Fast charging station with unipolar dc voltage bus proposed in [ 63 ]

In [ 64 ] a bipolar dc bus architecture is proposed. The charging station here adopted is shown in Fig.  11 . The bipolar dc bus is necessary to feed the three-level dc–dc converter. On one side, this architecture increase the capability of the FC station and reduce the THD; on the other hand, this system requires a DC power balance management mechanism [ 65 ].

figure 11

Fast charging station with bipolar dc voltage bus proposed in [ 64 ]

4.2 Power Control Strategies

A number of studies have been carried out also on the control strategy of EV charging station [ 69 , 70 , 71 , 72 ]. The first difference among them concerns the choice of uncoordinated charging or coordinated charging. In uncoordinated charging scheme the EV battery starts charging immediately when it is plugged in or after a fixed start delay chosen by the user. If the EVs are charged according to this scheme, their impact on the grid will result in very high peak demand and hence in huge grid issues such as: feeders and transformers overloading, high losses, high voltage drops, and more cost [ 69 ]. For all the aforementioned reasons, the research has focused on EV coordinated charging strategies, in particular. Authors in [ 70 ] propose a coordinated control strategy for an FCS with an ESS, their principal aim is that of reducing the electricity purchase cost and flatten the peak load. In [ 71 ] the FCS contains both ESSs and RESs, the coordinated energy management proposed here is obtained by using a Fuzzy Logic based control defined by three inputs and two outputs. Other control strategies based on very different and numerous algorithms have been proposed in literature. Authors in [ 72 ] investigate and compare three different control strategies for a charging station. The charging strategies coordinate in different ways the energy flow of the ESS and PV panels. The results prove that the control strategy which uses the energy of the ESS with in response to the number of EVs plugged in and takes into account also the amount of generated renewable energy has the best performance. Furthermore according to this control method the SoC of the ESS is restored during night and during excess of solar energy.

5 Charging Methods

The advantages of a lithium-ion battery over other types of energy storage devices such as high energy and power density, low memory effect and resulting capacity loss, make this type of battery the best candidate for the field of electric vehicles [ 73 ]. However, li-ion battery charging must be carried out very carefully, since the charging method greatly affects how actively electrochemical side reactions occur inside the battery, and hence the cycle life of the battery itself [ 74 ]. For this reason, finding the optimal technique to charge a battery in the shortest period of time with high efficiency and without damaging the cells, has become a new challenge for many researchers [ 75 ].

In this paragraph the main charging methods for Li-ion batteries are discussed.

5.1 Constant Current-Constant Voltage (CC–CV)

In this method, represented in Fig.  12 , both an initial constant current and a final constant voltage are used. The charging process start with a constant current until a certain voltage value, known as cut-off voltage, is reached. For Li-ion with the traditional cathode materials of cobalt, nickel, manganese and aluminium typically the cut-off voltage value is around 4.20 V/cell. The tolerance is ± 50 mV/cell. Battery charging continues with a constant voltage just equal to the cut-off value. Full charge is reached when the current decreases to between 3 and 5% of the rated current. Trickle or float charge at full charge is not suitable for a Li-ion battery; since it would cause plating of metallic lithium and compromise safety. Instead of trickle charge, a topping charge can be applied when the voltage drops below a set value.

figure 12

CC–CV charge stages for a Li-ion battery

In [ 77 ], a little bit different variant is proposed, in fact the first stage consists now in a trickle charge. This stage is activated only if the battery is deeply discharged, i.e. the cell voltage is below 3.0 V and after it the CC–CV method keeps place with the aforementioned way.

5.2 Five-Step Charging Pattern

An alternative method is here described in order to obtain faster and safer charging and longer battery cycle life [ 77 , 78 ]. The five-step charging pattern consists in a multistage (five stages) constant-current charging method, in which the charging time is divided into five steps. In each stage, the charging current is set to a constant threshold value. During charging, the voltage of the battery will increase and when it reaches the pre-set limit voltage, the stage number will increase and a new charging current set value will be applied accordingly. This process will continue until the stage number reaches 5. Figure  13 illustrates the concept of the five-step constant current charging pattern.

figure 13

Five-step charging pattern for a Li-ion battery

To find the charging current in each steps different algorithms can be used, however it could be difficult and time ineffective find the optimal charging pattern.

5.3 Pulse Charging Method

With this charging strategy the charging current is injected into the battery in form of pulses, so that a rest period is provided for the ions to diffuse and neutralize. The charging rate, which depends on the average current, can be controlled by varying the width of the pulses. It is claimed [ 77 , 79 ] that this method can really speed up the charging process, slow down the polarization effect and increase life cycles. As shown in Fig.  14 , every pulse charge current that is applied to the battery is characterized by the following factors: peak amplitude I pk , a duty cycle D  =  t on /T p , and frequency f .

figure 14

Pulse charge current parameters

Two different pulse charging methods exist: duty-fixed and duty-varied pulse-charge strategy. According to [ 79 ], the duty-varied strategy can increase the charging speed and the charging efficiency with respect to the conventional duty-fixed method.

5.4 Charging Strategies Based on Battery Physical-Based Models

In all the charging methods reported so far, only current and voltage limits are considered. Anyway, these limits do not take into consideration the aging process and the side reactions occurring inside the battery; and hence they might result too conservative for new batteries and possibly dangerous for aged batteries due to the altered behavior and characteristics. This, hence, motivates the development and the research of innovative charging algorithms which tackle the issue of the charging impact on battery state-of-health (SoH) and aging [ 76 , 80 ]. In fact, recently, many researches are focusing on develop new charging methods which minimize the charging time and extend the battery life at the same time [ 75 , 80 , 81 , 82 , 83 ].

This new category of charging strategies employs the electrochemical lithium-ion battery models to calculate quantitatively and almost precisely the amount of battery aging and to directly minimize the aging in a given charge time. The electrochemical battery model estimates the internal states of a battery in a more accurate way since it is built starting from the internal microstructure of the lithium-ion battery. However, it is very arduous to precisely estimate the parameters because the electrochemical model is composed of complicated coupled partial differential equations (PDEs), and involving a large number of parameters and boundary conditions it requires a large computation burden [ 82 ]. Therefore the complexity of these models often leads to the necessity for more memory and computational effort and thus they may not be practically implemented in the fast and real-time computations of EV BMS [ 86 ].

In recent years, many works focus on identifying simplified internal electrochemical models of the battery for efficiently determining the battery operation and hence finding the optimal charging profile.

For example, in [ 75 ] Klein et al. aim at achieving a fast charging rate while not excessively aging the cell, using nonlinear model predictive control (NMPC) techniques founded on a single particle model (SPM) of the lithium ion battery. In the single particle model (SPM) the electrodes are represented by two single spherical porous particles in which intercalation and de-intercalation phenomena take place. However, the variations in the electrolyte con-centration and in the potential are ignored. This electrochemical model is exhaustively explained in [ 83 , 84 ].

