Electric Vehicles Research

Thermal Material Trends Driven by SiC Adoption in EV Power Electronics

Membrane Materials for PEM Fuel Cells in the Post-PFAS World

Membrane Materials for PEM Fuel Cells in the Post-PFAS World

Here's Why the SDV Market Will Be Worth US$700 Billion by 2034

Here's Why the SDV Market Will Be Worth US$700 Billion by 2034

Upcoming Webinar on Hydrogen and Solar Charged Electric Vehicles

Upcoming Webinar on Hydrogen and Solar Charged Electric Vehicles

Safety in Electric Vehicles - Exploring Fire Protection Materials

Safety in Electric Vehicles - Exploring Fire Protection Materials

Upcoming Webinar on eVTOLs and Urban Air Mobility

Upcoming Webinar on eVTOLs and Urban Air Mobility

Unlocking Tomorrow's Vision: Exploring Lidar Technology Trends in 2024

Unlocking Tomorrow's Vision: Exploring Lidar Technology Trends in 2024

Upcoming Webinar - What's Next for the Quantum Computing Market?

Upcoming Webinar - What's Next for the Quantum Computing Market?

EV Charging Beyond the Utility Grid: A US$16Billion Market by 2034

EV Charging Beyond the Utility Grid: A US$16Billion Market by 2034

IDTechEx Release Thermal Management for EV Power Electronics Report

IDTechEx Release Thermal Management for EV Power Electronics Report

Webinar - Insights into Future Lidars for Automotive Applications

Webinar - Insights into Future Lidars for Automotive Applications

Taxiing For Takeoff: The Promise of eVTOLs

Taxiing For Takeoff: The Promise of eVTOLs

The Sustainability of Electric Vehicle Fire Protection Materials

The Sustainability of Electric Vehicle Fire Protection Materials

IDTechEx Release New Global eVTOL Aircraft Market Report

IDTechEx Release New Global eVTOL Aircraft Market Report

EVs Will Make Mining Cleaner, Cheaper, and More Productive

EVs Will Make Mining Cleaner, Cheaper, and More Productive

Webinar on the Evolution & Opportunities in Automotive HUD Technology

Webinar on the Evolution & Opportunities in Automotive HUD Technology

Which High-Tech Industries Need to Find Alternatives for PFAS?

Which High-Tech Industries Need to Find Alternatives for PFAS?

Integration of DMS & OMS Offers Advancements for In-Cabin Monitoring

Integration of DMS & OMS Offers Advancements for In-Cabin Monitoring

The Opportunities & Barriers Facing 3D Printing for Electric Vehicles

The Opportunities & Barriers Facing 3D Printing for Electric Vehicles

Metal vs Graphite Bipolar Plates: A Key PEM Fuel Cells Debate

Metal vs Graphite Bipolar Plates: A Key PEM Fuel Cells Debate

  • 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.

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

70417 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'.

Suggestions or feedback?

MIT News | Massachusetts Institute of Technology

  • Machine learning
  • Social justice
  • Black holes
  • Classes and programs

Departments

  • Aeronautics and Astronautics
  • Brain and Cognitive Sciences
  • Architecture
  • Political Science
  • Mechanical Engineering

Centers, Labs, & Programs

  • Abdul Latif Jameel Poverty Action Lab (J-PAL)
  • Picower Institute for Learning and Memory
  • Lincoln Laboratory
  • School of Architecture + Planning
  • School of Engineering
  • School of Humanities, Arts, and Social Sciences
  • Sloan School of Management
  • School of Science
  • MIT Schwarzman College of Computing

Designing better batteries for electric vehicles

Press contact :.

Photograph of electric vehicle battery.

Previous image Next image

The urgent need to cut carbon emissions is prompting a rapid move toward electrified mobility and expanded deployment of solar and wind on the electric grid. If those trends escalate as expected, the need for better methods of storing electrical energy will intensify.

“We need all the strategies we can get to address the threat of climate change,” says Elsa Olivetti PhD ’07, the Esther and Harold E. Edgerton Associate Professor in Materials Science and Engineering. “Obviously, developing technologies for grid-based storage at a large scale is critical. But for mobile applications — in particular, transportation — much research is focusing on adapting today’s lithium-ion battery to make versions that are safer, smaller, and can store more energy for their size and weight.”

Traditional lithium-ion batteries continue to improve, but they have limitations that persist, in part because of their structure. A lithium-ion battery consists of two electrodes — one positive and one negative — sandwiched around an organic (carbon-containing) liquid. As the battery is charged and discharged, electrically charged particles (or ions) of lithium pass from one electrode to the other through the liquid electrolyte.

One problem with that design is that at certain voltages and temperatures, the liquid electrolyte can become volatile and catch fire. “Batteries are generally safe under normal usage, but the risk is still there,” says Kevin Huang PhD ’15, a research scientist in Olivetti’s group.

Another problem is that lithium-ion batteries are not well-suited for use in vehicles. Large, heavy battery packs take up space and increase a vehicle’s overall weight, reducing fuel efficiency. But it’s proving difficult to make today’s lithium-ion batteries smaller and lighter while maintaining their energy density — that is, the amount of energy they store per gram of weight.

To solve those problems, researchers are changing key features of the lithium-ion battery to make an all-solid, or “solid-state,” version. They replace the liquid electrolyte in the middle with a thin, solid electrolyte that’s stable at a wide range of voltages and temperatures. With that solid electrolyte, they use a high-capacity positive electrode and a high-capacity, lithium metal negative electrode that’s far thinner than the usual layer of porous carbon. Those changes make it possible to shrink the overall battery considerably while maintaining its energy-storage capacity, thereby achieving a higher energy density.

“Those features — enhanced safety and greater energy density — are probably the two most-often-touted advantages of a potential solid-state battery,” says Huang. He then quickly clarifies that “all of these things are prospective, hoped-for, and not necessarily realized.” Nevertheless, the possibility has many researchers scrambling to find materials and designs that can deliver on that promise.

Thinking beyond the lab

Researchers have come up with many intriguing options that look promising — in the lab. But Olivetti and Huang believe that additional practical considerations may be important, given the urgency of the climate change challenge. “There are always metrics that we researchers use in the lab to evaluate possible materials and processes,” says Olivetti. Examples might include energy-storage capacity and charge/discharge rate. When performing basic research — which she deems both necessary and important — those metrics are appropriate. “But if the aim is implementation, we suggest adding a few metrics that specifically address the potential for rapid scaling,” she says.

Based on industry’s experience with current lithium-ion batteries, the MIT researchers and their colleague Gerbrand Ceder, the Daniel M. Tellep Distinguished Professor of Engineering at the University of California at Berkeley, suggest three broad questions that can help identify potential constraints on future scale-up as a result of materials selection. First, with this battery design, could materials availability, supply chains, or price volatility become a problem as production scales up? (Note that the environmental and other concerns raised by expanded mining are outside the scope of this study.) Second, will fabricating batteries from these materials involve difficult manufacturing steps during which parts are likely to fail? And third, do manufacturing measures needed to ensure a high-performance product based on these materials ultimately lower or raise the cost of the batteries produced?

To demonstrate their approach, Olivetti, Ceder, and Huang examined some of the electrolyte chemistries and battery structures now being investigated by researchers. To select their examples, they turned to previous work in which they and their collaborators used text- and data-mining techniques to gather information on materials and processing details reported in the literature. From that database, they selected a few frequently reported options that represent a range of possibilities.

Materials and availability

In the world of solid inorganic electrolytes, there are two main classes of materials — the oxides, which contain oxygen, and the sulfides, which contain sulfur. Olivetti, Ceder, and Huang focused on one promising electrolyte option in each class and examined key elements of concern for each of them.

The sulfide they considered was LGPS, which combines lithium, germanium, phosphorus, and sulfur. Based on availability considerations, they focused on the germanium, an element that raises concerns in part because it’s not generally mined on its own. Instead, it’s a byproduct produced during the mining of coal and zinc.

To investigate its availability, the researchers looked at how much germanium was produced annually in the past six decades during coal and zinc mining and then at how much could have been produced. The outcome suggested that 100 times more germanium could have been produced, even in recent years. Given that supply potential, the availability of germanium is not likely to constrain the scale-up of a solid-state battery based on an LGPS electrolyte.

The situation looked less promising with the researchers’ selected oxide, LLZO, which consists of lithium, lanthanum, zirconium, and oxygen. Extraction and processing of lanthanum are largely concentrated in China, and there’s limited data available, so the researchers didn’t try to analyze its availability. The other three elements are abundantly available. However, in practice, a small quantity of another element — called a dopant — must be added to make LLZO easy to process. So the team focused on tantalum, the most frequently used dopant, as the main element of concern for LLZO.

Tantalum is produced as a byproduct of tin and niobium mining. Historical data show that the amount of tantalum produced during tin and niobium mining was much closer to the potential maximum than was the case with germanium. So the availability of tantalum is more of a concern for the possible scale-up of an LLZO-based battery.

But knowing the availability of an element in the ground doesn’t address the steps required to get it to a manufacturer. So the researchers investigated a follow-on question concerning the supply chains for critical elements — mining, processing, refining, shipping, and so on. Assuming that abundant supplies are available, can the supply chains that deliver those materials expand quickly enough to meet the growing demand for batteries?

In sample analyses, they looked at how much supply chains for germanium and tantalum would need to grow year to year to provide batteries for a projected fleet of electric vehicles in 2030. As an example, an electric vehicle fleet often cited as a goal for 2030 would require production of enough batteries to deliver a total of 100 gigawatt hours of energy. To meet that goal using just LGPS batteries, the supply chain for germanium would need to grow by 50 percent from year to year — a stretch, since the maximum growth rate in the past has been about 7 percent. Using just LLZO batteries, the supply chain for tantalum would need to grow by about 30 percent — a growth rate well above the historical high of about 10 percent.

Those examples demonstrate the importance of considering both materials availability and supply chains when evaluating different solid electrolytes for their scale-up potential. “Even when the quantity of a material available isn’t a concern, as is the case with germanium, scaling all the steps in the supply chain to match the future production of electric vehicles may require a growth rate that’s literally unprecedented,” says Huang.

Materials and processing

In assessing the potential for scale-up of a battery design, another factor to consider is the difficulty of the manufacturing process and how it may impact cost. Fabricating a solid-state battery inevitably involves many steps, and a failure at any step raises the cost of each battery successfully produced. As Huang explains, “You’re not shipping those failed batteries; you’re throwing them away. But you’ve still spent money on the materials and time and processing.”

As a proxy for manufacturing difficulty, Olivetti, Ceder, and Huang explored the impact of failure rate on overall cost for selected solid-state battery designs in their database. In one example, they focused on the oxide LLZO. LLZO is extremely brittle, and at the high temperatures involved in manufacturing, a large sheet that’s thin enough to use in a high-performance solid-state battery is likely to crack or warp.

To determine the impact of such failures on cost, they modeled four key processing steps in assembling LLZO-based batteries. At each step, they calculated cost based on an assumed yield — that is, the fraction of total units that were successfully processed without failing. With the LLZO, the yield was far lower than with the other designs they examined; and, as the yield went down, the cost of each kilowatt-hour (kWh) of battery energy went up significantly. For example, when 5 percent more units failed during the final cathode heating step, cost increased by about $30/kWh — a nontrivial change considering that a commonly accepted target cost for such batteries is $100/kWh. Clearly, manufacturing difficulties can have a profound impact on the viability of a design for large-scale adoption.

Materials and performance

One of the main challenges in designing an all-solid battery comes from “interfaces” — that is, where one component meets another. During manufacturing or operation, materials at those interfaces can become unstable. “Atoms start going places that they shouldn’t, and battery performance declines,” says Huang.

As a result, much research is devoted to coming up with methods of stabilizing interfaces in different battery designs. Many of the methods proposed do increase performance; and as a result, the cost of the battery in dollars per kWh goes down. But implementing such solutions generally involves added materials and time, increasing the cost per kWh during large-scale manufacturing.

To illustrate that trade-off, the researchers first examined their oxide, LLZO. Here, the goal is to stabilize the interface between the LLZO electrolyte and the negative electrode by inserting a thin layer of tin between the two. They analyzed the impacts — both positive and negative — on cost of implementing that solution. They found that adding the tin separator increases energy-storage capacity and improves performance, which reduces the unit cost in dollars/kWh. But the cost of including the tin layer exceeds the savings so that the final cost is higher than the original cost.

In another analysis, they looked at a sulfide electrolyte called LPSCl, which consists of lithium, phosphorus, and sulfur with a bit of added chlorine. In this case, the positive electrode incorporates particles of the electrolyte material — a method of ensuring that the lithium ions can find a pathway through the electrolyte to the other electrode. However, the added electrolyte particles are not compatible with other particles in the positive electrode — another interface problem. In this case, a standard solution is to add a “binder,” another material that makes the particles stick together.

Their analysis confirmed that without the binder, performance is poor, and the cost of the LPSCl-based battery is more than $500/kWh. Adding the binder improves performance significantly, and the cost drops by almost $300/kWh. In this case, the cost of adding the binder during manufacturing is so low that essentially all the of the cost decrease from adding the binder is realized. Here, the method implemented to solve the interface problem pays off in lower costs.

The researchers performed similar studies of other promising solid-state batteries reported in the literature, and their results were consistent: The choice of battery materials and processes can affect not only near-term outcomes in the lab but also the feasibility and cost of manufacturing the proposed solid-state battery at the scale needed to meet future demand. The results also showed that considering all three factors together — availability, processing needs, and battery performance — is important because there may be collective effects and trade-offs involved.

Olivetti is proud of the range of concerns the team’s approach can probe. But she stresses that it’s not meant to replace traditional metrics used to guide materials and processing choices in the lab. “Instead, it’s meant to complement those metrics by also looking broadly at the sorts of things that could get in the way of scaling” — an important consideration given what Huang calls “the urgent ticking clock” of clean energy and climate change.

This research was supported by the  Seed Fund Program  of the MIT Energy Initiative (MITEI)  Low-Carbon Energy Center for Energy Storage ; by Shell, a founding member of MITEI; and by the U.S. Department of Energy’s Office of Energy Efficiency and Renewable Energy, Vehicle Technologies Office, under the Advanced Battery Materials Research Program. The text mining work was supported by the National Science Foundation, the Office of Naval Research, and MITEI.

This article appears in the Spring 2021 issue of Energy Futures , the magazine of the MIT Energy Initiative.

Share this news article on:

Related links.

  • Elsa Olivetti
  • MIT Energy Initiative
  • Department of Materials Science and Engineering
  • Energy Futures magazine

Related Topics

  • Climate change
  • Electric vehicles
  • Automobiles
  • Transportation
  • Energy storage
  • Lithium-ion
  • Renewable energy
  • Sustainability
  • Supply chains
  • Manufacturing

Related Articles

Close-up photograph of a charger plugged into an electric vehicle.

China’s transition to electric vehicles

particle in one electrode of a battery cell

Design could enable longer lasting, more powerful lithium batteries

Photo of a charging plug connected to an electric car

To boost emissions reductions from electric vehicles, know when to charge

View of driver and passenger looking out windshield of a car

3 Questions: The price of privacy in ride-sharing app performance

Using a novel methodology, MITEI researcher Joanna Moody and Associate Professor Jinhua Zhao uncovered patterns in the development trends and transportation policies of China’s 287 cities — including Fengcheng, shown here — that may help decision-makers learn from one another.

Transportation policymaking in Chinese cities

Previous item Next item

More MIT News

Illustration of bok choy has, on left, leaves being attacked by aphids, and on right, leaves burned by the sun’s heat. Two word balloons show the plant is responding with alarm: “!!!”

