Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport (Green Energy and Technology) 3030966739, 9783030966737

This book discusses the importance of transitioning from conventionally fuelled, electric and hydrogen personal vehicles

99 56

English Pages 129 [124] Year 2022

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Acknowledgements
Contents
Abbreviations
1 Introduction
1.1 General Introduction
1.2 Greenhouse Gas Emissions and Global Policy
1.3 Transport and Environmental Boundaries
1.3.1 Transport Emissions
1.3.2 Low Carbon Transport Alternative Options
1.3.3 The Role of Public Transport
1.4 Meeting the Sustainable Development Goals When Transitioning to Low Carbon Transport
1.4.1 Energy Implications from Transport
1.5 Inferences for Natural Capital and Ecosystem Services
1.6 Brief Outline
References
2 Passenger Land-Based Road Transport
2.1 Personal Vehicles
2.1.1 Travel Behaviour
2.1.2 Integrating Low Emission Personal Vehicles
2.2 Battery Electric Vehicles
2.2.1 Hybrid and Plug-In Hybrid Electric Vehicles
2.3 Hydrogen Vehicles
2.4 Barriers for Low Emission Transport
2.4.1 Associated Costs
2.4.2 Range Anxiety and Charging Infrastructure
2.5 Incentives for Low Emission Transport
2.5.1 Monetary Incentives
2.5.2 Non-monetary Incentives
2.6 Key Findings
References
3 Buses
3.1 Introduction
3.2 Conventionally Fuelled Buses
3.3 Electric Buses
3.4 Hydrogen Buses
3.4.1 Hydrogen Generation
3.5 Other Low Emission Alternatives
3.6 Key Findings
References
4 Trains
4.1 Introduction
4.2 Conventionally Fuelled Trains
4.3 Electric Trains
4.4 Hydrogen Trains
4.5 High-Speed Rail
4.6 Key Findings
References
5 Challenges of Implementing Electric and Hydrogen Public Transport
5.1 Introduction
5.2 Factors Influencing Public Transport Use
5.3 Public Acceptance of New Technologies
5.4 Technology and Infrastructure Challenges for Electric and Hydrogen Public Transport
5.4.1 Electric Public Transport
5.4.2 Hydrogen Public Transport
5.5 Costs of Electric and Hydrogen Public Transport
5.5.1 Electric Buses
5.5.2 Electric Trains
5.5.3 Hydrogen Buses
5.5.4 Hydrogen Trains
5.6 Electricity Demands for Low Carbon Transport Integration
5.6.1 Reducing Emissions and Demand on the Grid
5.6.2 Energy Storage Technologies
5.7 Environmental Implications
5.8 Key Findings
References
6 Low Carbon Public Transport and the Competition with Aviation
6.1 Introduction
6.2 Aviation Emission Policy
6.2.1 Responsibility for Aviation Emissions
6.3 Policies to Reduce Emissions from Aviation
6.3.1 Taxes
6.3.2 Emission Trading Scheme (ETS)
6.3.3 Phasing Out Short Haul Flights
6.3.4 Airport Surface Access Strategies
6.4 Conclusions
References
7 Freight
7.1 Emissions from Freight
7.2 Freight Shipping
7.2.1 Alternatives to Conventionally Fuelled Shipping
7.3 Freight Trucks
7.4 Alternatives to Land-Based Freight Movements
7.5 Key Findings
References
8 Low Carbon Transport for a Modern Working Environment
8.1 Coronavirus and Emissions
8.2 The ‘New Working Normal’
8.3 Private Vehicles and Public Transport
8.4 Active Travel
8.5 Aviation
8.6 Future Considerations
References
9 Policy Recommendations
9.1 Environmental Impact
9.2 Recommendations Within the Transport Sector
9.3 Conclusions
References
Recommend Papers

Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport (Green Energy and Technology)
 3030966739, 9783030966737

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Green Energy and Technology

Kathryn G. Logan Astley Hastings John D. Nelson

Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport

Green Energy and Technology

Climate change, environmental impact and the limited natural resources urge scientific research and novel technical solutions. The monograph series Green Energy and Technology serves as a publishing platform for scientific and technological approaches to “green”—i.e. environmentally friendly and sustainable—technologies. While a focus lies on energy and power supply, it also covers “green” solutions in industrial engineering and engineering design. Green Energy and Technology addresses researchers, advanced students, technical consultants as well as decision makers in industries and politics. Hence, the level of presentation spans from instructional to highly technical. **Indexed in Scopus**. **Indexed in Ei Compendex**.

More information about this series at https://link.springer.com/bookseries/8059

Kathryn G. Logan · Astley Hastings · John D. Nelson

Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport

Kathryn G. Logan Arizona Institute for Resilient Environments and Societies University of Arizona Tucson, AZ, USA Energy Institute University College Dublin Dublin, Ireland

Astley Hastings The School of Biological Sciences Institute of Biological and Environmental Science University of Aberdeen Aberdeen, Scotland

The School of Biological Sciences Institute of Biological and Environmental Science University of Aberdeen Aberdeen, Scotland John D. Nelson The University of Sydney Business School Institute of Transport and Logistics Studies University of Sydney Sydney, NSW, Australia

ISSN 1865-3529 ISSN 1865-3537 (electronic) Green Energy and Technology ISBN 978-3-030-96673-7 ISBN 978-3-030-96674-4 (eBook) https://doi.org/10.1007/978-3-030-96674-4 © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

To all those affiliated at the University of Aberdeen whose contribution to saving the planet takes us closer each day.

Acknowledgements

This research was carried out under the UK Energy Research Centre (UKERC) as part of the ADdressing Valuation of Energy and Nature Together (ADVENT) and UKERC-4 funded projects. Funding was received from the Natural Environment Research Council (NE/M019691/1), UK, and the School of Biological Sciences, University of Aberdeen, UK.

vii

Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 General Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Greenhouse Gas Emissions and Global Policy . . . . . . . . . . . . . . . . . . 1.3 Transport and Environmental Boundaries . . . . . . . . . . . . . . . . . . . . . . 1.3.1 Transport Emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.2 Low Carbon Transport Alternative Options . . . . . . . . . . . . . . 1.3.3 The Role of Public Transport . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Meeting the Sustainable Development Goals When Transitioning to Low Carbon Transport . . . . . . . . . . . . . . . . . . . . . . . . 1.4.1 Energy Implications from Transport . . . . . . . . . . . . . . . . . . . . 1.5 Inferences for Natural Capital and Ecosystem Services . . . . . . . . . . . 1.6 Brief Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1 1 2 4 4 5 6 7 8 9 11 11

2 Passenger Land-Based Road Transport . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Personal Vehicles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.1 Travel Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.2 Integrating Low Emission Personal Vehicles . . . . . . . . . . . . . 2.2 Battery Electric Vehicles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Hybrid and Plug-In Hybrid Electric Vehicles . . . . . . . . . . . . . 2.3 Hydrogen Vehicles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Barriers for Low Emission Transport . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.1 Associated Costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.2 Range Anxiety and Charging Infrastructure . . . . . . . . . . . . . . 2.5 Incentives for Low Emission Transport . . . . . . . . . . . . . . . . . . . . . . . . 2.5.1 Monetary Incentives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.2 Non-monetary Incentives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6 Key Findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19 19 20 22 23 24 25 26 26 27 27 28 29 29 30

ix

x

Contents

3 Buses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Conventionally Fuelled Buses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Electric Buses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Hydrogen Buses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.1 Hydrogen Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5 Other Low Emission Alternatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6 Key Findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

37 37 38 41 42 43 43 44 45

4 Trains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Conventionally Fuelled Trains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Electric Trains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Hydrogen Trains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5 High-Speed Rail . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6 Key Findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

51 51 53 53 54 55 56 57

5 Challenges of Implementing Electric and Hydrogen Public Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Factors Influencing Public Transport Use . . . . . . . . . . . . . . . . . . . . . . 5.3 Public Acceptance of New Technologies . . . . . . . . . . . . . . . . . . . . . . . 5.4 Technology and Infrastructure Challenges for Electric and Hydrogen Public Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.1 Electric Public Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.2 Hydrogen Public Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5 Costs of Electric and Hydrogen Public Transport . . . . . . . . . . . . . . . . 5.5.1 Electric Buses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5.2 Electric Trains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5.3 Hydrogen Buses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5.4 Hydrogen Trains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6 Electricity Demands for Low Carbon Transport Integration . . . . . . . 5.6.1 Reducing Emissions and Demand on the Grid . . . . . . . . . . . . 5.6.2 Energy Storage Technologies . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7 Environmental Implications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.8 Key Findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

64 65 67 68 68 69 69 69 70 71 72 73 74 74

6 Low Carbon Public Transport and the Competition with Aviation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2 Aviation Emission Policy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.1 Responsibility for Aviation Emissions . . . . . . . . . . . . . . . . . . .

81 81 82 83

59 59 60 63

Contents

xi

6.3 Policies to Reduce Emissions from Aviation . . . . . . . . . . . . . . . . . . . . 6.3.1 Taxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3.2 Emission Trading Scheme (ETS) . . . . . . . . . . . . . . . . . . . . . . . 6.3.3 Phasing Out Short Haul Flights . . . . . . . . . . . . . . . . . . . . . . . . 6.3.4 Airport Surface Access Strategies . . . . . . . . . . . . . . . . . . . . . . 6.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

84 84 85 86 87 87 88

7 Freight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1 Emissions from Freight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2 Freight Shipping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.1 Alternatives to Conventionally Fuelled Shipping . . . . . . . . . . 7.3 Freight Trucks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 Alternatives to Land-Based Freight Movements . . . . . . . . . . . . . . . . . 7.5 Key Findings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

91 91 92 93 97 98 99 99

8 Low Carbon Transport for a Modern Working Environment . . . . . . . 8.1 Coronavirus and Emissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2 The ‘New Working Normal’ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3 Private Vehicles and Public Transport . . . . . . . . . . . . . . . . . . . . . . . . . 8.4 Active Travel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5 Aviation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.6 Future Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

103 103 104 105 106 107 107 107

9 Policy Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.1 Environmental Impact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2 Recommendations Within the Transport Sector . . . . . . . . . . . . . . . . . 9.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

111 111 112 114 115

Abbreviations

AIS BEV BRT CCS CFB CFT CFV CNG CO2 CORSIA DAC EB ES ET EU EV FC GDP GHG H2 HB HEV HFO HSR HT HV ICAO ICE ICEV IEA IMO

Automatic identification system Battery electric vehicle Bus rapid transit Carbon capture and storage Conventionally fuelled bus Conventionally fuelled train Conventionally fuelled vehicle Compressed natural gas Carbon dioxide emissions Carbon Offsetting and Reduction Scheme for International Aviation Direct air capture Electric bus Ecosystem service Electric train European Union Electric vehicle Fuel cell Gross domestic product Greenhouse gas Hydrogen Hydrogen bus Hybrid electric vehicle Heavy fuel oil High-speed rail Hydrogen train Hydrogen vehicle International Civil Aviation Organization Internal combustion engine Internal combustion engine vehicle International Energy Agency International Maritime Organization xiii

xiv

IPCC ITF LCA LNG N2 O NC NCC NDC NHSR NOx OCED PEM PEMFC PHEV PM QRA SDGs SMR SOx TDM TOCs TVG UAV UF UNFCCC USA WFH WLTP

Abbreviations

Intergovernmental Panel on Climate Change International Transport Forum Life cycle assessment Liquefied natural gas Nitrous dioxide Natural capital Natural Capital Committee Nationally determined contribution Non-high-speed rail Nitrous oxide Organisation for Economic Co-operation and Development Proton-exchange membrane Polymer electrolyte membrane fuel cell Plug-in hybrid electric vehicle Particulate matter Quantitative risk assessment Sustainable development goals Steam methane reform Sulphur dioxide Travel demand management Total cost of ownership Train à Grande Vitesse Unmanned aerial vehicles Utility factor United Nations Framework Convention on Climate Change United States of America Work from home Worldwide Harmonized Light Vehicle Test Procedure

Chapter 1

Introduction

Abstract To reduce greenhouse gas (GHG) emissions and meet Paris Agreement targets, nationally determined contributions have been made by every country in the world. Many countries are aiming for a net zero emission target, however, net zero has been defined differently country to country making unified emission reductions more difficult. Transport and energy generation remain the two largest global emitting sectors and substantial transformation will be required to meet emission reductions. Internal combustion engine vehicles for personal use remain the highest emitting transport type, which has led to governments and policymakers introducing legislation to ban and phase out their sale over the coming decades in favour of low carbon alternatives, including electric and hydrogen fuelled vehicles. Although electric and hydrogen transport are considered ‘zero emission’ at their point of use, their true environmental impact is determined by the source of the electricity used to ‘fuel’ these vehicles. Therefore, an integrated and interdisciplinary approach to meet net zero will be required as there will need to be trade-offs between GHG emission reductions, climate regulation and the potential impact upon ecosystem services. By integrating alternative fuels and encouraging travel behaviour to support public transport, which has a lower level of emission per person per kilometre travelled, there is the potential to have a significant impact on emission level reduction. Taking into consideration experience from different countries that have successfully implemented pathways towards low carbon transport, lessons can be learnt from the best policies and decarbonise both the transport and energy sectors.

1.1 General Introduction The transport sector remains a focal point of any debate regarding energy conservation due to the current reliance on the fossil fuel industry for both passenger and freight road, rail, sea, or air travel [1]. One proposed solution to mitigate global transport greenhouse gas (GHG) emissions is to develop and deploy cleaner low carbon technologies [2], including low emission fuelled transport such as electric and hydrogen alternatives. As personal vehicles are the largest contributor to landbased transport type for emissions, focus is often placed here. However, studies © The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_1

1

2

1 Introduction

have highlighted that transitioning to electric or hydrogen personal vehicles to meet climate emission targets will not be enough and significant behavioural changes to encourage greater uptake of low carbon electric and hydrogen fuelled public transport will be necessary [3, 4]. This book will discuss the emission levels of different transport types and the need to transition towards low carbon public transport if emission reduction targets are to be met. The contemporary challenges, including behavioural, infrastructure, cost, and the impact that COVID-19 has had within the transport industry, as well as the consequences for the environment in terms of trade-offs between emission reduction targets and infrastructure changes.

1.2 Greenhouse Gas Emissions and Global Policy Climate change has become a topic of global importance with the urgency to reduce anthropogenic GHG emissions widely acknowledged in literature and international policy objectives [5–8]. To reduce the pressing nature of the global ‘climate emergency’, three key international protocols have been introduced following the initiation of the Kyoto Protocol in 1997 [9]. Firstly, the 2030 Sustainable Development Goals (SDGs), which act as climate change mitigation measures at local, subnational/national, regional, and global levels [10]. The SDGs represent an interconnected, complex network of interactions through 17 SDG goals, with 169 associated targets, addressing extreme poverty, socio-economic inclusion, and ecological sustainability [11]. Whilst the SDGs are often siloed, there are interlinkages and feedback (positive and negative) between the economic, social, and environmental dimensions of the SDGs that are often unaccounted for [12–18]. Secondly, assessment reports made by the Intergovernmental Panel on Climate Change (IPCC), have provided evaluations of the rising temperature and possible risks in the climate system under various emission scenarios. Under the IPCC’s fifth assessment report in 2016, researchers involved conducted a comprehensive assessment of the climate system change, risks, emission budget, and mitigation pathway choice of 2 °C global warming based on the research results available [19]. Finally, through the IPCC assessment and a series of political pushes, ‘The Paris Agreement’, was introduced in 2015, by the United Nations Framework Convention on Climate Change (UNFCCC). The Paris Agreement aims to keep the average global temperature to ‘well below 2 °C’ above pre-industrial levels and to pursue efforts to limit the temperature increase to 1.5 °C above pre-industrial levels [20–22]. This was further emphasised when the Heads of Government attended the 26th United Nations Climate Change Conference of the Parties (COP26) in Scotland in 2021, seeking to secure global net zero by mid-century and to keep the 1.5º C target within reach. Under ratification of the Paris Agreement, and to meet the SDGs and targets in the IPCC report, many countries have set their own nationally determined contributions (NDCs) through limitations and targets to reduce emission levels [23]. These NDCs are often hard to compare as more than 100 national governments (e.g., China, Japan, EU, UK, etc.) and over 800 cities [24] have set or are considering setting a

1.2 Greenhouse Gas Emissions and Global Policy

3

‘net zero’ target [25, 26]. However, the specific details, including the dates to meet these targets, differ between countries with some targets focusing on primary carbon dioxide (CO2 ) emissions, whilst others cover all GHG emissions [26]. Furthermore, this focus on CO2 emissions exclusively would mean that concentrations of other GHGs could be rising even if the goal of net zero carbon emissions is attained [26]. In addition, some of these targets do not aim to reduce emissions but compensate them with offsets. This idea of a carbon offset does not necessarily address the problem at hand as it allows a business, government or an individual to pay another party to cut or remove a quantity of GHG emissions from the atmosphere [27]. This does not aid equality as it allows anyone the opportunity to offset their emissions if they have the financial capabilities and emission offsets are often outsourced to third world countries i.e., building forests to reduce air quality in a different country that where the offsets are purchased [28]. For example, to reduce emissions from a coal power station, the (European) power station could mitigate emissions by buying offsets that could take the form of channelling money to countries like Brazil, Indonesia, and South Africa to plant trees and restore forests. Although these offsets can help ‘buy time’, they do not directly address the economic, political, and technical work required to address climate change at a global scale. Alternatively, new technologies are being introduced, such as carbon capture and storage (CCS) and direct air capture (DAC) which aim to ‘trap’ CO2 emissions and store them underground in deep geological formations, therefore achieving negative emissions and carbon removal [29]. Both CCS and DAC technologies can act as an interim solutions whilst other challenges including energy storage and 100% renewable energy generation is achieved, however require substantial financial investment. The differences within these NDC targets, and what they mean, are increasingly important, as according to UN estimates, global GHG emissions in 2030 will be some 15 billion tonnes carbon dioxide equivalent (Gt CO2 e) higher than required under a 2 °C stabilisation path [30], which will have global detrimental impacts. The Climate Change Laws of the World database, which has tracked global legislation from every country in the world over the past thirty years has records of ~2601 climate change laws and policies worldwide, covering both mitigation and adaptation (as of March 2022) [31]. The database incorporates laws (e.g., UK’s Climate Change Act 2008 (2050 Target Amendment Order 2019)), dedicated climate measures (e.g., New Zealand’s Climate Change Response (Emissions Trading) Amendment), and sector policies (e.g., Finland’s Act 478/2017 on the distribution of alternative fuels for transport) [30]. Furthermore, the database also includes other laws that focus on adaptation, either exclusively (e.g., Japan’s Climate Change Adaptation Act) or laws that act in conjunction with wider climate or environmental objectives (e.g., the 2050 Climate Strategy of the Marshall Islands) [30]. It is noteworthy that there is no country in the world that does not have at least one climate change law in place [32]. Although laws have been implemented globally, ensuring they are met through emission reductions is imperative. As low carbon transport and energy generation are interlinked, these sectors rely on one another if net zero targets are to be met and cannot be considered in silo if a just transition is to be made. This has been further emphasised as over half of the climate laws already in place contain provisions regarding energy supply, including

4

1 Introduction

the promotion of renewables, with over 40% including demand-side energy efficiency [30]. Through analysis of both these sectors, conclusions can be drawn on the best strategies to reduce GHG emissions from transport and meet both NDCs and the Paris Agreements on a global scale.