In [ 81 ], a method to optimize the charging current for Li-ion batteries using a soft actor-critic (SAC) reinforcement learning algorithm is proposed. As battery aging model Kim et al. use the single particle model augmented by electrolyte and thermal dynamics (SPMeT). Anyway, even if this model is more exhaustive than the SPM one, it is subject to a complicated mathematical structure including PDEs, ordinary-differential equations and algebraic equations. Therefore the corresponding model based algorithms can be computationally too arduous to be implemented in typical battery management systems (BMSs) [ 80 ]. In light of this, in [ 80 ], Zou et al. propose and validate a simplified ODE-based SPMeT model to achieve fast charging control. Once validated the reduced model, the authors present a charging strategy based on model predictive control (MPC) techniques which aims to charge the battery in the fastest possible manner, without excessively degrading the battery’s SOH.

Another widely employed electrochemical battery model is the pseudo two-dimensional model (P2D). This model is constructed based on the assumption that electrodes are seen as an aggregation of spherical particles (2D representation) in which the Li + ions are inserted. The first spatial dimension of this model, represented by variable x, is the horizontal axis. The second spatial dimension is the particle radius r. The cell is comprised of three regions that imply four distinct boundaries. The full description of this model can be found in [ 85 ]. Finally, in [ 82 ], the authors present an effective method to estimate the parameters of the P2D model using a neural network-based estimation scheme.

6 Genetic Algorithm for Optimal Charging Technology Size and Forecasting of Future Trends

Concerning the onboard charger, as seen in paragraph II, different technologies exist in literature and in industrial applications. The cost, dimension and weight of this product is strongly influenced by its power rating; nowadays many sizes are proposed such as: 0.75, 3.3, 6.6, 11 and 22 kW. Many different power ratings are described also for off-board charger and consequently the cost varies as a function of the power and quality.

Generally, the wide range of power ratings, the fast improvements and the lack of standardization in e-mobility technologies make quite difficult estimating the actual optimal power rate of charging technology and forecasting future tendencies for large interval of time.

In this paragraph, initially, using a genetic algorithm (GA) single-objective model, the best size of both the onboard and off-board charging technology is estimated; then, through a sensitive analysis, we try to evaluate the possible future trends in the e-mobility as the costs of the battery and of the charging technology vary.

GA is an heuristic search algorithm, which mimics the natural selection process and it is used to solve optimization problems. GAs are being used to solve a wide variety of optimization problems in the field of EVs such as: charging scheduling [ 87 , 88 , 89 ], charging station planning and location [ 90 , 91 ] and drive train control strategy implementation [92–93].

6.1 Objective Function

For managing the EV charging technology, a single-objective optimization is used to determine the optimal size of the charging technology both on-board and off-board and to determine a suitable battery capacity. The proposed optimization allows to find the optimal trade-off between the onboard and off-board charger power rate.

The aim of the sizing procedure is to minimize the overall costs coming from the charging of electric vehicles. Therefore, the optimization problem is modeled in the expressions ( 1 ), ( 2 ) and ( 3 ).

First of all, the objective function representing the cost of the charging system is computed in ( 1 )

The coefficients \({k}_{i}\) represent the cost of the i-th component allocated to the vehicle, while coefficients \({\beta }_{i}\) embody the benefits. The coefficient \({k}_{bat}\) represents the cost of the battery per kWh, more precisely it also includes an amount which consider the costs arising from the space occupied by the battery inside the vehicle, its positioning and all the other aspects which are influenced by the battery capacity. Therefore, the first factor of the function describes the cost trend as a function of the battery size. This behaviour is not perfectly linear since we assume the fact that as the battery capacity increases the payload available inside the vehicle decreases. Coefficient \({k}_{on}\) includes the costs in € per kW of all the components influenced by the rated power of the onboard charger: the cost of power electronics, cost of space, cost of control and so on. Also for the cost trend of the onboard battery charger a non-linear behaviour has been assumed for the same reasons explained for the battery. In particular for this component, it can be seen that a minimum level of 6.5 kW is taken, since we assume that a smaller battery charger will not be used anymore. The computation of the cost coming from the off-board charger power rating follows the same concept used for the onboard charger, with the difference that a minimum value has not been imposed.

Once the cost of each single technology has been estimated, the next passage consists in weighing its benefits, which will be added in the cost function with a negative sign and two coefficients: \({\beta }_{on}\) and \({\beta }_{bat-off}\) , which respectively represent the benefit in terms of €/kW for the on-board battery charger and the benefits in terms of €/(kW kWh). The coefficient related to the capacity of the battery and to the power rating of the off-board charger is expected to assume high values, since it concerns the reduction of the driving range anxiety, which we know to be one of the major cause limiting the penetration of the EV worldwide. Starting from the onboard charger, for this parameter a logarithm behaviour has been assumed, since we suppose that an increasing of the \({P}_{on}\) specially takes advantage passing from 6.5 kW to values such as 22 kW, an additional increase doesn’t weigh so much on the overall system. Finally, the benefits of the off-board charger power rating \({P}_{off}\) and of the battery capacity \({C}_{bat}\) are introduced. The advantages brought by these two variable are linked together and considered in the last term of the function. An increase in the capacity of the battery should be compensated by an increase in the rating of the off-board charger, otherwise it is not fully appreciated, and vice versa if the off-board charger becomes more powerful, the capacity of the battery has to limit the stresses to which the battery is exposed.

6.2 Optimal Size of the Charging Technology

As aforementioned, finding the minimum of the function \({f}_{c}\) ( 2 ) means find those values of \({C}_{bat}, {P}_{on}\) and \({P}_{off}\) which are such as to minimise as much as possible the cost of the overall charging system. In particular, the minimum of the cost function has been performed through the use of genetic algorithms coded in Matlab. The bounds listed in ( 3 ) have been used.

Figures  15 , 16 and 17 given below report the minimum of the function once the value of the coefficients changes. The starting condition considers the actual prices and benefits reported in Table 2 . By using these values for the coefficients, a battery capacity of 60 kWh, an off-board charger of 170 kW and an onboard one of 14 kW are obtained from the optimization of the cost function.

figure 15

\({C}_{bat},{P}_{on}\) \({\mathrm{C}}_{\mathrm{bat}},{\mathrm{P}}_{\mathrm{on}}\) \({\mathrm{C}}_{\mathrm{bat}},{\mathrm{P}}_{\mathrm{on}}\) \({\mathrm{C}}_{\mathrm{bat}},{\mathrm{P}}_{\mathrm{on}}\) and \({P}_{off}\) trends as the coefficient \({k}_{bat}\) varies

figure 16

\({C}_{bat},{P}_{on}\) and \({P}_{off}\) trends as the coefficient \({k}_{on}\) varies

figure 17

\({C}_{bat},{P}_{on}\) and \({P}_{off}\) trends as the coefficient \({k}_{off}\) varies

6.3 Forecast of Future Trends

In the following second step of the study, the values of coefficients are varied, one at a time, to perform a sensitivity analysis.

It is possible to notice that the capacity of the battery and the power rating of the off-board charger are not influenced by the variation of the cost of the onboard charger per kW. The power rating of the onboard charger seems to stay quite constant around values of 14 kW, however if the cost fell below 100 € each kW, \({P}_{on}\) would start to increase until it reaches 18 kW at 25 €/kW.