Plant sensors could act as an early warning system for farmers

Read full story →

A lab technician standing over a piece of equipment, resembling a dryer, with a cloud of vapor coming out of it

A home where world-changing innovations take flight

A man moves three large boxes on a handtruck while a woman standing in back of an open van takes inventory

3 Questions: Enhancing last-mile logistics with machine learning

Four women sit on a stage, one with a raised fist, in front of a projected slide headlined "Women in STEM."

Women in STEM — A celebration of excellence and curiosity

Stylized drawing of a computer monitor with a black screen, surrounded by green beams of light and a completed task list on each side. Behind these objects are two IBM quantum computers, shown as cylinders connected to wires

A blueprint for making quantum computers easier to program

A diagram shows a box of rows of long silver tubes stacked on top of each other. Tiny brown objects representing carbon nanotubes are in between the layers. An inset enlarges the brown objects and they are an array of tree-like scaffolding.

“Nanostitches” enable lighter and tougher composite materials

  • More news on MIT News homepage →

Massachusetts Institute of Technology 77 Massachusetts Avenue, Cambridge, MA, USA

  • Map (opens in new window)
  • Events (opens in new window)
  • People (opens in new window)
  • Careers (opens in new window)
  • Accessibility
  • Social Media Hub
  • MIT on Facebook
  • MIT on YouTube
  • MIT on Instagram
  • Future Energy Systems Center
  • Studies and reports
  • Funding opportunities
  • Basic energy science
  • Built environment and infrastructure
  • Climate and environment
  • Conventional energy
  • Developing world
  • Energy efficiency
  • Nuclear energy
  • Policy and economics
  • Power distribution and energy storage
  • Renewable energy
  • Transportation
  • Undergraduate education
  • Graduate & postdoctoral
  • Online education
  • Education research
  • Current members
  • Energy Futures
  • In the media
  • Affiliations

Cost-competitive electric vehicles that go the distance

Detailed look at 125 u.s. auto models finds those emitting less carbon are the least expensive to drive.

If you look in aggregate at the most popular vehicles on the market today, one doesn’t have to pay more for a lower carbon-emitting vehicle. Jessika Trancik, Energy Studies Professor 

Researchers at MIT have just completed the most comprehensive study yet to address whether existing electric vehicles, despite their limited driving range, could bring about a meaningful reduction in the greenhouse-gas emissions that are causing global climate change. The four-year project integrated two enormous datasets: one highly detailed set of second-by-second driving behavior based on GPS data, and another broader, more comprehensive set of national data based on travel surveys. Together, the two datasets encompass millions of trips made by drivers all around the country. According to study author Jessika Trancik, “roughly 90 percent of the personal vehicles on the road daily could be replaced by a low-cost electric vehicle available on the market today, even if the cars can only charge overnight.”

This research was supported in part by the MIT Energy Initiative.

Related publications

Personal Vehicles Evaluated Against Climate Change Mitigation Targets

September 2016

Research Team

electric vehicles research

Institute for Data, Systems, and Society

Related news

How Many of Our Vehicles Could Be Electric? How Does 87% Strike You?

electric vehicles research

Create an account

Create a free IEA account to download our reports or subcribe to a paid service.

Trends and developments in electric vehicle markets

  • Electric Vehicles Initiative
  • Electric Vehicles Initiative campaigns

Trends and developments in electric light-duty vehicles

Trends and developments in electric heavy-duty vehicles, private sector commitment and other electrification trends, deployment of vehicle-charging infrastructure.

  • Are we entering the era of the electric vehicle?
  • Policies affecting the electric light-duty vehicle market
  • Policies affecting the electric heavy-duty vehicle market
  • Outlook for electric mobility
  • Charging infrastructure
  • Implications for electric mobility
  • References for figures

Cite report

IEA (2021), Global EV Outlook 2021 , IEA, Paris https://www.iea.org/reports/global-ev-outlook-2021, Licence: CC BY 4.0

Share this report

  • Share on Twitter Twitter
  • Share on Facebook Facebook
  • Share on LinkedIn LinkedIn
  • Share on Email Email
  • Share on Print Print

Report options

More than 10 million electric cars were on the world’s roads in 2020 with battery electric models driving the expansion, global electric passenger car stock, 2010-2020, electric car registrations increased in major markets in 2020 despite the covid pandemic, global electric car registrations and market share, 2015-2020, electric car registrations and market share in north-western european region, 2015-2020, electric car registrations and market share in selected countries, 2015-2020, electric cars had a record year in 2020, with europe overtaking china as the biggest market.

After a decade of rapid growth, in 2020 the global electric car stock hit the 10 million mark, a 43% increase over 2019, and representing a 1% stock share. Battery electric vehicles (BEVs) accounted for two-thirds of new electric car registrations and two-thirds of the stock in 2020. China, with 4.5 million electric cars, has the largest fleet, though in 2020 Europe had the largest annual increase to reach 3.2 million.

Overall the global market for all types of cars was significantly affected by the economic repercussions of the Covid-19 pandemic. The first part of 2020 saw new car registrations drop about one-third from the preceding year. This was partially offset by stronger activity in the second-half, resulting in a 16% drop overall year-on-year. Notably, with conventional and overall new car registrations falling, global electric car sales share rose 70% to a record 4.6% in 2020.

About 3 million new electric cars were registered in 2020. For the first time, Europe led with 1.4 million new registrations. China followed with 1.2 million registrations and the United States registered 295 000 new electric cars.

Numerous factors contributed to increased electric car registrations in 2020. Notably, electric cars are gradually becoming more competitive in some countries on a total cost of ownership basis. Several governments provided or extended fiscal incentives that buffered electric car purchases from the downturn in car markets. 

Overall Europe’s car market contracted 22% in 2020. Yet, new electric car registrations more than doubled to 1.4 million representing a sales share of 10%. In the large markets, Germany registered 395 000 new electric cars and France registered 185 000.  The United Kingdom more than doubled registrations to reach 176 000. Electric cars in Norway reached a record high sales share of 75%, up about one-third from 2019. Sales shares of electric cars exceeded 50% in Iceland, 30% in Sweden and reached 25% in the Netherlands. 

This surge in electric car registrations in Europe despite the economic slump reflect two policy measures. First, 2020 was the target year for the European Union’s CO 2 emissions standards that limit the average carbon dioxide (CO 2 ) emissions per kilometre driven for new cars. Second, many European governments increased subsidy schemes for EVs as part of stimulus packages to counter the effects of the pandemic.

In European countries, BEV registrations accounted for 54% of electric car registrations in 2020, continuing to exceed those of plug-in hybrid electric vehicles (PHEVs). However, the BEV registration level doubled from the previous year while the PHEV level thripled. The share of BEVs was particularly high in the Netherlands (82% of all electric car registrations), Norway (73%), United Kingdom (62%) and France (60%).

The overall car market in China was impacted by the panademic less than other regions. Total new car registrations were down about 9%.

Registration of new electric cars was lower than the overall car market in the first-half of 2020. This trend reversed in the second-half as China constrained the panademic. The result was a sales share of 5.7%, up from 4.8% in 2019. BEVs were about 80% of new electric cars registered.  

Key policy actions muted the incentives for the electric car market in China. Purchase subsidies were initially due to expire at the end of 2020, but following signals that they would be phased out more gradually prior to the pandemic, by April 2020 and in the midst of the pandemic, they were instead cut by 10% and exended through 2022. Reflecting economic concerns related to the pandemic, several cities relaxed car licence policies , allowing for more internal combustion engines vehicles to be registered to support local car industries. 

United States

The US car market declined 23% in 2020, though electric car registrations fell less than the overall market. In 2020, 295 000 new electric cars were registered, of which about 78% were BEVs, down from 327 000 in 2019. Their sales share nudged up to 2%. Federal incentives decreased in 2020 due to the federal tax credits for Tesla and General Motors, which account for the majority of electric car registrations, reaching their limit .

Other countries

Electric car markets in other countries were resilent in 2020. For example, in Canada the new car market shrunk 21% while new electric car registrations were broadly unchanged from the previous year at 51 000.

New Zealand is a notable exception. In spite of its strong pandemic response, it saw a decline of 22% in new electric car registrations in 2020, in line with a car market decline of 21%. The decline seems to be largely related to exceptionally low EV registrations in April 2020 when New Zealand was in lockdown.

Another exception is Japan, where the overall new car market contracted 11% from the 2019 level while electric car registrations declined 25% in 2020. The electric car market in Japan has fallen in absolute and relative terms every year since 2017, when it peaked at 54 000 registrations and a 1% sales share. In 2020, there were 29 000 registrations and a 0.6% sales share.

Consumer spending on EVs continues to rise, while government support stabilises

Consumer spending

Consumers spent USD 120 billion on electric car purchases in 2020, a 50% increase from 2019, which breaks down to a 41% increase in sales and a 6% rise in average prices. The rise in average prices reflects that Europe, where prices are higher on average than in Asia, accounted for a bigger proportion of new electric car registrations. In 2020, the global average BEV price was around USD 40 000 and around USD 50 000 for a PHEV.

Government spending

Governments across the world spent USD 14 billion on direct purchase incentives and tax deductions for electric cars in 2020, a 25% rise year-on-year. Despite this, the share of government incentives in total spending on EVs has been on a downward slide from roughly 20% in 2015 to 10% in 2020.

All the increase in government spending was in Europe, where many countries responded to the pandemic -induced economic downturn with incentive schemes that boosted electric car sales. In China, government spending decreased as the eligibility requirements for incentive programmes tightened.

An important novelty in subsidy schemes was the introduction of price caps in Europe and China , i.e. no subsidy given for vehicles with prices above a certain threshold. This might be responsible for average electric car price falling in Europe and China: BEV cars sold in China were 3% cheaper in 2020 than in 2019, while PHEV cars in Europe were 8% cheaper.

Consumer and government spending on electric cars, 2015-2020

More electric car models are available; ranges start to plateau, electric car models available globally and average range, 2015-2020, electric car models available by region, 2020, automakers entice customers with a wide menu including electric suv models.

Worldwide about 370 electric car models were available in 2020, a 40% increase from 2019. China has the widest offering, reflecting its  less consolidated automotive sector and that it is the world’s largest EV market. But in 2020 the biggest increase in number of models was in Europe where it  more than doubled.

BEV models are offered in most vehicle segments in all regions; PHEVs are skewed towards larger vehicle segments. Sport utility vehicle (SUV) models account for half of the available electric car models in all markets. China has nearly twice as many electric car models available as the European Union, which has more than twice as many electric models as the United States. This difference can partially be explained by the comparatively lower maturity of the US EV market, reflecting its weaker regulations and incentives at the national level.

The average driving range of new BEVs has been steadily increasing. In 2020, the weighted average range for a new battery electric car was about 350 kilometres (km), up from 200 km in 2015.The weighted average range of electric cars in the United States tends to be higher than in China because of a bigger share of small urban electric cars in China.The average electric range of PHEVs has remained relatively constant about 50 km over the past few years.

The widest variety of models and the biggest expansion in 2020 was in the SUV segment. More than 55% of announced models worldwide are SUVs and pick-ups. Original equipment manufacturers (OEMs) may be moving to electrify this segment for the following reasons:

  • SUVs are the fastest growing market segment in Europe and China, and by far the largest market share in the United States.
  • SUVs command higher prices and generally offer higher profit margins than smaller vehicles. This means OEMs find it easier to bear the extra costs of electrification for SUVs since the powertrain accounts for a smaller share of the total cost compared with a small car.
  • Electrifying the heaviest and most fuel consuming vehicles goes further toward meeting emissions targets than electrifying a small car.
  • In Europe, the ZLEV credit scheme in the most recent CO 2 emissions standards offers strong incentives for selling electric SUVs from 2025, as it relaxes emissions standards in proportion to their potential to reduce specific CO 2 emissions. In fact, in Europe, the share of electric SUV models is higher than for the overall market.

China leads in electric LCV sales with Europe not far behind and Korea entering the market

Global electric light-commercial vehicle (LCV) stock numbers about 435 000 units. About a third of these are in Europe where new electric LCV registrations in 2020 were only 5% below those in China, which is the world leader.

Electric LCV registrations in China in 2020 were 3 400 units below the previous year and slightly less than half of the peak in 2018. The bulk of the electric LCV registrations are BEVs, with PHEVs accounting for less than 10%.

In Europe, electric LCV registrations jumped almost 40% in 2020 from the prior year to exceed 37 000 units. Though that was less impressive than the more than doubling of electric car registrations. New EV registrations in Europe are being driven by economic stimulus packages and by CO 2 standards that limit emissions per kilometre driven. However, current standards for LCVs are not stringent enough to warrant large-scale electrification, as they do for passenger cars.

Registration of electric LCVs in 2020 in the rest of the world were about 19 000 units. Most of these were in Korea, reflecting the launch of two new BEV LCV models, but Canada also added to the stock of electric LCVs. Other markets around the world have yet to see much uptake of electric LCVs.

The explosion of home deliveries during the Covid-19 pandemic further boosted the electric LCV expansion in some countries. Increased deliveries raised concerns about air pollution , particularly in urban areas. In response, a number of companies announced  plans to electrify delivery fleets .

Electric LCVs registrations by region, 2015-2020

18 of the 20 largest oems have committed to increase the offer and sales of evs, original equipment manufacturer announcements related to electric light-duty vehicles.

Original equipment manufacturer announcements related to electric light-duty vehicles

BMW (2021); BJEV-BAIC (2021); BYD (2021); Chery (2021); Changan Automobile (2021); Daimler (2021); Dongfeng (2021); FAW (2021); Ford (2021); GAC; General Motors; Honda (2021); Hyundai (2020); Mazda (2021); Renault-Nissan (2019); Maruti Suzuki (2019); SAIC (2021); Stellantis (2021); Toyota (2021); Volkswagen (2021).

This table is based on the authors’ understanding of OEM announcements and may not be complete. It includes only announcements related to electric light-duty vehicles (PHEVs and BEVs) and it excludes announcements related to hybrid vehicles and those that do not provide a clear indication of the EV share.

Manufacturers’ electrification targets align with the IEA’s Sustainable Development Scenario

OEMs are expected to embrace electric mobility more widely in the 2020s. Notably 18 of the 20 largest OEMs (in terms of vehicles sold in 2020), which combined accounted for almost 90% of all worldwide new car registrations in 2020, have announced intentions to increase the number of available models and boost production of electric light-duty vehicles (LDVs).

A number of manufacturers have raised the bar to go beyond previous announcements related to EVs with an outlook beyond 2025. More than ten of the largest OEMs worldwide have declared electrification targets for 2030 and beyond.

Significantly, some OEMs plan to reconfigure their product lines to produce only electric vehicles. In the first-trimester of 2021 these announcements included: Volvo will only sell electric cars from 2030 ; Ford will only electric car sales in Europe from 2030 ; General Motors plans to offer only electric LDVs by 2035 ; Volkswagen aims for 70% electric car sales in Europe, and 50% in China and the United States by 2030 ; and Stellantis aims for 70% electric cars sales in Europe and 35% in the United States .

Overall, the announcements by the OEMs translate to estimated cumulative sales of electric LDVs of 55-72 million by 2025. In the short term (2021-2022), the estimated cumulative sales align closely with the electric LDV projections in the IEA’s Stated Policies Scenario . By 2025, the estimated cumulative sales based on the OEMs announcements are aligned with the trajectories of IEA Sustainable Development Scenario. 

OEMs’ announcements compared to electric LDVs stock projections, 2021-2025

Electric bus and truck registrations expanded in major markets in 2020.

Electric bus and electric heavy-duty truck (HDT) registrations increased in 2020 in China, Europe and North America. The global electric bus stock was 600 000 in 2020 and the electric HDT stock was 31 000.

Bus registrations

China continues to dominate the electric bus market , with registration of 78 000 new vehicles in 2020, up 9% on the year to reach a sales share of 27%. Local policies to curb air pollution are the driving force.