1.3 Transport and Environmental Boundaries 1.3.1 Transport Emissions Despite global initiatives focused on the decarbonisation of transport, transport carbon emissions have been growing in both absolute and relative terms since 1990 [33]. To meet the Paris Agreement target, urgent and deep decarbonisation of the transport sector will be necessary, although this remains challenging [34]. Landbased transport, combined with national shipping and aviation, contributes to ~19% of global emissions, whilst international shipping and aviation contributes to an additional ~3.5% [35]. With a growing demand for mobility and private vehicle ownership [36], GHG emissions from transport have continued to rise [37]; despite many countries adopting more efficient vehicles and promoting low carbon transport alternatives [38]. Moreover, the International Energy Agency (IEA) has reported that global transport is responsible for a quarter of combustion emissions, with road transport alone accounting for a majority of these [39, 40]. This is due to internal combustion engine vehicles (ICEVs) not being energy efficient, for example, in the average passenger car, only 21% of the fuel is used to move the vehicle and the remaining 79% accounts for energy losses [41]. The IPCC has determined that whilst following current trends, GHG emissions from transport could increase at a faster rate than emissions from the other energy end-use sectors, emitting around 12 Gt CO2 eqyr−1 by 2050 [42]. Whilst focusing on reducing transport emissions, ‘decoupling’ the link between economic growth, transport growth and transport energy consumption is an important concept which will likely result in the delinking of environmental costs from economic development [43, 44]. This allows governments and policymakers to focus on both their economic development whilst suppressing transport-related carbon emissions and ensuring targets in the SDGs are met [33]. For example, a global analysis of transport decoupling between 1990 and 2015 highlighted that most nations only obtained a ‘stable’ decoupling pathway without consistent improvements to absolute decoupling [45]. In some economically developing countries, this concept can allow for ‘leapfrogging’ which is the process of skipping stages in development to avoid pollution-intensive stages by jumping ahead to be a leading innovator and utiliser of the new technology, i.e., encouraging first time vehicle consumers to purchase an electric personal vehicle instead of a combustion engine personal vehicle [46–49]. To ensure advancements to meet the Paris Agreement targets and for countries to remain in line with their NDCs, approaches to lower GHG emissions from the

1.3 Transport and Environmental Boundaries

5

transport sector can be grouped into four categories. Firstly, technical, such as the integration of low emission alternatives such as electric/hydrogen public transport. Secondly, legislative, such as the introduction of a carbon or fuel tax whilst subsidising public transport to reduce costs for the individual. Thirdly, infrastructural, focusing on the extensive development of methods such as the integration of cycling lanes to promote active travel and designing towns to reduce the need to travel for work, shopping, and leisure (the 15-min city concept). Finally, behavioural, through the introduction of travel demand management initiatives (TDM) to active ‘push’ and ‘pull’ individuals to choose sustainable travel options. Examples of ‘pull’ measures include the promotion of car sharing [50, 51] or enhanced facilities for cyclists at the workplace. ‘Push’ measures include pricing levers such as parking management or road user charges. These four categories will need to be integrated simultaneously to achieve the largest impact of emission reduction. By addressing these categories, there is hope that emission reduction targets can be met, however a unified approach across these categories will be necessary.

1.3.2 Low Carbon Transport Alternative Options The introduction of electric and hydrogen alternative transport options is necessary as even with the technological advances of ICEVs, emission reduction is not enough to meet most countries NDCs. This has led to many countries deciding to ban and phase out the sale of ICEVs and integrate low emission transport alternatives. Although electric and hydrogen transport are considered ‘zero emission’ at the point of use as they do not produce tailpipe emissions, the carbon intensity of electricity generation will determine the environmental impact of these transport types [52]; to date, uptake remains slow [53]. This is because numerous studies have highlighted that transitioning to electric and hydrogen alternatives may not be enough to have a significant impact on emission levels, especially due to the vehicle turnover rates [54, 55]. As a result, a growing consensus that this technological transition will not be sufficient, or fast enough, to transform the transport system means that a modal shift to alternatives needs to be considered. Since widespread electrification is proving to be a slow process and is likely to be too relaxed to contribute towards meaningful ambitious climate change mitigation targets [56], there must be recognition that many of the challenges in solving transport problems relate to affecting behavioural change rather than just technical solutions. Despite encouragement through economic incentives to replace car trips with alternative low carbon sustainable transport modes, research and practice have found a substantial resistance from individuals to reduce car use [57–59]. Increased public transport uptake would help to mitigate emission levels as there is a higher carrying capacity on buses and trains than with personal vehicles. For example, it was estimated that a typical passenger car carrying one individual emits ~40.4 kg of CO2 per 100 passenger miles compared to a conventional bus whose capacity of 70 passengers emits only ~6.3 kg [60]. However, several studies have highlighted that individuals

6

1 Introduction

believe that private vehicles not only provide status to their owners including representing their career achievements, but also ‘speed, home, safety, sexual success, freedom, family, masculinity and genetic breeding’ [61–65]. This has contributed to an on-going decline in public transport patronage in many affluent countries since the 1960s. Yet, it is an important fact that emission levels from public transport remains lower than personal transport and that if public transport is decarbonised, emission levels are likely to be lower. Therefore, if measures are put in place that actively encourage the use of public transport, transport emissions are likely to decline.

1.3.3 The Role of Public Transport Public transport is collective transport accessible by the general public, often provided on fixed routes (whether on-road or tracks). It includes transport by trains, trams, bus (including bus rapid transit (BRT)), light rapid transit, coaches, aircraft, taxis, and the newer forms of ridesharing such as Uber and Lyft [66]. Whilst the bus was arguably the “workhorse” of transport in cities for many years, public transport has come under pressure to maintain its market share, particularly within rural areas [67]. This is primarily due to demographic changes (i.e., ageing populations), higher fuel costs and the increased usage of personal vehicles and (more recently) micro-mobility options such as e-bikes or e-scooters [68]. Over time public transport has become more expensive to provide and increasingly unprofitable in some countries, particularly in Europe [67]. However, there is potential to tackle problems related to accessibility and mobility [67, 69] by introducing measures to encourage a modal shift to public transport that optimises the environmental, social, and cost benefits of use [70]. By encouraging public transport use that supports the needs of the individual, for example through the provision of flexible and demand responsive forms of transport [71], coupled with the integration of alternative fuels such as electric and hydrogen, it is arguable that greater decarbonisation, energy security and urban air quality improvements could occur [72]. Encouraging public transport uptake will remain a challenge due to the impacts of the novel coronavirus (COVID-19), or SARS-CoV-2, pandemic which has already influenced the energy and transport sectors across the world [73]. Due to lockdowns and restrictions, many countries experienced a sudden decrease in both GHG emissions and air pollutants; for example, between March 2019 and March 2020, Europe saw a decrease of between 20 and 30% [41, 42]. Furthermore, a sharp reduction in public transport demands have been observed with individuals opting to use their own personal vehicles as they believe this will reduce their risk of getting the virus [74]. The COVID-19 pandemic has greatly impacted public transport in three important ways [66]. Firstly, in that many people have been furloughed or are working from home (WFH) and as such have had no need to use public transport. Secondly, the need to self-distance to avoid infection, which has posed several challenges to the traditional operation and use of public transport, especially as a mode of mass

1.3 Transport and Environmental Boundaries

7

transit. Finally, in relation to aviation (which often provides “lifeline services” to remote locations), there has been a need to prevent travel to seek and halt the spread of COVID-19, with additional knock-on effects for airport access strategies. Therefore, not only will policymakers need to consider how to encourage individuals to use public transport, but also encourage individuals who have previously used public transport to continue or return.

1.4 Meeting the Sustainable Development Goals When Transitioning to Low Carbon Transport Together, the Paris Agreement and the SDGs provide a blueprint towards a sustainable, low carbon and more equitable global future [11]. However, evaluating the success of these targets also remains a key challenge to best determine their attainment and whether implementation of similar targets should be considered at other regional, national, and country wide scenarios for different sectors [75]. For example, although there have been tools developed to measure the success of decarbonisation, in particular GHG emission calculators, these often overlook social and political considerations [76, 77]. This often results in decarbonisation focusing on resolving emissions i.e., transitioning to battery electric vehicles, rather than engaging with a contested political process or behavioural changes [75, 78]. Although coupling technological and socio-economic perspectives is necessary to identify technically feasible, financially viable, and socially equitable transition scenarios [79, 80], these need to be done in partnership between academia and policymakers to ensure climate change targets are met. SDG7 aims to ‘ensure access to affordable, reliable, sustainable and modern energy for all’, with SDG 7.2 highlighting the importance of substantially increasing the share of renewable energy within the total energy mix by 2030 to support low carbon energy transitions. To achieve this target, a low carbon energy transition towards renewables will need to occur, on a local, regional, and national level [81]. An energy transition refers to the shift from one dominant energy source, or set of sources, to another. In this case, from fossil fuels to renewable energy technologies, that leads to lower GHG emissions being emitted [82–84]. As energy transitions are embedded within a wider political, social, and economic context, there is scope that these transitions have potential to worsen or exacerbate existing inequalities or introduce new vulnerabilities into communities [85, 86]. The progress of these transitions relies on multiple parameters including key stakeholders and how outcomes address new paths and opportunities [11, 87, 88]. This will be particularly important for the transition towards lower emission transport as if the energy generated is not from fossil fuels and the environmental benefits of this transition will be negated. These trade-offs and synergies will require an interdisciplinary approach to reduce social resistance, lack of awareness about current (and new) technologies or negative social or environmental impacts, either real or perceived [89–92]. Several studies have focused on social acceptance or community acceptance of renewable energy

8

1 Introduction

technologies, policies or methods that help to enable a renewable energy transition [93–95]. Although beyond out with the scope of this book, further research will need to focus on social acceptance and behavioural change to ensure a smooth transition towards low carbon transport.

1.4.1 Energy Implications from Transport Energy policy fluctuates around the ‘energy trilemma’ which focuses on three fundamental objectives for an affordable, secure, and sustainable energy system [96–98]. However, transitioning towards net zero allows focus to be placed on a low carbon energy network that is affordable for individuals to meet their daily needs. Net zero emission targets, adopted by national governments, often focus on decarbonising the whole economy, including energy supply and demand, as well as other emission sources including transport, industrial processes, and agriculture etc. [99]. The transport and energy sectors work in partnership as the way in which energy is generated can influence the level of emissions within the transport sector. Transport relies heavily on liquid fuels derived from oil and consumes ~55% of the world’s total liquid fuels [38]. This high dependence on resources raises significant concerns for energy security, which has become an important topic economically, socially, and politically on a global scale [100, 101]. Although focus is often placed on a country’s dependence on oil and non-renewable resources, energy security encompasses modern technologies such as renewables [102]. Energy security, one strand of the trilemma, is a complex and multi-dimensional concept, with international energy security influencing other sectors and becoming a focal point for policymakers [103, 104]. Therefore, protecting the energy security of a system using appropriate energy policies and technologies is necessary [105, 106]. The definition of energy security varies in the literature with some definitions focusing on the security of supply such as the availability and costs, whereas other researchers argue for a more comprehensive definition that includes downstream impacts on the economy and social welfare [107–109]. However, these definitions of energy security evolve as circumstances change [110]. To provide energy security to a user, it has been argued that four key conditions need to be met as outlined in the 4As framework [102, 111, 112]. Firstly, the ‘availability’ of a sustainable, natural, extractable, or renewable energy source. Through the diversification of energy supply, countries can reduce and mitigate the risk of any energy disruptions. For example, Shanghai, China is considered an importdependent mega-city as it remains heavily dependent on oil and coal for primary energy consumption, however, diversifying this with renewables will reduce the potential for energy shortages or blackouts [107, 113, 114]. Secondly, ‘applicability’ of technology and infrastructure to extract and harness available energy. This remains essential for uninterrupted energy within the market to provide a stable energy supply. Thirdly, ‘acceptability’ of the energy sources’ environmental and social impacts. This is to reduce societal concerns where individuals could be denied basic energy

1.4 Meeting the Sustainable Development Goals …

9

services. However, as a key aim of energy security is to ensure against energy import disruptions for adequate access to energy sources to sustainable social and economic welfare to an acceptable level, this remains important to consider [115]. Finally, the ‘affordability’ of the energy sources for the end user which needs to consider several factors. For example, the absolute energy price, price volatility and the level of competition within the energy market [107]. In addition to this, the currency in which the amenity is traded in also should be taken into consideration as exchange rates and purchasing power of different currencies can determine how much a country should pay for their imports [107]. Most studies on energy security are country-specific, and often do not consider that energy security remains a concern for both developed and developing countries [110]. By applying energy security indicators, countries with a high level of gross domestic product (GDP) tend to have a low level of energy security performance [116]. Therefore, the historic development of energy security highlights that developing countries must first increase their energy intensity whilst developing modern infrastructure, whereas a developed country needs to focus on less energy intensive investments [117, 118]. This has led Germany to introduce the Energy Security of Supply Act, which restricts sales, purchases, or use of goods to be used for certain priority purchases as part of their energy transition [119]. Similarly, Russia has introduced the Energy Security Doctrine of the Russian Federation, as part of their national security, that includes assuring quantity, quality, and efficiency of their energy supply to consumers [119]. This type of legislation is fundamental to see economic growth and social welfare through an energy secure system [120]. For decarbonised transport, ensuring the 4As framework is met within the energy sector is important to ensure a successful transition as fluctuations in energy prices will likely see fluctuating transport costs [100, 109]. This has led to a divide within transport use as individuals who have money are able to travel more frequently in higher emitting transport types such as airplanes [121].

1.5 Inferences for Natural Capital and Ecosystem Services Although transport brings several advantages to society, both from a personal perspective and economically, these benefits can result in undesirable outcomes including climate change, air pollution, water pollution, congestion, accidents, and noise etc. [122]. With the transition towards electric and hydrogen transport, additional challenges will need to be addressed, including the impact that this will have on the environment. This is because this transition will likely require trade-offs between GHG emissions and infrastructure changes, including materials required for vehicle construction and charging infrastructure. Through this transition towards low carbon transport, it remains inevitable there will be an impact on natural capital (NC) and ecosystem services (ES) through additional energy generation, and the required infrastructure (generation source, power distribution, charging infrastructure and from the vehicles themselves). The Natural

10

1 Introduction

Capital Committee (NCC) defines NC as ‘the elements of nature that produce value or benefits to people (directly and indirectly), such as the stock of forests, rivers, land, minerals and oceans, as well as the natural processes and functions that underpin their operation’ (NCC, 2014). This definition therefore suggests that NC includes the interactions and processes that are involved in nature’s capacity to persist based on physical, biological, and chemical processes. NC comes in two broad types: renewables and non-renewables. Renewables are NC that nature provides for ‘free’ and keeps giving for free, provided it does not deplete below its threshold for sustainable reproduction [123]. Alternatively, non-renewables are objects that nature provides for free, which do not regenerate and tend to be inanimate such as fuels and mineral ores including, although not limited to, oil, gas, coal, copper, lithium, cobalt, and lead [123]. ES are the direct and indirect contributions of ecosystems to human well-being and can be split into three categories [124–127]. Firstly, provisioning services which incorporates goods obtained from the environment such as water, food, fibre, and fodder. Secondly, regulating and maintenance services which includes benefits obtained from regulating ecosystem processes such as soil formation, habitats, water, air quality and climate, and finally, cultural services such as spiritual and intellectual interactions [124–126]. Even with the additional energy generation required to meet transport demands during this transition to low carbon, the environmental impact of electric and hydrogen transport, particularly from low carbon public transport, remains significantly lower than that of conventionally fuelled alternatives. Reduced reliance on fossil fuels removes the environmental risks that are associated with extraction, with incidents such as oil spills often having decade long consequences for the surrounding ecosystem. Further downstream, impacts of fossil fuel usage such as ocean acidification and global temperature rises causing sea ice melt are all factors that impact the environment and are therefore have a value from a NC and ES viewpoint. This is not to say that all renewable generation methods do not have any quantifiably negative effect on the local environment, however the impacts are generally small when compared to no implementation [128, 129]. In addition, other factors including where the vehicle is constructed and how it got to the user should also be considered with local manufacturing a priority. Furthermore, renewable energy sources and charging stations should be installed near where vehicles are likely to be recharged. This will reduce transmission and distribution losses, reducing the additional energy that may need to be generated. This means that whilst public transport may be more difficult to implement in rural areas, the supply of local clean energy as fuel, either directly as electricity or via hydrogen production, may be more easily implemented and have a reduced impact on ES and NC. As technology continues to improve, the total number of onshore or offshore wind turbines or solar panels will be less as they can generate more energy and will require less land and sea area, reducing the impact on ES and NC. Ensuring that electric and hydrogen transport is charged at varying times throughout the day and night may also be necessary to ensure there is reduced impact

1.5 Inferences for Natural Capital and Ecosystem Services

11

on the grid infrastructure. By ensuring low emission vehicles are charged during offpeak hours, primarily during the night, and by taking advantage of vehicle to grid technology, there is likely to be a reduction in the increased peak generation capacity required which will allow for a smoother transition towards low carbon transport [130]. This will also allow for a decreased reliance on fossil fuels during the transition towards low carbon energy generation. One solution to limit this is to integrate ‘packages’ which allow consumers to either purchase a BEV battery outright or to pay a mileage fee and return the battery after use. This would allow battery manufacturing companies or governments to manage and reduce the potential impact on the electricity supply infrastructure [131].

1.6 Brief Outline This book will be organised in the following chapters: Chapter 2 will discuss the emissions from ICEVs and discuss low carbon alternatives, including battery electric vehicles (BEVs) and hydrogen vehicles. Chapters 3 and 4 will discuss current bus and rail use within a global context, low carbon alternatives including electric and hydrogen, factors influencing bus and rail usage and policy recommendations to encourage uptake. Chapter 5 will discuss the key challenges of implementing low carbon public transport, including infrastructure challenges, electricity demand and environmental impact. Chapter 6 will discuss the challenges of encouraging low carbon public transport with aviation. Chapter 7 will discuss the challenges that the freight sector currently faces at incorporating low carbon alternatives. Chapter 8 will discuss the challenges of incorporating low carbon transport within a modern working environment, taking into consideration the impact of COVID-19. Finally, Chap. 9 will discuss key conclusions and make policy recommendations on how we can globally achieve the Paris Agreement targets. This brief will utilise examples from countries that have already adopted or are in the process of adopting electric and hydrogen transport to better understand the lessons learnt for policymakers across the world. This can lead to ‘leapfrogging’, which has potential for both the transport and the energy sectors to ensure decarbonised electricity for low carbon transport. A key example of this has been Japan, world leaders in electric train technology, with the world’s first high-speed bullet train, which has allowed other countries to adapt to low emission trains when transitioning away from diesel trains, to meet emission reduction targets.