As expected, the value of the power of the onboard charger does not change as \({k}_{bat}\) changes since the benefits and the costs of this parameter has not be tied to the cost of the battery technology.

The coefficient \({k}_{off}\) , represented in Fig.  17 , includes all the costs related to the off-board charger weighed from the point of view of the owner. The real cost per kW of this technology is currently around 2500 €, by considering all the aspects such as: power electronics, control, connection to MV grid, connection and communication with the vehicle, installation and maintenance.

Finally, a forecast is done about the semiconductor devices. In this field, two wide bandgap materials are showing great promise for the future: Gallium Nitride (GaN) and Silicon Carbide (SiC). Thanks to all their characteristics, compared in Table 3 , wide bandgap semiconductors allow for greater power efficiency, smaller size, lighter weight and lower overall cost. In particular, by comparing SiC and GaN properties, it can be concluded that GaN is the most suitable for low power, high frequency application while SiC material is more appropriate for high power high frequency applications [ 92 ].

7 Conclusion

The first part of this paper reviews the current conditions of EV battery charging technologies. The most common topologies which are suitable candidates for each level of an EV charger have been presented. Level 1 and level 2 charger topologies are usually mounted inside the vehicle forming in this way the so called onboard charger. On the other side level 3 chargers are installed off board the vehicles, in this way their collection leads to the creation of the so called FCSs which are promising candidates for future EV high penetration. The most common technologies used in a FCS are multilevel converters which have a high power density and a lower current harmonic distortion. To reduce the impact on the grid, almost all the FCSs are integrated with RESs and ESSs.

Li-ion batteries can be recharged according many different charging techniques which can more or less complicate the charger architecture and control. In particular, the standard charging strategy are simplest since they don’t require model information to charge the battery. Furthermore, they can be realized with very basic circuits, keeping the costs of the charger to a minimum. On the other hand, the charging strategies based on electrochemical models, taking into account the internal dynamics of the battery, consider also the aging of the battery and other constraints, hence resulting in greater accuracy and. All this is at the expense of cost and computational difficulty.

A long-term forecasting of future trends in the field of EV charging systems is a very tough task for different many reasons including the lack of standards and the continuous improvements. In the last section of this paper a possible forecasting and estimation based on GA is described. According to the results, with the actual costs, the capacity of the battery of an electric vehicle should be around 60 kWh, the power rating of the onboard charger around 14 kW and that of the off-board fast charger around 170 kW.

Finally, the silicon switching devices are expected to be replaced by wide bandgap silicon carbide (SiC) devices in order to allow a remarkable reduction in charger’s weight and volume.

Rizvi SAA, Xin A, Masood A, Iqbal S, Ullah Jan M, Rehman H (2018) Electric vehicles and their impacts on integration into power grid: a review. In: Presented at the 2nd IEEE Conference on Energy Internet and Energy System Integration (EI2), BeiJing, China, Oct. 20–22, 2018

Papadopoulos P, Cipcigan LM, Jenkins N (2009) Distribution networks with electric vehicles. In: Presented at the 44th International Universities Power Engineering Conference (UPEC), Glasgow, UK, Sept. 1–4, 2009

Longo M, Foiadelli F, Yaïci W (2019) Electric vehicles integrated with renewable energy sources for sustainable mobility. In: Martinez LR (ed) New trends in electrical vehicle powertrains. IntechOpen, UK

Google Scholar  

Ces T (2009) Transportation options in a carbon constrained world: hybrids, plug-in hybrids, biofuels, fuel cell electric vehicles, and battery electric vehicles. Int J Hydrogen Energy 34(23):9279–9296. https://doi.org/10.1016/j.ijhydene.2009.09.058

Article   Google Scholar  

Liu L, Kong F, Liu X, Peng Y, Wang Q (2015) A review on electric vehicles interacting with renewable energy in smart grid. Renew Sustain Energy Rev 51:648–661. https://doi.org/10.1016/j.rser.2015.06.036

Pellitteri F, Caruso M, Castiglia V, Di Tommaso AO, Miceli R, Schirone L (2016) An inductive charger for automotive applications. In: Presented at the 42nd Annual Conference of the IEEE Industrial Electronics Society IECON, Florence, Italy, Oct. 23–26, 2016

Panchal C, Stegen S, Lu J (2018) Review of static and dynamic wireless electric vehicle charging system. Eng Sci Technol Intern J 21(5):922–937. https://doi.org/10.1016/j.jestch.2018.06.015

Sunab L, Maa D, Tanga H (2018) A review of recent trends in wireless power transfer technology and its applications in electric vehicle wireless charging. Renew Sustain Energy Rev 91:490–503

Yilmaz M, Krein PT (2012) Review of charging power levels and infrastructure for plug-in electric and hybrid vehicles. In: Presented at the IEEE International Electric Vehicle Conference, Greenville, SC, USA, Mar. 4–8, 2012

Tran VT, Sutanto D, Muttaqi KM (2017) The state of the art of battery charging infrastructure for electrical vehicles: topologies, power control strategies, and future trend. In: Presented at the Australasian Universities Power Engineering Conference (AUPEC), Melbourne, VIC, Australia, Nov. 19–22, 2017

A novel low cost integrated on-board charger topology for electric vehicles and plug-in hybrid electric vehicles. In: Presented at 2012 27th Annual IEEE Applied Power Electronics Conference and Exposition (APEC), Orlando, FL, USA, Feb. 5–9, 2012

Ammous M, Khater M, AlMuhaini M (2017) impact of vehicle-to-grid technology on the reliability of distribution systems. In: Presented at the 9th IEEE-GCC Conference and Exhibition (GCCCE), Manama, Bahrain, May 8–11, 2017

Zhou X, Zou L, Ma Y, Gao Z (2017) Research on impacts of the electric vehicles charging and discharging on power grid. In: Presented at the 29th Chinese Control And Decision Conference (CCDC), Chongqing, China, May 28–30, 2017

Jiang Z, Tian H, Beshir MJ, Sibagatullin R, Mazloomzadeh A (2016) Statistical analysis of electric vehicles charging, station usage and impact on the grid. In: Presented at the IEEE Power & Energy Society Innovative Smart Grid Technologies Conference (ISGT), Minneapolis, MN, USA, Sept. 6–9, 2016

Yilmaz M, Krein PT (2013) Review of the impact of vehicle-to-grid technologies on distribution systems and utility interfaces. IEEE Trans Power Electron 28(12):5673–5689. https://doi.org/10.1109/TPEL.2012.2227500

Tan KM, Ramachandaramurthy VK, Yong JY (2016) Integration of electric vehicles in smart grid: a review on vehicle to grid technologies and optimization techniques. Renew Sustain Energy Rev 53:720–732. https://doi.org/10.1016/j.rser.2015.09.012

Abousleiman R, Al-Refai A, Rawashdeh O (2013) Charge capacity versus charge time in CC–CV and pulse charging of Li-ion batteries. SAE Tech Pap. https://doi.org/10.4271/2013-01-1546

Majid N, Hafiz S, Arianto S, Yuono R, Astuti ET, Prihandoko B (2017) Analysis of effective pulse current charging method for lithium ion battery. J Phys Conf Ser 817:012008. https://doi.org/10.1088/1742-6596/817/1/012008

Smith KA, Rahn CD, Wang C (2008) Model-based electrochemical estimation of lithium-ion batteries. In:2008 IEEE International Conference on Control Applications, San Antonio, TX, pp. 714–719. https://doi.org/10.1109/CCA.2008.4629589 .