Electric bus registrations in Europe were 2 100, an increase of around 7%, well below the doubling in registrations seen in 2019. Electric buses now make up 4% of all new bus registrations in Europe. It is too early to see the effect of the non-binding European Clean Bus Deployment Initiative and demand may be still largely driven by muncipal level policies.

In North America, there were 580 new electric bus registrations in 2020, down almost 15% from the prior year. In the United States, electric bus deployment primarly reflects polices in California, which is the location of most of the current e-bus stock. In South America, Chile leads the way registering 400 electric buses in 2020 for a total stock of more than 800. India increased electric bus registrations 34% to 600 in 2020.

Heavy-duty truck registrations

Global electric HDT registrations were 7 400 in 2020, up 10% on the previous year. The global stock of electric HDTs numbers 31 000. China continues to dominate the category, with 6 700 new registrations in 2020, up 10% though much lower than the fourfold increase in 2019. Electric HDT registrations in Europe rose 23% to about 450 vehicles and in the United States increased to 240 vehicles. Electric trucks are still below 1% of sales in both. 

Electric truck registrations by region, 2015-2020

Electric bus registrations by region, 2015-2020, electric heavy-duty vehicle models are broadening.

The availability of electric heavy-duty vehicles (HDVs) models is expanding in leading global markets. 1 Buses were the earliest and most successful case of electrification in the HDV market, but the growing demand for electric trucks is pushing manufacturers to broaden product lines. Nevertheless, model availability is not the only indicator of a healthy market – fewer total models may reflect the reliability and broad applicability of existing designs, whereas more diversity of models may reflect the need to tailor products for specific needs and operations.

The growth in electric model availability from 2020 to 2023 across segments – bus, medium freight truck (MFT), heavy freight truck (HFT) and others – demonstrates manufacturers’ commitments to electrification. Truck makers such as Daimler , MAN , Renault , Scania and Volvo have indicated they see an all-electric future. The broadening range of available zero-emission HDVs, particularly in the HFT segment, demonstrates the commitment to provide fleets the flexibility to meet operational needs.

The HDV segment includes a wide variety of vehicle types, e.g. from long-haul freight to garbage collection trucks. China has the most variety in available electric bus models. The availablity of MFT models is broadest in the United States. For HFTs – the segment where the EV model offer is expected to the grow the most – Europe offers the widest selection of models.

Number of electric HDVs models available by segment and year, 2020-2023

Types of zero-emission hdvs expand, and driving range lengthens, current and announced zero emission hdv models by segment, release year and powertrain in major markets, 2020-2023.

Current And Announced Zero Emission Hdv Models By Segment Release Year And Powertrain In Major Markets 2020 2023

IEA analysis based on the Global Drive to Zero ZETI tool.

Data are derived from CALSTART’s Zero-Emission Technology Inventory. Although the inventory is continuously updated, this snapshot may be not fully comprehensive due to new model announcements and small manufacturers not yet captured in the inventory. The term zero-emission vehicle includes BEVs, PHEVs and FCEVs. Other includes garbage, bucket, concrete mixer, mobile commercial and street sweeper trucks. Years after 2021 include announced models.

Private sector demand for zero-emission commercial vehicles amplifies market signals for OEMs to develop EVs

Notes: Based on authors understanding of private sector announcements and may not be comprehensive. Sources: Amazon (2020); Anheuser-Busch (2019); DHL Group (2019); FedEx (2021); H2 Mobility Association (2019); Ingka Group (2018); Japan Post (2019); JD (2017); SF Express (2018); Suning (2018); UPS (2019); Various companies (2017) (2020) and Walmart (2020).

Climate Group’s EV100 Initiative update on private sector commitments

Despite a turbulent year, major companies around the world are accelerating the transition to electric mobility by shifting fleets to electric vehicles and installing charging stations.

The Climate Group’s EV100 Initiative brings together over 100 companies in 80 markets committed to making electric transport the new normal by 2030. This equates to 4.8 million vehicles switched to EVs and chargers installed in 6 500 locations by 2030.

Collectively, by 2020 EV100 members had already deployed 169 000 zero-emission vehicles, double the previous year. Even though companies identify commercial vans and heavy-duty vehicles as the most difficult EVs to find, the number of commercial electric vehicles rose 23% in 2020, including a threefold increase in electric trucks.

EV100 members are also expanding the availability of charging infrastructure for staff and customers, with 16 900 charging points installed at 2 100 locations worldwide. Over half of EV100 members are using renewables to power all their charging operations.

Significant barriers to EV adoption remain. EV100 members reported the lack of charging infrastructure as the top barrier (especially in the United States and United Kingdom). Lack of availability of appropriate vehicle types was also highlighted by the companies as a persistant obstacle. The purchase price of EVs remains an important hurdle despite many companies acknowledge the significant cost savings over the lifetime of a vehicle due to lower fuel and maintenance costs.

To help overcome these barriers, 71% of EV100 members support more favourable EV procurement tax benefits and 70% favour more supportive policies at state, regional and city government levels. Sixty percent of the member companies support government targets to phase out petrol and diesel vehicles.

Top 5 barriers to EV adoption reported by EV100 member companies

Battery demand lagged ev sales in 2020; europe sees highest rise in demand.

Automotive lithium-ion (Li-Ion) battery production was 160 gigawatt-hours (GWh) in 2020, up 33% from 2019. The increase reflects a 41% increase in electric car registrations and a constant average battery capacity of 55 kilowatt-hours (kWh) for BEVs and 14 kWh for PHEVs. Battery demand for other transport modes increased 10%. Battery production continues to be dominated by China, which accounts for over 70% of global battery cell production capacity.

China accounted for the largest share of battery demand at almost 80 GWh in 2020, while Europe had the largest percentage increase at 110% to reach 52 GWh. Demand in the United States was stable at 19 GWh.

Nickel-manganese-cobalt continues to be the dominant chemistry for Li-ion batteries, with around 71% sales share and nickel-cobalt-aluminium accounting for most of the rest. Lithium-iron-phosphate battery chemistry has regained sales share but is still under 4% for the electric car market.

According to the BNEF’s yearly survey of battery prices, the weighted average cost of automotive batteries declined 13% in 2020 from 2019, reaching USD 137/kWh at a pack level. Lower prices are offered for high volume purchases, confirmed by teardown analysis of a VW ID3 showing an estimated cost of USD 100/kWh for its battery cells.

In Europe, demand for batteries in 2020 exceeded domestic production capacity. Today Europe’s main battery factories are located in Poland and Hungary. Production capacity is roughly 35 GWh per year, but announced capacity could yield up to 400 GWh by 2025 . Momentum was evident in 2020 in Europe with many new battery plants announced or under construction with support from the European Investment Bank . In the United States, both Korean and domestic battery manufacturers have signalled large investments in a market currently dominated by a Tesla-Panasonic joint venture.

Battery demand by region, 2015-2020

Battery demand by mode, 2015-2020, pandemic spreads popularity of electric micromobility.

Electric micromobility surged in the second-half of 2020, one of the consumer trends that accelerated during the Covid-19 pandemic, further boosted by the construction of bike lanes and other measures to promote mobility. Sales of private e-bikes in the United States more than doubled in 2020, outpacing sales of all bikes which were up an already healthy 65%.

Many shared micromobility operators reduced or suspended services during the height of the second-quarter 2020 Covid-19 lockdowns. But as confinements were eased, services rebounded strongly, with 270 cities worldwide relaunching operations . As of February 2021, around 650 cities have shared micromobility services. In Europe, e-scooter services have increased rapidly, with more than 100 cities adding operations since July 2020.

Preliminary data from operators indicate average trip distances on e-scooters have increased by around 25% relative to before the pandemic . Operators are increasingly offering more powerful e-bikes with plans to expand into electric mopeds , which could further displace longer trips currently completed by car or public transit.

Several major operators are introducing swappable batteries to improve operational efficiency and reduce emissions. Although the use of swappable batteries increases the number of total batteries needed to support a fleet, it can significantly reduce operational emissions and enable longer lifetime of vehicles.

Privately owned electric two/three-wheelers (which include motorised vehicles such as motorcycles and mopeds but exclude micromobility solutions) are concentrated in Asia, with China accounting for 99% of registrations. The global stock of electric two/three-wheelers is now around 290 million. Electric two/three-wheelers account for one-third of all two/three-wheeler sales. While current sales are dominated by Asia, the market is growing rapidly in Europe, rising by 30% in 2020, benefitting from wider model availability and continued incentives.

Availability of dockless shared micromobility services in Europe and Central Asia, 2019-2021

Korea takes a lead in deploying fuel cell electric vehicles.

Fuel cell electric vehicles (FCEVs) are zero-emission vehicles that convert hydrogen stored on-board using a fuel cell to power an electric motor. FCEV cars became commercially available in 2014, though registrations remain three orders of magnitude lower than EVs as hydrogen refuelling stations (HRS) are not widely available and unlike EVs cannot be charged at home. Few commercial FCEV models are available and with high fuel cost and purchase prices result in a higher total cost of ownership than EVs.

To address the chicken-and-egg problem for FCEVs a number of goverments have funded the construction of HRS and have deployed public buses and trucks, such as garbage trucks, to provide a certain level of station utilisation. Today, there are approximately 540 HRS globally that provide fuel for almost 35 000 FCEVs. Approximately three-quarters of the FCEVs are LDVs, 15% are buses and 10% are trucks.

In 2020, Korea took the lead in FCEVs, surpassing the United States and China, to reach more than 10 000 vehicles. To support these FCEVs, the number of HRS in Korea increased by 50%, with 18 new stations in 2020. FCEVs in China are almost exclusively buses and trucks, unlike most other countries where cars are dominant. China accounts for 94% of global fuel cell buses and 99% of fuel cell trucks.

In 2020, the global FCEV stock increased 40%, with Korea contributing half and doubling its total FCEV stock. Japan and China increased the number of HRS, each opening about 25 stations in 2020. Worldwide the number of HRS increased 15%. 

Fuel cell vehicles and hydrogen refueling station stock by region, 2020

Fuel cell electric vehicles stock by region and by mode, 2020, publicly accessible slow and fast chargers increased to 1.3 million in 2020, stock of slow public electric light duty vehicles chargers, 2015-2020, stock of fast public electric light duty vehicles chargers, 2015-2020, installation of publicly accessible chargers expanded sevenfold in the last five years; covid-19 muted the pace in 2020 while china still leads.

While most charging of EVs is done at home and work, roll-out of publicly accessible charging will be critical as countries leading in EV deployment enter a stage where simpler and improved autonomy will be demanded by EV owners. Publicly accessible chargers reached 1.3 million units in 2020, of which 30% are fast chargers. Installation of publicly accessible chargers was up 45%, a slower pace than the 85% in 2019, likely because work was interrupted in key markets due to the pandemic. China leads the world in availability of both slow and fast publicly accessible chargers.

Slow chargers

The pace of slow charger (charging power below 22 kW) installations in China in 2020 increased by 65% to about 500 000 publicly accessible slow chargers. This represents more than half of the world’s stock of slow chargers.

Europe is second with around 250 000 slow chargers, with installtions increasing one-third in 2020. The Netherlands leads in Europe with more than 63 000 slow chargers. Sweden, Finland and Iceland doubled their stock of slow chargers in 2020.

Installation of slow chargers in the United States increased 28% in 2020 from the prior year to total 82 000. The number of slow chargers installed in Korea rose 45% in 2020 to 54 000, putting it in second place.

Fast chargers

The pace of fast charger (charging power more than 22 kW) installations in China in 2020 increased by 44% to almost 310 000 fast chargers, slower than the 93% pace of annual growth in 2019. The relatively high number of publically available fast chargers in China is to compensate for a paucity of private charging options and to facilitate achievement of goals for rapid EV deployment.

In Europe, fast chargers are being rolled out at a higher rate than slow ones. There are now more than 38 000 public fast chargers, up 55% in 2020, including nearly 7 500 in Germany, 6 200 in the United Kingdom, 4 000 in France and 2 000 in the Netherlands. The United States counts 17 000 fast chargers, of which nearly 60% are Tesla superchargers. Korea has 9 800 fast chargers.

Publicly accessible fast chargers facilitate longer journeys. As they are increasingly deployed, they will enable longer trips and encourage late adopters without access to private charging to purchase an electric vehicle.

Most countries in Europe did not achieve 2020 AFID targets for publicly accessible chargers

European countries for the most part failed to meet the recommended electric vehicle supply equipment (EVSE) per EV 2020 targets for publicly accessible chargers set by the Alternative Fuel Infrastructure Directive (AFID). However, there are wide disparities between countries.

AFID, the key policy regulating the deployment of public electric EVSE in the European Union, recommended that member states aim for 1 public charger per 10 EVs, a ratio of 0.1 in 2020.

In the European Union, the average public EVSE per EV ratio was 0.09 at the end of 2020. But that is not the whole story. The Netherlands and Italy are above the target at 0.22 and 0.13 respectively, with almost all being slow chargers, though fast chargers are 3% of the installations in the Netherlands and 9% in Italy.

Countries with the highest EV penetration tend to have the lowest EVSE per EV ratios, such as Norway (0.03), Iceland (0.03) and Denmark (0.05). In these sparsely populated countries with many detached houses and private parking spaces, most EV owners can largely use private home charging . To a lesser extent, it also refects that the Nordic countries have a higher proportion of fast chargers, with shares of 40% in Iceland, 31% in Norway and 17% in Denmark. 

Ratio of public chargers per EV stock by country, 2020

Planning needs to start now for megachargers to enable long-distance trucking.

The roll-out of public charging infrastructure has so far mostly focused on serving electric light-duty vehicles. The electrification of heavy freight trucks (HFTs) is a longer term endeavour, with less than 40 electric HFTs on the road in 2020.

HFTs require batteries with high capacity to meet their needs for heavy-duty cycles and long-range operations, and consequently they require high power charging. So far charging options for HFTs have tended to be early stage demonstrations, proof-of-concept activities and efforts to faciliate standardisation .   

Megachargers of 1 megawatt (MW) or more would be capable of charging trucks operating over long distances reasonably quickly. Long-term planning for megacharger infastructure is needed now to avoid negative impacts on the electrical grid. Some impact to grids is inevitable given the high power requirements of megachargers. Significant investment may be needed for grid reinforcements, modernisation, storage and integration with power systems. Planning and co-ordination among electricity generators, distribution system operators and megacharging operations are needed.

Some efforts are underway to develop standards for megachargers. Working jointly, the CHAdeMO association and the China Electricity Council have developed an ultra-high power charging standard (up to 900 kW), called ChaoJi. A version up to 1.8 MW, called Ultra ChaoJi, is under development. In parallel, the CharIN initiative established a task force called the Megawatt Charging System Taskforce which aims to develop a new high power standard above 1 MW by 2023 for charging heavy-duty trucks, based on the combined charging system (CCS) standard . Prototype testing started in September 2020 . Tesla announced in late 2020 that it is working with third-parties to develop a standard for megachargers that can be provided to Semi truck owners. Tesla is one of five to have submitted a design to CharIN.

Industry experts addressing international standardisation are evaluating avenues to harmonise megacharger standards for mutual compatibility, in order to facilitate the roll-out of electric HFTs.

There are also regional efforts to develop megacharging infrastructure. Underpinned with stimulus funding , Iberdrola, a Spanish multinational electric utility, has expressed interest in installing megacharger infrastructure in heavy-duty freight truck corridors in Spain by 2025. ElaadNL (EV knowledge centre of Dutch grid operators), along with local and national government entities, in September 2021 launched an open-access test centre for companies and academia that offers test facilities for megachargers. In the United States, the West Coast Clean Transit Corridor Initiative aims to install charging sites capable of charging HDTs at 2 MW along key transit corridors from Mexico to the boder with Canada by 2030.