References 1. Brand C, Götschi T, Dons E et al (2021) The climate change mitigation impacts of active travel: evidence from a longitudinal panel study in seven European cities. Glob Environ Chang 67:102224

12

1 Introduction

2. Gómez Vilchez JJ, Jochem P (2020) Powertrain technologies and their impact on greenhouse gas emissions in key car markets. Transp Res Part D Transp Environ 80:102214 3. Logan KG, Nelson JD, Brand C, Hastings A (2021) Phasing in electric vehicles: does policy focusing on operating emission achieve net zero emissions reduction objectives? Transp Res Part A Policy Pract 152:100–114 4. Logan KG, Nelson JD, Hastings A (2021) Low emission vehicle integration: will National Grid electricity generation mix meet UK net zero? Proc Inst Mech Eng Part A J Power Energ. https://doi.org/10.1177/09576509211015472 5. Banister D (2019) The climate crisis and transport. Transp Rev 39:565–568 6. Küfeo˘glu S, Hong DKK, Hong DKK (2020) Emissions performance of electric vehicles: a case study from the United Kingdom. Appl Energ 260:114241 7. Cadez S, Czerny A, Letmathe P (2019) Stakeholder pressures and corporate climate change mitigation strategies. Bus Strateg Environ 28:1–14 8. Yildiz I (2019) Review of climate change issues: a forcing function perspective in agricultural and energy innovation. Int J Energ Res 43:2200–2215 9. Helmers E, Leitão J, Tietge U, Butler T (2019) CO2-equivalent emissions from European passenger vehicles in the years 1995–2015 based on real-world use: assessing the climate benefit of the European “diesel boom.” Atmos Environ 198:122–132 10. Kavvada A, Metternicht G, Kerblat F, Mudau N, Haldorson M, Laldaparsad S, Friedl L, Held A, Chuvieco E (2020) Towards delivering on the sustainable development goals using earth observations. Remote Sens Environ 247:111930 11. Delafield G, Donnison C, Roddis P, Arvanitopoulos T, Sfyridis A, Dunnett S, Ball T, Logan KG (2021) Conceptual framework for balancing society and nature in net-zero energy transitions. Environ Sci Policy 125:189–201 12. Bali Swain R, Ranganathan S (2021) Modeling interlinkages between sustainable development goals using network analysis. World Dev 138:105136 13. Spaiser V, Ranganathan S, Swain RB, Sumpter DJT (2017) The sustainable development oxymoron: quantifying and modelling the incompatibility of sustainable development goals. Int J Sustain Dev World Ecol 24:457–470 14. Schmidt-Traub G, Kroll C, Teksoz K, Durand-Delacre D, Sachs JD (2017) National baselines for the sustainable development goals assessed in the SDG index and dashboards. Nat Geosci 10:547–555 15. Saito O, Managi S, Kanie N, Kauffman J, Takeuchi K (2017) Sustainability science and implementing the sustainable development goals. Sustain Sci 12:907–910 16. Allen C, Metternicht G, Wiedmann T (2019) Prioritising SDG targets: assessing baselines, gaps and interlinkages. Sustain Sci 14:421–438 17. Allen C, Metternicht G, Wiedmann T (2017) An iterative framework for national scenario modelling for the sustainable development goals (SDGs). Sustain Dev 25:372–385 18. Weitz N, Carlsen H, Nilsson M, Skånberg K (2018) Towards systemic and contextual priority setting for implementing the 2030 Agenda. Sustain Sci 13:531–548 19. Gao Y, Gao X, Zhang X (2017) The 2 °C global temperature target and the evolution of the long-term goal of addressing climate change—from the United Nations framework convention on climate change to the paris agreement. Engineering 3:272–278 20. Rogelj J, Den Elzen M, Höhne N, Fransen T, Fekete H, Winkler H, Schaeffer R, Sha F, Riahi K, Meinshausen M (2016) Paris Agreement climate proposals need a boost to keep warming well below 2 °C. Nature 534:631–639 21. UNFCCC (2010) Report of the conference of the parties on its fifteenth session, held in Copenhagen from 7 to 19 December 2009, Addendum, Part Two: Action taken by the Conference of the Parties at its fifteenth session. Unfccc 1–43 22. Röck M, Saade MRM, Balouktsi M, Rasmussen FN, Birgisdottir H, Frischknecht R, Habert G, Lützkendorf T, Passer A (2020) Embodied GHG emissions of buildings—The hidden challenge for effective climate change mitigation. Appl Energy 258:114107 23. Roelfsema M, van Soest HL, Harmsen M et al (2020) Taking stock of national climate policies to evaluate implementation of the Paris agreement. Nat Commun 11:2096

References

13

24. Höhne N, den Elzen M, Rogelj J et al (2020) Emissions: world has four times the work or one-third of the time. Nature 579:25–28 25. van Soest HL, den Elzen MGJ, van Vuuren DP (2021) Net-zero emission targets for major emitting countries consistent with the Paris agreement. Nat Commun 12:2140 26. Rogelj J, Geden O, Cowie A, Reisinger A (2021) Net-zero emissions targets are vague: three ways to fix. Nature 265–368 27. Günther SA, Staake T, Schöb S, Tiefenbeck V (2020) The behavioral response to a corporate carbon offset program: a field experiment on adverse effects and mitigation strategies. Glob Environ Chang 64:102123 28. Kornek U, Klenert D, Edenhofer O, Fleurbaey M (2021) The social cost of carbon and inequality: when local redistribution shapes global carbon prices. J Environ Econ Manage 107:102450 29. Mitchell-larson E, Allen M (2021) Prosets: making continued use of fossil fuels compatible with a credible transition to net zero. Clim Chang 1–9 30. Eskander SMSU, Fankhauser S (2020) Reduction in greenhouse gas emissions from national climate legislation. Nat Clim Chang 10:750–756 31. Grantham Research Institute (2021) Climate change Laws of the world 32. Eskander S, Fankhauser S, Setzer J (2020) Global lessons from climate change legislation and litigation. Environ Energ Policy Econ 2:pp 44–82, University of Chicago Press 33. Tsoi KH, Loo BPY, Banister D (2021) “Mind the (Policy-Implementation) Gap”: Transport decarbonisation policies and performances of leading global economies (1990–2018). Glob Environ Chang 68:102250 34. Dhar S, Pathak M, Shukla PR (2018) Transformation of India’s transport sector under global warming of 2 °C and 1.5 °C scenario. J Clean Prod 172:417–427 35. Peters GP, Andrew RM, Canadell JG, Friedlingstein P, Jackson RB, Korsbakken JI, Le Quéré C, Peregon A (2020) Carbon dioxide emissions continue to grow amidst slowly emerging climate policies. Nat Clim Chang 10:3–6 36. Cavoli C (2021) Accelerating sustainable mobility and land-use transitions in rapidly growing cities: Identifying common patterns and enabling factors. J Transp Geogr 94:103093 37. Menezes E, Maia AG, de Carvalho CS (2017) Effectiveness of low-carbon development strategies: evaluation of policy scenarios for the urban transport sector in a Brazilian megacity. Technol Forecast Soc Change 114:226–241 38. de Blas I, Mediavilla M, Capellán-Pérez I, Duce C (2020) The limits of transport decarbonization under the current growth paradigm. Energ Strateg Rev 32:100543 39. Elzinga D, Bennett S, Best D et al (2015) Energy technology perspectives 2015: mobilising innovation to accelerate climate action. Paris Int. Energ Agency 40. Serrenho AC, Norman JB, Allwood JM (2017) The impact of reducing car weight on global emissions: the future fleet in Great Britain. Philos Trans R Soc A Math Phys Eng Sci. https:// doi.org/10.1098/rsta.2016.0364 41. Holmberg K, Erdemir A (2019) The impact of tribology on energy use and CO2 emission globally and in combustion engine and electric cars. Tribol Int 135:389–396 42. IPCC (2014) Climate change 2014 synthesis report summary chapter for policymakers. https:// doi.org/10.1017/CBO9781107415324 43. Loo BPY, Banister D (2016) Decoupling transport from economic growth: extending the debate to include environmental and social externalities. J Transp Geogr 57:134–144 44. Stead D, Banister D (2003) Decoupling the link between economic growth, transport growth and transport energy consumption in Europe. ECEE Summer Study Time turn Down Energ Demand Panel 3:515–526 45. Loo BPY, Tsoi KH, Banister D (2020) Recent experiences and divergent pathways to transport decoupling. J Transp Geogr 88:102826 46. Goldemberg J (1998) Leapfrog energy technologies. Energ Policy 26:729–741 47. Munasinghe M (1999) Is environmental degradation an inevitable consequence of economic growth: tunneling through the environmental Kuznets curve. Ecol Econ 29:89–109

14

1 Introduction

48. Schroeder PM, Chapman RB (2014) Renewable energy leapfrogging in China’s urban development? Current status and outlook. Sustain Cities Soc 11:31–39 49. Gallagher KS (2006) Limits to leapfrogging in energy technologies? Evidence from the Chinese automobile industry. Energ Policy 34:383–394 50. Amatuni L, Ottelin J, Steubing B, Mogollón JM (2020) Does car sharing reduce greenhouse gas emissions? Assessing the modal shift and lifetime shift rebound effects from a life cycle perspective. J Clean Prod 266:121869 51. Temenos C, Nikolaeva A, Schwanen T, Cresswell T, Sengers F, Watson M, Sheller M (2017) Theorizing mobility transitions: an interdisciplinary conversation. Transfers 7:113–129 52. Weiss M, Dekker P, Moro A, Scholz H, Patel MK (2015) On the electrification of road transportation—a review of the environmental, economic, and social performance of electric two-wheelers. Transp Res Part D Transp Environ 41:348–366 53. Bastida-Molina P, Hurtado-Pérez E, Peñalvo-López E, Cristina Moros-Gómez M (2020) Assessing transport emissions reduction while increasing electric vehicles and renewable generation levels. Transp Res Part D Transp Environ 88:102560 54. Logan KG, Nelson JD, Hastings A (2020) Electric and hydrogen buses: shifting from conventionally fuelled cars in the UK. Transp Res Part D Transp Environ. https://doi.org/10.1016/j. trd.2020.102350 55. Logan KG, Nelson JD, McLellan BC, Hasting A, Hastings A (2020) Towards electric and hydrogen rail: potential contribution to net zero. Transp Res Part D Transp Environ 87:102523 56. Brand C, Anable J, Ketsopoulou I, Watson J (2020) Road to zero or road to nowhere? Disrupting transport and energy in a zero-carbon world. Energ Policy 139:111334 57. Lattarulo P, Masucci V, Pazienza MG (2019) Resistance to change: car use and routines. Transp Policy 74:63–72 58. Innocenti A, Lattarulo P, Pazienza MG (2013) Car stickiness: heuristics and biases in travel choice. Transp Policy 25:158–168 59. Andersson A, Winslott Hiselius L, Adell E (2020) The effect of marketing messages on the motivation to reduce private car use in different segments. Transp Policy 90:22–30 60. Lowe M, Aytekin B, Gereffi G (2009) Public transit buses: a green choice gets greener. Cent Globalization Governance Competitiveness 61. Urry J (2016) Mobilities: new perspectives on transport and society. Routledge 62. Mattioli G, Roberts C, Steinberger JK, Brown A (2020) The political economy of car dependence: a systems of provision approach. Energy Res Soc Sci 66:101486 63. Lucas K (2012) Transport and social exclusion: where are we now? Transp Policy 20:105–113 64. Lucas K, Mattioli G, Verlinghieri E, Guzman A (2016) Transport poverty and its adverse social consequences. Proc Inst Civ Eng Transp 169:353–365 65. Mattioli G, Colleoni M (2016) Transport disadvantage, car dependence and urban form. In: Pucci P, Colleoni M (eds) Understanding mobilities for designing contemporary cities. Springer International Publishing, Cham, pp 171–190 66. Mulley C, Nelson J, Ison S (eds) (2021) The Routledge handbook of public transport. Routledge 67. te Morsche W, La Paix PL, Geurs KT (2019) Potential uptake of adaptive transport services: an exploration of service attributes and attitudes. Transp Policy 84:1–11 68. Flores PJ, Jansson J (2021) The role of consumer innovativeness and green perceptions on green innovation use: the case of shared e-bikes and e-scooters. J Consum Behav. https://doi. org/10.1002/cb.1957 69. Currie G (2010) Quantifying spatial gaps in public transport supply based on social needs. J Transp Geogr 18:31–41 70. Redman L, Friman M, Gärling T, Hartig T (2013) Quality attributes of public transport that attract car users: a research review. Transp Policy 25:119–127 71. Nelson JD, Wright S (2021) Flexible Transport Services. In: Mulley C, Nelson JD, Ison S (eds) Handbook of public transport. Routledge, Abingdon, Oxon UK, pp 224–235 72. Offer GJ, Howey D, Contestabile M, Clague R, Brandon NP (2010) Comparative analysis of battery electric, hydrogen fuel cell and hybrid vehicles in a future sustainable road transport system. Energ Policy. https://doi.org/10.1016/j.enpol.2009.08.040

References

15

73. Zhang J, Hayashi Y, Frank LD (2021) COVID-19 and transport: findings from a world-wide expert survey. Transp Policy 103:68–85 74. Sui Y, Zhang H, Shang W, Sun R, Wang C, Ji J, Song X, Shao F (2020) Mining urban sustainable performance: Spatio-temporal emission potential changes of urban transit buses in post-COVID-19 future. Appl Energ 280:115966 75. Tozer L (2020) Catalyzing political momentum for the effective implementation of decarbonization for urban buildings. Energ Policy 136:111042 76. Fuller S (2017) Configuring climate responsibility in the city: carbon footprints and climate justice in Hong Kong. Area 49:519–525 77. Rice JL (2010) Climate, carbon, and territory: greenhouse gas mitigation in seattle, Washington. Ann Assoc Am Geogr 100:929–937 78. van der Ven H, Bernstein S, Hoffmann M (2017) Valuing the contributions of nonstate and subnational actors to climate governance. Glob Environ Polit 17:1–20 79. Hooper T, Austen MC, Beaumont N, Heptonstall P, Holland RA, Ketsopoulou I, Taylor G, Watson J, Winskel M (2018) Do energy scenarios pay sufficient attention to the environment? Lessons from the UK to support improved policy outcomes. Energ Policy 115:397–408 80. Patrizio P, Pratama YW, Mac DN (2020) Socially equitable energy system transitions. Joule 4:1700–1713 81. Shyu C-W (2021) A framework for ‘right to energy’ to meet UN SDG7: policy implications to meet basic human energy needs, eradicate energy poverty, enhance energy justice, and uphold energy democracy. Energ Res Soc Sci 79:102199 82. Child M, Breyer C (2017) Transition and transformation: a review of the concept of change in the progress towards future sustainable energy systems. Energ Policy 107:11–26 83. Sovacool BK, Geels FW (2016) Further reflections on the temporality of energy transitions: a response to critics. Energ Res Soc Sci 22:232–237 84. Sovacool BK, Griffiths S (2020) Culture and low-carbon energy transitions. Nat Sustain 3:685–693 85. Sovacool BK, Turnheim B, Hook A, Brock A, Martiskainen M (2021) Dispossessed by decarbonisation: reducing vulnerability, injustice, and inequality in the lived experience of low-carbon pathways. World Dev 137:105116 86. Sovacool BK, Hess DJ, Cantoni R (2021) Energy transitions from the cradle to the grave: a meta-theoretical framework integrating responsible innovation, social practices, and energy justice. Energ Res Soc Sci 75:102027 87. Kern F, Rogge KS (2016) The pace of governed energy transitions: agency, international dynamics and the global Paris agreement accelerating decarbonisation processes? Energ Res Soc Sci 22:13–17 88. Geels FW, Sovacool BK, Schwanen T, Sorrell S (2017) The socio-technical dynamics of low-carbon transitions. Joule 1:463–479 89. Levenda AM, Behrsin I, Disano F (2021) Renewable energy for whom? A global systematic review of the environmental justice implications of renewable energy technologies. Energ Res Soc Sci 71:101837 90. Del Bene D, Scheidel A, Temper L (2018) More dams, more violence? A global analysis on resistances and repression around conflictive dams through co-produced knowledge. Sustain Sci 13:617–633 91. Fraune C, Knodt M (2018) Sustainable energy transformations in an age of populism, posttruth politics, and local resistance. Energ Res Soc Sci 43:1–7 92. Fast S (2013) Social acceptance of renewable energy: trends, concepts, and geographies. Geogr Compass 7:853–866 93. Devine-Wright P, Batel S, Aas O, Sovacool B, Labelle MC, Ruud A (2017) A conceptual framework for understanding the social acceptance of energy infrastructure: Insights from energy storage. Energ Policy 107:27–31 94. Wüstenhagen R, Wolsink M, Bürer MJ (2007) Social acceptance of renewable energy innovation: an introduction to the concept. Energ Policy 35:2683–2691

16

1 Introduction

95. Sovacool BK, Lakshmi Ratan P (2012) Conceptualizing the acceptance of wind and solar electricity. Renew Sustain Energ Rev 16:5268–5279 96. Hammond GP, O’Grady Á (2017) The potential environmental consequences of shifts in UK energy policy that impact on electricity generation. Proc Inst Mech Eng Part A J Power Energ 231:535–550 97. Foxon T (2013) Transition pathways for a UK low carbon electricity future. Energ Policy 52:10–24 98. Jenkins K, Heffron RJ, McCauley D (2016) The political economy of energy justice: a nuclear energy perspective BT—The Palgrave handbook of the international political economy of energy. In: Van de Graaf T, Sovacool BK, Ghosh A, Kern F, Klare MT (eds) Palgrave Macmillan UK, London, pp 661–682 99. Pye S, Broad O, Bataille C et al (2021) Modelling net-zero emissions energy systems requires a change in approach. Clim Policy 21:222–231 100. Le T-H, Chang Y, Taghizadeh-Hesary F, Yoshino N (2019) Energy insecurity in Asia: a multi-dimensional analysis. Econ Model 83:84–95 101. Gasser P (2020) A review on energy security indices to compare country performances. Energ Policy 139:111339 102. Cherp A, Jewell J (2014) The concept of energy security: beyond the four as. Energ Policy 75:415–421 103. Solaymani S (2020) A demand-side assessment of sustainable energy security in Iran. Int J Energy Water Resour 4:307–320 104. Prasad S, Sheetal KR, Venkatramanan V, Kumar S, Kannojia S (2019) Sustainable energy: Challenges and perspectives. In: Shah S, Venkatramanan V, Prasad R (eds) Sustainable green technologies for environmental management. Springer Singapore, Singapore, pp 175–197 105. Raghoo P, Surroop D, Wolf F, Leal Filho W, Jeetah P, Delakowitz B (2018) Dimensions of energy security in small Island developing States. Util Policy 53:94–101 106. Ranjan A, Hughes L (2014) Energy security and the diversity of energy flows in an energy system. Energy 73:137–144 107. Spanjer A (2007) Russian gas price reform and the EU–Russia gas relationship: Incentives, consequences and European security of supply. Energ Policy 35:2889–2898 108. Jamasb T, Pollitt M (2008) Security of supply and regulation of energy networks. Energ Policy 36:4584–4589 109. Vivoda V (2010) Evaluating energy security in the Asia-Pacific region: a novel methodological approach. Energ Policy 38:5258–5263 110. Ang BW, Choong WL, Ng TS (2015) Energy security: definitions, dimensions and indexes. Renew Sustain Energy Rev 42:1077–1093 111. Deese DA (1979) Energy: economics, politics, and security. Int Secur 4:140–153 112. Yergin D (1988) Energy security in the 1990s. Foreign Aff 67:110–132 113. Shao S, Guo L, Yu M, Yang L, Guan D (2019) Does the rebound effect matter in energy import-dependent mega-cities? Evidence from Shanghai (China). Appl Energy 241:212–228 114. Shao S, Yang L, Yu M, Yu M (2011) Estimation, characteristics, and determinants of energyrelated industrial CO2 emissions in Shanghai (China), 1994–2009. Energ Policy 39:6476– 6494 115. Lesbirel SH (2004) Diversification and energy security risks: the Japanese case. Jpn J Polit Sci 5:1–22 116. Matsumoto K, Shiraki H (2018) Energy security performance in Japan under different socioeconomic and energy conditions. Renew Sustain Energ Rev 90:391–401 117. Azzuni A, Breyer C (2018) Definitions and dimensions of energy security: a literature review. WIREs Energy Environ 7:e268 118. Global Energy Assessment Writing Team (2012) Global Energy Assessment: toward a sustainable future. https://doi.org/10.1017/CBO9780511793677 119. Proskuryakova L (2018) Updating energy security and environmental policy: energy security theories revisited. J Environ Manage 223:203–214