Byrne DM, Oliner SD, Sichel DE (2017) How fast are semiconductor prices falling? Finance and Economics Discussion Series. Washington: Board of Governors of the Federal Reserve System. 2017(005). https://doi.org/10.17016/FEDS.2017.005 .

Goldie-Scot L (2020) A behind the scenes take on lithium-ion battery prices. BloombergNEF, 2019. Accessed 30 June 2020. [Online]. Available https://about.bnef.com/blog/behind-scenes-take-lithium-ion-battery-prices/

Zhuang W, Lu S, Lu H (2014) Progress in materials for lithium-ion power batteries. In: 2014 International Conference on Intelligent Green Building and Smart Grid (IGBSG), Taipei, 2014, pp. 1–2, doi: https://doi.org/10.1109/IGBSG.2014.6835262

K. Fahem, D. E. Chariag, L. Sbita, “On-board bidirectional battery chargers topologies for plug-in hybrid electric vehicles”, presented at the International Conference on Green Energy Conversion Systems (GECS), Hammamet, Tunisia, Mar. 23–25, 2017.

L. Cao ; H. Li ; H. Zhang, “Model-free power control of front-end PFC AC/DC converter for on-board charger”, presented at the IEEE 8th International Power Electronics and Motion Control Conference (IPEMC-ECCE Asia), Hefei, China, May 22–26, 2016.

F. Musavi; M. Edington; W. Eberle ; W. G. Dunford, “Energy efficiency in plug-in hybrid electric vehicle chargers: Evaluation and comparison of front end AC–DC topologies”, presented at the IEEE Energy Conversion Congress and Exposition, Phoenix, AZ, USA, Sept. 17–22, 2011.

Lee PW, Lee YS, Cheng DKW, Liu XC (2000) Steady-state analysis of an interleaved boost converter with coupled inductors. IEEE Trans Industr Electron 47(4):787–795. https://doi.org/10.1109/41.857959

Kanimozhi G, Kumar SS, Likhitha K (2016) Battery charger for automotive applications. In: Presented at the 10th International Conference on Intelligent Systems and Control (ISCO), Coimbatore, India, Jan. 7–8, 2016

Gautam D, Musavi F, Edington M, Eberle W, Dunford WG (2011) An automotive on-board 3.3 kW battery charger for PHEV application. In: Presented at the IEEE Vehicle Power and Propulsion Conference, Chicago, IL, USA, Sept. 6–9, 2011

Wang H, Dusmez S, Khaligh A (2014) Design and analysis of a full-bridge LLC-based PEV charger optimized for wide battery voltage range. IEEE Trans Veh Technol 63(4):1603–1613. https://doi.org/10.1109/TVT.2013.2288772

Yan X, Li L, Gao Y, Tao Z (2017) Design methodology of LLC converters based on mode analysis for battery charging applications. In: Presented at the 43rd Annual Conference of the IEEE Industrial Electronics Society (IECON), Beijing, China, 29 Oct–1 Nov 2017

Hu S, Deng J, Mi C, Zhang M (2013) LLC resonant converters for PHEV battery chargers. In: Presented at the 28th Annual IEEE Applied Power Electronics Conference and Exposition (APEC), Long Beach, CA, USA, Mar 17–21, 2013

Singh AK, Pathak MK, Rao YS (2017) A new two-stage converter with reduction of DC-link capacitor for plug-in electric vehicle battery charger.In: Presented at the 3rd International Conference on Computational Intelligence & Communication Technology (CICT), Ghaziabad, India, Feb 9–10, 2017

Lee JY, Chae HJ (2014) 6.6-kW onboard charger design using DCM PFC converter with harmonic modulation technique and two-stage DC/DC converter. IEEE Trans Indus Electron 61(3):1243–1252. https://doi.org/10.1109/TIE.2013.2262749

Kurokawa F, Hattori S (2015) Single stage AD–DC full-bridge converter for battery charger. In: Presented at the IEEE International Telecommunications Energy Conference (INTELEC), Osaka, Japan, Oct. 18–22, 2015

Kim B, Kim M, Choi S (2016) Single-stage electrolytic capacitor-less AC–DC converter with high frequency isolation for EV charger. In: Presented at the IEEE 8th International Power Electronics and Motion Control Conference (IPEMC-ECCE Asia), Hefei, China, May 22–26, 2016

Li S, Deng J, Mi CC (2013) Single-stage resonant battery charger with inherent power factor correction for electric vehicles. IEEE Trans Veh Technol 62(9):4336–4344. https://doi.org/10.1109/TVT.2013.2265704

Woo D, Joo D, Lee B (2015) On the feasibility of integrated battery charger utilizing traction motor and inverter in plug-in hybrid electric vehicles. IEEE Trans Power Electron 30(12):7270–7281. https://doi.org/10.1109/TPTEL.2015.2396200

Wang L, Liang J, Xu G, Xu K, Song Z (2012) A novel battery charger for plug-in hybrid electric vehicles. In: Presented at the IEEE International Conference on Information and Automation, Shenyang, 2012, pp. 168–173. https://doi.org/10.1109/ICInfA.2012.6246802

Haghbin S, Carlson O (2014) Integrated motor drive and non-isolated battery charger based on the split-phase PM motors for plug-in vehicles. J Eng 2014(6):275–283. https://doi.org/10.1049/joe.2014.0126

Shi C, Tang Y, Khaligh A (2017) A single-phase integrated onboard battery charger using propulsion system for plug-in electric vehicles. IEEE Trans Veh Technol 66(12):10899–10910. https://doi.org/10.1109/TVT.2017.2729345

Kim S, Kang F (2015) Multifunctional onboard battery charger for plug-in electric vehicles. IEEE Trans Industr Electron 62(6):3460–3472. https://doi.org/10.1109/TIE.2014.237687

Nguyen HV, Lee D (2018) Single-phase multifunctional onboard battery chargers with active power decoupling capability. In: Presented at the IEEE Applied Power Electronics Conference and Exposition (APEC), San Antonio, TX, 2018, pp. 3434–3439

Aggeler D, Canales F, Zelaya-De La Parra H, Coccia A, Butcher N, Apeldoorn O (2010) Ultra-fast DC-charge infrastructures for EV-mobility and future smart grids. In: Presented at the IEEE PES Innovative Smart Grid Technologies Conference Europe (ISGT Europe), Gothenberg, 2010, pp. 1–8. https://doi.org/10.1109/ISGTEUROPE.2010.5638899

Tu H, Feng H, Srdic S, Lukic S (2019) Extreme fast charging of electric vehicles: a technology overview. IEEE Trans Transport Electrific 5(4):861–878. https://doi.org/10.1109/TTE.2019.2958709