Electric HDVs data are derived from the Global Drive to Zero’s Zero Emission Technology Inventory (ZETI) which is a regularly updated tool that offers a detailed glimpse of announced OEM production model timelines. ZETI data are meant to support fleet operators and policy makers and should not be construed as representative of the entire vehicle market.

Reference 1

Subscription successful.

Thank you for subscribing. You can unsubscribe at any time by clicking the link at the bottom of any IEA newsletter.

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 31 January 2022

A circular economy approach is needed for electric vehicles

  • Jessika Luth Richter   ORCID: orcid.org/0000-0001-5786-5927 1  

Nature Electronics volume  5 ,  pages 5–7 ( 2022 ) Cite this article

15k Accesses

32 Citations

119 Altmetric

Metrics details

  • Interdisciplinary studies
  • Sustainability

Electric vehicles could help reduce greenhouse gas emissions and deliver a sustainable transport system. But the full life cycle of electric vehicles needs to be considered in order to avoid creating resource issues while trying to achieve the necessary climate goals.

The future of mobility — and road transport in particular — is a topic of increasing focus for governments across the globe as they seek to reduce greenhouse gas emissions in compliance with the Paris Agreement and limit other pollutants that cause negative health effects. The transition to more sustainable transportation, which is required to achieve such reductions, is fuelling a rise in electric vehicle sales that is likely to accelerate in the coming years. But while electrification of transport is certainly needed, manufacturing electric vehicles is inherently resource intensive and it remains unclear whether the necessary resources will be available, at least in the short term, to meet this increasing demand.

The steel and aluminium used in the transport sector represent around 17% 1 and 27% 2 of the global use of these materials respectively, most of it in vehicles; and the share of plastic in vehicle compositions continues to increase, bringing with it problems related to fossil fuel feedstocks 3 . Vehicles are also typically integrated with complex electronics (which is why the semiconductor chip shortage during the COVID-19 pandemic has led to delays in vehicle production). With electric vehicles, rare-earth magnets are required for many of the motor technologies, and other critical raw materials — including lithium, cobalt and graphite — are required for the batteries. Batteries account for a substantial proportion of the environmental impact of electric vehicles over their life cycle 4 : materials such as lithium require extraction from sensitive and unique environments, while materials such as cobalt have social risks linked to the mining, such as child labour and conflicts 5 .

How best to deal with vehicles when their driving days are done is also a concern. The high metal content makes a good business case for vehicle end-of-life recycling for the metals. Processes are also now refined enough to recover and reuse a high amount of the materials in traditional vehicles and their components, including batteries 6 . However, price fluctuations for secondary materials influence the business case and the increasing amount of residual material — which includes plastics, textiles, glass and critical metals — means that some material is still often wasted; the hazardous residual materials also pose risks if not properly managed.

To meet the resource and recycling challenges created by electric vehicles, the management of the end of life of such vehicles needs to start from the very beginning. A circular economy approach is, in particular, needed for electric vehicles in order to reduce their environmental impact and ensure that trade-offs are minimized in achieving the necessary climate goals.

Current strategies and policies

Circular strategies to deal with waste include reusing, repairing, refurbishing, remanufacturing and recycling (Fig. 1 ). There are already well-developed secondhand parts and car markets that can prolong vehicles’ lifetimes, and a key question will be about optimal lifetimes and circular strategies, the nuances of which will change over time. There is a need to balance the environmental costs of extending the lifetime of internal combustion engine vehicles and less efficient electric vehicles with the environmental gains of sending these to recycling earlier to reuse components and materials to create more efficient electric vehicles. And as with other products, extending the lifetime of existing vehicles risks a delay in the take-up of new electric vehicles 7 . As well as strategies for repair and maintenance, there are also strategies for intensifying the first use of electric vehicles by using the idle capacity of the vehicle through car sharing schemes.

figure 1

The useful lifetime of electric vehicles can be extended through repair and reuse. Refurbishing and remanufacturing can have lower environmental impact compared with manufacturing, and recycling materials can decrease the demand for mining of new primary materials.

Recycling strategies for electric vehicles, and the policies driving them, are only just beginning to deal with changing vehicle technologies. It has been over 20 years since the end-of-life vehicle directive was introduced in the European Union, which created extended producer responsibility for vehicle manufacturers 8 . In practice, extended producer responsibility requires producers to take back their products for reuse, recycling or remanufacturing. It aims to reduce the life-cycle environmental impact of waste vehicles and also incentivize eco-designs that consider reuse and recycling from the start. Beyond the European Union, countries including Japan, Korea and China also have highly centralized, regulated systems for managing end-of-life vehicles.

Batteries, and components containing critical raw materials, are a key focus in developing circular economy strategies. There are reuse strategies for extending the life of electric vehicle batteries, following a hierarchy of strategies to optimize life-cycle value through direct reuse, repurposing, refurbishment and remanufacturing 9 . Also, there are already markets for repurposing batteries in energy storage systems.

Challenges with circularity

Product policies to incentivize eco-design need to account for multiple, and sometimes conflicting, circular strategies in order to address resource issues. For example, substituting and decreasing the amount of critical materials can reduce the need for extraction, but it is also more difficult to recover small amounts of material later in recycling. Changing designs also creates a moving target, and there is a need for monitoring and flexibility of specific requirements; a key challenge for such policies is to balance a level of certainty with an ability to adapt to technological and market developments 10 .

So far, product policies have mainly focused on hardware products, but with increased electronic integration and vehicle autonomy, software can — and should — also support circularity measures including longer lifetimes, repair and maintenance. The increasing need for specialized diagnostic tools and software related to hardware has, for instance, been noted as an impediment to repair and maintenance, if they are not easily accessible at reasonable cost 11 .

Extended producer responsibility and waste management policies have focused on managing hazardous waste and general recycling at least costs. Having targets only on recycling has sometimes resulted in incentivizing recycling over reuse strategies. The focus on economies of scale and cost reduction has also resulted in trade-offs with the quality of recycled materials. This could be the case with electric vehicles, where the demand for materials for new vehicles results in the recycling of these materials rather than prolonging their use in existing electric vehicles.

Even with extended-producer-responsibility systems in place, the data available are insufficient to fully understand the fate of the components and materials from vehicle waste, and it is estimated that there is still a considerable amount of secondhand cars and waste exported outside countries and regions with extended-producer-responsibility policies 12 . Electric vehicles and their components will also likely be exported to capture their substantial reuse and remanufacturing potential in lower-income countries 13 . This enables longer lifetimes and access to electric vehicles in lower-income countries, but it can also result in waste where waste management infrastructure is poorly developed — and parallels can be drawn to electronic waste. The anticipated demand for the materials — potentially also driven by regulation — could drive investment in better waste management in low-income countries or incentivize the re-export of the waste to recycling facilities that can extract the valuable materials.

Waste electric vehicle batteries pose challenges in terms of fires and hazardous contamination, and the recovery of resources requires environmentally sound recycling 10 . Under idealized conditions, it has been estimated that recycling end-of-life electric vehicle batteries could provide 60% of cobalt, 53% of lithium, 57% of manganese and 53% of nickel needed globally in 2040 14 . But we are currently far from such an ideal scenario. Unlike lead acid batteries, which are profitable for recycling, the recycling processes for electric vehicle batteries are still developing, and this, combined with current low volumes, means that recycling is mainly driven by regulation. Extended producer responsibility in the European Union, and waste management regulations in countries including China, Japan and India, have also specifically targeted electric vehicle batteries, but there remains a lack of effective policy in much of the world 14 .

In 2020, the European Union proposed a new battery regulation 15 that seeks to address many of the issues surrounding electric vehicles and represents the most ambitious policy to date. It aims to increase transparency, traceability and accountability across the battery life-cycle, it requires access to battery management systems, and it mandates digital passports, carbon footprint declarations and maximum thresholds. In addition to traditional extended-producer-responsibility targets for collection and recycling, it has specific recycling rates for lithium, cobalt and nickel, while also specifying targets for the use of recycled materials in new batteries to incentivize demand. Imposing specific material recycling has also been suggested for extended-producer-responsibility legislation for end-of-life vehicles as a whole 16 .

It has been suggested that such ambitious rules for electric vehicle batteries may increase costs and slow the adoption of electric cars 17 . But, arguably, increased supply-chain transparency should be incentivized generally. In addition, the transition to electric vehicles should still be driven, in large part, by policies phasing out fossil fuel vehicles and these potential trade-offs highlight the need for policy mixes and a systems approach. Increased costs could also incentivize sufficiency measures in transport (that is, discussion of mobility needs and what needs are met by electric vehicles versus other transport options) and minimize rebound effects in the transition to electric vehicles 18 . Circular economy discussions have been criticized for not adequately considering sufficiency measures, costs and just distribution of these costs 19 — and these points will need to be carefully considered going forward.

Distribution of value, as well as distribution of costs, also needs to be discussed when designing circular economy policies and strategies for electric vehicles. Digital product passports are on political agendas and could be a key part of supplying better data, making supply chains more transparent and enabling collaboration between value-chain actors. Passports could potentially create recycling fees to follow the vehicle or component wherever it ends as waste and help with investment in sound waste and resource management systems. This could ensure that longer lifetimes for vehicles are not a trade-off with increased environmental impacts from poor waste management.

Circular economy strategies, in turn, also need to consider goals of climate policies. Limiting cross-border flows of used and waste vehicles in order to address leakage of circular economy value should be considered with regard to the needs of all countries to transition from fossil fuels. Such transfers of value can also help the development of sustainable mobility systems aligned with climate commitments. Specific country and regional economic competitiveness gains from circular economy strategies need to be balanced with the goals of addressing climate change globally. Ultimately, the transition to electric vehicles is likely to be more sustainable if approached from both climate and circular economy perspectives.

World Steel Association. Steel Facts (2019); https://go.nature.com/3qYV2WD

All About Aluminium. Aluminium Applications: Aluminium in Transport ; https://go.nature.com/3HG67mr

Esteva, L. C. A., Kasliwal, A., Kinzler, M. S., Kim, H. C. & Keoleian, G. A. J. Ind. Ecol. 25 , 877–889 (2021).

Article   Google Scholar  

Harper, G. et al. Nature 575 , 75–86 (2019).

Lèbre, É. et al. Nat. Commun. 11 , 4823 (2020).

Despeisse, M., Kishita, Y., Nakano, M. & Barwood, M. Procedia CIRP 29 , 668–673 (2015).

Marcus, J. S., Zachmann, G., Gardner, S., Tagliapietra, S. & Lykogianni, E. Promoting Product Longevity: How Can the EU Product Safety and Compliance Framework Help Promote Product Durability and Tackle Planned Obsolescence, Foster the Production of More Sustainable Products, and Achieve More Transparent Supply Chains for Consumers? (European Parliament, 2020).

European Parliament and Council. Directive 2000/53/EC of the European Parliament and of the Council of 18 September 2000 on End-of-Life Vehicles (2000); https://go.nature.com/3I2nf68

Albertsen, L., Richter, J. L., Peck, P., Dalhammar, C. & Plepys, A. Resour. Conserv. Recycl. 172 , 105658 (2021).

Baars, J., Domenech, T., Bleischwitz, R., Melin, H. E. & Heidrich, O. Nat. Sustain. 4 , 71–79 (2021).

Svensson-Hoglund, S. et al. J. Clean. Prod. 288 , 125488 (2021).

Sakai, S. et al. J. Mater. Cycles Waste Manag. 16 , 1–20 (2014).

Nasr, N. et al. Re-defining Value—The Manufacturing Revolution. Remanufacturing, Refurbishment, Repair and Direct Reuse in the Circular Economy (UN Environment Programme, 2018); https://go.nature.com/337U4PJ

Dunn, J., Slattery, M., Kendall, A., Ambrose, H. & Shen, S. Environ. Sci. Technol. 55 , 5189–5198 (2021).

EU Commission. Proposal for a Regulation of the European Parliament and of the Council Concerning Batteries and Waste Batteries, Repealing Directive 2006/66/EC and Amending Regulation (EU) No 2019/1020 (2020); https://go.nature.com/3Fo2wYF

Bhari, B., Yano, J. & Sakai, S. J. Mater. Cycles Waste Manag. 23 , 644–663 (2021).

Melin, H. E. et al. Science 373 , 384–387 (2021).

Font Vivanco, D., Freire-González, J., Kemp, R. & van der Voet, E. Environ. Sci. Technol. 48 , 12063–12072 (2014).

Corvellec, H., Stowell, A. F. & Johansson, N. J. Ind. Ecol. https://doi.org/10.1111/jiec.13187 (2021).

Download references

Acknowledgements

This work is funded by the Swedish Research Council for Sustainable Development (Formas) project ‘Circular Economy: capturing value in waste through extended producer responsibility policies’ (grant number 2017-01037).

Author information

Authors and affiliations.

Lund University, Lund, Sweden

Jessika Luth Richter

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Jessika Luth Richter .

Ethics declarations

Competing interests.

The author declares no competing interests.

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Richter, J.L. A circular economy approach is needed for electric vehicles. Nat Electron 5 , 5–7 (2022). https://doi.org/10.1038/s41928-021-00711-9

Download citation

Published : 31 January 2022

Issue Date : January 2022

DOI : https://doi.org/10.1038/s41928-021-00711-9

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

This article is cited by

Sustainable futures: from causes of environmental degradation to solutions.

  • Carla Sofia Ferreira Fernandes
  • Fátima Alves
  • João Loureiro

Discover Sustainability (2024)

Circular economy based approach for green energy transitions and climate change benefits

  • Amol Niwalkar
  • Tushar Indorkar
  • Rakesh Kumar

Proceedings of the Indian National Science Academy (2023)

Circular economy strategies for combating climate change and other environmental issues

  • Mingyu Yang
  • Pow-Seng Yap

Environmental Chemistry Letters (2023)

Policy recommendations to enhance circular economy of LIBs in an emerging economy

  • Tushar Gahlaut
  • Gourav Dwivedi

Environment Systems and Decisions (2023)

Impacts of shared mobility on vehicle lifetimes and on the carbon footprint of electric vehicles

  • Johannes Morfeldt
  • Daniel J. A. Johansson

Nature Communications (2022)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

electric vehicles research

  • Follow us on Facebook
  • Follow us on Twitter
  • Criminal Justice
  • Environment
  • Politics & Government
  • Race & Gender

Expert Commentary

Electric vehicles: 5 recent studies to inform your coverage

Long touted as a way to curb carbon emissions, electric vehicles are still a long way from dominating personal transportation in the U.S. Recent research highlighted here explores challenges ahead for the electric vehicle market.

electric vehicle

Republish this article

Creative Commons License

This work is licensed under a Creative Commons Attribution-NoDerivatives 4.0 International License .

by Clark Merrefield, The Journalist's Resource August 29, 2022

This <a target="_blank" href="https://journalistsresource.org/environment/electric-vehicles-research-roundup/">article</a> first appeared on <a target="_blank" href="https://journalistsresource.org">The Journalist's Resource</a> and is republished here under a Creative Commons license.<img src="https://journalistsresource.org/wp-content/uploads/2020/11/cropped-jr-favicon-150x150.png" style="width:1em;height:1em;margin-left:10px;">

With extreme drought in the Western U.S. causing water levels in the Colorado River to fall to historic lows and the Northeast also experiencing an extraordinarily dry summer, it’s a good time to explore recent research about electric vehicles — long touted as a key way to turn back the clock on climate change .

Carbon emissions increase the likelihood of extreme weather events, such as drought and wildfires. Electric vehicles “have large potential to reduce land-based transport [greenhouse gas] emissions,” according to a 2022 report from the Intergovernmental Panel on Climate Change, a scientific analysis organization headquartered in Geneva and endorsed by the United Nations.