References

17

120. Bompard E, Carpignano A, Erriquez M, Grosso D, Pession M, Profumo F (2017) National energy security assessment in a geopolitical perspective. Energy 130:144–154 121. Galvani A, Lew AA, Perez MS (2020) COVID-19 is expanding global consciousness and the sustainability of travel and tourism. Tour Geogr 22:567–576 122. Mostert M, Caris A, Limbourg S (2017) Road and intermodal transport performance: the impact of operational costs and air pollution external costs. Res Transp Bus Manag 23:75–85 123. Helm D (2019) Natural capital: assets, systems, and policies. Oxford Rev Econ Policy 35:1–13 124. Raudsepp-Hearne C, Peterson GD, Tengö M, Bennett EM, Holland T, Benessaiah K, MacDonald GK, Pfeifer L (2010) Untangling the environmentalist’s paradox: why is human well-being increasing as ecosystem services degrade? Bioscience 60:576–589 125. Holland RA, Scott K, Hinton ED et al (2016) Bridging the gap between energy and the environment. Energ Policy 92:181–189 126. Bateman IJ, Harwood AR, Mace GM et al (2013) Bringing ecosystem services into economic decision-making: land use in the United Kingdom. Science (80- ) 341:45–50 127. Haines-Young R, Potschin M (2012) CICES version 4: response to Consultation. Eur Environ Agency 1–17 128. Borja Á, Galparsoro I, Irigoien X et al (2011) Implementation of the European marine strategy framework directive: a methodological approach for the assessment of environmental status, from the Basque Country (Bay of Biscay). Mar Pollut Bull. https://doi.org/10.1016/j.marpol bul.2011.03.031 129. De Dominicis M, Wolf J, O’Hara Murray R (2018) Comparative effects of climate change and tidal stream energy extraction in a shelf sea. J Geophys Res Ocean 123:5041–5067 130. Brown S, Pyke D, Steenhof P (2010) Electric vehicles: the role and importance of standards in an emerging market. Energ Policy 38:3797–3806 131. Andersen PH, Mathews JA, Rask M (2009) Integrating private transport into renewable energy policy: the strategy of creating intelligent recharging grids for electric vehicles. Energ Policy 37:2481–2486

Chapter 2

Passenger Land-Based Road Transport

Abstract Even with technological advances, internal combustion engine vehicles (ICEVs) are unlikely to meet net zero targets, whilst emitting high levels of greenhouse gas emissions in addition to impacting public health. Technological improvements of ICEVs are not enough to meet targets. Therefore, phasing out and banning the sale of new ICEVs as soon as possible could provide a stronger impetus to reduce transport emissions. The integration of low emission vehicles including battery electric vehicles, plug-in hybrid electric vehicles and hybrid electric vehicles is often seen as a method to reduce transport emissions. Although these vehicles are often considered ‘zero emission’ at their point of use, their true environmental impact is dependent on the carbon intensity of electricity used to ‘fuel’ the vehicle. Therefore, without the decarbonisation of electricity generation, environmental benefits of low emission transport will be diminished. This chapter focuses on private vehicles and shows that transitioning to low emission transport faces many barriers including cost, range anxiety and charging infrastructure distribution, which need to be overcome for an effective transition to low emission vehicles. This has resulted in numerous monetary and non-monetary incentives being introduced to encourage this transition. However even with this transition, emission levels will remain high per person per kilometre travelled and other low carbon alternatives need to be considered.

2.1 Personal Vehicles Reducing transport emissions, in particular from personal vehicles, will play a central role in mitigating climate change, with numbers of light duty vehicles projected to increase from 900 million in 2012 to 1.7 billion in 2035 [1, 2]. Internal combustion engine vehicles (ICEVs) remain a universal aspect of most societies often viewed as a necessity [3, 4], however they negatively contribute to the environment through greenhouse gas (GHG), nitrogen oxide (NOx ) and particulate matter (PM) emissions. In rapidly industrialising countries, such as China, private vehicle ownership continues to increase rapidly [5]. In 2016, China reported that the number of vehicles

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_2

19

20

2 Passenger Land-Based Road Transport

had increased by 26.15 times since 2000, with the average distance travelled also increasing dramatically [5]. The negative environmental impact of ICEVs is manifested by high energy consumption, air pollution [6] with country specific impacts such as the photochemical smog [7] in China, traffic congestion and traffic noise etc. [8–10]), in addition to public health impacts [11]. There is an increased awareness of the adverse health impacts associated with a range of transport-related exposures and practices [12], including premature mortality rates [13–15]. For example, a recent study in Barcelona, Spain assessed the health impacts of urban transport-related exposures (including air pollution, noise, heat, green space, and physical activity), with results suggesting that 20% of premature mortality may be preventable by changing current urban transport practices to more sustainable transport measures [16]. In December 2020, a London coroner made legal history by ruling that air pollution exposure linked to traffic was a cause of the death of a nine-year-old girl in February 2013 [17]. The average new passenger car purchased in the EU, Iceland, Norway, and the UK exhibited a steady decline of almost 22 g of carbon dioxide per kilometre (gCO2 km−1 ) between 2010 and 2016. However, average emissions increased by 2.8 gCO2 km−1 in 2017 and 2018, and by 1.6 gCO2 km−1 in 2019, with the average car in 2021 emitting 122.3 gCO2 km−1 [18]. Although this remains below the EU target of 130 gCO2 km−1 that was set for 2019, the EU’s new target is that new vehicles to emit less than 95 gCO2 km−1 from 2021 [19]. The recent gradual increase in emission levels is likely due to the increased uptake of petrol fuelled sports utility vehicles (SUVs) [19–21]. SUVs represented 38% of new car registrations and an average emission of 134 gCO2 km−1 , approximately 13 gCO2 km−1 higher than the average emissions of new petrol cars [19].

2.1.1 Travel Behaviour To reduce the environmental and health concerns associated with private vehicle use, and to understand the reasons behind why individuals chose to travel, it is useful to contextualise why certain transport modes are chosen within society. Travel behaviour is in part based on individual preference, attitudes, and perceptions of the different transport modes; however various factors can influence this [22]. For example, traffic levels, traffic density and parking availability and public activity levels, all influence choice. In addition, some individuals may be influenced by climate change and the carbon emissions produced from fuel when the vehicle is in use [23, 24]. Lifestyle and socio-cultural factors including expenditure patterns, localism, multiple car ownership, (un)acceptability of air travel, social norms, habits and, changes in the number of and composition of the population can also influence transport uptake [23]. Furthermore, social changes such as an increasing and ageing population can lead to changes in personal mobility with active travel (such as walking and cycling) often decreasing with age and the uptake of car use simultaneously

2.1 Personal Vehicles

21

increasing [25]. Alternatively, millennials (those born in the 1980s and 1990s) own fewer vehicles than previous generations when they were young; however, when millennials become more economically independent, young adults they may own slightly more vehicles [26]. Residential location and distance to work and leisure amenities also plays a key role in transport choice. For example, individuals who live in a more suburban-style neighbourhood (i.e., with low density and diversity) with limited public transport and active travel opportunities are more likely to travel by personal vehicle and have relatively long travel distances [27], which can be partially explained by differences in household car ownership [28–30]. Individuals who live in rural areas may also be forced to purchase multiple cars (i.e., become “captive” car owners) as public transport is often lacking and common destinations are beyond a walking and cycle distance [28]. Alternatively, individuals who live in more urban locations may not own a personal vehicle (or multiple vehicles) as destinations are often nearby and are more likely to have public transport available [28]. Furthermore, poor accessibility and mobility can result in social exclusion, where an individual is unable to participate effectively within society [31]. For example, a study in Brisbane, Australia, studies highlighted that transport choices influence social participation and the daily life of older individuals (aged 57–87) [32]. This further emphasises that social exclusion within any age group restricts socioeconomic participation, ultimately impacting an individual’s health, life quality, cohesion, and equity in society [33]. With many countries experiencing an aging population, the mobility needs of older individuals has been drawing increasing attention, particularly as the cost and convenience of personal vehicles remains a popular method of transport for the older population. However, trends towards car dependence may not be positive within an ageing society [34] as this may contribute towards road traffic accidents, congestion, and environmental pollution [35, 36]. For example, in Japan, the Japanese National Police Agency reported that drivers aged 75 or older caused 460 fatal road accidents in 2018, increasing from 8.7% to 14.8% within a ten-year period [35].

2.1.1.1

Personal Vehicle Use

Personal vehicles offer an individual greater privacy, protection, autonomy, freedom, and control compared to public transport whilst also allowing individual expression of personal identity, status, and maturity [4, 37, 38]. Furthermore, personal vehicle travel is also often cheaper, more flexible, convenient, and predictable than other alternatives [37]. These are important factors for the individual who takes on the family caring role as cost, convenience, and contingency for both routine needs and emergencies make personal vehicle use more favourable. Car users often attribute social status and personal image to using and owning a car, also known as ‘car pride’ [39]. A study of New York and Houston, USA demonstrated that an individual with a higher ‘car pride’ is strongly predictive of a greater likelihood of owning a vehicle [40]. Other socio-demographic factors have been proven to have considerable impact on personal vehicle ownership [7]. For

22

2 Passenger Land-Based Road Transport

example, income plays a key role in personal vehicle ownership, with high-income households more likely to own at least one vehicle than low-income households [41]. In addition, larger households, gender, number of children and a high number of licensed drivers and workers all contribute to higher vehicle ownership [41, 42]. Educational level also plays a key role as this is directly related to income, with a higher level of education likely to increase income and result in higher vehicle ownership [29]. As private vehicles are often used for commuting and leisure purposes, they are often not at full capacity. For example, in the UK, the average vehicle only has ~1.6 passengers per trip [43]. This means that the emission levels per person per kilometre travelled remain high compared to an individual using active travel. Furthermore, cars are parked for ~95% of their lifetime, which often results in the large numbers of poorly utilised vehicles in cities [44, 45]. As a result, some countries have introduced car sharing schemes to reduce car usage by emphasising the benefits of access to a vehicle without the associated costs and responsibilities of ownership. The goal is to see an increased usage in the vehicle, decreasing the emissions produced per person per kilometre travelled [46, 47]. For example, in the Netherlands, one study highlighted that ~40% of car drivers would be willing to replace some of their private car trips by carsharing with an ~20% indicating that they would consider not purchasing a new vehicle if carsharing was available near to them [48]. This demonstrates that there is potential to decrease car ownership. However, this still highlights that ~80% of individuals would not be willing to give up their personal vehicle, therefore alternatives need to be considered.

2.1.2 Integrating Low Emission Personal Vehicles Recent technological improvements in vehicle efficiency have reduced the energy requirements of road transport, causing a reduction in (CO2 ) emissions [49]. However, numerous studies have highlighted that these improvements alone will not be enough to decrease emission levels to meet Paris Agreement targets [50, 51]. Thus, a modal shift away from ICEVs towards low-carbon alternatives holds considerable potential to mitigate carbon emissions [52]. Given the limitations of these proposed technological advances, many countries have decided to ban and phase out the sale of ICEVs and encourage low emission vehicles purchase. For example, in Asia (India by 2030, Israel by 2030), across Europe (France by 2025, Ireland by 2030, Netherlands by 2030, Norway by 2025, Sweden by 2030, UK by 2030) and in South America (Costa Rica by 2021) [53]. Whilst most countries have taken this approach, China has become an early adopter of battery electric vehicles (BEVs) through early adoption, vehicle production and implementation of associated charging infrastructure through a ‘leapfrogging’ approach [54]. Leapfrogging allows countries to skip stages in development and avoid pollution stages in development whilst ‘leaping’ ahead to become a leading technological innovator and utiliser [55–59].

2.1 Personal Vehicles

23

Low emission vehicle adoption is currently highly dependent on a strong demandside orientation of policies in place [60–63], which range from road privileges (travelling within bus lanes or preferential parking) to financial benefits (tax deductions and purchase subsidies) [64, 65]. Most studies focus on BEV adoption is in areas which have already implemented strong policies, for example China, Germany, Norway etc., however BEV uptake remains low where BEV policies are lacking [63]. This is because human technology systems possess inertia and display resistance to change which makes them durable for multiple social, economic, and technical reasons [66]. Transitioning to low emission vehicle fleets will require significant trade-offs between policies [67]. Therefore, understanding the drivers and barriers of the adoption of BEVs is essential to better understand what can accelerate the adoption of low emission vehicles over ICEVs.

2.2 Battery Electric Vehicles With ICEVs being phased out across the world and the ever-growing climate change concerns, technological improvements and cost reductions, there has been a growth in the integration of low carbon alternatives including BEVs [68]. However, there is a growing consensus that this technological transition will not be sufficient or fast enough to transform the transport system [69, 70]. Even with the technical difficulties being mostly resolved for BEVs, and policy incentives introduced, the target penetration in most countries will not been achieved. The International Energy Agency’s EV 30@30 scenario highlights that there may be as many as 44 million BEV sales per year by 2030, the equivalent of ~30% sales share for BEVs by 2030 among the participating countries [71]. Electric transport is not a new technology and dates back to the nineteenth century in the UK, when Thomas Parker, an inventor and innovator, was responsible for electrifying the London Underground, overhead tramways in Liverpool and Birmingham, and claimed to have perfected the first EV as early as 1884 [72]. However, production of the internal combustion engine (ICE) powered by petrol and diesel was considered easier to produce on mass production lines, making this more affordable to the average consumer. BEVs differ from ICEVs as they replace the ICE with an electric motor, the mechanical transmission system is simplified, electricity storage, charging and control systems are added, and brake energy recover systems are used [73–78]. Although climate change and vehicle-based air pollution has become central within the public debate about transport, there remain challenges that need to be addressed when estimating emission levels to ensure targets are going to be met [52]. BEVs are considered ‘zero emission’ at their point of use as they do not produce tailpipe emissions [79]. Therefore, the carbon intensity of electricity generation will determine the environmental impact of BEVs in the transport network [80]. Numerous studies have been used to investigate the emission levels between ICEVs and BEVs (including plug-in hybrid electric vehicles (PHEVs) and hybrid electric vehicles (HEVs)), with results varying based on model used and differences

24

2 Passenger Land-Based Road Transport

for countries/regions [81]. These studies have all highlighted that the deep decarbonisation of the electricity network away from fossil fuels towards renewables will be necessary [82], otherwise the environmental benefits of electric transport will be negated [83]. For example, Varga (2013) analysed and compared the CO2 emissions from BEVs and ICEVs in Romania, considering the electricity generation mix [84]. Their study highlighted that if BEV market penetration were to increase in line with the Romanian Government’s BEV policy, emissions would not decrease. Similarly, Faria et al. (2013) analysed BEV and ICEV emissions in Poland, France and Portugal who are heavily reliant on fossil fuels, nuclear energy, and renewables respectively in their generation mixes for BEVs demonstrating different outcomes for car emissions [75]. Furthermore, many studies have highlighted that BEV policies should consider the size and type of the vehicle as many studies focus on standard vehicle sizes whereas SUVs have begun to be more popular.

2.2.1 Hybrid and Plug-In Hybrid Electric Vehicles PHEVs and HEVs have been introduced into the transport network as they consume less energy than ICEVs and could act as a ‘transition technology’ towards battery BEVs. This is because these transport types have an ICE as well as the ability to rely on electricity, with most PHEVs having an all-electric range of ~50 kms [85]. Furthermore, in the UK, the Department of Transport (DfT) predicts that PHEVs will replace ICEVs, before a full transition to BEVs by 2050 due to factors such as range anxiety, lack of charging infrastructure and time required for charging [86, 87]. Although PHEVs and HEVs look like promising alternatives to BEVs, as they could reduce range anxiety, they remain a controversial transport type. This is primarily due to the ‘dieselgate’ scandal in 2015, uncovered by the US Environmental Protection Agency, who discovered a software in diesel car engines allowing them to cheat on emissions tests [88]. This software was installed in over 11 million vehicles produced by several car manufacturers including the German Volkswagen AG (VW) and its subsidiary companies. This has resulted in an increased divergence, or ‘gap’ between ‘real world’ and ‘official’ energy use and air pollutant emissions of road vehicles [89, 90]. This ‘divergence’ is often linked to differences in the driving cycle (i.e., speed, acceleration, and altitude profile), vehicle conditions (i.e., test mass, driving resistances, start conditions, etc.) and the optimisations of the vehicle control strategies for the type of approval test (if applied) compared with ‘real world’ driving [91]. To estimate the official vehicle emission measurement for the purpose of certification, in the EU the Worldwide Harmonized Light Vehicle Test Procedure (WLTP) (Regulation ((EU) 2018/1832), was made mandatory from 2018 onwards [92]. This will see passenger car emissions reduce by 15% by 2025 and 37.5% by 2030 relative to 2021 standards [92]. The WLTP uses a utility factor (UF) which determines the share of the test completed by the PHEV using electricity (versus fuel) [93]. The UF ranges from 0, i.e., an ICEV or hybrid electric vehicle that only drives on a conventional fuel, to 1, i.e., a PHEV and a BEV that only drives electric

2.2 Battery Electric Vehicles

25

[94]. Research indicates that there is heterogeneity in the value of UF compared with the values used in the WLTP test among different users. This ‘divergence’ between the certification value and real-world emissions raises scepticism at multiple levels: policy, industry, market [95] with many countries transitioning away from PHEVs and HEVs in favour of BEVs. This has primarily been through a combination of push and pull measures. For example, deterring the use of ICEVs, PHEVs and HEVs through higher taxes, whilst simultaneously encouraging the sales of BEVs purchase through grants/subsidies for purchase costs and charging infrastructure, scrappage schemes for old vehicles and other benefits such as reduced toll fees, driving in bus lanes and free charging.