Verma A, Singh B (2017) Three phase off-board bi-directional charger for EV with V2G functionality. In: 2017 7th International Conference on Power Systems (ICPS), Pune, 2017, pp. 145–150. https://doi.org/10.1109/ICPES.2017.8387283

Fang Y, Cao S, Xie Y, Wheeler P (2016) Study on bidirectional-charger for electric vehicle applied to power dispatching in smart grid. In: Presented at the IEEE 8th International Power Electronics and Motion Control Conference (IPEMC-ECCE Asia), Hefei, 2016, pp. 2709–2713

Mortezaei A, Abdul-Hak M, Simoes MG (2018) A Bidirectional NPC-based level 3 EV charging system with added active filter functionality in smart grid applications. In: 2018 IEEE Transportation Electrification Conference and Expo (ITEC), Long Beach, CA, 2018, pp. 201–206. https://doi.org/10.1109/ITEC.2018.8450196

Kim J, Lee J, Eom T, Bae K, Shin M, Won C (2018) Design and control method of 25 kW high efficient EV fast charger. In: Presented at the 21st International Conference on Electrical Machines and Systems (ICEMS), Jeju, 2018, pp. 2603–2607

Rajendran G, Vaithilingam C, Prakash O (2019) Modeling of Vienna rectifier with PFC controller for electric vehicle charging stations. In: AIP Conference Proceedings. 2137(1). https://doi.org/10.1063/1.5120996

Chen S, Yu W, Meyer D (2019) Design and implementation of forced air-cooled, 140 kHz, 20 kW SiC MOSFET based Vienna PFC. In: Proc. IEEE Appl. Power Electron. Conf. Expo. (APEC), Mar 2019, pp. 1196–1203

Taghizadeh S, Hossain MJ, Lu J (2015) Bidirectional isolated vehice to grid (V2G) system: an optimized implementation and approach. In: 2015 IEEE PES Asia-Pacific Power and Energy Engineering Conference (APPEEC), Brisbane, QLD, 2015, pp. 1–5. https://doi.org/10.1109/APPEEC.2015.7380912

Xue L, Shen Z, Boroyevich D, Mattavelli P, Diaz D (2015) Dual active bridge-based battery charger for plug-in hybrid electric vehicle with charging current containing low frequency ripple. IEEE Trans Power Electron 30(12):7299–7307. https://doi.org/10.1109/TPEL.2015.2413815

Zahid ZU, Dalala ZM, Chen R, Chen B, Lai J (2015) Design of bidirectional DC–DC resonant converter for vehicle-to-grid (V2G) applications. IEEE Trans Transport Electrifi 1(3):232–244. https://doi.org/10.1109/TTE.2015.2476035

Moonem MA, Krishnaswami H (2012) Analysis and control of multi-level dual active bridge DC–DC converter. In: 2012 IEEE Energy Conversion Congress and Exposition (ECCE), Raleigh, NC, 2012, pp. 1556–1561. https://doi.org/10.1109/ECCE.2012.6342628

Akagi H, Yamagishi T, Tan NML, Kinouchi S, Miyazaki Y, Koyama M (2015) Power-loss breakdown of a 750-V 100-kW 20-kHz bidirectional isolated DC–DC converter using SiC-MOSFET/SBD dual modules. IEEE Trans Indus Appl 51(1):420–428. https://doi.org/10.1109/TIA.2014.2331426

Tan L, Wu B, Rivera S (2015) A bipolar-DC-bus EV fast charging station with intrinsic DC-bus voltages equalization and minimized voltage ripples. In: IECON 2015-41st Annual Conference of the IEEE Industrial Electronics Society, Yokohama, 2015, pp. 002190–002195. https://doi.org/10.1109/IECON.2015.7392426

Haritha AS, Jithin KJ (2019) An efficient resonant converter based charging scheme for electric vehicle. In: 2019 5th International Conference on Advanced Computing and Communication Systems (ICACCS), Coimbatore, India, 2019, pp. 599–603

Zhao H, Wang L, Chen Z, He X (2019) Challenges of fast charging for electric vehicles and the role of red phosphorus as anode material: review. Energies 12:3897. https://doi.org/10.3390/en12203897

Dusmez S, Cook A, Khaligh A (2011) Comprehensive analysis of high quality power converters for level 3 off-board chargers. In: 2011 IEEE Vehicle Power and Propulsion Conference, Chicago, IL, 2011, pp. 1–10. https://doi.org/10.1109/VPPC.2011.6043096

Feizi M, Beiranvand R (2020) An improved phase-shifted full-bridge converter with extended ZVS operation range for EV battery charger applications. In: 2020 11th Power Electronics, Drive Systems, and Technologies Conference (PEDSTC), Tehran, Iran, 2020, pp. 1–6. https://doi.org/10.1109/PEDSTC49159.2020.9088444

Feizi M, Beiranvand R (2020) Simulation of a high power self-equalized battery charger using voltage multiplier and phase-shifted full bridge converter for lithium-ion batteries. In: 2020 11th Power Electronics, Drive Systems, and Technologies Conference (PEDSTC), Tehran, Iran, 2020, pp. 1–6. https://doi.org/10.1109/PEDSTC49159.2020.9088454

Ota Y, Taniguchi H, Suzuki H, Nakajima T, Baba J, Yokoyama A (2012) Implementation of grid-friendly charging scheme to electric vehicle off-board charger for V2G. In: 2012 3rd IEEE PES Innovative Smart Grid Technologies Europe (ISGT Europe), Berlin, 2012, pp. 1–6

Vasiladiotis M, Rufer A, Béguin A (2012) Modular converter architecture for medium voltage ultra fast EV charging stations: global system considerations. In: Presented a the IEEE International Electric Vehicle Conference, Greenville, SC, 2012, pp. 1–7

Rivera S, Wu B, Kouro S, Yaramasu V, Wang J (2015) Electric vehicle charging station using a neutral point clamped converter with bipolar DC bus. IEEE Trans Industr Electron 62(4):1999–2009

Tan L, Wu B, Rivera S, Yaramasu V (2016) Comprehensive DC power balance management in high-power three-level DC–DC converter for electric vehicle fast charging. IEEE Trans Power Electron 31(1):89–100

Gjelaj M, Træholt C, Hashemi S, Andersen PB (2017) Optimal design of DC fast-charging stations for EVs in low voltage grids. In: Presented at the IEEE Transportation Electrification Conference and Expo (ITEC), Chicago, IL, USA, Jun. 22–24, 2017

Mauri G, Bertini D, Fasciolo E, Fratti S (2013) The impact of EV’s fast charging stations on the MV distribution grids of the Milan metropolitan area. In: Presented at the 22nd International Conference and Exhibition on Electricity Distribution (CIRED 2013), Stockholm, Sweden, Jun. 10–13, 2013

Shiramagond T, Lee W (2018) Integration of renewable energy into electric vehicle charging infrastructure. In: Presented at the 2018 IEEE International Smart Cities Conference (ISC2), Kansas City, MO, USA, Sept. 16–19, 2018, pp. 1–7. https://doi.org/10.1109/ISC2.2018.8656981