The first electric cars date to the 1830s, though they weren’t ready for everyday use until the 1870s. By the early 1900s, electric vehicles made up one-third of all vehicles in America, according to the U.S. Department of Energy. The rise of Henry Ford’s gas-fueled Model T in the early 1910s spelled an end to the electric car for decades. The 1970s oil crisis renewed interest in electric vehicles. In 1996 General Motors released the EV1, a mass-produced all-electric car. In 2000 Toyota announced the U.S. launch of its Prius, the first mass produced hybrid vehicle, running on both gas and electric power. Tesla began delivering its first electric car, a sports car called the Roadster, in 2009, followed by the Model S luxury sedan in 2012.

Electric vehicles make up a small fraction of the roughly 261 million light-duty vehicles, including motorcycles, on U.S. roads. The Environmental Protection Agency defines light-duty vehicles as those weighing less than 8,500 pounds. Most passenger cars, SUVs, motorcycles and pickup trucks are considered light duty. Passenger cars and light-duty trucks account for 68% of U.S. road vehicle emissions and the transportation sector as a whole is the biggest greenhouse gas polluter, according to the latest federal government data.

There were about 2 million registered hybrid and fully electric vehicles in the U.S. in 2021, according to a 2022 report from the International Energy Agency, an independent intergovernmental organization based in Paris. “Some of the main drivers underpinning growth in the United States in 2021 were the increased production of Tesla models and the availability of new generation electric models by incumbent automakers,” according to the report. Combined sales of fully and partially electric vehicles doubled from 295,000 in 2020 to 631,000 last year — representing 5% of new vehicle sales, according to the agency.

Globally, hybrid and fully electric vehicles make up less than 1% of the 1.3 billion light-duty vehicles on the road today, according to the U.S. Energy Information Administration, which collects and publishes energy statistics and analysis. The administration projects that by 2050, there will be 672 million plug-in vehicles around the world, making up about one-third of the global vehicle fleet.

One barrier to widespread electric vehicle adoption is fleet turnover, which simply means new vehicles replacing old ones. New gas-powered cars can last years, even decades, before they are replaced by lower- or no-emissions vehicles. As the authors of one of the papers included below write, “even if [electric vehicle] market share jumped dramatically, it would take decades to replace the existing vehicle fleet, during which time vehicle [greenhouse gas] emissions would continue, worsening climate change.”

To reduce electric vehicle costs and speed production, the Inflation Reduction Act, which President Joe Biden signed into law on Aug. 16, provides tax credits to battery makers, potentially bringing down battery production costs by about one-third. Some states are also attempting to accelerate fleet turnover through policy action. California, for example, by 2035 will ban sales of new cars that run solely on gasoline.

The five studies featured here explore innovative ways to spur electric vehicle purchases, such as a revamped cash-for-clunkers program , which originally made national headlines in the U.S. during the Great Recession. Other topics include overviews of current charging infrastructure needs, considerations related to social and economic inequality, and the prospect of reusing car batteries to store solar power.

Inequality and the Future of Electric Mobility in 36 U.S. Cities: An Innovative Methodology and Comparative Assessment Patricia Romero-Lankao, Alana Wilson and Daniel Zimny-Schmitt. Energy Research & Social Science, September 2022.

The study: The authors explore social and economic inequalities in 36 U.S. cities, then discuss how those inequities can inform the rollout of transportation technologies, such as electric vehicles, so the broadest number of people can take advantage of them. They specifically explore how inequities play out for “wealthy, urban disadvantaged, urban renters, middle-class homeowners, and rural/exurban” groups.

The findings: The authors focus on 20 large metropolitan areas — each with a population of more than 1.5 million people — and 16 medium metropolitan areas — each with 500,000 to 1.5 million people — in California, Colorado, Illinois, Missouri, New York, Ohio, Oregon, Pennsylvania, Texas and the District of Columbia. They conclude that in order for electric vehicle rollouts to be successful, government policies will need to be tailored to the specific needs of each group. “For instance, rural/exurban populations might require electric carpooling,” the authors write. City dwellers, however, especially renters and households with lower incomes, would benefit more from electrified transit options rather than individually owned electric vehicles.

In the authors’ words: “Offering a variety of electrified options would provide benefits to the largest percentage of people and have more potential to sustainably decrease greenhouse gas emissions, reduce tailpipe emissions, and improve the health of people and ecosystems.”

Accelerating Vehicle Fleet Turnover to Achieve Sustainable Mobility Goals Sergey Naumov, David Keith and John Sterman. Journal of Operations Management, March 2022.

The study: The authors examine potential designs for cash-for-clunkers programs that would help turn over more of the nation’s vehicle fleet to low- or no-emissions vehicles. The most notable cash-for-clunkers program in the U.S., called the Car Allowance Rebate System, ran during the summer of 2009 with $3 billion in congressional allocations. People with eligible gas-guzzling cars received a credit to purchase newer, more fuel efficient vehicles. Engines in the old cars were disabled and the cars scrapped. The authors use a model of light duty vehicle turnover in the U.S., developed as part of prior research from two of the authors, and simulate cash-for-clunkers policies that would achieve significant vehicle turnover by 2050.  

The findings: Despite thousands of dollars in existing tax credits at the federal level and in some states for people buying zero-emissions vehicles, “alternative fuel vehicles have only achieved low single-digit market share in the United States to date,” the authors write. Their model assumes that patterns of car turnover will remain largely the same in the coming decades — with the average light vehicle having a useful life of 17 to 30 years — though they acknowledge that technological advances, like self-driving cars, could change those patterns. They explore several different program designs, including those in which people can get a cash credit for the purchase of an electric vehicle or traditional combustion engine vehicle; an electric or hybrid; or electric only. The authors find the design that would work best at turning over the existing vehicle fleet and reducing emissions is one that makes all gas-powered vehicles eligible, regardless of their age or fuel efficiency, and the credit reserved for those who replace their old car with a fully electric one. Government vehicle fleets being made fully electric would further spur the no-emissions vehicle market, they add.

In the authors’ words: “[Cash-for-clunkers] programs will also primarily benefit more affluent individuals who buy the majority of new cars, while low-income individuals tend to purchase used vehicles or forgo car ownership altogether, instead relying on public transportation. However, by accelerating fleet turnover, [cash-for-clunkers] policies speed reductions in harmful tailpipe emissions. The adverse health impacts of these emissions are disproportionately borne by the poor and especially by people of color.”

The State of Play in Electric Vehicle Charging Services — A Review of Infrastructure Provision, Players, and Policies Sarah LaMonaca and Lisa Ryan. Renewable and Sustainable Energy Reviews, February 2022.

The study: The authors explore current charging options for electric vehicle owners and policies that could spur the development of more charging infrastructure. They also discuss whether charging infrastructure is “a public good or private asset,” considering that “both the public and private sectors have been involved in the provision of charging stations and it is timely to consider the roles of the different actors in this market.”

The findings: Charging services have different cost structures, which make it hard for consumers to compare whether it’s cheaper to charge at home or at a public charging station. Some charging services charge by the minute, the hour, by energy used, or even offer subscriptions. The authors suggest that legislation could standardize how charging pricing is presented and “could help to improve the customer experience when searching out charging services.” To promote electric vehicle usage, federal and state governments could subsidize charging infrastructure, which is still expensive to build and maintain. Better data is needed on how and when people use public chargers, which federal or state governments could compel private providers to make public.

In the authors’ words: “Because the likely social benefits are not limited to those who can pay, decisions about this infrastructure are an important public policy concern and should not be just a matter for private firms and investors; it is therefore rarely fully privately-funded or owned.”

Quantifying the Emissions Impact of Repurposed Electric Vehicle Battery Packs in Residential Settings Alizer Khowaja, Matthew Dean and Kara Kockelman. Journal of Energy Storage, November 2021.

The study: The authors use detailed electrical grid generation data and electricity demand data from homes with solar power in Austin, Texas, to explore the potential emissions benefits of adding the storage power of old electric vehicle batteries to energy efficient homes. Based on previous environmental research, the authors note that drawbacks crop up throughout the life of an electric vehicle battery, including “the depletion of water tables during mining, high [greenhouse gas] emissions during battery manufacturing, e-waste due to a small percentage of batteries being recycled after operation, and contamination or exposure of toxic chemicals after disposal.”

The findings: Used electric vehicle batteries can hold from 60% to 80% of their original capacity, “which under favorable conditions could provide up to 10 years of second-life stationary [battery storage systems] at an economic savings of up to 60% compared to new storage systems,” the authors write. Assuming processes are in place to distribute and install old electric vehicle batteries to store solar power for homes, the authors find that such setups could reduce carbon emissions by one ton each year for each equipped house. By comparison, the average car emits roughly 4.6 tons of carbon per year, according to the Environmental Protection Agency.

In the authors’ words: “While the [greenhouse gas] savings for each household may appear negligible, the power of scale and rising social cost of CO2 may allow communities to considerably reduce their carbon footprint and transition both the power and transportation sector away from traditional fuel sources.”

There’s No Place like Home: Residential Parking, Electrical Access, and Implications for the Future of Electric Vehicle Charging Infrastructure Yanbo Ge, Christina Simeone, Andrew Duvall, and Eric Wood. National Renewable Energy Laboratory Technical Report, October 2021.

The report: The authors examine the types of homes that currently have charging access, noting that “there is uncertainty about how effectively home charging can scale as the primary charging location for electric vehicle owners.” They conducted an online survey from May 13 to May 31, 2020, asking 5,250 adults across the U.S. about their access to parking and electrical outlets. The authors use the survey results to project what the future of charging infrastructure might look like.

The findings: The authors find that one quarter of those surveyed have driveways or garages with electrical access, with another quarter reporting they could have such access installed. In one future scenario, the authors project that if every car on the road were electric, about one quarter of vehicles would lack at-home charging. They note that for electric vehicle ownership to extend beyond “high-income, single-family homes that have access to off-street parking,” city planners will need to consider how to provide charging infrastructure for people living in multi-family housing, such as apartment buildings, where overnight, off-street parking may not be easily available.

In the authors’ words: “In situations where residential off-street charging access is unattainable, a portfolio of solutions may be possible, including providing access to public charging in residential neighborhoods (on street), at workplaces, at commonly visited public locations, and (when necessary) at centralized locations via high power fast charging infrastructure (similar to existing gas stations).”

About The Author

' src=

Clark Merrefield

Read our research on: Gun Policy | International Conflict | Election 2024

Regions & Countries

How americans view electric vehicles.

A row of electric vehicles charge at a public station in San Ramon, California, in 2023. (Smith Collection/Gado via Getty Images)

About four-in-ten Americans (38%) say they’re very or somewhat likely to seriously consider an electric vehicle (EV) for their next vehicle purchase, according to a recent Pew Research Center survey .

Pew Research Center conducted this study to understand Americans’ views on electric vehicles. We surveyed 10,329 U.S. adults from May 30 to June 4, 2023.

Everyone who took part in the survey is a member of the Center’s American Trends Panel (ATP), an online survey panel that is recruited through national, random sampling of residential addresses. This way, nearly all U.S. adults have a chance of selection. The survey is weighted to be representative of the U.S. adult population by gender, race, ethnicity, partisan affiliation, education and other categories. Read more about the ATP’s methodology .

Here are the questions used for this analysis, along with responses, and its methodology .

A bar chart showing that Democrats, younger adults and urban residents are more open to purchasing an electric vehicle.

Half of U.S. adults say they are not too or not at all likely to consider purchasing an EV, while another 13% say they do not plan to purchase a vehicle. The share of the public interested in purchasing an EV is down 4 percentage points from May 2022.

Over the past year, the Biden administration has announced a range of measures aimed at increasing EV adoption, including tax credits for EV buyers and emissions limits for car manufacturers. Major automakers are increasing EV production , and electric vehicles’ share of all new U.S. car sales rose sharply over the past two years, to 8.5%.

Democrats and Democratic-leaning independents, younger adults, and people living in urban areas are among the most likely to say they would consider purchasing an EV. The 9% of U.S. adults who currently own a hybrid or electric vehicle are also particularly likely to consider an EV for their next purchase. A majority of this group (68%) says they are very or somewhat likely to seriously consider it.

Among those who would consider purchasing an EV, about seven-in-ten say helping the environment (72%) and saving money on gas (70%) are major reasons why. A small share (12%) cite keeping up with the latest trends in vehicles as a major reason.

Expectations for future electric vehicle infrastructure  

One potential obstacle to greater EV adoption is the availability of public charging stations.

A bar chart showing that Americans have low levels of confidence that the U.S. will build necessary EV infrastructure.

Currently, most EV owners charge their vehicles at home . Some who have used public chargers find that they are unreliable or limited in number. In September 2022, the Biden administration set aside $5 billion to create a network of EV charging stations .

Americans express limited confidence that the country will build the necessary infrastructure to support large numbers of EVs on the roads. Some 17% say they are extremely or very confident this will happen, while 30% are somewhat confident. And 53% are not too or not at all confident.

Republicans and GOP leaners are especially likely to doubt that the U.S. will build the charging stations and infrastructure needed to support EVs: 74% say they have not too much or no confidence at all in this. By comparison, 34% of Democrats and Democratic leaners say the same.

A bar chart showing that people who are confident U.S. will build charging infrastructure are more likely to consider an EV purchase.

Americans who are confident the country will build the necessary infrastructure are more likely than others to say they would consider purchasing an EV.

Among those who are extremely or very confident that the U.S. will build the infrastructure needed to support EVs, 68% say they would be at least somewhat likely to consider purchasing an EV.

Just 19% of those who are not too or not at all confident in future EV infrastructure say they are at least somewhat likely to consider purchasing an EV.

Views on phasing out gasoline cars and trucks

Line charts showing that support for phasing out new gasoline vehicles by 2035 has decreased over the past 2 years.

Accelerating the transition to EVs is a central part of President Joe Biden’s climate agenda. The administration has proposed new emission limits for automakers that would reduce the number of gas-powered cars and trucks they could sell. Some states have gone further, with plans to ban new gas-powered car sales by 2035. 

However, the idea of phasing out the production of new gas-powered vehicles by 2035 faces more public opposition than support. About six-in-ten Americans (59%) say they oppose this, while 40% favor it.

The share of Americans who favor phasing out gas-powered vehicles has declined 7 points since 2021. Support is down among both Democrats and Republicans.

A bar chart showing that 73% of Republicans say they would be upset if the U.S. stopped making new gasoline vehicles.

Currently, a majority of Democrats (64%) favor phasing out production of gas-powered vehicles by 2035, but 84% of Republicans oppose this.

Partisans also have different emotional reactions to the idea of ending gas-powered vehicle production. A clear majority of Republicans (73%) say they would feel upset about it, but views among Democrats are more mixed. Some 37% say they would feel excited, while 43% would feel neutral and 20% would be upset.

Note: This is an update to a post originally published June 3, 2021. Here are the questions used for this analysis, along with responses, and its methodology .

electric vehicles research

Sign up for our weekly newsletter

Fresh data delivered Saturday mornings

Home solar panel adoption continues to rise in the U.S.

Most americans support expanding solar and wind energy, but republican support has dropped, natural gas viewed more positively than other fossil fuels across 20 global publics, fast facts about u.s. views on oil and gas production as white house moves to open alaska refuge to drilling, renewable energy is growing fast in the u.s., but fossil fuels still dominate, most popular.

About Pew Research Center Pew Research Center is a nonpartisan fact tank that informs the public about the issues, attitudes and trends shaping the world. It conducts public opinion polling, demographic research, media content analysis and other empirical social science research. Pew Research Center does not take policy positions. It is a subsidiary of The Pew Charitable Trusts .