2.3 Hydrogen Vehicles Over the last decade, there have been a growing number of studies focusing on hydrogen playing a crucial role in the global sustainability strategy within the transport sector [96–98]. This is because hydrogen fuel cell vehicles, if introduced with decarbonised electricity, could reduce GHG pollution by ~80% in 2100 compared to the 1990 levels in the USA [99]. In 2020, there were three light-duty hydrogen fuel cell electric vehicles commercially available in the USA: the Toyota Mirai, Hyundai Nexo, and the Honda Clarity [100]. Outside the USA, China has been embracing a hydrogen future with the aim of one million hydrogen vehicles on the road by 2029 [100]. Similar to electric transport, hydrogen vehicles are considered ‘zero emission’ at their point of use, however hydrogen requires electricity and again the true environmental costs are determined by the carbon intensity of electricity generation. Hydrogen has gained significant attention as a promising energy vector for renewable-rich energy due to the high gravimetric energy density that ensures it remains desirable for both stationary and mobile applications [101]. Almost 90% of hydrogen is produced through natural gas reform which results in CO2 emissions being produced [102], however hydrogen generated electrochemically through water electrolysis is the most environmentally friendly method with only oxygen produced as a by-product [103]. This process accounts for ~4% of global hydrogen production [104]. Hydrogen generated through electrolysis has limited deployment due to the economic barriers associated with production, often dependent on the electricity costs [105]. If the water electrolysers are powered by fluctuating renewables, unsustainable gas crossover between the cathodic and anodic chambers can result in an explosive mixture of oxygen and hydrogen being produced [106, 107]. Therefore, if hydrogen technology is to advance, ensuring that the production of hydrogen remains sustainable will be necessary to keep emission levels low. Hydrogen transport offers an alternative to battery electric transport’s generally limited range and long charging times, but research has remained several years behind battery electric transport [102], and there are significantly higher capital costs [108]. Despite hydrogen vehicles becoming commercially available, from an economic

26

2 Passenger Land-Based Road Transport

stance, BEVs currently dominate the market because they are more cost competitive to ICEVs [109]. Similar to electric transport, there remains the ‘chicken and egg’ scenario as consumer demands for hydrogen remain low and it is often not considered economically feasible to build large-scale hydrogen fuelling stations [96]. Therefore, integration will require significant nationwide infrastructure requirements to ensure market progression [108]. Due to the advantages and disadvantages of both electric and hydrogen transport types for reducing transport emissions, provided they are fuelled with decarbonised electricity, a combination of both transport types will most likely be needed to meet demand [110]. If electricity generation is decarbonised, it should therefore be considered that ‘transitioning’ towards hydrogen alternatives should be favoured over PHEVs as these provide the range required with lower emission levels.

2.4 Barriers for Low Emission Transport Although there are environmental benefits of transitioning to low emission vehicles in comparison with ICEVs, there remain several barriers for widescale integration. These include costs (i.e., upfront costs, operation and maintenance, insurance, resale value etc.), range anxiety, lack of charging infrastructure and time required for charging [111]. These barriers to widespread integration need to be overcome or emission levels from transport will remain static.

2.4.1 Associated Costs Although the average BEV has a higher upfront cost at ~1.4 times higher than an average ICEV [112], governments and manufacturers are striving to reduce the purchase price [113–117]. With that said, the overall vehicle lifetime running costs are often lower for BEVs, for example, BEVs cost ~$0.02 per mile travelled compared with ICEVs which cost ~$0.12 per mile [118]. The cost difference between fossil fuel and electrical energy is maintained by higher taxes (in the form of fuel duty and VAT) on fossil fuels than electricity. Furthermore, battery costs remain a determinant in terms of the economic feasibility of BEV uptake [119], with Bloomberg estimating costs to be as low as $100/kWh by 2023 [120]. Without this decrease in battery costs, upfront costs will not decrease as individuals do not want to pay for a higher initial cost, even if savings are likely to be seen during the vehicle’s lifetime [121, 122]. In addition, there remains high uncertainty for BEVs in regard to maintenance costs and the depreciation of BEVs which may hinder adoption [123]. Furthermore, recent purchase subsidies in Italy highlight that BEVs have become cost competitive in terms of the total cost of ownership (TCO) in comparison to diesel vehicles but not with petrol vehicles, unless there is a high annual distance driven (more than 12,500 km annually) [124]. Although the TCO is not often readily

2.4 Barriers for Low Emission Transport

27

available to potential buyers, this may not be a primary reason that can influence a buyer’s decision as TOCs are generally underestimated by individuals [125].

2.4.2 Range Anxiety and Charging Infrastructure Multiple studies have highlighted that the limited range of BEVs, long charging times and public infrastructure availability remains a significant obstacle which can cause what is known as ‘range anxiety’ [119, 126–128]. Several studies have highlighted that current BEV range satisfies the range needed for most users, with increased BEV experience lowering this anxiety [129, 130]. For example, a study using a high-resolution dataset of petrol users in 13 counties in Atlanta, Georgia in the USA demonstrated that a significant (although low) fraction of transport needs could be met with BEVs without adaptation [131]. Therefore, an important, but not sole, cause for range anxiety is that it will likely force a driver to change their driving pattern [132]. As BEV are more limited than ICEVs and are sometimes required for long distance trips, the availability of charging points is essential and remains a significant barrier concerning BEV consumers [133]. This is a ‘chicken and egg’ scenario as without increased BEV uptake, stakeholders are often unwilling to invest in additional infrastructure for recharging BEVs. This will be particularly important in rural locations as many consumers require assurance that they will be able to travel the distance needed. Several studies have highlighted that charging infrastructure was the best predictor for the penetration for low emission vehicles [60, 134]. Furthermore, the existence of fast public charging infrastructure has the ability to enhance the possibilities for driving longer distances and reduce the related social costs [135, 136]. As charging time remains long, introduction of fast-charging infrastructure has seen consumers willing to pay a higher price for fast-charging [137]. However, in an Austrian study, ~88% of BEV users are likely to charge their vehicle at home, ~8.8% of users are likely to charge their vehicle at work with ~1.7% and ~1.5% occurring at public or fast-charging stations [138]. This highlights that public charging infrastructure is only used for a small proportion of charges and may only be required to reduce individual range anxiety, (potentially) resulting in a waste of funds. In addition, if the driving range of a vehicle increases then there is the possibility of a trade-off between high battery capacity and a high density of charging infrastructures as significant investment in public charging infrastructure may not be required [139].

2.5 Incentives for Low Emission Transport To encourage a shift towards low carbon private vehicles, several incentives have been introduced to lower the costs of BEVs, HEVs and PHEVs in most countries [140, 141]. Monetary incentives are the most popular for incentivising consumers to

28

2 Passenger Land-Based Road Transport

adopt low emission vehicles, however many countries have introduced non-monetary assistances that consumers will benefit from.

2.5.1 Monetary Incentives Monetary incentives such as purchase incentives at the point-of-sale grants, VAT and purchase tax exemptions, post purchase rebates and income tax credits are very common to encourage low emission vehicle uptake [142]. Point of sale grants reduce the upfront cost of a new low emission vehicle when a consumer purchases in the form of a government discount or grant at the point of purchase [142]. For example, the PHEV grant [143] and the one-off purchase grant of £4,500 in the UK. Similarly, the Green Vehicle Purchasing Promotion and the Clean Energy Subsidy programme in Japan allow consumers to get a discount on new purchases with the monetary value higher for BEVs, followed by PHEVs and a smaller discount for diesel passenger vehicles [144, 145]. However, these fiscal measures often differ between counties and within countries such as in Canada which offers purchase grants administered at the Provincial level in British Columbia, Quebec and Ontario which range between CA$5,000–8,500. This is similar to Belgium which offers financial incentives based on a regional basis, and the USA which differs between states [146]. VAT and purchase tax exemptions allow consumers to purchase a low emission vehicle at lower or zero VAT or pay no purchase tax [142]. However, the way in which VAT exemption is implemented varies widely. For example, in the Netherlands individuals who purchase a BEV do not pay a purchase tax, however the tax of other vehicles is calculated based on the CO2 emissions produced. Portugal has a VAT reduction for PHEVs if the vehicle value is less than e50,000. Similarly, in the Canary Islands in Spain, PHEVs are VAT exempt, however the vehicle has to emit less than 110 gCO2 km−1 . Some countries including Brazil, Hong Kong, the Netherlands, Norway, Switzerland, and Taiwan do not offer any national subsidies but different forms of tax benefits [146]. Post purchase rebates are given to the consumer after the low emission vehicle has been purchased [142]. This is a common incentive in the USA, in use in several states including California where BEV buyers can apply for a $2,500 rebate and PHEV buyers can apply for a $1,500 rebate. Income tax credits are the least common financial incentive to encourage low emission vehicle uptake which allows consumers to pay a reduced income tax bill at the end of the financial year [142]. In the USA, consumers who have purchased a BEV can pay $7,500 less in tax, however if the consumer does not have a tax liability of this amount, then they can only claim up to their level of liability [142]. Finally, other incentives have been introduced such as in Brazil where BEV users are exempt from import duty [147].

2.5 Incentives for Low Emission Transport

29

2.5.2 Non-monetary Incentives Other non-fiscal incentive schemes exist across the world. For example, in Norway, BEV and PHEV users are exempted from toll charges. Further analysis highlighted that toll exceptions and the right to use bus designated lanes, did have a statistically significant impact on the sale of BEVs in Norway. However, their results could be influenced by neighbouring major cities not having these incentives [148]. Conflictingly, another study highlighted that 53% of the respondents in the Norwegian Electric Vehicle Association survey from 2016 stated that road tolls were an important incentive, with 14% reporting free parking and 12% reporting access to bus lanes important [149]. In 2017, Greece abolished the luxury tax for low emitting PHEVs, which was an additional tax for passenger cars based on the wholesale price of a vehicle. The tax was specifically aimed at vehicles with a price over e20,000. These have since been replaced with a 15% cashback for PHEVs that emit ≤50 gCO2 km−1 of up to e8,000, plus an extra e2,500 if an old taxi is scrapped. In addition, consumers received 15% cashback for vans (up to e4,000 for PHEVs), plus e1,000 for scrapping. A number of countries including Denmark, Germany, Netherlands, Italy, and the UK have introduced various ultra-low emission zones with standards applied to passenger cars and vans (as well as other transport types) to be able to travel within the zones [150]. As most low carbon transport produces lower grams of carbon dioxide than ICEVs, they are often allowed to travel within these zones. Vehicles who do not meet these standards either get charged a fee or a fine if they do not comply, e.g., in London, UK [151].

2.6 Key Findings Although there are clear environmental and health benefits associated with transitioning from an ICEV to a BEV, as this is the lowest emission alternative vehicle, travel behaviour choices indicate that individuals opt for ICEVs for multiple reasons including: cost, convenience, and status. However, over the last decade many countries have introduced new legislation to phase out and ban the sale of ICEVs in favour of low emission alternatives. However, earlier integration will decrease emission levels, making Paris Agreement targets more likely to be achieved. There remain several barriers for widespread integration that need to be overcome. A number of these barriers, including cost, are likely to decrease over time as technology advances making widespread integration much more plausible. Range anxiety may be able to be mitigated using hydrogen vehicles and as batteries advance, the need for public charging infrastructure decreases as vehicle range increases. Although hydrogen vehicles are a new technology and do not yet offer the same availability and choice as a BEV, this low carbon alternative could reduce this barrier. Encouraging this transition has led to numerous monetary and non-monetary incentives to be introduced with purchase grants the most common incentive.

30

2 Passenger Land-Based Road Transport

As this chapter highlights, even with this transition towards low emission personal vehicles, other alternatives need to be considered as emissions will remain high during this transition period. Therefore, consideration needs to be placed on low carbon public transport which is the focus of the following chapters.

References 1. Lam A, Mercure J-F (2021) Which policy mixes are best for decarbonising passenger cars? Simulating interactions among taxes, subsidies and regulations for the United Kingdom, the United States, Japan, China, and India. Energy Res Soc Sci 75:101951 2. IEA (2012) World Energy Outlook 2012. 3. Mann E, Abraham C (2006) The role of affect in UK commuters’ travel mode choices: an interpretative phenomenological analysis. Br J Psychol 97:155–176 4. Woods R, Masthoff J (2017) A comparison of car driving, public transport and cycling experiences in three European cities. Transp Res Part A Policy Pract 103:211–222 5. Ao Y, Yang D, Chen C, Wang Y (2019) Exploring the effects of the rural built environment on household car ownership after controlling for preference and attitude: evidence from Sichuan, China. J Transp Geogr 74:24–36 6. Johansson C, Lövenheim B, Schantz P, Wahlgren L, Almström P, Markstedt A, Strömgren M, Forsberg B, Sommar JN (2017) Impacts on air pollution and health by changing commuting from car to bicycle. Sci Total Environ 584–585:55–63 7. Li S, Zhao P (2017) Exploring car ownership and car use in neighborhoods near metro stations in Beijing: does the neighborhood built environment matter? Transp Res Part D Transp Environ 56:1–17 8. Miedema HME (2007) Adverse effects of traffic noise. In: Gärling T, Steg L (eds) Threats from car traffic to the quality of urban life. Emerald Group Publishing Limited, pp 53–77 9. Levy JI, Buonocore JJ, von Stackelberg K (2010) Evaluation of the public health impacts of traffic congestion: a health risk assessment. Environ Heal 9:65 10. Hennessy D, Wiesenthal D (1997) The relationship between traffic congestion, driver stress and direct versus indirect coping behaviours. Ergonomics 40:348–361 11. Milakis D, van Arem B, van Wee B (2017) Policy and society related implications of automated driving: a review of literature and directions for future research. J Intell Transp Syst 21:324– 348 12. Khreis H, May AD, Nieuwenhuijsen MJ (2017) Health impacts of urban transport policy measures: a guidance note for practice. J Transp Heal 6:209–227 13. Pucher J, Buehler R, Bassett DR, Dannenberg AL (2010) Walking and cycling to health: a comparative analysis of City, State, and International Data. Am J Public Health 100:1986– 1992 14. Mueller N, Rojas-Rueda D, Cole-Hunter T, de Nazelle A, Dons E, Gerike R, Götschi T, Int Panis L, Kahlmeier S, Nieuwenhuijsen M (2015) Health impact assessment of active transportation: a systematic review. Prev Med (Baltim) 76:103–114 15. Oja P, Titze S, Bauman A, de Geus B, Krenn P, Reger-Nash B, Kohlberger T (2011) Health benefits of cycling: a systematic review. Scand J Med Sci Sports 21:496–509 16. Mueller N, Rojas-Rueda D, Basagaña X et al (2017) Urban and transport planning related exposures and mortality: a health impact assessment for cities. Environ Health Perspect 125:89–96 17. Taylor L (2021) WHO cuts air pollution limits to save millions of lives. Br Med J (Online) 374 18. EEA (2021) CO2 performance of new passenger cars in Europe. https://www.eea.europa.eu/ ims/co2-performance-of-new-passenger

References

31

19. EEA (2020) Average CO2 emissions from new cars and new vans increased again in 2019 20. Brand C, Anable J, Ketsopoulou I, Watson J (2020) Road to zero or road to nowhere? Disrupting transport and energy in a zero carbon world. Energy Policy 139:111334 21. Watson J, Bradshaw M, Froggat A et al (2019) Review of Energy Policy 2019 22. Gardner B, Abraham C (2008) Psychological correlates of car use: a meta-analysis. Transp Res Part F Traffic Psychol Behav 11:300–311 23. Brand C, Anable J, Morton C (2019) Lifestyle, efficiency and limits: modelling transport energy and emissions using a socio-technical approach. Energy Effic 12:187–207 24. Yang C, McCollum D, McCarthy R, Leighty W (2009) Meeting an 80% reduction in greenhouse gas emissions from transportation by 2050: a case study in California. Transp Res Part D Transp Environ 14:147–156 25. Shergold I, Lyons G, Hubers C (2015) Future mobility in an ageing society—Where are we heading? J Transp Heal 2:86–94 26. Klein NJ, Smart MJ (2017) Millennials and car ownership: Less money, fewer cars. Transp Policy 53:20–29 27. Ewing R, Cervero R (2010) Travel and the built environment. J Am Plan Assoc 76:265–294 28. De Vos J, Cheng L, Witlox F (2020) Do changes in the residential location lead to changes in travel attitudes? A structural equation modeling approach. Transportation (Amst). https:// doi.org/10.1007/s11116-020-10119-7 29. Ding C, Wang D, Liu C, Zhang Y, Yang J (2017) Exploring the influence of built environment on travel mode choice considering the mediating effects of car ownership and travel distance. Transp Res Part A Policy Pract 100:65–80 30. Van Acker V, Witlox F (2010) Car ownership as a mediating variable in car travel behaviour research using a structural equation modelling approach to identify its dual relationship. J Transp Geogr 18:65–74 31. Mariotti I, Burlando C, Landi S (2021) Is local public transport unsuitable for elderly? Exploring the cases of two Italian cities. Res Transp Bus Manag 100643 32. Zeitler E, Buys L (2015) Mobility and out-of-home activities of older people living in suburban environments: ‘because I’m a driver, I don’t have a problem”. Ageing Soc 35:785–808 33. Al-Rashid MA, Goh HC, Harumain YAS, Ali Z, Campisi T, Mahmood T (2021) Psychosocial barriers of public transport use and social exclusion among older adults: empirical evidence from Lahore, Pakistan. Int J Environ Res Public Health. https://doi.org/10.3390/ijerph180 10185 34. Currie G, Delbosc A (2010) Exploring public transport usage trends in an ageing population. Transportation (Amst) 37:151–164 35. Liu S, Yamamoto T, Yao E, Nakamura T (2021) Examining public transport usage by older adults with smart card data: a longitudinal study in Japan. J Transp Geogr 93:103046 36. Rosenbloom S (2001) Sustainability and automobility among the elderly: an international assessment. Transportation (Amst) 28:375–408 37. Anable J, Gatersleben B (2005) All work and no play? The role of instrumental and affective factors in work and leisure journeys by different travel modes. Transp Res Part A Policy Pract 39:163–181 38. Beirão G, Sarsfield Cabral JA (2007) Understanding attitudes towards public transport and private car: a qualitative study. Transp Policy 14:478–489 39. Moody J, Zhao J (2020) Travel behavior as a driver of attitude: car use and car pride in U.S. cities. Transp Res Part F Traffic Psychol Behav 74:225–236 40. Moody J, Zhao J (2019) Car pride and its bidirectional relations with car ownership: case studies in New York City and Houston. Transp Res Part A Policy Pract 124:334–353 41. Potoglou D, Kanaroglou PS (2008) Modelling car ownership in urban areas: a case study of Hamilton, Canada. J Transp Geogr 16:42–54 42. Tyrinopoulos Y, Antoniou C (2013) Factors affecting modal choice in urban mobility. Eur Transp Res Rev 5:27–39 43. DfT (2019) Vehicle mileage and occupancy (NTS0905: Car occupancy, England: since 2002)