Abul’Wafa AR, El’Garably A, Mohamed WAF (2017) Uncoordinated vs coordinated charging of electric vehicles in distribution systems performance. Intern J Eng Inform Syst (IJEAIS) 1(6):54–65

Huajie D, Zechun H, Yonghua S, Xiaorui H, Yongxiang L (2014) Coordinated control strategy of energy storage system with electric vehicle charging station. In: Presented at the IEEE Conference and Expo Transportation Electrification Asia-Pacific (ITEC Asia-Pacific), Beijing, China, 31 Aug–3 Sept 2014

García-Triviño P, Fernández-Ramírez LM, Torreglosa JP, Jurado F (2016) Fuzzy logic control for an electric vehicles fast charging station. In: Presented at the International Symposium on Power Electronics, Electrical Drives, Automation and Motion (SPEEDAM), Anacapri, Italy, Jun 22–24, 2016

Castello CC, LaClair TJ, Maxey LC (2014) Control strategies for electric vehicle (EV) charging using renewables and local storage. In: Presented at the IEEE Transportation Electrification Conference and Expo (ITEC), Dearborn, MI, USA, Jun 18–15, 2014

Ayoub E, Karami N (2015) Review on the charging techniques of a Li-ion Battery. In: Presented at the Third Annual International Conference on Technological Advances in Electrical, Electronic, and Computer Engineering (TAEECE2015), Beirut, 2015, pp. 50–55

Gao Y, Jiang J, Zhang C, Zhang W, Ma Z, Jiang Y (2017) Lithium-ion battery aging mechanisms and life model under different charging stresses. J Power Sour 356:103–114. https://doi.org/10.1016/j.jpowsour.2017.04.084

Klein R, Chaturvedi NA, Christensen J, Ahmed J, Findeisen R, Kojic A (2011) Optimal charging strategies in lithium-ion battery. In: Proceedings of the 2011 American Control Conference, San Francisco, CA, 2011, pp. 382–387. https://doi.org/10.1109/ACC.2011.5991497

Zhang SS (2006) The effect of the charging protocol on the cycle life of a Li-ion battery. J Power Sour 161(2):1385–1391

Ayoub E, Karami N (2015) Review on the charging techniques of a Li-Ion battery. In: Presented at the Third International Conference on Technological Advances in Electrical, Electronics and Computer Engineering (TAEECE), Beirut, 2015, pp. 50–55

Liu Y, Hsieh C, Luo Y (2011) Search for an optimal five-step charging pattern for Li-ion batteries using consecutive orthogonal arrays. IEEE Trans Energy Convers 26(2):654–661

Chen L (2009) Design of duty-varied voltage pulse charger for improving Li-ion battery-charging response. IEEE Trans Industr Electron 56(2):480–487

Zou C, Hu X, Wei Z, Wik T, Egardt B (2018) Electrochemical estimation and control for lithium-ion battery health-aware fast charging. IEEE Trans Industr Electron 65(8):6635–6645. https://doi.org/10.1109/TIE.2017.2772154

Kim M, Baek J, Han S (2020) Optimal charging method for effective Li-ion battery life extension based on reinforcement learning. arXiv e-prints

Chun H, Kim J, Yu J, Han S (2020) Real-time parameter estimation of an electrochemical lithium-ion battery model using a long short-term memory network. IEEE Access 8:81789–81799. https://doi.org/10.1109/ACCESS.2020.2991124

Santhanagopalan S, Guo Q, Ramadass P, White RE (2006) Review of models for predicting the cycling performance of lithium ion batteries. J Power Sour 156(2):620–628

Perez H, Dey S, Hu X, Moura S (2017) Optimal charging of Li-ion batteries via a single particle model with electrolyte and thermal dynamics. J Electrochem Soc 164(7):A1679. https://doi.org/10.1149/2.1301707jes

Jokar A, Rajabloo B, Désilets M, Lacroix M (2016) Review of simplified pseudo-two-dimensional models of lithium-ion batteries. J Power Sour 327:44–55

Du H, Cao D, Zhang H (2017) Modeling, dynamics, and control of electrified vehicles. Woodhead Publishing, UK

Elmehdi M, Abdelilah M (2019) Genetic algorithm for optimal charge scheduling of electric vehicle fleet. In: Presented at NISS19: Proceedings of the 2nd International Conference on Networking, Information Systems and Security,Rabat (Morocco), Mar. 2019. https://doi.org/10.1145/3320326.3320329

Korotunov S, Tabunshchyk G, Okhmak V (2020) Genetic algorithms as an optimization approach for managing electric vehicles charging in the smart grid. CMIS

Alonso M, Amaris H, Germain JG, Galan JM (2014) Optimal charging scheduling of electric vehicles in smart grids by heuristic algorithms. Energies 7(4):2449. https://doi.org/10.3390/en7042449

Yan X, Duan C, Chen X, Duan Z (2014) Planning of Electric Vehicle charging station based on hierarchic genetic algorithm. In: 2014 IEEE Conference and Expo Transportation Electrification Asia-Pacific (ITEC Asia-Pacific), Beijing, 2014, pp. 1–5. https://doi.org/10.1109/ITEC-AP.2014.6941087

Chen S, Shi Y, Chen X, Qi F (2015) Optimal location of electric vehicle charging stations using genetic algorithm. In: 2015 17th Asia-Pacific Network Operations and Management Symposium (APNOMS), Busan, 2015, pp. 372–375. https://doi.org/10.1109/APNOMS.2015.7275344

Gallium nitride (GaN) versus silicon carbide (SiC) in the high frequency (RF) and power switching applications. Microsemi Corporation, Aliso Viejo, California, USA. Accessed 17 Apr 2019. [Online]. Available: https://www.richardsonrfpd.com/resources/RellDocuments/SYS26/Microsemi-A-Comparison-of-Gallium-Nitride-Versus-Silicon-Carbide.pdf

Download references

Acknowledgements

This research is part of the Energy for Motion project of the Department of Energy of Politecnico di Milano, funded by the Italian Ministry of University and Research (MIUR) through the Department of Excellence Grant 2018-2022.

Open access funding provided by Politecnico di Milano within the CRUI-CARE Agreement.

Author information

Authors and affiliations.

Department of Energy, Politecnico Di Milano, Milan, Italy

Morris Brenna, Federica Foiadelli, Carola Leone & Michela Longo

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Carola Leone .

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Brenna, M., Foiadelli, F., Leone, C. et al. Electric Vehicles Charging Technology Review and Optimal Size Estimation. J. Electr. Eng. Technol. 15 , 2539–2552 (2020). https://doi.org/10.1007/s42835-020-00547-x

Download citation

Received : 08 April 2020

Revised : 07 July 2020

Accepted : 17 September 2020

Published : 02 October 2020

Issue Date : November 2020

DOI : https://doi.org/10.1007/s42835-020-00547-x

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Electric vehicle
  • Plug-in electric vehicles
  • Battery charger
  • Charging infrastructure
  • Vehicle-to-grid
  • Grid-to-vehicle
  • Charging methods
  • Find a journal
  • Publish with us
  • Track your research
  • Skip to main content
  • Keyboard shortcuts for audio player

Life Kit

  • LISTEN & FOLLOW
  • Apple Podcasts
  • Google Podcasts
  • Amazon Music

Your support helps make our show possible and unlocks access to our sponsor-free feed.