Site Logo

Recent Publications

Consumer response to transformative transportation innovations: Advancing the Reflexive Participant Approach to survey research https://authors.elsevier.com/sd/article/S2590-1982(23)00006-4

Patterns of Electric Vehicle Charging on Transportation Network Companies in the US https://authors.elsevier.com/a/1gYZK_UzsnI6C6

The costs and challenges of installing corridor DC Fast Chargers in California https://www.sciencedirect.com/science/article/pii/S2213624X23000238

Electrified autonomous freight benefit analysis on fleet, infrastructure and grid leveraging Grid-Electrified Mobility (GEM) model https://www.sciencedirect.com/science/article/pii/S0306261923001241

View more recent publications

ITS Updates

{{ item.title }}.

Electric Vehicles Research Service

North American OEMs Launch Their Own Charging Network to Fight Against a Tesla Monopoly

Latest Research

The EV Market Slowdown

The EV Market Slowdown

Securing the EV Supply Chain: Battery Recycling in the United States

Securing the EV Supply Chain: Battery Recycling in the United States

Key Takeaways From CES 2024

Key Takeaways From CES 2024

82 Technology Trends That Will—And Will Not—Shape 2024

82 Technology Trends That Will—And Will Not—Shape 2024

55 Technology Leaders To Watch In 2024

55 Technology Leaders To Watch In 2024

Related spotlights.

Spotlight

View All Research

Companies Covered

Battery technologies inc.

SAIC

Analog Devices Inc

European Commission

Ford Motor Company

Volkswagen group.

UPS

ADS-TEC Energy

General motors corporation.

ChargePoint

Analyst Support

Bonte-headshot

Electric Vehicles Coverage

Service Coverage Shield

Coverage Areas Include

  • Lithium Metal Battery Chemistries and Novel Alternatives
  • Silicon Anode
  • Battery Management Systems
  • Removing Rare Earth Metals and Scarce Materials
  • Scaling EV Battery Production and Cost Reduction
  • EV Propulsion
  • Smart Charging, Vehicle-to-Grid (V2G), and Bi-Directional Energy Flow
  • PEV versus Hydrogen Fuel Cell (HFC)-EV in Commercial Vehicles
  • Electrified Micromobility
  • Battery Lifecycle Management and Safety
  • Wireless Charging
  • Buffer Battery Charging and Microgrids
  • EV Supply Chain Consolidation
  • Battery Swapping

Deliverables

Latest news & resources for electric vehicles, filter by type, reports & data.

Research Report | 2Q 2024 | AN-6131

Research Report | 1Q 2024 | AN-5803

Whitepaper | 1Q 2024 | WP-1037

Whitepaper | 1Q 2024 | WP-1035

Whitepaper | 4Q 2023 | WP-1034

Accelerating EV Charging Rates

Research Report | 4Q 2023 | AN-5810

Electric Vehicle Charging Infrastructure

Market Data | 4Q 2023 | MD-EVCI-23

The Sustainable Future is Here. Are You Ready?

Whitepaper | 4Q 2023 | WP-1026

Electric Vehicles

Market Data | 3Q 2023 | MD-EV-102

Next-Generation Electric Vehicle Batteries

Research Report | 3Q 2023 | AN-5808

Chinese Electric Vehicle OEMs

Research Report | 3Q 2023 | AN-5809

Battery Management Systems for Electric Vehicles

Research Report | 2Q 2023 | AN-5807

Zero-Emission Vehicle Adoption in Fleets: Investments, Use Cases, Reduced Costs, Regulations, and Infrastructure

Research Report | 2Q 2023 | AN-5483

Market Data | 1Q 2023 | MD-EV-101

37 Technology Stats You Need to Know for 2023

Whitepaper | 1Q 2023 | WP-1013

Electric Vehicle Smart Charging Platforms

Research Report | 4Q 2022 | AN-5545

Bringing Electric Vehicles into the Mainstream

Research Report | 1Q 2022 | AN-5480

Connected Services for Electric Vehicles

Presentation | 2Q 2020 | PT-2331

Electric Vehicle Battery and Charging Technologies

Research Report | 3Q 2019 | AN-5152

Vehicle-to-Grid Technologies and Applications

Research Report | 3Q 2018 | AN-4995

North American OEMs Launch Their Own Charging Network to Fight Against a Tesla Monopoly

Insight | 2Q 2024 | IN-7297

Modern Wheels Coupled with Outdated Infrastructure: EVs Facing Increasing Cybersecurity Vulnerabilities

Insight | 1Q 2024 | IN-7259

VinFast Makes Headlines at CES 2024, but Supplier Beware

Insight | 1Q 2024 | IN-7228

CES 2024: Key Highlights from A Supply Chain Lens

Insight | 1Q 2024 | IN-7224

BYD Overtakes Tesla, Presenting a Growing Threat to Legacy OEMs

Insight | 1Q 2024 | IN-7218

The EV Market in Indonesia: A Push to Become a Leading Manufacturing Base in the Southeast Asian Region

Insight | 1Q 2024 | IN-7188

Stellantis Follows Volkswagen by Buying a Stake in a Chinese OEM, but Not for the Same Reasons

Insight | 4Q 2023 | IN-7165

Tesla Revamps the Model 3, Staying Ahead of the Competition in the West but Fighting for Market Share in China

Insight | 4Q 2023 | IN-7090

CarbonScape Invests US$18 Million in Biographite Production in Europe and the United States to Tackle Material Scarcity for EVs and Battery Storage

Insight | 3Q 2023 | IN-7087

CCS Is Dead in North America as Tesla’s NACS Plug Takes Over

Insight | 3Q 2023 | IN-7005

Dassault Systèmes’ 2023 Analyst Days: The Strategy Is to Focus on Experiences, Outcomes, and Health

Insight | 2Q 2023 | IN-6976

JLR Partners with Elektrobit to Enable Cutting-Edge In-Car Digital Experiences in Its Next-Generation Vehicles

Insight | 2Q 2023 | IN-6955

NIO Takes On Tesla and Aims to Disrupt the Global Fast Charging Market with Its Latest Battery Swap Technology

Insight | 2Q 2023 | IN-6908

The European Union Will Ban All Polluting Cars from 2035. Who’s Ready for It?

Insight | 1Q 2023 | IN-6868

Supply Chain Challenges to Curtail Electric Vehicle Production over the Next Decade

Insight | 3Q 2022 | IN-6625

Why are Electric Vehicles at the Forefront of Vehicle Tech?

Insight | 2Q 2022 | IN-6570

Identifying and Addressing the Challenges to Mainstream Electric Vehicles

Insight | 1Q 2022 | IN-6399

How Electric Vehicle Charging Systems are Accelerating the Clean Energy Revolution

Insight | 1Q 2021 | IN-6034

Subscribe now for data-driven insight and expert analysis to fuel your business.

twitter

Cyber & Digital Security

Industrial & manufacturing, sustainable technologies, all other services.

  • Resource Center
  • Journalist Inquiry
  • ABI Research News
  • Management Team
  • Office Locations

©2024 Allied Business Intelligence, Inc. All Rights Reserved

  • 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
  • What's My Car Worth?
  • Buyer's Guide

2024 Fiat 500e Might Be All the EV You Need

In its role as an urban commuter or as a complement to a larger SUV, the 500e hits its mark with solid dynamics and a panoply of standard equipment.

2024 fiat 500e

Staggeringly Normal Specs

Most modern EVs make enough horsepower to put generation-old sports cars to shame, and that's not exactly what everyone wants or needs. The 2024 500e is more of an anachronism in this sense, returning to the days of small, inexpensive cars that provide tepid acceleration. Here, a single permanent-magnet electric motor powers the front axle and produces just 117 horsepower and 162 lb-ft of torque. Fiat claims the 500e will meander its way to 60 mph in 8.5 seconds before topping out at 94 mph. And while that pales next to the dual-motor Hyundai Ioniq 5, which gets to 60 in 4.5 seconds , it is quicker than a Chevy Trax or a Kia Sportage .

2024 fiat 500e

Under the body lies a similarly modest lithium-ion battery, which we estimate has roughly 37 kWh of usable capacity and will be good for 140 to 150 miles of range. Our projection would've been lower, but the 500e's compact dimensions mean Fiat was able to keep the curb weight at a respectable claimed 2952 pounds, a far cry from those 5000-plus-pound Chunka Lunkas rolling around. The battery will charge from empty to 100 percent in a claimed six hours on a 6.6-kW Level 2 charger or in less than 4.5 hours on an 11.0-kW connection, and at its max DC fast-charge rate of 85 kW, it will refill to 80 percent in 35 minutes, Fiat says.

Aesthetically, this little one is clearly a Fiat 500. The exterior picks up some LED lighting front and rear, as well as flush electronic door handles. But the flashiest of the new stuff as well as the throwbacks lie in the cabin. The dashboard trim, rounded gauge cluster, and two-spoke steering wheel are meant to evoke the OG 1957 Cinquecento . We dig the dedicated wireless-charging nook just below the (physical!) climate controls and 10.3-inch center display, which runs the latest version of Stellantis's Uconnect 5 software. A 7.0-inch digital gauge cluster is a nice thing to see, especially as fellow small-car manufacturer Mini seems intent on eliminating that feature in favor of a cheapo head-up display.

2024 fiat 500e

While the eyes may deceive, the 500e is a bit larger than it was before, ringing in nearly an inch longer in both wheelbase and overall length and also 2.2 inches wider. An additional 1.7 inches of shoulder room in the front row keeps the Fiat from feeling truly cramped, and the cargo hold will swallow eight cubic feet of stuff, or a few backpacks or several bags of groceries, with ease. The back seat will fit an adult, though not for long journeys—but then, you're not going on a long journey, are you? Behind a seat set for a six-foot-tall driver, there was decent headroom, but the legroom was more reminiscent of a budget European airline than a car.

Despite the small accommodations, the center console has a decently sized hidey-hole that could hold a small tablet. Two USB-A ports and another USB-C reside in various crannies. There's only one cupholder that isn't integrated into a door panel, but it can be folded and stowed when not needed, opening space in the cabin's lower half.

In addition to the tech, there's a bunch of other standard kit in this package. LED headlights, automatic climate control, keyless entry and start, wireless device charging and smartphone mirroring, and rain-sensing wipers all are included. Also standard is the buyer's choice of a Level 2 home charger or fast-charge credits through the company's Free2move Charge program.

2024 fiat 500e

Staggeringly Normal Demeanor

While the Fiat 500e's vibe might fit our Miami drive location well, this locale is not exactly suited for, well, driving. Endless parades of stoplights, confused drivers piloting rented convertibles (with the top still up, natch), strips of asphalt where lunar-rover testing clearly takes place (or should)—Vice City has it all. The 500e's ride was, as we expected, a little on the flinty side, but we found very little unwanted interior noise until higher speeds. Chucking this little guy into corners at normal speeds made for a fun urban jaunt, with an appropriate amount of body roll for a small but kind of tall car.

It's fortunate that the 500e feels zippy at a relatively modest pace, because that's all the pace the Italian jelly bean can muster. There's a good bit of right-pedal sensitivity at lower speeds in both the standard Normal mode and the more efficient Range mode, and the zero- to 30-mph span is what matters most in a car of this ilk. Switch into Range mode, and the pedal does require a smidge more prodding to get going, but the increased regenerative braking permits one-pedal driving, so it's the mode we preferred. If you're not a fan of regen, Normal mode's coasting and braking feel like any other small car's. There's also a Sherpa mode that limits the top speed to 50 mph and cuts max motor output to 76 horsepower. It was explained as an "Oh crap, I have less battery than I thought and need to get home" mode, not one for normal use.

2024 fiat 500e

Dollars and Sense

Achieving a $34,095 base price is not the easiest thing to do, especially at a time when the average transaction price for a new car is approaching the $50,000 mark, and that becomes clear in places like the torsion-beam rear axle and rear drum brakes. Unlike with Fiat's last foray into battery-electric models, parent company Stellantis probably won't lose the farm on every car it sells this time, despite that the new 500e's MSRP is lower than its 2019 forebear's, even with years having passed between their debuts.

While it won't solve the problem of charging access for those who park on the street, the 2024 Fiat 500e does help address one issue the EV space needs to work on: affordable variety. It's a true city car, with the thrift and capability needed for most weekly forays, and it works well as a second around-town-mobile when long trips aren't on the docket. It's a turtle among hares, but if that's all you need, why go overboard?

Specifications

2024 Fiat 500e Vehicle Type: front-motor, front-wheel-drive, 4-passenger, 2-door hatchback

PRICE Base: Inspi(Red), $34,095; Music, $37,595; Beauty, $37,595

POWERTRAIN Motor: permanent-magnet synchronous AC Power: 117 hp Torque: 162 lb-ft Battery Pack: liquid-cooled lithium-ion, 37.0 kWh ( C/D est) Onboard Charger: 11.0 kW Peak DC Fast-Charge Rate: 85 kW Transmission: direct-drive

DIMENSIONS Wheelbase: 91.4 in Length: 143.0 in Width: 66.3 in Height: 60.1 in Passenger Volume, F/R: 48/29 ft 3 Cargo Volume: 8 ft 3 Curb Weight ( C/D est): 2950 lb

PERFORMANCE ( C/D EST) 60 mph: 8.2 sec 1/4-Mile: 16.5 sec Top Speed: 94 mph

EPA FUEL ECONOMY ( C/D EST) Combined: 116–120 MPGe Range: 140–150 mi

Headshot of Andrew Krok

Cars are Andrew Krok’s jam, along with boysenberry. After graduating with a degree in English from the University of Illinois at Urbana-Champaign in 2009, Andrew cut his teeth writing freelance magazine features, and now he has a decade of full-time review experience under his belt. A Chicagoan by birth, he has been a Detroit resident since 2015. Maybe one day he’ll do something about that half-finished engineering degree.

preview for HDM All sections playlist - Car & Driver US:

.css-190qir1:before{background-color:#000000;color:#fff;left:0;width:50%;border:0 solid transparent;bottom:48%;height:0.125rem;content:'';position:absolute;z-index:-10;} First Drives .css-188buow:after{background-color:#000000;color:#fff;right:0;width:50%;border:0 solid transparent;bottom:48%;height:0.125rem;content:'';position:absolute;z-index:-10;}

2024 toyota land cruiser

Aston Martin DB12 Volante: Sunshine and Substance

2025 porsche taycan turbo gt

2025 Porsche Taycan Turbo GT Goes to Extremes

2025 porsche taycan

2025 Porsche Taycan Sees Big Range and Power Boost

2024 jeep gladiator mojave x

2024 Jeep Gladiator Mojave Keeps On Keepin' On

2025 volvo ex30 ev

The 2025 Volvo EX30 and the Scandinavian Flick

maserati gt2

Maserati's GT2 Race Car Will Be Your Best Friend

2024 mercedes benz s350d

Forgotten Star: 2024 Mercedes-Benz S-Class Diesel

2025 honda cr v efcev

Honda CR-V e:FCEV Is a Novel Take on Hydrogen

2024 kia sorento x pro

2024 Kia Sorento X-Pro Is Movin' On Up

2025 maserati grecale folgore ev

2025 Maserati Grecale Folgore EV: More Mainstream

2024 lincoln nautilus

2024 Lincoln Nautilus Is a Quiet Revolution

  • POWER Plant ID
  • POWER Events
  • Connected Plant
  • Distributed Energy
  • International
  • COVID-19 Coverage
  • Carbon Capture
  • Climate change
  • Cybersecurity
  • Distributed Power
  • Electric Vehicles
  • Energy Storage
  • Environmental
  • Instrumentation & Controls
  • Legal & Regulatory
  • Legislative
  • Ocean/Marine
  • Physical security
  • Plant Design
  • Power Demand
  • Research and Development
  • Supply Chains
  • Tidal Power
  • Waste to Energy
  • About POWER
  • Privacy Policy
  • Diversity, Equity, Inclusion & Belonging
  • Accessibility Statement

New modification of Russian VVER-440 fuel loaded at Paks NPP in Hungary

DECEMBER 14, 2020 — After the recent refueling at power unit 3 of the Hungarian Paks NPP, its VVER-440 reactor has been loaded with a batch of fresh fuel including 18 fuel bundles of the new modification. The new fuel will be introduced at all four operating power units of the Paks NPP, and the amount of new-modification bundles in each refueling will be increased gradually.