32

2 Passenger Land-Based Road Transport

44. Berg J, Henriksson M, Ihlström J (2019) Comfort first! vehicle-sharing systems in urban residential areas: the importance for everyday mobility and reduction of car use among pilot users. Sustainability. https://doi.org/10.3390/su11092521 45. Shoup DC (1997) The high cost of free parking. J Plan Educ Res 17:3–20 46. Priya Uteng T, Julsrud TE, George C (2019) The role of life events and context in type of car share uptake: Comparing users of peer-to-peer and cooperative programs in Oslo, Norway. Transp Res Part D Transp Environ 71:186–206 47. Shaheen SA, Cohen AP, Chung MS (2009) North American carsharing: 10-year retrospective. Transp Res Rec 2110:35–44 48. Liao F, Molin E, Timmermans H, van Wee B (2020) Carsharing: the impact of system characteristics on its potential to replace private car trips and reduce car ownership. Transportation (Amst) 47:935–970 49. Alam MS, Hyde B, Duffy P, McNabola A (2017) Assessment of pathways to reduce CO2 emissions from passenger car fleets: case study in Ireland. Appl Energ 189:283–300 50. Logan KG, Nelson JD, Hastings A (2020) Electric and hydrogen buses: shifting from conventionally fuelled cars in the UK. Transp Res Part D Transp Environ. https://doi.org/10.1016/j. trd.2020.102350 51. Logan KG, Nelson JD, McLellan BC, Hasting A, Hastings A (2020) Towards electric and hydrogen rail: potential contribution to net zero. Transp Res Part D Transp Environ 87:102523 52. Cuenot F, Fulton L, Staub J (2012) The prospect for modal shifts in passenger transport worldwide and impacts on energy use and CO2 . Energ Policy 41:98–106 53. Senecal PK, Leach F (2019) Diversity in transportation: why a mix of propulsion technologies is the way forward for the future fleet. Results Eng 4:100060 54. Gupta R, Mejia C, Gianchandani Y, Kajikawa Y (2020) Analysis on formation of emerging business ecosystems from deals activities of global electric vehicles hub firms. Energ Policy 145:111532 55. Goldemberg J (1998) Leapfrog energy technologies. Energ Policy 26:729–741 56. Munasinghe M (1999) Is environmental degradation an inevitable consequence of economic growth: tunneling through the environmental Kuznets curve. Ecol Econ 29:89–109 57. Schroeder PM, Chapman RB (2014) Renewable energy leapfrogging in China’s urban development? Current status and outlook. Sustain Cities Soc 11:31–39 58. Logan KG, Nelson JD, Lu X, Hastings A (2020) UK and China: Will electric vehicle integration meet Paris agreement targets? Transp Res Interdiscip Perspect. https://doi.org/10.1016/ j.trip.2020.100245 59. Gallagher KS (2006) Limits to leapfrogging in energy technologies? Evidence from the Chinese automobile industry. Energ Policy 34:383–394 60. Sierzchula W, Bakker S, Maat K, van Wee B (2014) The influence of financial incentives and other socio-economic factors on electric vehicle adoption. Energ Policy 68:183–194 61. Lévay PZ, Drossinos Y, Thiel C (2017) The effect of fiscal incentives on market penetration of electric vehicles: a pairwise comparison of total cost of ownership. Energ Policy 105:524–533 62. Münzel C, Plötz P, Sprei F, Gnann T (2019) How large is the effect of financial incentives on electric vehicle sales?—A global review and European analysis. Energ Econ 84:104493 63. Brückmann G, Willibald F, Blanco V (2021) Battery electric vehicle adoption in regions without strong policies. Transp Res Part D Transp Environ 90:102615 64. Hardman S (2019) Understanding the impact of reoccurring and non-financial incentives on plug-in electric vehicle adoption—a review. Transp Res Part A Policy Pract 119:1–14 65. Brückmann G, Bernauer T (2020) What drives public support for policies to enhance electric vehicle adoption? Environ Res Lett. https://doi.org/10.1088/1748-9326/ab90a5 66. Geels FW (2002) Technological transitions as evolutionary reconfiguration processes: a multilevel perspective and a case-study. Res Policy 31:1257–1274 67. Bhardwaj C, Axsen J, Kern F, McCollum D (2020) Why have multiple climate policies for light duty vehicles? Policy mix rationales, interactions and research gaps. Transp Res Part A Policy Pract 135:309–326

References

33

68. Calearo L, Marinelli M, Ziras C (2021) A review of data sources for electric vehicle integration studies. Renew Sustain Energy Rev 151:111518 69. Brand C, Götschi T, Dons E et al (2021) The climate change mitigation impacts of active travel: evidence from a longitudinal panel study in seven European cities. Glob Environ Chang 67:102224 70. Creutzig F, Roy J, Lamb WF et al (2018) Towards demand-side solutions for mitigating climate change. Nat Clim Chang 8:260–263 71. IEA (2019) Global EV outlook 2019 72. Chan CC (2013) The rise fall of electric vehicles in 1828–1930: Lessons learned [scanning our past]. Proc IEEE 101:206–212 73. Apostolaki-Iosifidou E, Codani P, Kempton W (2017) Measurement of power loss during electric vehicle charging and discharging. Energy 127:730–742 74. Emadi A, Lee YJ, Rajashekara K (2008) Power Electronics and Motor Drives in Electric, Hybrid Electric, and Plug-In Hybrid Electric Vehicles. IEEE Trans Ind Electron 55:2237–2245 75. Faria R, Marques P, Moura P, Freire F, Delgado J, de Almeida AT (2013) Impact of the electricity mix and use profile in the life cycle assessment of electric vehicles. Renew Sustain Energ Rev 24:271–287 76. Tie SF, Tan CW (2013) A review of energy sources and energy management system in electric vehicles. Renew Sustain Energ Rev 20:82–102 77. Zhang H, Tolbert LM, Ozpineci B (2011) Impact of SiC devices on hybrid electric and plug-in hybrid electric vehicles. IEEE Trans Ind Appl 47:912–921 78. Holmberg K, Erdemir A (2019) The impact of tribology on energy use and CO2 emission globally and in combustion engine and electric cars. Tribol Int 135:389–396 79. Bastida-Molina P, Hurtado-Pérez E, Peñalvo-López E, Cristina Moros-Gómez M (2020) Assessing transport emissions reduction while increasing electric vehicles and renewable generation levels. Transp Res Part D Transp Environ 88:102560 80. Weiss M, Dekker P, Moro A, Scholz H, Patel MK (2015) On the electrification of road transportation—a review of the environmental, economic, and social performance of electric two-wheelers. Transp Res Part D Transp Environ 41:348–366 81. Peng T, Ou X, Yan X (2018) Development and application of an electric vehicles life cycle energy consumption and greenhouse gas emissions analysis model. Chem Eng Res Des 131:699–708 82. Creutzig F, Jochem P, Edelenbosch OY et al (2015) Transport: a roadblock to climate change mitigation? Science 350:911–912 83. Woo J, Choi H, Ahn J (2017) Well-to-wheel analysis of greenhouse gas emissions for electric vehicles based on electricity generation mix: a global perspective. Transp Res Part D Transp Environ 51:340–350 84. Varga BO (2013) Electric vehicles, primary energy sources and CO2 emissions: Romanian case study. Energy 49:61–70 85. Adnan N, Md Nordin S, Hadi Amini M, Langove N (2018) What make consumer sign up to PHEVs? Predicting Malaysian consumer behavior in adoption of PHEVs. Transp Res Part A Policy Pract 113:259–278 86. Küfeo˘glu S, Hong DKK, Hong DKK (2020) Emissions performance of electric vehicles: a case study from the United Kingdom. Appl Energy 260:114241 87. Hu X, Sun F, Zou Y, Peng H (2011) Online estimation of an electric vehicle lithium-ion battery using recursive least squares with forgetting. In: Proceedings of the 2011 American Control Conference, pp 935–940 88. Peitzmeier C, Loschke C, Wiedenhaus H, Klemm O (2017) Real-world vehicle emissions as measured by in situ analysis of exhaust plumes. Environ Sci Pollut Res 24:23279–23289 89. Ligterink NE, Eijk ARA (2014) Update analysis of real-world fuel consumption of business passenger cars based on Travelcard Nederland fuelpass data. 25 90. Tietge U, Mock P, Franco V, Zacharof N (2017) From laboratory to road: modeling the divergence between official and real-world fuel consumption and CO2 emission values in the German passenger car market for the years 2001–2014. Energ Policy 103:212–222

34

2 Passenger Land-Based Road Transport

91. Weller K, Lipp S, Röck M, Matzer C, Bittermann A, Hausberger S (2019) Real world fuel consumption and emissions from LDVs and HDVs. 5:1–22 92. Ehrenberger SI, Konrad M, Philipps F (2020) Pollutant emissions analysis of three plug-in hybrid electric vehicles using different modes of operation and driving conditions. Atmos Environ 234:117612 93. Wu X, Aviquzzaman M, Lin Z (2015) Analysis of plug-in hybrid electric vehicles’ utility factors using GPS-based longitudinal travel data. Transp Res Part C Emerg Technol 57:1–12 94. Paffumi E, De Gennaro M, Martini G (2018) Alternative utility factor versus the SAE J2841 standard method for PHEV and BEV applications. Transp Policy 68:80–97 95. Fontaras G, Zacharof N-G, Ciuffo B (2017) Fuel consumption and CO2 emissions from passenger cars in Europe—Laboratory versus real-world emissions. Prog Energy Combust Sci 60:97–131 96. Alazemi J, Andrews J (2015) Automotive hydrogen fuelling stations: an international review. Renew Sustain Energ Rev 48:483–499 97. Ball M, Weeda M (2015) The hydrogen economy—Vision or reality? Int J Hydrogen Energ 40:7903–7919 98. Moliner R, Lázaro MJ, Suelves I (2016) Analysis of the strategies for bridging the gap towards the hydrogen economy. Int J Hydrogen Energ 41:19500–19508 99. Dougherty W, Kartha S, Rajan C, Lazarus M, Bailie A, Runkle B, Fencl A (2009) Greenhouse gas reduction benefits and costs of a large-scale transition to hydrogen in the USA. Energ Policy 37:56–67 100. Zhang C, Greenblatt JB, Wei M, Eichman J, Saxena S, Muratori M, Guerra OJ (2020) Flexible grid-based electrolysis hydrogen production for fuel cell vehicles reduces costs and greenhouse gas emissions. Appl Energ 278:115651 101. Frey D, Kim J, Dvorkin Y, Modestino MA (2020) Spatiotemporal decoupling of water electrolysis for dual-use grid energy storage and hydrogen generation. Cell Rep Phys Sci 1:100226 102. Kolbe K, Lechtenböhmer S, Fischedick M (2019) Hydrogen derived from algae and cyanobacteria as a decentralized fueling option for hydrogen powered cars: size, space, and cost characteristics of potential bioreactors. Int J Sustain Transp 0:325–334 103. Dincer I, Acar C (2015) Review and evaluation of hydrogen production methods for better sustainability. Int J Hydrogen Energ 40:11094–11111 104. Koj JC, Wulf C, Schreiber A, Zapp P (2017) Site-dependent environmental impacts of industrial hydrogen production by Alkaline water electrolysis. Energies. https://doi.org/10.3390/ en10070860 105. Schaber K, Steinke F, Hamacher T (2012) Transmission grid extensions for the integration of variable renewable energies in Europe: who benefits where? Energ Policy 43:123–135 106. Symes MD, Cronin L (2013) Decoupling hydrogen and oxygen evolution during electrolytic water splitting using an electron-coupled-proton buffer. Nat Chem 5:403–409 107. Mergel J, Carmo M, Fritz D (2013) Status on technologies for hydrogen production by water electrolysis. In: Transition to renewable energy systems . Wiley Online Library, pp 425–450 108. Ajanovic A, Haas R (2019) Economic and environmental prospects for battery electric- and fuel cell vehicles: a review. Fuel Cells 19:515–529 109. Edwards RL, Font-Palma C, Howe J (2021) The status of hydrogen technologies in the UK: a multi-disciplinary review. Sustain Energy Technol Assessments 43:100901 110. Staffell I, Scamman D, Velazquez Abad A, Balcombe P, Dodds PE, Ekins P, Shah N, Ward KR (2019) The role of hydrogen and fuel cells in the global energy system. Energy Environ Sci 12:463–491 111. Mukherjee SC, Ryan L (2020) Factors influencing early battery electric vehicle adoption in Ireland. Renew Sustain Energy Rev 118:109504 112. Weldon P, Morrissey P, O’Mahony M (2018) Long-term cost of ownership comparative analysis between electric vehicles and internal combustion engine vehicles. Sustain Cities Soc 39:578–591

References

35

113. Haustein S, Jensen AF (2018) Factors of electric vehicle adoption: a comparison of conventional and electric car users based on an extended theory of planned behavior. Int J Sustain Transp 12:484–496 114. Vassileva I, Campillo J (2017) Adoption barriers for electric vehicles: Experiences from early adopters in Sweden. Energy 120:632–641 115. Hardman S, Shiu E, Steinberger-Wilckens R (2016) Comparing high-end and low-end early adopters of battery electric vehicles. Transp Res Part A Policy Pract 88:40–57 116. Plötz P, Schneider U, Globisch J, Dütschke E (2014) Who will buy electric vehicles? Identifying early adopters in Germany. Transp Res Part A Policy Pract 67:96–109 117. Bahamonde-Birke FJ, Hanappi T (2016) The potential of electromobility in Austria: evidence from hybrid choice models under the presence of unreported information. Transp Res Part A Policy Pract 83:30–41 118. Wilberforce T, El-Hassan Z, Khatib FN, Al Makky A, Baroutaji A, Carton JG, Olabi AG (2017) Developments of electric cars and fuel cell hydrogen electric cars. Int J Hydrogen Energy 42:25695–25734 119. Egbue O, Long S (2012) Barriers to widespread adoption of electric vehicles: an analysis of consumer attitudes and perceptions. Energ Policy 48:717–729 120. Bloomberg NEF (2019) Battery pack prices fall as market ramps up with market average At $156/kWh In 2019 121. Berkeley N, Jarvis D, Jones A (2018) Analysing the take up of battery electric vehicles: An investigation of barriers amongst drivers in the UK. Transp Res Part D Transp Environ 63:466–481 122. Hidrue MK, Parsons GR, Kempton W, Gardner MP (2011) Willingness to pay for electric vehicles and their attributes. Resour Energy Econ 33:686–705 123. Habla W, Huwe V, Kesternich M (2021) Electric and conventional vehicle usage in private and car sharing fleets in Germany. Transp Res Part D Transp Environ 93:102729 124. Scorrano M, Danielis R, Giansoldati M (2020) Dissecting the total cost of ownership of fully electric cars in Italy: the impact of annual distance travelled, home charging and urban driving. Res Transp Econ 80:100799 125. Andor MA, Gerster A, Gillingham KT, Horvath M (2020) Running a car costs much more than people think—stalling the uptake of green travel. Nature 580:453–455 126. Rezvani Z, Jansson J, Bodin J (2015) Advances in consumer electric vehicle adoption research: a review and research agenda. Transp Res Part D Transp Environ 34:122–136 127. She Z-Y, Sun Q, Ma J-J, Xie B-C (2017) What are the barriers to widespread adoption of battery electric vehicles? A survey of public perception in Tianjin, China. Transp Policy 56:29–40 128. Lim MK, Mak H-Y, Rong Y (2015) Toward mass adoption of electric vehicles: impact of the range and resale anxieties. Manuf Serv Oper Manag 17:101–119 129. Needell ZA, McNerney J, Chang MT, Trancik JE (2016) Potential for widespread electrification of personal vehicle travel in the United States. Nat Energy 1:16112 130. Franke T, Krems JF (2013) What drives range preferences in electric vehicle users? Transp Policy 30:56–62 131. Pearre NS, Kempton W, Guensler RL, Elango VV (2011) Electric vehicles: how much range is required for a day’s driving? Transp Res Part C Emerg Technol 19:1171–1184 132. Melliger MA, van Vliet OPR, Liimatainen H (2018) Anxiety versus reality—Sufficiency of battery electric vehicle range in Switzerland and Finland. Transp Res Part D Transp Environ 65:101–115 133. Leitinger C, Schuster A, Litzlbauer M (2011) Smart electric mobility—speichereinsatz für regenerative elektrische Mobilität und Netzstabilität. na 134. Kihm A, Trommer S (2014) The new car market for electric vehicles and the potential for fuel substitution. Energ Policy 73:147–157 135. Dong J, Lin Z (2012) Within-day recharge of plug-in hybrid electric vehicles: Energy impact of public charging infrastructure. Transp Res Part D Transp Environ 17:405–412

36

2 Passenger Land-Based Road Transport

136. Xi X, Sioshansi R, Marano V (2013) Simulation–optimization model for location of a public electric vehicle charging infrastructure. Transp Res Part D Transp Environ 22:60–69 137. Zhang Q, Li H, Zhu L, Campana PE, Lu H, Wallin F, Sun Q (2018) Factors influencing the economics of public charging infrastructures for EV—a review. Renew Sustain Energy Rev 94:500–509 138. Baresch M, Moser S (2019) Allocation of e-car charging: assessing the utilization of charging infrastructures by location. Transp Res Part A Policy Pract 124:388–395 139. Nie Y (Marco), Ghamami M (2013) A corridor-centric approach to planning electric vehicle charging infrastructure. Transp Res Part B Methodol 57:172–190 140. Skeete J-P, Wells P, Dong X, Heidrich O, Harper G (2020) Beyond the EVent horizon: Battery waste, recycling, and sustainability in the United Kingdom electric vehicle transition. Energy Res Soc Sci 69:101581 141. Grote M, Preston J, Cherrett T, Tuck N (2019) Locating residential on-street electric vehicle charging infrastructure: a practical methodology. Transp Res Part D Transp Environ 74:15–27 142. Hardman S, Chandan A, Tal G, Turrentine T (2017) The effectiveness of financial purchase incentives for battery electric vehicles—A review of the evidence. Renew Sustain Energ Rev 80:1100–1111 143. Skippon S, Chappell J (2019) Fleets’ motivations for plug-in vehicle adoption and usage: UK case studies. Transp Res Part D Transp Environ 71:67–84 144. Hao H, Ou X, Du J, Wang H, Ouyang M (2014) China’s electric vehicle subsidy scheme: rationale and impacts. Energ Policy 73:722–732 145. Japan Automobile Manufacturers Association (2009) Fact Sheet: Japanese government incentives for the purchase of environmentally friendly vehicles 146. Rietmann N, Lieven T (2019) A comparison of policy measures promoting electric vehicles in 20 countries. In: Finger M, Audouin M (eds) The Governance of Smart Transportation Systems. Towar. New Organ. Struct DevShared, Autom Electr Integr Mobil. Springer International Publishing, Cham, pp 125–145 147. Benvenutti LMM, Ribeiro AB, Forcellini FA, Maldonado MU (2016) The effectiveness of tax incentive policies in the diffusion of electric and hybrid cars in Brazil. 41o Congr Latinoam Din Sist 1–11 148. Mersky AC, Sprei F, Samaras C, Qian Z (Sean) (2016) Effectiveness of incentives on electric vehicle adoption in Norway. Transp Res Part D Transp Environ 46:56–68 149. Ingeborgrud L, Ryghaug M (2019) The role of practical, cognitive and symbolic factors in the successful implementation of battery electric vehicles in Norway. Transp Res Part A Policy Pract 130:507–516 150. Holman C, Harrison R, Querol X (2015) Review of the efficacy of low emission zones to improve urban air quality in European cities. Atmos Environ 111:161–169 151. Morton C, Lovelace R, Anable J (2017) Exploring the effect of local transport policies on the adoption of low emission vehicles: evidence from the London congestion charge and hybrid electric vehicles. Transp Policy 60:34–46

Chapter 3

Buses

Abstract If net zero targets are to be achieved, policymakers will need to encourage a modal shift from personal vehicles towards low carbon public transport, for example electric and hydrogen buses. This is because emissions per person per kilometre travelled remains significantly lower for bus users. Conventionally fuelled buses will not meet emission reduction targets, even with technological advancements. Therefore, integration of electric buses (EBs) and hydrogen buses (HBs) will need to be introduced, as both emit lower levels of emissions provided, they are fuelled from renewable resources. Although EBs produce the lowest level of emissions, it is likely that a combination of both EBs and HBs will be required as EBs are more suited to urban areas due to their current range and the frequency to be charged. HBs have a larger range and require additional infrastructure, making them more suited to more rural areas and long-distance routes. Policymakers need to focus on encouraging a modal shift from personal transport towards sustainable public transport as simple electrification of personal vehicles will not meet the required targets.