You asked, we answered: Your questions about electric vehicles

Camila Domonoske square 2017

Camila Domonoske

Photograph of a person's hand plugging in their electric vehicle to charge with an EV charger. The hand is cut out of the photograph and part of a collage with green rectangles, question marks, and graph paper.

If you're thinking about getting an electric vehicle, you're not alone.

People in the U.S. buy more than a million new cars every month, and as of March, less than 10% of those are electric vehicles. But more than half of car shoppers are at least considering battery-powered cars and SUVs, according to multiple studies .

And shoppers have lots of questions. In January, The Sunday Story, an NPR podcast, asked listeners for their EV questions. More than 60 listeners sent in queries, and The Sunday Story and Life Kit teamed up to answer them. The listener questions have been edited for length and clarity.

Are EVs truly better for the planet, even with mining for batteries and fossil-fuel -based electricity to charge them? (This was the No. 1 question asked by our listeners.)

The answer is yes . Many researchers have confirmed it , and online tools let you compare the impacts for yourself. One of the most recent analyses comes from Corey Cantor with the energy research company BloombergNEF, who headlined his report last month: "No Doubt About It: EVs Really Are Cleaner Than Gas Cars."

"Big picture, moving away from spewing more CO2 into the atmosphere is a good thing for the climate," he says. And the environmental benefits of EVs are getting bigger over time as grids get cleaner.

The Electric Car Race! Vroom, Vroom!

Short Wave

  • Amazon Alexa

Is it better from an environmental standpoint to buy an electric vehicle now, or keep driving the gas car you have until you need a new car? –Ali Mercural, Portland, Ore.

For the climate, there's a strong case for switching now.

Yes, creating that new EV — getting the materials to build it from scratch — is resource-intensive. But the climate impact of a gas-powered car increases every single day you drive it.

To be precise, more than 85% of a gas-powered vehicle's lifetime emissions come from using the car, not from building the car. That's according to researchers at Argonne National Laboratory. And that means the new EV, despite its manufacturing costs, will be cleaner over time.

Jessika Trancik , a professor at the Institute for Data, Systems and Society at Massachusetts Institute of Technology, suggests taking the long view on decisions like these. Think not just about emissions right now but over the entire time you'll own a vehicle.

"Generally speaking, switching to that electric vehicle is going to provide a benefit over the lifetime of the car," she says.

I'm not proud, but I've run out of gas twice in my life. Luckily, I had friends nearby to bring me a gallon of gas. What would happen if I ran out of charge in an EV? Would a tow truck come to charge me up? How long would that take? And how embarrassing would that be? –Robin Rzechula, Chicago

We can't promise it won't be embarrassing, but a tow truck could tow you to a charger. In some cities, AAA will bring a mobile charger to you.

Overall, charging is a different experience than fueling up. With a combustion engine, you have to regularly make a stop at a gas station to fill up. With an EV, for daily driving, most people charge at home overnight – which drivers frequently cite as a major perk of EV ownership. (This does require the ability to charge at home).

For road trips, on the other hand, many parts of the country still have limited availability of fast chargers, which are high-speed chargers designed for use in the middle of a trip. Charger speeds and reliability at public charging stations vary, and charging takes much longer than filling up at a gas station.

So charging takes less work day-to-day, but more planning on long trips. Map out chargers on your route so you won't find yourself calling AAA.

Does leasing an electric car come with the same perks (like tax rebates) as buying an electric car? –Hallie Andrews, Washington, D.C.

The same or better.

There's a federal $7,500 tax credit for purchasing an EV, now available as an up-front credit toward the cost of the car. But the list of vehicles that qualify is short because of requirements meant to support U.S. jobs and supply chains. Buyers also have to be under an income cap.

Leased electric vehicles all qualify for a $7,500 credit – no matter where they're built, with no income cap. Check your lease paperwork to confirm that the credit is being fully passed along to you.

Efforts underway to make cities more EV-friendly

Wouldn't it be better to design cities around mass transit and use mass transit than get everyone to convert to electric vehicles? – Thomas Guffey, Los Angeles

Yes, designing cities to encourage mass transit – and to make them more walkable and bikeable – has a lower carbon footprint than relying on electric vehicles, in addition to other benefits . Electric bikes also have a fraction of the environmental footprint of EVs.

Switching to EVs is an important part of fighting climate change, but far from the only change that needs to happen.

The digital story was edited by Malaka Gharib. The visual editor is Beck Harlan. We'd love to hear from you. Leave us a voicemail at 202-216-9823, or email us at [email protected].

Listen to Life Kit on Apple Podcasts and Spotify , and sign up for our newsletter .

  • Life Kit: Sustainability
  • electric vehicles
  • electric vehicle
  • mass transit

IMAGES

  1. (PDF) ELECTRIC AND HYBRID ELECTRIC VEHICLES

    research paper on electric vehicles pdf

  2. (PDF) Use of Solar Energy in Electric Vehicles

    research paper on electric vehicles pdf

  3. Research Papers On Cars

    research paper on electric vehicles pdf

  4. (PDF) Electric Vehicles Beyond Energy Storage and Modern Power Networks

    research paper on electric vehicles pdf

  5. (PDF) Hybrid Electric Vehicle Design

    research paper on electric vehicles pdf

  6. (PDF) A REVIEW ON CHALLENGES AND OPPORTUNITIES OF ELECTRIC VEHICLES (EVS)

    research paper on electric vehicles pdf

VIDEO

  1. European Studies: Electric Car Fires

  2. New Report, Electric vehicles more polluting than ICE cars & EVs are inflating Motor insurance costs

  3. Electric vehicles on the rise in India

COMMENTS

  1. PDF Charging the Future: Challenges and Opportunities for Electric Vehicle

    Electric vehicles (EVs) have advanced significantly this decade, owing in part to decreasing battery costs. Yet EVs remain more costly than gasoline fueled vehicles over their useful life. This paper analyzes the additional advances that will be needed, if electric vehicles are to sig-nificantly penetrate the passenger vehicle fleet. Battery Prices

  2. (PDF) The electric vehicle: a review

    Abstract: Electric vehicles (EV), as a promising way to reduce the greenhouse. effect, have been researched extensively. With im provements in the areas. of power electrics, energy storage and ...

  3. The rise of electric vehicles—2020 status and future expectations

    Nakada T, Ishikawa S and Oki S 2014 Development of an electric motor for a newly developed electric vehicle SAE Technical Paper 2014-01-1879. Google Scholar; Narassimhan E and Johnson C 2018 The role of demand-side incentives and charging infrastructure on plug-in electric vehicle adoption: analysis of US States Environ. Res. Lett. 13 074032 ...

  4. An Overview on Why Electric Cars Are the Future of Transportation

    estimated to be from Electric and Hybrids ("The Future is Electric," 2020). Countries are all working towards a plan to only Electric or Hybrid vehicles by. 2050 and China is leading the way with a total of 1.3 million sales in 2020. Researchers at JP Morgan estimate that it would increase upto 55% by 2025.