Development of the new VVER-440 fuel modification was completed in 2020 under the contract between TVEL JSC and MVM Paks NPP Ltd. Its introduction would optimize the hydro-uranium ratio in the reactor core, enabling to increase the efficiency of fuel usage and advance the economic performance of the power plant operation. All VVER-440 fuel modifications are manufactured at the Elemash Machine-Building Plant, a facility of TVEL Fuel Company in Elektrostal, Moscow Region.

Paks Nuclear Power Plant

“Introduction of a new fuel is an option to improve technical and economic performance of a nuclear power plant without substantial investment. We are actively engaged in development of new models and modifications of VVER-440 fuel for power plants in Europe. The projects of the new fuels for Loviisa NPP in Finland, Dukovany NPP in the Czech Republic, Mochovce and Bohunice NPPs in Slovakia, are currently at various stages of implementation. Despite the same reactor model, these projects are quite different technically and conceptually, since we take into account the individual needs and requirements of our customers,” commented Natalia Nikipelova, President of TVEL JSC.

For reference:

The project of development and validation of the new fuel has been accomplished with participation of a number of Russian nuclear industry enterprises, such as OKB Gidropress (a part of Rosatom machine-building division Atomenergomash), Bochvar Institute (material science research facility of TVEL Fuel Company), Elemash Machine-building plant and Kurchatov Institute national research center. At the site of OKB Gidropress research and experiment facility, the new fuel passed a range of hydraulic, longevity and vibration tests.

Paks NPP is the only functioning nuclear power plant in Hungary with total installed capacity 2000 MWe. It operates four similar units powered by VVER-440 reactors and commissioned one by one in 1982-1987. Currently, Paks NPP is the only VVER-440 plant in the world operating in extended 15-monthes fuel cycle. The power plant produces about 15 bln kWh annually, about a half of electric power generation in Hungary. In 2018, the project of increasing the duration of Paks NPP fuel cycle won the European competition Quality Innovation Award in the nomination “Innovations of large enterprises”. Russian engineers from TVEL JSC, Kurchatov Institute, OKB Gidropress, Bochvar Institute and Elemash Machine-building plant provided assistance to the Hungarian colleagues in accomplishment of the project.

  TVEL Fuel Company of Rosatom incorporates enterprises for the fabrication of nuclear fuel, conversion and enrichment of uranium, production of gas centrifuges, as well as research and design organizations. It is the only supplier of nuclear fuel for Russian nuclear power plants. TVEL Fuel Company of Rosatom provides nuclear fuel for 73 power reactors in 13 countries worldwide, research reactors in eight countries, as well as transport reactors of the Russian nuclear fleet. Every sixth power reactor in the world operates on fuel manufactured by TVEL.  www.tvel.ru  

  • Share full article

Advertisement

Supported by

Guest Essay

EU-India join forces to promote start-up collaboration on E-Vehicles Batteries under Trade and Technology Council

The European Union (EU) and India today launched expressions of interest for start-ups working in Electric Vehicles (EVs) Battery Recycling to participate in a matchmaking event. This initiative takes place under the India-EU Trade and Technology Council (TTC) announced by the Prime Minister of India, Shri Narendra Modi and Ms Ursula von der Leyen President of the European Commission, at their meeting in New Delhi in April 2022. It aims to enhance the cooperation between European and Indian SMEs and start-ups in the clean and green technologies sector. The intended exchange of knowledge and expertise will be instrumental in advancing the circularity of rare materials and transitioning towards carbon-neutrality in both India and the EU.

Funded by the EU and supported by the Office of the Principal Scientific Adviser to the Government of India, the proposed matchmaking initiative seeks to facilitate cooperation and partnerships between solution providers and adopters from both regions.

Open to start-up innovators from the European Union and from India, both calls for Expression of Interest close on 30 April. 

Twelve startups from Europe and India (six each) will be selected and get a pitching opportunity during a joint EU-India matchmaking event in June 2024. The online event will take place in the presence of Professor Ajay Kumar Sood, the Principal Scientific Adviser to the Government of India, and Mr. Marc Lemaître, Director-General for Research and Innovation at the European Commission. Six finalists (3 from the EU and 3 from India) out of the twelve shortlisted candidates will be selected following their pitching presentations in the matchmaking events. The six EU and Indian selected start-ups will be awarded with a visit to India and the EU respectively, to explore market opportunities with their counterparts. 

This initiative is a direct outcome of the EU-India TTC, a collaboration platform to address key trade, trusted technology, and security challenges. Under the Green and clean energy technologies Working Group, recycling of batteries for e-vehicles and standards through pre-normative research is one of the key identified areas for enhanced cooperation.

Mr. Marc Lemaitre , emphasized the need for collaboration dedicated to the EV Battery Recycling industry:

"The matchmaking event is a step-ahead to unlock innovative possibilities leading to a green and circular economy. We encourage innovators from the EU to seize this opportunity and explore potential collaborations with their Indian counterparts."

Prof. Ajay Kumar Sood  said:

“The matchmaking event under India-EU TTC Working Group 2 offers Indian startups an exclusive platform to demonstrate their expertise in battery recycling technologies. It provides a chance for Indian innovators to establish strategic alliances with their counterparts in the EU, accelerating the development of advanced battery recycling techniques focused on waste minimization and resource sustainability. Our objective is to harmonize efforts with EU innovators to jointly develop battery recycling solutions that drive industry expansion. We are dedicated to fostering a collaborative environment where sustainability and innovation form the cornerstone of a flourishing circular economy.”

Mr. Hervé Delphin , Ambassador to the EU Delegation to India remarked:

"Cooperation on research and innovation is key to unlock new and sustainable technologies to the market.  The development of recycling battery technologies for E-Vehicles presents opportunities to reduce CO2 emissions and supply chains dependencies. This initiative under the Trade and Technology Council will mobilise EU and India vibrant start-ups ecosystems to identify and design innovative solutions. This collaboration is a concrete example of what the TTC can deliver for the benefit of EU and India.”

Mr. Saurabh Kumar , Ambassador of India to Belgium, Ambassador to Luxembourg and the European Union said:

“This matchmaking event provides an outstanding platform to dynamic start-ups on both sides to forge strategic partnerships and explore investment avenues in EV Battery Recycling Technologies. It reflects the shared and deep commitment of India and the EU to a sustainable and greener future. The event is part of a broader effort to promote innovation and forge stronger economic relations under the India-EU Trade and Technology Council. It is my belief that this initiative will be a foundation stone for many more collaborations between India and the EU.”

This initiative underscores the broader commitment of the European Union to fostering global partnerships for a greener future.

The TTC is a key forum to deepen the strategic partnership on trade and technology between the two partners. Geostrategic challenges have reinforced the EU and India's common interest in ensuring security, prosperity and sustainable development based on shared values. It will help increase EU-India bilateral trade, which is at historical highs, with €120 billion worth of goods traded in 2022. In 2022, €17 billion of digital products and services were traded.

The TTC consists of three Working Groups:

  • Working Group 1 on Strategic Technologies, Digital Governance and Digital Connectivity
  • Working Group 2 on Green and Clean Energy Technologies
  • Working Group 3 on Trade, Investment and Resilient Value Chains

The first ministerial meeting of the TTC took place in Brussels on 16 May 2023, co-chaired by Executive Vice-Presidents Margrethe Vestager and Valdis Dombrovskis on the EU side, and on the Indian side by Subrahmanyam Jaishankar, Minister of External Affairs; Piyush Goyal, Minister of Commerce and Industry; and Rajeev Chandrasekhar, Minister of State for Skill Development and Entrepreneurship and Electronics and Information Technology. They were joined by High Representative/Vice-President Josep Borrell, as well as Commissioner for Internal Market, Thierry Breton.

Working Groups are now jointly working to advance identified objectives and key actions. The matchmaking event launched today is one of the agreed short-term actions under Working Group 2 on Green and Clean Energy Technologies.

More information

Calls for expression of interest

International cooperation with India

Press contact:

EC Spokesperson for Research, Science and Innovation

Share this page

Realtor.com Economic Research

  • Data library

2024 Realtor.com Housing Market and Electric Vehicle Report

Jiayi Xu

  • Markets with higher electric vehicle ownership rates see higher shares of EV-friendly homes for sale. San Jose, CA, saw the highest share of EV-friendly listings on Realtor.com® in 2023 (4.9%), five times as high as the national share of 0.9%. 
  • Markets with higher EV ownership rates and crowded public charging facilities will see higher demand for EV-friendly homes, such as Oxnard, CA; Stockton, CA; Riverside, CA; Urban Honolulu, HI; and Portland, OR.
  • Markets with a higher share of EV-friendly listings and less crowded public charging facilities are considered EV-friendly housing markets. Specifically, San Jose, CA; Salt Lake City, UT; San Francisco, CA; Boston, MA; and Seattle, WA, topped the list of most EV-friendly housing markets. 
  • 31.7% of EV-friendly homes listed on Realtor.com in 2023 were built after 2020, but a notable share of 17.8% were built before 1970.
  • 29.6% of EV-friendly listings are homes without garages. 
  • Renters living in multifamily dwellings face bigger challenges in getting access to near-home chargers, but luxury rental communities are more likely to be EV-friendly.

Electric vehicles (EVs) are important for tackling climate risks because they provide a cleaner and greener alternative to regular cars. 1   As consumer demand for EVs has risen significantly over the past few years , it will not be surprising to expect an increasing demand for EV-friendly homes in the housing market.

In honor of Earth Day 2024, Realtor.com® collaborated with Cox Automotive Inc. to study the current landscape of EV-friendly homes and to identify markets with higher demands for EV-friendly homes. For this research, we define EV-friendly homes as those featuring at-home EV chargers or equivalent attributes explicitly in their listing descriptions on Realtor.com.

Markets with higher EV ownership rates see higher share of EV-friendly homes for sale

An important f actor determining the number of EV-friendly homes for sale is the ownership rates of electric vehicles (Figure 1).

For example, San Jose-Sunnyvale-Santa Clara, CA , where 1 in 5 households has an electric vehicle, saw the highest share of EV-friendly homes listed in 2023 (4.9%) among all the metros. In addition, in various California metros such as San Francisco, Los Angeles, Oxnard, San Diego, and Santa Cruz, approximately 1 in 10 households owns an electric vehicle, and EV-friendly home listings are also above average.

Figure 1: Markets with higher EV ownership rates see higher share of EV-friendly home listings, top 300 metros 

electric vehicles research

In fact, 0.9% of for-sale homes listed on Realtor.com in 2023 were described as EV-friendly. While the share was slightly below 1%, it has been growing rapidly as the rate was only 0.1% five years ago (Figure 2).

In addition to Californian metros as mentioned above, Boulder, CO (3.4%); Seattle, WA (3.3%); Bloomington, IL (2.2%); Urban Honolulu, HI (2.1%); Bend, OR (2.1%); Trenton, NJ (2.0%); and Austin, TX (2.0%), all saw a higher than average share of EV-friendly homes listed for sale. However, large metros such as Birmingham, AL (0.1%); Milwaukee, WI (0.2%); and Memphis, TN (0.2%), saw a lower than average share of EV-friendly listings in 2023.

Figure 2: Share of EV-friendly homes among homes listed on Realtor.com

electric vehicles research

Markets with crowded public charging stations could see higher demand for EV-friendly homes

In addition to the growing use of electric vehicles, the limited availability of public charging stations will drive the demand for EV-friendly homes. Therefore, we consider public charging stations as potential alternatives for at-home charging. These include charging stations with ports accessible to the general public, such as those in shopping mall parking lots and those at workplace settings where only employees have access.

Specifically, if public charging stations are frequently congested, electric vehicle owners are more likely to demand at-home charging to avoid long waiting times. To measure this effect, we calculated the ratio of EVs to public charging ports within each metro, interpreted as a congestion index. 2 A higher index value indicates more crowded public charging stations, strengthening the case for prioritizing in-home charging and potential higher demand for EV-friendly homes.

Within the top 100 metros, the congestion index varies from 4 to 62, with an average index of 19. Notably, Albany, NY, exhibits the lowest ratio, with just 4 electric vehicles per public charging port, while McAllen, TX, experiences the highest ratio, with 62 EVs per public charging port.

Table 1 shows the list of metros with the highest congestion index, suggesting higher potential demand for at-home charging. 

Table 1: Metros with the most crowded public charging ports, top 100 metros

However, some of the markets with the most crowded public charging infrastructure have relatively low overall levels of EV ownership.

Specifically, i t is interesting to see metros such as Oxnard, CA; Stockton, CA; Riverside, CA; Urban Honolulu, HI; and Portland, OR, pop up on the list where we see higher EV ownership rates and crowded public charging facilities, suggesting higher demand for at-home chargers.

Top EV-friendly housing markets

Table 2 shows a list of the top EV-friendly housing markets among the top 100 metros, characterized by a favorable combination of EV-friendly housing supply and demand.

The supply of EV-friendly housing is measured by the share of EV-friendly homes listed on Realtor.com in 2023, and the demand is assessed by the congestion index as calculated by EVs per public port. A higher share of EV-friendly listings implies greater availability of homes equipped with at-home chargers on the market. However, a higher congestion index signals a demand for more EV-friendly homes due to crowded public charging facilities, indicating demand surpasses supply.

The top EV-friendly markets were determined by calculating percentile levels for each metric and subsequently deriving a weighted average between the two to come up with an EV-friendly housing score for each market, forming the basis for market rankings.

Table 2: Top EV-friendly housing markets

electric vehicles research

It is very interesting to see non-California markets like Salt Lake City, UT, near the top of the list. The higher share of EV-friendly listings and the less crowded public charging facilities suggest that the housing market in Salt Lake City has a great combination of supply and demand for EV-friendly homes.

In fact, Salt Lake City, UT; Denver, CO; and Durham, NC, are the metros with the highest share of young households (with head of household aged 25 to 34). In addition, many of these top markets either are important tech hubs or have rising tech sectors. 

At-home EV charging is accessible to many older homes

The significance of having an at-home EV charger is growing for both homebuyers and sellers, particularly in markets with a higher prevalence of EVs but limited public charging infrastructure. For buyers, the availability of home charging options can save considerable time and money, and enhance overall convenience. Sellers stand to benefit by showcasing an EV home charger as a notable feature, attracting a broader pool of potential buyers. However, a common misconception about home chargers is that they are exclusively applicable to recently built homes. 

It is generally true that at-home EV chargers are more common among newer homes. The typical EV-friendly home was just 6 years old versus 29 years for the typical home without such a feature. 3 In addition, among all the EV-friendly homes listed in 2023, 31.7% of them were built after 2020 and 47.3% of them were built in or after 2010. Put another way, nearly one-third of EV-friendly home listings in 2023 were built within the past four years while nearly half were built in the past 14 years.

Specifically, among all these EV-friendly homes built in or after 2010, the greatest share of them were in Riverside, CA (5.6%); Atlanta, GA (5.1%); Los Angeles, CA (5.0%); and Seattle, WA (4.9%).

Looking ahead, there will be more new homes that are EV-friendly, especially driven by supportive policies promoting EV adoption. For example, as of September 2022, California building codes require that “new one- and two-unit single family dwellings or townhouses with attached private garages must have electrical conduit installed that is capable of supporting a Level 2 EV charging station.”

Figure 3: While more common among newer homes, properties of any age can be EV-friendly

electric vehicles research

Nevertheless, it does not mean homeowners of older homes will not have access to home charging. In fact, 17.8% of EV-friendly homes listed on Realtor.com in 2023 were built before 1970 (i.e., 54 years ago).