3.1 Introduction Across the world, Governments are increasingly challenging citizens to make a modal transition from their personal vehicles towards public transport [1, 2]. Public transport remains an important aspect in the development of socially, environmentally, economically and sustainable communities [3]. For decades buses have been the backbone of the public transport systems of cities around the world [4, 5]. For example, in the nineteenth century in London, horse-drawn public transport was a popular method of transport [6]. This was soon replaced with non-horse powered public transport in the early twentieth century, such as with the introduction of trolley buses and trams with and without rails [6]. More recently, petrol and diesel public transport has been used in cities and is now being replaced with hybrid electrical and hydrogen alternatives. The energy and environmental implications of buses cannot be ignored [7]. A modal shift from personal vehicles towards public transport and active travel has © The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_3

37

38

3 Buses

been deemed necessary for lowering greenhouse gas (GHG) emissions [8], improving public health [9] and the liveability of communities [10, 11]. This is because, unlike personal vehicles, high capacity public transport systems (such as BRT [12]) can be used for the mass transportation of people, therefore at full capacity, emissions per person per kilometre travelled, are lower for buses than personal vehicles [4, 13, 14]. For example, by assuming the maximum capacity of an average bus is 80 and a small personal vehicle is four, at least 20 personal vehicles could be taken off the road if buses were fully utilised, significantly reducing road transport emissions and congestion [14]. When examining the role of buses within an integrated low carbon transport system, fuel consumption and efficiency are highly influenced by several operational factors or conditions that need to be considered when making emission projections. For example, buses in local service operation behave considerably differently to coaches, trucks and light duty vehicles as they have to stop frequently to serve passengers [7]. Furthermore, the operating characteristics such as the bus design (i.e. type of bus and fuel type etc.), driver characteristics (i.e. behaviour when accelerating, decelerating, stopped etc.), operating conditions (i.e. passenger load etc.), traffic flow (i.e. congestion during rush hour or frequency of traffic interruptions etc.), road conditions (i.e. road grade and type etc.), environmental (i.e. weather etc.) and the number of bus stops etc., all need to be considered as this will influence emissions from buses regardless of their ‘fuel’ type [15]. Furthermore, urban buses are likely to travel at a lower average speed, exhibit high frequency of acceleration and deceleration, and long idle times, all of which can directly increase total GHG emissions, including a lower nitrous oxide (NOx) conversion efficiency in the after treatment [16, 17]. Thus the characteristics of urban bus operation can lead to a higher level of energy consumption than other types of road transport [7]. Previous research indicates that the passenger load, road grade, and traffic congestion significantly affect the performance of buses [15]. To achieve a reduction in emission levels will require the large-scale introduction of low emission buses. This chapter discusses bus-based mass transit in urban and rural areas (freight will be discussed in Chap. 7). Section 3.2 compares the emissions performance of conventionally fuelled buses compared to personal vehicles. Section 3.3 discusses EBs; Sect. 3.4 discusses hydrogen buses and Sect. 3.5 discusses other alterative low emission buses. Section 3.6 draws key conclusions regarding the integration of low emission bus use within an integrated transport system. This chapter highlights the emissions produced, the efficiency of these bus types and how buses should be integrated into the transport network to meet net zero targets. Chapter 5 will discuss the barriers and challenges of wide scale integration of low emission public transport.

3.2 Conventionally Fuelled Buses Depending on the operation, technology, age and fuel type, conventionally fuelled buses (CFBs) can produce significant levels of pollutants mainly including carbon

3.2 Conventionally Fuelled Buses

39

monoxide, nitrous oxide (NOx ) and carbon dioxide (CO2 ) [18–21]. Despite the continual developments of emission standards and technological developments of diesel buses they remain a serious air polluter in urban areas [22, 23]. Although buses are larger and therefore emit more emissions, the emissions are less per passenger per kilometre travelled for all pollutants. Recent studies have highlighted that CFBs emit higher levels of emissions per passenger per kilometre travelled compared to private vehicles [24], therefore a rapid move towards zero emission buses will be necessary to reduce emissions if net zero targets are to be met. Diesel fuelled buses have experienced a rapid and significant technological evolution due to demanding emission standards over the last 20 years, for example from EURO I to EURO VI in Europe (see Table 3.1) [25]. This has led to many advanced technologies being introduced to diesel engines such as variable valve timing, exhaust gas recirculation and high boost pressures [26] through the introduction of diesel hybrid buses. Diesel hybrid buses include an electrical energy storage system in conjunction with non-electrical motor generator to facilitate regenerative breaking and allow reducing idling time and have been introduced in various countries and cities [27]. This has led to a decrease in emissions, for example, in Europe, between EURO I and EURO VI regulations, NOx emissions limits for diesel bus engines have reduced by up to 95% (see Table 3.1) [28]. Since the EURO IV regulation, most diesel buses have been equipped with a selective catalytic reduction system to reduce NOx emissions [29]. Furthermore, the average age of a bus in the EU is ~11.7 years therefore not all retrofitted to fit the latest EURO standards (as described below), apart from in areas where legislation dictates that it is necessary. Today, 80% of new buses worldwide are sold in regions that enforce EURO V or earlier standards [30]. Furthermore, in the EU, if the engine emission standards were the same for buses and cars, then there would be less emissions. However, in the new proposed ultra-low emission zones, diesel cars are EURO 6, petrol cars are EURO 5 but buses remain at EURO VI standard [31]. As with the ‘dieselgate’ scandal, several studies have focused on the ‘real world’ and ‘official’ emissions of diesel buses as they have been shown to differ. Emission certification tests are a method to compare the emissions of vehicles and determine whether they remain within certain limits [32]. However, for heavy duty vehicles, decisions have previously been made without data from on-road operation or even framework studies and have considered standalone engine test bench measurements [33]. However, this presents significant disadvantages, for example when the fuel saving mode is in use, it often refers to the ‘off-cycle’ mode. This means the electronic control module allows emissions in excess of certification standard to be produced to obtain better fuel economy, but does not turn on during engine certification standard test [32]. This therefore can result in emission levels produced in road operation being higher than estimated from the standard test. Considering these emission projections, Governments and policymakers have begun introducing measures to limit and regulate emissions. For example, in Madrid, Spain, one study focused on the diesel EURO V buses under urban off-cycle conditions [7]. Their results indicated that the buses operated with an air brake thermal efficiency of 41%, break specific fuel consumption of 205 g/kWh and a CO2 energy

October 2008

December 2012

EURO V

EURO VI

1.5 4.0

WHSC WHTC

1.5

1.5

2.1

0.16

0.13

0.46

0.46

0.66

0.25

1.1

1.1

1.1

1.1

HC

0.46

0.4

2.0

3.5

5.0

2.0

7.0

7.0

8.0

8.0

NOx

10

10

NH3 (ppm)

0.01

0.01

0.02

0.02

0.10 0.13**

0.02

0.15

0.25

0.36

0.612

PM

Source Nylund et al. [28] * EEVs are enhanced environmentally friendly vehicles ** Engines less than 0.75 dm3 swept volume per cylinder and a rated power speed of more than 3000 per minute

October 2005

October 2000

1.5

4.0

October 1999 (EEVs* only)

October 1997

4.5

4.5

CO

4.0 ESC and ELR

ECE R49

Test cycle

October 1995

1992, > 85 kW

1992, < 85 kW

EURO IV

EURO III

EURO II

EURO I

Date

Table 3.1 Overview of the European Emission Standards for heavy duty diesel engines (g/kWh)

8 * 1011

8 * 1011

PN (#/kWh)

0.5

0.5

0.8

0.15

Smoke (m−1 )

40 3 Buses

3.2 Conventionally Fuelled Buses

41

emission factor of 637 g/kWh [7]. In addition, the NOx energy emission factor was 80% higher than the levels in the EURO V standard. Although this could be partially explained by the high frequency of fuel stops and other characteristics of urban buses, results highlight that there remains a significant difference between ‘real world’ and ‘official emissions’ that needs to be addressed if net zero targets are to be met [7]. Similarly, in Brazil, their current emission standards are equivalent to EURO V. The national government in São Paulo’s environmental agency (CETESB) are working towards new emission standards equivalent to EURO VI with São Paulo independently working to set legally binding targets for a fossil free bus fleet [30]. The barriers including cost and infrastructure for implementation are further discussed in Chap. 5. One method to reduce the emission levels from buses is to replace CFBs with electric or hydrogen alternative buses.

3.3 Electric Buses The first electric style buses were trolleybuses operating in a battery mode using overhead contact wires [34]. Alternately, trolleybuses can pass through sections of their route without overhead contact wires using two different methods. Firstly, trolleybuses can have additional energy sources installed either through a diesel engine that serves as an electric generator for the main electric traction motor which is switched-on during parts of the route with no contact wires; or secondly, through the addition of a battery that can be recharged on the electrified sections of the route [34]. The second method is considered the more environmentally friendly option; however, it has a significantly lower capacity. This is because in some countries, such as Poland, there remains an unfavourable energy balance where it remains less harmful to use diesel buses with a high exhaust emission standard as electricity is primarily generated using coal [35, 36]. Therefore, ensuring that EBs are fuelled using renewables is a key concern. In addition, the cost of the trolleybus infrastructure, including the overhead contact wires and their maintenance, remains significant therefore alternatives using a fast-charging battery concept that does not require overhead contact wires are cheaper. Although trolley buses fit a purpose in the presence of cheap coal fired electric fuelled transport, to meet net zero emission reductions battery technology EBs will need to be integrated into the transport network. EBs, which are referred to here as buses fuelled using electricity and stored using a battery, have recently become commercially available as battery technology has developed to allow EBs to become a viable solution for mass public transport [37– 40] due to their significant environmental, operational and energy related advantages [41–47]. There has been increased support for EBs from both Governments and policymakers at a country level through the introduction of various grants and subsidies to support EB uptake [48]. For example, through the Transportation Investment Generating Economic Recovery programme in the USA, the Green Bus Fund programme in the UK and the Ten Cities and Thousand Vehicle Programme in China [49]. A recent study predicted that over 47% of the world’s total city bus fleet will be replaced

42

3 Buses

by EBs by 2025 [42], with the majority of this transition being in China where over 1,000 EBs are in service in Tianjin, whilst Shenzhen has replaced all diesel city buses with EBs [42, 48]. EBs have also been introduced elsewhere in Asia (i.e. Korea [38], Sri Lanka [50] etc.), Europe (i.e. Germany [51], Finland [52], France [50], Sweden [53], the UK [14], etc.), the Middle East (i.e. Israel [41]), North America (i.e. Canada [54], the USA [50, 55] etc.), South America (i.e. Brazil [56], Colombia [50], Uruguay [50] etc.) and Oceania (i.e. New Zealand [50] etc.). EBs have many advantages, including low environmental emissions, energy conservation and minimal noise, making them an ideal technology for urban areas [43, 55]. Furthermore, EBs are particularly well suited to local bus services which operate at low speeds and have frequent stops [55]. However, the true environmental impact is dependent on the electricity generation mix and type of EB [57]. For example, in California, assuming a 12 year lifetime, a well-to-wheel emission analysis on EBs was completed which demonstrated that EBs emit approximately 0.35– 0.40 kg CO2 /km, compared to a diesel CFBs travelling on the same routes which emitted 1.65–2.00 kg CO2 /km [43]. Furthermore, another study using the same 12year lifetime, estimated that EBs release on average 543–1004 tonnes CO2 -eq over the course of the buses lifetime, compared to 1,446–2,284 tonnes CO2 -eq for a diesel CFB [58]. However, the actual values that EBs operate at are dependent upon the different degrees of electrification that depend on the propulsion system, including hybrid electric, fuel cell electric and battery electric vehicle types [22, 59–62]. EBs have been found to have a fuel efficiency ranging between 0.76 and 2.79 kWh km−1 with an average of 1.65 kWh km−1 [57]. This is approximately three times more fuel efficient than CFBs (although this does not consider the thermal efficiency of electricity generation), however this range can fluctuate depending on weather conditions, e.g., extreme low conditions [57]. This is because cold temperatures were found to reduce bus range by approximately 40% and increase the start-up times due to battery efficiency [63, 64]. Therefore, EBs are not suited in particularly cold countries as this significantly impacts the bus range, causing additional charging and increasing the GHG emissions associated. Progress is currently being made in Finland and Sweden regarding their use as some of the first cold weather trials were completed. This is through water-based heating systems within bus terminals with buses connected to a warm water exchanger which heats the fluids of the engine/battery cooling system and interior heating system. This also allows the driver and passengers to be warm when travelling, ensuring that the temperatures do not act as a deterrent to travel [65]. Alternatively, trolley buses work under these cooler conditions due to their battery management and external power source, such as the Solaris trolley bus which are prepared for difficult winter conditions.

3.4 Hydrogen Buses Hydrogen buses (HBs) offer similar benefits to EBs in terms of achieving low environmental emissions, energy conservation and generating minimal noise [43, 55],

3.4 Hydrogen Buses

43

however an increased advantage is providing further route flexibility as they do not require en-route recharging. This is because they have a significantly larger range of ~ 450 km range, typically taking 7–10 min for a full charge [66]. Hydrogen is considered a secondary energy resource with a higher energy density per unit mass than many fuels [67–69]. Depending on how the hydrogen is generated, HBs are often considered more fuel-efficient ranging from 2.40 to 4.22 kWh km−1 compared to 3.83–6.03 kWh km−1 for a diesel CFB when hydrogen is generated through electrolysis. Although HBs are still a relatively new technology, they have been introduced in Asia (i.e. China [70], Korea [71], Taiwan [72]), Europe (i.e. Czech Republic [73], Denmark [68], Netherlands [74], Romania [75], the UK [14] etc.), North America (i.e. the USA [76] etc.) and South America (i.e. Argentina [77], Brazil [78] etc.).

3.4.1 Hydrogen Generation As discussed in Chap. 2, there are several methods of hydrogen generation: through electrolytic processes, thermochemical processes (i.e., steam methane reform (SMR) either with natural gas reforming, coal gasification, biomass gasification, biomassderived liquid reforming, solar thermochemical hydrogen (STCH)), direct solar water splitting process (i.e., photoelectrochemical (PEC) or photobiological) and biological (i.e., microbial; biomass conversion or photobiological) processes. For hydrogen transport, thermochemical processes, especially SMR with natural gas, goal gasification or electrolysis are currently the most popular methods of hydrogen generation. Although electrolysis, also known as ‘green hydrogen’, is the most ‘sustainable’ method of hydrogen generation if the electricity is generated from renewable or nuclear generators, this process is still on a relatively small scale and will need to be scaled up for more environmentally hydrogen generation [79]. Hydrogen generation through electrolysis has been growing and is expected to be price competitive with hydrogen produced from fossil fuels within a decade [79]. Hydrogen produced via alkaline water electrolysis is now considered a mature technology with megawatt scale installations commercially available [79].

3.5 Other Low Emission Alternatives Further research has been completed to better understand other local emission alternatives to diesel and petrol buses, including the use of compressed natural gas (CNG) or liquified natural gas, each of which has its own advantages and disadvantages [80]. This is because natural gas emits fewer air pollutants and produces ~ 25% less CO2 per unit of energy than diesel fuel [81].

44

3 Buses

A recent study in Poland evaluated the environmental impacts of replacing diesel buses with CNG buses in urban environments [82]. The results indicated that replacing diesel buses in Radom city with a CNG bus would result in the reduction of GHG emissions and particulate emissions by 400 and 52.5 kg respectively throughout the buses life cycle. Furthermore, in 2002, Delhi in India, transformed all diesel buses and chuc-chuc taxis to CNG reducing their CO2 , particulate matter and sulphur dioxide emissions [83]. Removing chuc-chuc taxis in Delhi played a significant role as this removed the two-stroke petrol oil engines from the bus-based public transport system which is now transitioning to battery electric with a battery exchange system being implemented. Although CNG has a significant price advantage over gas and petrol and diesel fuels [80], CNG powered vehicles require significantly larger storage capacity as the gas needs to be compressed at high pressure at a ambient temperature and stored in a tank [84]. This results in a considerable barrier to widescale integration [85]. Alternatively, LNG has to be cooled around −162 °C [86] which is ~ 600 times denser [87] than natural gas and is therefore considered more convenient to store and transport within large quantities. Methane used for transportation can also be produced from Biogenic sources using municipal waste sewage, animal manure, bioenergy crops or other organic material. This is compressed and used in thermal engines or fuel cells/electric motors as a gaseous biofuel and is low carbon.