  5. PDF The Economics of Electric Vehicles National Bureau of Economic Research

    The Economics of Electric Vehicles David S. Rapson and Erich Muehlegger NBER Working Paper No. 29093 July 2021, Revised July 2022 JEL No. Q54,Q55,Q58,R4 ABSTRACT We examine the private and public economics of electric vehicles (EVs) and discuss when market forces will produce the optimal path of EV adoption. Privately, consumer cost savings ...

  6. Electric vehicles: the future we made and the problem of unmaking it

    1. Introduction. According to the UK Society of Motor Manufacturers and Traders (SMMT), the Tesla Model 3 sold 2,685 units in December 2019, making it the 9th best-selling car in the country in that month (by new registrations; in August, a typically slow month for sales, it had been 3rd with 2,082 units sold; Lea, 2019; SMMT, 2019).As of early 2020, battery electric vehicles (BEVs) such as ...

  7. PDF A Review on Electric Vehicles: Technologies and Challenges

    Abstract:Electric Vehicles (EVs) are gaining momentum due to several factors, including the price reduction as well as the climate and environmental awareness. This paper reviews the advances of EVs regarding battery technology trends, charging methods, as well as new research challenges and open opportunities.

  8. Global Perspectives on and Research Challenges for Electric Vehicles

    This paper describes the characteristics of worldwide scientific contributions to the field of electric vehicles (EVs) from 1955 to 2021. For this purpose, a search within the Scopus database was conducted using "Electric Vehicle" as the keyword. As a result, 50,195 documents were obtained through analytical and bibliometric techniques and classified into six communities according to the ...

  9. A Review on Electric Vehicles: Technologies and Challenges

    Electric Vehicles (EVs) are gaining momentum due to several factors, including the price reduction as well as the climate and environmental awareness. This paper reviews the advances of EVs regarding battery technology trends, charging methods, as well as new research challenges and open opportunities. More specifically, an analysis of the worldwide market situation of EVs and their future ...

  10. A review of electric vehicle technology: Architectures, battery

    Electric vehicles are a relatively green area with much potential for research. Recently, several research papers have been published in this area to present the new research and challenges to mature EV technology to the next level.

  11. The current research on electric vehicle

    In 21st century, Electric vehicle(EV) as a green transportation tool, EV has won great attention from researchers, leading to a series of intensive and extensive studies. With the development of high-storage battery and EV, the number of the EV will rise dramatically, so that the random charging and discharging behavior will bring new challenges to safe and stable operation of power grids. The ...

  12. (PDF) Recent development on electric vehicles

    Recent Development on Electric Vehicles. K.W .E CHENG. 1. Department of Electrical Engineering, The Hong Kong Polytechnic University, Hong Kong. Email: [email protected]. Abstract - This ...

  13. (PDF) Electric cars. Advantages and disadvantages

    But it functions on electricity which can be produced from numerous renewable sources such as hydro, solar, or even nuclear power. An EV can convert 59-64% of the electrical energy into usable ...

  14. Review and Development of Electric Motor Systems and Electric

    This paper presents a review on the recent research and technical progress of electric motor systems and electric powertrains for new energy vehicles. Through the analysis and comparison of direct current motor, induction motor, and synchronous motor, it is found that permanent magnet synchronous motor has better overall performance; by comparison with converters with Si-based IGBTs, it is ...

  15. Energy and battery management systems for electrical vehicles: A

    A battery is a type of electrical energy storage device that has a large quantity of long-term energy capacity. A control branch known as a "Battery Management System (BMS)" is modeled to verify the operational lifetime of the battery system pack (Pop et al., 2008; Sung and Shin, 2015).For the purposes of safety, fair balancing among the cells of the battery package has to be under ...

  16. A review on barrier and challenges of electric vehicle in India and

    The Society of Indian Automobile (SIAM) along with other automobile manufacturers aim in achieving selling of hundred percent pure EVs (battery electric and fuel cell vehicles) for intra-city public transport fleets by 2030 [90].Under this scheme, i) 40% of new electric vehicle sale is expected to put on the market by 2030 and ii) 60% of new ...

  17. Electric Vehicles Charging Technology Review and Optimal Size

    Many different types of electric vehicle (EV) charging technologies are described in literature and implemented in practical applications. This paper presents an overview of the existing and proposed EV charging technologies in terms of converter topologies, power levels, power flow directions and charging control strategies. An overview of the main charging methods is presented as well ...

  18. (PDF) Overview of Electric and Hybrid Vehicles

    The two EV technologies. currently remain at the top are the battery electric vehicles. (BEVs) and plug-in hybrid electric vehicles (PHEV). This paper. gives an overview of various EV technologies ...

  19. A Study on the Adoption of Electric Vehicles in India: The Mediating

    These vehicles are also known as plug-in vehicles, EVs or the battery electric vehicles (BEVs). The transportation sector contributes about a quarter of GHG emissions. Automobiles are the primary source of GHG emission world over with China emitting 25.9 per cent, the USA 13.87 per cent followed by India 7.45 per cent.

  20. Electric Vehicles Security and Privacy: Challenges, Solutions, and

    vehicle by remotely controlling it through the infotainment system. Furthermore, most existing vehicles exploit Controller Area Network (CAN) as in-vehicle network architecture, which has already been proved as non-secure [6] and, there-fore, may be vulnerable to potential cyberattacks. Lastly, privacy shall also be guaranteed to prevent ...

  21. Wireless charging technologies for electric vehicles: Inductive

    2.4.1 Challenges associated with stationary inductive charging for electric vehicles Power transfer efficiency. Inductive charging systems for electric vehicles often encounter energy losses during the charging process, primarily due to factors such as distance between the charging pad and the vehicle, alignment, and electromagnetic interference.

  22. (PDF) Electric Vehicles for India: Overview and Challenges

    [email protected], [email protected]. Abstract. The global pollution is on rise and every effort made, being to reduce the CO2 emissions and save the planet. One such. effort is the introduction of ...

  23. Thinking about getting an electric vehicle? Here's what to know

    Switching to EVs is an important part of fighting climate change, but far from the only change that needs to happen. The digital story was edited by Malaka Gharib. The visual editor is Beck Harlan ...

  24. Electronics

    With the continuous development of renewable energy technologies, both domestically and internationally, the focus of energy research has gradually shifted towards renewable energy directions such as distributed photovoltaics and wind power. The penetration rate of renewable energy generation is constantly increasing, at the same time, the elements in the grid are becoming increasingly complex ...

  25. (PDF) ELECTRIC AND HYBRID ELECTRIC VEHICLES

    Major car manufacturers across t he globe hav e at least one model that uses hybrid t echnology or is completely converted to. electric. Electric Vehi cles (EVs) are experien cing rapid growth ...

  26. A Comprehensive Review on G2V,V2G and Optimization ...

    Request PDF | On Oct 16, 2023, Riddhi Thorat and others published A Comprehensive Review on G2V,V2G and Optimization Techniques for Grid Integration of Electric Vehicles | Find, read and cite all ...