Specifically, among all these older but EV-friendly homes, the greatest shares were in Los Angeles, CA (14.5%); San Francisco, CA (8.3% ); New York, NY (7.0%); and Washington, DC (5.3%). While it might be expensive to upgrade the electricity system in older homes, many states provide funds or rebates for energy improvements , including purchasing and installing electric vehicle supply equipment (EVSE), which could help homeowners save some costs. 

Garages are common for EV-friendly homes but are not a must

Owning a home with a garage appears to enhance the practicality of having an in-home charger, with approximately 70.4% of homes deemed EV-friendly on Realtor.com in 2023 featuring “garage” in their listing descriptions.

Notably, the significance of having a garage is more pronounced in EV-friendly homes located in markets with colder climates, such as Colorado Springs, CO; Chicago, IL; and Denver, CO, where over 90% of the listed EV-friendly properties include garages. Meanwhile, warmer metros such as Cape Coral, FL; Oxnard, CA; Phoenix, AZ; and Riverside, CA, also see around 80% of EV-friendly listings equipped with garages.

Figure 4: Markets where EV-friendly listings often include garages

electric vehicles research

Nonetheless, the absence of a garage does not imply a lack of access to at-home chargers, as 29.6% of EV-friendly listings are without a garage.

Specifically, homes without garages accounted for 57.5% of EV-friendly listings in Urban Honolulu, HI; 52.5% in Miami, FL; 46.4% in Seattle, WA; 45.7% in San Antonio, TX; 41.9% in New York, NY; and 41.5% in Boston, MA.

Figure 5: Markets where EV-friendly listings are often without garages 

electric vehicles research

Renters face bigger challenges in accessing EV charging facilities

Renters (0.25%) are more than three times less likely than homeowners (0.87%) to own an electric vehicle, according to an empirical study from the University of California, Berkeley.

One important factor leading to this gap is the accessibility of charging facilities. Unlike the homeowners who can charge at home, renters, especially those who live in apartment buildings, largely rely on public charging facilities. While chargers at workplaces and public parking lots could be good options for many, a recent University of California, Los Angeles report found that residents who live in multifamily dwellings showed strong preferences for near-home charging. 

Some rental communities are responding to consumer demand for EV charging facilities. Renters in Austin, TX (1.7%), and Los Angeles, CA (1.7%), faced the largest share of EV-friendly rental community listings on Realtor.com in 2023 across the top 50 metros, followed by San Jose, CA (1.4%); Orlando, FL (1.4%); Atlanta, GA (1.3%); Phoenix, AZ (1.1%); San Francisco, CA (1.0%); and San Diego, CA (1.0%). However, there are few details about how many charging ports or stations are available within these rental communities, which limits the ability of potential tenants to use this information to make decisions. In addition, EV-friendliness is a more popular feature among rental communities with “luxury” included in their listing descriptions.

For example, 6.7% of luxury rental communities in Los Angeles, CA, listed in 2023 were considered EV-friendly, and the EV-friendly share was 5.5% among luxury rental communities in Austin, TX, and San Francisco, CA.

Meanwhile, things might get slightly easier for renters living in new multifamily homes. As of September 2022, building codes require new multifamily dwellings in California to have 10% of parking spaces be EV-capable and 25% of parking spaces be EV-ready. New multifamily dwellings with more than 20 units must install Level 2 EV charging stations in 5% of all parking spaces.

Methodology

The homes considered in this study are single-family homes and condos/townhomes/row homes/co-ops listed for sale on Realtor.com in 2018–23. EV-friendly homes are properties featuring terms such as “electric vehicles” and “240-volt outlet” in the listing descriptions.

To understand the relations between EV ownership rates and EV-friendly homes, property-level data was consolidated into metro-level and matched to electric vehicle and public charging ports data from S&P Global and U.S. Department of Energy Alternatives Fuel Data Center , both aggregated from the ZIP code level to t he metro level. 

Top EV-friendly markets were first determined by calculating percentile levels for 1) share of EV-friendly listings and 2) congestion index (EVs per public port). Then a weighted average was taken between the two metrics to come up with an EV-friendly housing score for each market, forming the basis for market rankings. As, according to a study from the U.S. Department of Energy, 80% of EV charging is done at home, we assign 0.8 as the weight to the home metric and 0.2 to the public charging metric.

Rental communities refer to rental properties managed by property management companies listed on Realtor.com in 2023. EV-friendly rental communities are those featuring terms such as “electric vehicles” and “240-volt outlet” in the listing descriptions. For this research, we focus only on rental communities within the top 50 metros. 

  • For this research, electric vehicles include battery and plug-in hybrid models.
  • We use public ports in the calculation because a public station can have multiple charging ports. More details can be found here .
  • Among all the homes listed on Realtor.com in 2023, the built year for a typical EV-friendly home is 2017 while the built year for a typical non-EV-friendly home is 1994.

Sign up for updates

Join our mailing list to receive the latest data and research.

electric vehicles research

NS Energy is using cookies

We use them to give you the best experience. If you continue using our website, we'll assume that you are happy to receive all cookies on this website.

NS_Energy_logo_new

We have recently upgraded our technology platform. Due to this change if you are seeing this message for the first time please make sure you reset your password using the Forgot your password Link .

Russia’s TVEL to introduce new fuel design at Dukovany nuclear plant in Czech Republic

Nuclear Power Reactor

By NS Energy Staff Writer    30 Jul 2019

TVEL will conduct a number of pre-irradiation test for the new fuel design aimed at qualification for a licence

Project Gallery

1d9dea056f1462b792491080efac0e62

Image: The Dukovany nuclear power plant in Czech Republic. Photo: courtesy of The State Atomic Energy Corporation ROSATOM.

TVEL, a Russian nuclear fuel cycle company and a subsidiary of Rosatom, has signed a contract with Czech national power company ČEZ to introduce the new RK 3+ modification of its VVER-440 fuel at the 2,040MW Dukovany nuclear power plant (NPP).

Prior to the introduction, TVEL will carry out a number of pre-irradiation tests for the new fuel design aimed at qualification for a licence from the Czech Republic’s State Office for Nuclear Safety.

TVEL said in a statement: “The project will also consider the operations experience of the prototype of such fuel at power unit 4 of Kola NPP in Murmansk region (Russia’s North-West, Kola Peninsula), where it has been successfully used since 2010.”

Compared to the previous generations of VVER-440 fuel, the RK 3+ design has enhanced physical and thermo-hydraulic properties.

Additionally, the new fuel design features longer fuel rod pitch, which will streamline the water-uranium ratio in the reactor core and increase the efficiency of the fuel.

The design provides an option to the engineers with the elongation of the fuel cycle at the power plant.

TVEL JSC research and development vice-president Alexander Ugryumov said: “The new project of switching Dukovany NPP to RK 3+ fuel will facilitate the optimization of the plant fuel cycle strategy, also increasing safety and economic efficiency of the power units operation.”

Dukovany nuclear power plant was commissioned in 1985

Commissioned in 1985, the Dukovany NPP has four power units powered by VVER-440 reactors. Each of the reactors has a heat capacity of 1,375MW and an electric capacity of 510MW.

The power plant generates about 13 billion kWh of electricity per annum, meeting approximately 20% of power needs in the Czech Republic.

The country also operates Temelin NPP, featuring two units with VVER-1000. The two power plants account for about 35% of the country’s total power generation capacity.

The two nuclear power plants in the Czech Republic have been operated with Russian nuclear fuel manufactured by TVEL’s Machine-Building Plant (Elektrostal, Moscow region).

Recently, TVEL secured a contract to supply nuclear fuel to the future power units 7 and 8 of the Tianwan nuclear power plant (NPP) in China.

IMAGES

  1. The Race For The Electric Car

    electric vehicles research

  2. What new technologies will electric vehicles bring?

    electric vehicles research

  3. Electric Vehicles: A Short History Lesson

    electric vehicles research

  4. Electric Car Research & Technology

    electric vehicles research

  5. Electric Vehicle Technology & Development

    electric vehicles research

  6. LION Smart

    electric vehicles research

COMMENTS

  1. Electric Vehicles Research by IDTechEx

    IDTechEx predicts that their market value will surpass US$2.5 billion by 2034. The choice between metal and graphite materials for bipolar plates is a significant consideration, with each option having distinct advantages depending on the specific application. Electric Vehicles Research. Apr 15, 2024.

  2. Trends in electric vehicles research

    Abstract. Electrification of vehicles has been recognised as a key part of meeting global climate change targets and a key aspect of sustainable transport. Here, an integrative and bird's-eye view of scholarly research on Electric Vehicles (EV) is provided with a focus on an objective and quantitative determination of research trends.

  3. 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 ...

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

    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.

  5. Electric vehicles

    Electric car markets are seeing exponential growth as sales exceeded 10 million in 2022. The share of electric cars in total sales has more than tripled in three years, from around 4% in 2020 to 14% in 2022. EV sales are expected to continue strongly through 2023. Over 2.3 million electric cars were sold in the first quarter, about 25% more ...

  6. 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

  7. The new car batteries that could power the electric vehicle revolution

    Source: Adapted from G. Harper et al. Nature 575, 75-86 (2019) and G. Offer et al. Nature 582, 485-487 (2020) Today, most electric cars run on some variant of a lithium-ion battery. Lithium is ...

  8. Electric Vehicle Research and Development

    By 2045, HEVs could achieve a 43% to 81% improvement in fuel economy and PHEVs could achieve a 73% to 96% improvement in fuel economy. In its National Economic Value Assessment of Plug-In Electric Vehicles, NREL uses a scenario approach to estimate costs and benefits of increased EV market growth across the United States.

  9. Designing better batteries for electric vehicles

    Caption. Solid-state batteries now being developed could be key to achieving the widespread adoption of electric vehicles — potentially a major step toward a carbon-free transportation sector. A team of researchers from MIT and the University of California at Berkeley has demonstrated the importance of keeping future low-cost, large-scale ...

  10. Cost-competitive electric vehicles that go the distance

    According to study author Jessika Trancik, "roughly 90 percent of the personal vehicles on the road daily could be replaced by a low-cost electric vehicle available on the market today, even if the cars can only charge overnight.". This research was supported in part by the MIT Energy Initiative. Electric power Transportation.

  11. Trends and developments in electric vehicle markets

    After a decade of rapid growth, in 2020 the global electric car stock hit the 10 million mark, a 43% increase over 2019, and representing a 1% stock share. Battery electric vehicles (BEVs) accounted for two-thirds of new electric car registrations and two-thirds of the stock in 2020. China, with 4.5 million electric cars, has the largest fleet ...

  12. A circular economy approach is needed for electric vehicles

    Batteries, and components containing critical raw materials, are a key focus in developing circular economy strategies. There are reuse strategies for extending the life of electric vehicle ...

  13. Electric vehicles: 5 recent studies to inform your coverage

    The first electric cars date to the 1830s, though they weren't ready for everyday use until the 1870s. By the early 1900s, electric vehicles made up one-third of all vehicles in America, according to the U.S. Department of Energy. The rise of Henry Ford's gas-fueled Model T in the early 1910s spelled an end to the electric car for decades.

  14. How Americans view electric vehicles

    Pew Research Center conducted this study to understand Americans' views on electric vehicles. We surveyed 10,329 U.S. adults from May 30 to June 4, 2023. Everyone who took part in the survey is a member of the Center's American Trends Panel (ATP), an online survey panel that is recruited through national, random sampling of residential ...

  15. Electric Vehicle Research Center

    GIS Toolbox. The GIS Toolbox for California is a public online tool that allows users to look at the number of electric vehicles, EV owning households, or chargers needed under various adoption scenarios. The Toolbox draws on the results from many of our projects, and we will continue to add variables and scenarios to paint an even more ...

  16. Electric Cars Are Better for the Planet

    The new estimates update a study published in 2016 and add to a growing body of research underscoring the potential lifetime savings of electric cars. Comparing individual cars can be useful ...

  17. Electric Vehicles

    Electric vehicles. Evanthia A. Nanaki, in Electric Vehicles for Smart Cities, 2021 Abstract. Electric vehicles (EVs) are a promising technology for achieving a sustainable transport sector in the future, due to their very low to zero carbon emissions, low noise, high efficiency, and flexibility in grid operation and integration. This chapter includes an overview of electric vehicle ...

  18. (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 ...

  19. Electric Vehicles

    ABI Research's Electric Vehicles Research Service focuses on Electric Vehicle (EV) trends and technologies in passenger, commercial, and two-wheel vehicles. From scaling production to optimizing operation and facilitating end-of-life, the Electric Vehicle Research Service will provide holistic coverage of the new materials, chemistries, software, and services powering the zero-emission ...

  20. 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 ...

  21. EV Ownership Ticks Up, but Fewer Nonowners Want to Buy One

    Close to half now say they would not buy an EV. WASHINGTON, D.C. -- Seven percent of Americans, up from 4% a year ago, report that they own an electric vehicle. That increase is matched by an equal decline in the percentage saying they are seriously considering buying one, from 12% to 9%. Meanwhile, fewer Americans -- 35%, down from 43% in 2023 ...

  22. 2024 Fiat 500e Might Be All the EV You Need

    Here, a single permanent-magnet electric motor powers the front axle and produces just 117 horsepower and 162 lb-ft of torque. Fiat claims the 500e will meander its way to 60 mph in 8.5 seconds ...

  23. New modification of Russian VVER-440 fuel loaded at Paks NPP in Hungary

    At the site of OKB Gidropress research and experiment facility, the new fuel passed a range of hydraulic, longevity and vibration tests. ... from electric vehicles to electrified bus and truck ...

  24. Opinion

    Hertz is selling off about a third of its electric cars and Audi is slowing its transition to E.V.s. There are plenty of obvious headwinds for E.V.s — cost, range, and charging infrastructure ...

  25. EU-India join forces to promote start-up collaboration on E-Vehicles

    Research and innovation news alert: ... (EU) and India today launched expressions of interest for start-ups working in Electric Vehicles (EVs) Battery Recycling to participate in a matchmaking event. This initiative takes place under the India-EU Trade and Technology Council (TTC) announced by the Prime Minister of India, Shri Narendra Modi and ...

  26. Plugin Hybrids Take Advantage Of European EV Pause, But Will ...

    In 2024's first quarter, EV sales in Europe rose 6% to around 460,000 compared with the same period last year, while PHEV sales accelerated by 14% to 260,000, according to data from Germany's ...

  27. 2024 Realtor.com Housing Market and Electric Vehicle Report

    As, according to a study from the U.S. Department of Energy, 80% of EV charging is done at home, we assign 0.8 as the weight to the home metric and 0.2 to the public charging metric. Rental ...

  28. Not Ready for an Electric Vehicle? Buy This Car in 2024

    As of March 29, 2024, Toyota Prius Plug-in Hybrids from model years 2017-2022 can qualify for up to $4,000 of used EV tax credits. The car must have a sale price of $25,000 or less, and you also ...

  29. Luxury PHEVs With the Longest Electric-Only Range

    Volvo XC60 Recharge - Electric-only range: 35 miles The Recharge is the more powerful powertrain option for Volvo's XC60 lineup. A 2.0-liter 4-cylinder turbocharged gas engine drives the front wheels while an electric motor with an 18.8-kWh battery drives the rear wheels. Similar to the V60, the combined output is 455 hp and 523 lb.-ft. of torque. . The EPA estimates 28 MPG for both city ...

  30. TVEL to introduce new fuel design at Dukovany NPP in Czech Republic

    TVEL, a Russian nuclear fuel cycle company and a subsidiary of Rosatom, has signed a contract with Czech national power company ČEZ to introduce the new RK 3+ modification of its VVER-440 fuel at the 2,040MW Dukovany nuclear power plant (NPP). Prior to the introduction, TVEL will carry out a number of pre-irradiation tests for the new fuel ...