3.6 Key Findings Encouraging a modal shift from personal vehicles towards public transport will be necessary to reduce transport emissions as emissions per person per kilometre travelled is significantly lower for bus users. However, even with technological improvements, CFBs will not meet emission reduction targets, therefore alternative low emission technologies need to be integrated. Transitioning a fleet to EBs can reduce the GHG emissions per person per kilometre travelled in comparison to hydrogen and CFBs, provided the electricity is generated from renewables (as discussed in Chap. 1 with the barriers and challenges of alternative bus implementation discussed in Chap. 5). As shown in this chapter, HBs produce the second lowest emission levels, however this is dependent on how the hydrogen is generated and has been further highlighted in other studies [14]. For example, hydrogen from electrolysis using electricity generated by renewable or nuclear generators produces the lowest levels of emissions, whereas hydrogen generated from SMR using natural gas or coal can negatively impact the environment. This will result in the need for alternative technologies, such as carbon capture and storage and direct air capture (as mentioned in Chap. 2) to capture SMR process emissions. Even though HBs produce higher levels of emissions than EBs, their emission levels remain significantly lower than CFBs assuming low carbon electricity. Due to the current operating range of EBs it is likely that a combination of both EBs and HBs will need to be integrated in the bus system. This is because EBs

3.6 Key Findings

45

are more suited to urban areas due to their current range and the frequency to be charged. Alternatively, HBs have a much larger range and require additional infrastructure, making them more suited to more rural areas and long-distance routes. It is also recognised that the various costs and indeed emissions levels may change with changing technology and scalability of sold volumes of buses [88]. To meet future travel demands within a low carbon environment, it is therefore likely that a combination of both EBs and HBs will be required. Chapter 4 will consider the implementation of conventionally fuelled electric and hydrogen trains as alternatives to personal vehicles for longer journeys.

References 1. Chowdhury S, Hadas Y, Gonzalez VA, Schot B (2018) Public transport users’ and policy makers’ perceptions of integrated public transport systems. Transp Policy 61:75–83 2. Rabl A, de Nazelle A (2012) Benefits of shift from car to active transport. Transp Policy 19:121–131 3. van Lierop D, Badami MG, El-Geneidy AM (2018) What influences satisfaction and loyalty in public transport? A review of the literature. Transp Rev 38:52–72 4. Holmberg K, Andersson P, Nylund N-O, Mäkelä K, Erdemir A (2014) Global energy consumption due to friction in trucks and buses. Tribol Int 78:94–114 5. Olyslagers F, Mulley C, Nelson JD (2021) Bus—from workhorse to thoroughbred. In: Mulley C, Nelson JD, Ison S (eds) Handbook of public transport. Routledge, Abingdon, pp 171–185 6. Giles-Corti B, Zapata-Diomedi B, Jafari A, Both A, Gunn L (2020) Could smart research ensure healthy people in disrupted cities? J Transp Heal 19:100931 7. Rosero F, Fonseca N, López J-M, Casanova J (2020) Real-world fuel efficiency and emissions from an urban diesel bus engine under transient operating conditions. Appl Energy 261:114442 8. Paulsson A, Isaksson K, Sørensen CH, Hrelja R, Rye T, Scholten C (2018) Collaboration in public transport planning—why, how and what? Res Transp Econ 69:377–385 9. Ming Wen L, Rissel C (2008) Inverse associations between cycling to work, public transport, and overweight and obesity: findings from a population based study in Australia. Prev Med (Baltim) 46:29–32 10. Kenworthy JR (2006) The eco-city: ten key transport and planning dimensions for sustainable city development. Environ Urban 18:67–85 11. Kenworthy J (2007) Urban planning and transport paradigm shifts for cities of the postpetroleum age. J Urban Technol 14:47–70 12. Ferbrache F (2019) Developing bus rapid transit: the value of BRT in urban spaces. Edward Elgar Publishing 13. Ruano-Daza E, Cobos C, Torres-Jimenez J, Mendoza M, Paz A (2018) A multiobjective bilevel approach based on global-best harmony search for defining optimal routes and frequencies for bus rapid transit systems. Appl Soft Comput 67:567–583 14. Logan KG, Nelson JD, Hastings A (2020) Electric and hydrogen buses: shifting from conventionally fuelled cars in the UK. Transp Res Part D Transp Environ. https://doi.org/10.1016/j. trd.2020.102350 15. Frey HC, Rouphail NM, Zhai H, Farias TL, Gonçalves GA (2007) Comparing real-world fuel consumption for diesel- and hydrogen-fueled transit buses and implication for emissions. Transp Res Part D Transp Environ 12:281–291 16. Rodman Oprešnik S, Seljak T, Vihar R, Gerbec M, Katrašnik T (2018) Real-world fuel consumption, fuel cost and exhaust emissions of different bus powertrain technologies. Energies. https://doi.org/10.3390/en11082160

46

3 Buses

17. Sowman J, Box S, Wong A, Grote M, Laila DS, Gillam G, Cruden AJ, Preston JM, Fussey P (2018) In-use emissions testing of diesel-driven buses in Southampton: is selective catalytic reduction as effective as fleet operators think? IET Intell Transp Syst 12:521–526 18. Pan Y, Chen S, Qiao F, Ukkusuri SV, Tang K (2019) Estimation of real-driving emissions for buses fueled with liquefied natural gas based on gradient boosted regression trees. Sci Total Environ 660:741–750 19. Wang A, Ge Y, Tan J, Fu M, Shah AN, Ding Y, Zhao H, Liang B (2011) On-road pollutant emission and fuel consumption characteristics of buses in Beijing. J Environ Sci 23:419–426 20. Yu Q, Li T, Li H (2016) Improving urban bus emission and fuel consumption modeling by incorporating passenger load factor for real world driving. Appl Energy 161:101–111 21. Zhang S, Yu L, Song G (2017) Emissions characteristics for heavy duty diesel trucks under different loads based on vehicle-specific power. Transp Res Rec 2627:77–85 22. Miles J, Potter S (2014) Developing a viable electric bus service: the Milton Keynes demonstration project. Res Transp Econ 48:357–363 23. Wang J, Rakha HA (2016) Fuel consumption model for conventional diesel buses. Appl Energy 170:394–402 24. Alam A, Hatzopoulou M (2014) Reducing transit bus emissions: alternative fuels or traffic operations? Atmos Environ 89:129–139 25. Guo J, Ge Y, Hao L, Tan J, Li J, Feng X (2014) On-road measurement of regulated pollutants from diesel and CNG buses with urea selective catalytic reduction systems. Atmos Environ 99:1–9 26. Knecht W (2008) Diesel engine development in view of reduced emission standards. Energy 33:264–271 27. Partridge J, Wu W, Bucknall R (2017) Development of bus drive technology towards zero emissions: a review. Intech 13 28. Nylund NO, Karvonen V, Kuutti H, Laurikko J (2014) Comparison of diesel and natural gas bus performance 29. Lowell D, Kamakaté F (2012) Urban off-cycle NOX emissions from EURO IV/V trucks and buses. Int Counc Clean Transp 30. ICCT (2017) Zero emission buses are worth reaching for, but emission performance standards are the low-hanging fruit. https://theicct.org/blog/staff/zeroemission-buses-are-worthreaching-for. Accessed 8 Oct 2021 31. ACEA (2021) Average age of the EU vehicle fleet, by country. Driv Mobil Eur. https://www. acea.auto/figure/average-age-of-eu-vehicle-fleet-by-country/. Accessed 6 Oct 2021 32. Liu Z, Ge Y, Johnson KC, Shah AN, Tan J, Wang C, Yu L (2011) Real-world operation conditions and on-road emissions of Beijing diesel buses measured by using portable emission measurement system and electric low-pressure impactor. Sci Total Environ 409:1476–1480 33. Pelkmans L, Debal P (2006) Comparison of on-road emissions with emissions measured on chassis dynamometer test cycles. Transp Res Part D Transp Environ 11:233–241 34. Grygar D, Koháni M, Štefún R, Drgoˇna P (2019) Analysis of limiting factors of battery assisted trolleybuses. Transp Res Procedia 40:229–235 35. Połom M, Wi´sniewski P (2021) Assessment of the emission of pollutants from public transport based on the example of diesel buses and trolleybuses in Gdynia and Sopot. Int J Environ Res Public Health. https://doi.org/10.3390/ijerph18168379 36. Połom M, Wi´sniewski P (2021) Implementing electromobility in public transport in Poland in 1990–2020. A review of experiences and evaluation of the current development directions. Sustainability. https://doi.org/10.3390/su13074009 37. Lajunen A (2014) Energy consumption and cost-benefit analysis of hybrid and electric city buses. Transp Res Part C Emerg Technol 38:1–15 38. Choi U-D, Jeong H-K, Jeong S-K (2012) Commercial operation of ultra-low floor electric bus for Seoul city route. In: 2012 IEEE vehicle power and propulsion conference, pp 1128–1133 39. Lajunen A (2018) Lifecycle costs and charging requirements of electric buses with different charging methods. J Clean Prod 172:56–67

References

47

40. Corazza MV, Guida U, Musso A, Tozzi M (2016) A European vision for more environmentally friendly buses. Transp Res Part D Transp Environ 45:48–63 41. Alwesabi Y, Liu Z, Kwon S, Wang Y (2021) A novel integration of scheduling and dynamic wireless charging planning models of battery electric buses. Energy 230:120806 42. An K (2020) Battery electric bus infrastructure planning under demand uncertainty. Transp Res Part C Emerg Technol 111:572–587 43. Lajunen A, Lipman T (2016) Lifecycle cost assessment and carbon dioxide emissions of diesel, natural gas, hybrid electric, fuel cell hybrid and electric transit buses. Energy 106:329–342 44. Ko YD, Jang YJ (2013) The optimal system design of the online electric vehicle utilizing wireless power transmission technology. IEEE Trans Intell Transp Syst 14:1255–1265 45. Iliopoulou C, Kepaptsoglou K (2019) Integrated transit route network design and infrastructure planning for online electric vehicles. Transp Res Part D Transp Environ 77:178–197 46. Pternea M, Kepaptsoglou K, Karlaftis MG (2015) Sustainable urban transit network design. Transp Res Part A Policy Pract 77:276–291 47. Helber S, Broihan J, Jang YJ, Hecker P, Feuerle T (2018) Location planning for dynamic wireless charging systems for electric airport passenger buses. Energies. https://doi.org/10. 3390/en11020258 48. Tang X, Lin X, He F (2019) Robust scheduling strategies of electric buses under stochastic traffic conditions. Transp Res Part C Emerg Technol 105:163–182 49. Li J-Q (2016) Battery-electric transit bus developments and operations: a review. Int J Sustain Transp 10:157–169 50. An K, Jing W, Kim I (2020) Battery-swapping facility planning for electric buses with local charging systems. Int J Sustain Transp 14:489–502 51. Rupp M, Handschuh N, Rieke C, Kuperjans I (2019) Contribution of country-specific electricity mix and charging time to environmental impact of battery electric vehicles: a case study of electric buses in Germany. Appl Energy 237:618–634 52. Laurikko J, Pihlatie M, Nylund NO, Halmeaho T, Kukkonen S, Lehtinen A, Karvonen V, Mäkinen R, Ahtiainen S (2015) Electric city bus and infrastructure demonstration environment in Espoo, Finland. In: 28th International Electric Vehicle Symposium and Exhibition 2015, EVS 2015 1–11 53. Xylia M, Leduc S, Patrizio P, Kraxner F, Silveira S (2017) Locating charging infrastructure for electric buses in Stockholm. Transp Res Part C Emerg Technol 78:183–200 54. Mohamed M, Ferguson M, Kanaroglou P (2018) What hinders adoption of the electric bus in Canadian transit? Perspectives of transit providers. Transp Res Part D Transp Environ 64:134– 149 55. Wang Y, Huang Y, Xu J, Barclay N (2017) Optimal recharging scheduling for urban electric buses: a case study in Davis. Transp Res Part E Logist Transp Rev 100:115–132 56. Velandia Vargas JE, Falco DG, da Silva Walter AC, Cavaliero CKN, Seabra JEA (2019) Life cycle assessment of electric vehicles and buses in Brazil: effects of local manufacturing, mass reduction, and energy consumption evolution. Int J Life Cycle Assess 24:1878–1897 57. Deliali A, Chhan D, Oliver J, Sayess R, Pollitt KJG, Christofa E (2021) Transitioning to zero emission bus fleets: state of practice of implementations in the United States. Transp Rev 41:164–191 58. Ercan T, Tatari O (2015) A hybrid life cycle assessment of public transportation buses with alternative fuel options. Int J Life Cycle Assess 20:1213–1231 59. Ça˘gatay Bayindir K, Gözüküçük MA, Teke A (2011) A comprehensive overview of hybrid electric vehicle: powertrain configurations, powertrain control techniques and electronic control units. Energy Convers Manag 52:1305–1313 60. Mahmoud M, Garnett R, Ferguson M, Kanaroglou P (2016) Electric buses: a review of alternative powertrains. Renew Sustain Energy Rev 62:673–684 61. Ribau JP, Silva CM, Sousa JMC (2014) Efficiency, cost and life cycle CO2 optimization of fuel cell hybrid and plug-in hybrid urban buses. Appl Energy 129:320–335 62. Yong JY, Ramachandaramurthy VK, Tan KM, Mithulananthan N (2015) A review on the stateof-the-art technologies of electric vehicle, its impacts and prospects. Renew Sustain Energy Rev 49:365–385

48

3 Buses

63. Prohaska R, Kelly K, Eudy L (2016) Fast charge battery electric transit bus in-use fleet evaluation. In: The 2016 IEEE transportation electrification conference and Expo, pp 1–6 64. Shirazi Y, Carr E, Knapp L (2015) A cost-benefit analysis of alternatively fueled buses with special considerations for V2G technology. Energy Policy 87:591–603 65. SOLARIS (2019) Solaris buses perfectly prepared for difficult winter conditions. https://www. solarisbus.com/en/busmania/solaris-buses-perfectly-prepared-for-difficult-winter-condition s-854. Accessed 6 Oct 2021 66. Pagliaro M, Meneguzzo F (2019) Electric bus: a critical overview on the dawn of its widespread uptake. Adv Sustain Syst 3:1800151 67. Liu J, Zheng S, Zhang Z, Zheng J, Zhao Y (2020) Numerical study on the fast filling of on bus gaseous hydrogen storage cylinder. Int J Hydrogen Energy 45:9241–9251 68. Sørensen B, Hauge Petersen A, Juhl C et al (2004) Hydrogen as an energy carrier: scenarios for future use of hydrogen in the Danish energy system. Int J Hydrogen Energy 29:23–32 69. Mori D, Hirose K (2009) Recent challenges of hydrogen storage technologies for fuel cell vehicles. Int J Hydrogen Energy 34:4569–4574 70. Zhang G, Zhang J, Xie T (2020) A solution to renewable hydrogen economy for fuel cell buses—a case study for Zhangjiakou in North China. Int J Hydrogen Energy 45:14603–14613 71. Kim H, Eom M, Kim B-I (2020) Development of strategic hydrogen refueling station deployment plan for Korea. Int J Hydrogen Energy 45:19900–19911 72. Chang C-C, Liao Y-T, Chang Y-W (2019) Life cycle assessment of alternative energy types— including hydrogen—for public city buses in Taiwan. Int J Hydrogen Energy 44:18472–18482 73. Stehlík K, Tkáˇc M, Bouzek K (2019) Recent advances in hydrogen technologies in the Czech Republic. Int J Hydrogen Energy 44:19055–19060 74. Bakker S, Konings R (2018) The transition to zero emission buses in public transport—the need for institutional innovation. Transp Res Part D Transp Environ 64:204–215 75. Pamucar D, Iordache M, Deveci M, Schitea D, Iordache I (2020) A new hybrid fuzzy multicriteria decision methodology model for prioritizing the alternatives of the hydrogen bus development: a case study from Romania. Int J Hydrogen Energy. https://doi.org/10.1016/j.ijhydene. 2020.10.172 76. Lee D-Y, Elgowainy A, Vijayagopal R (2019) Well-to-wheel environmental implications of fuel economy targets for hydrogen fuel cell electric buses in the United States. Energy Policy 128:565–583 77. Iannuzzi L, Hilbert JA, Silva Lora EE (2021) Life cycle assessment (LCA) for use on renewable sourced hydrogen fuel cell buses vs diesel engines buses in the city of Rosario, Argentina. Int J Hydrogen Energy. https://doi.org/10.1016/j.ijhydene.2021.01.065 78. de Miranda PEV, Carreira ES, Icardi UA, Nunes GS (2017) Brazilian hybrid electric-hydrogen fuel cell bus: improved on-board energy management system. Int J Hydrogen Energy 42:13949– 13959 79. Kovaˇc A, Paranos M, Marciuš D (2021) Hydrogen in energy transition: a review. Int J Hydrogen Energy 46:10016–10035 80. Pourahmadiyan A, Ahmadi P, Kjeang E (2021) Dynamic simulation and life cycle greenhouse gas impact assessment of CNG, LNG, and diesel-powered transit buses in British Columbia, Canada. Transp Res Part D Transp Environ 92:102724 81. Zhang S, Wu Y, Liu H, Huang R, Yang L, Li Z, Fu L, Hao J (2014) Real-world fuel consumption and CO2 emissions of urban public buses in Beijing. Appl Energy 113:1645–1655 82. Dyr T, Misiurski P, Ziółkowska K (2019) Costs and benefits of using buses fuelled by natural gas in public transport. J Clean Prod 225:1134–1146 83. Goyal P, Sidhartha, (2003) Present scenario of air quality in Delhi: a case study of CNG implementation. Atmos Environ 37:5423–5431 84. Van Basshuysen R (2015) Natural gas and renewable methane for powertrains. In: Van Basshuysen R (ed) Futur. Strateg. Clim. Mobil., 1st edn. Springer, Berlin, pp 973–978 85. Browne D, O’Mahony M, Caulfield B (2012) How should barriers to alternative fuels and vehicles be classified and potential policies to promote innovative technologies be evaluated? J Clean Prod 35:140–151

References

49

86. Mokhatab S, Mak JY, Valappil J V, Wood DA (2013) Handbook of liquefied natural gas. Gulf Professional Publishing 87. Constable GA, Gibson CJ, Gram A (1989) Use of LNG in heavy duty vehicles 88. Hensher DA, Wei E, Balbontin C (2022) Comparative assessment of zero emission electric and hydrogen buses in Australia. Transp Res Part D Transp Environ 102:103130

Chapter 4

Trains

Abstract For travel within urban and rural areas, there is a compelling case for policy makers to encourage a transition from using personal vehicles towards low emission rail. Although trains are often seen as a ‘green’ mode of transport, they represent ~ 9% of global passenger movement and only consume ~ 0.6% of global oil, therefore it remains important to consider methods for emission reduction. Without improvements to the current rolling stock, as well as the tracks and platforms, emissions could remain static at best. Electric trains (ETs) produce the lowest levels of emissions when compared to conventionally fuelled trains (CFTs) and hydrogen trains (HTs). However, emissions remain dependent on the electricity generation mix, and it is likely that in the future a combination of both HTs and ETs will be required to travel within urban and rural areas. For trains to be considered a real alternative to personal vehicles and support the broader objectives of integrated transport, early investment into new rolling stock will be required. This is particularly important as the average life of a train is ~ 20 years which means that early implementation will be necessary.

4.1 Introduction Rail is often perceived as a ‘green’ mode of transport, with several studies focusing on the greenhouse gas (GHG) emissions associated with the railway network [1–5]. Rail is responsible for ~ 9% of global motorised passenger movement and for ~ 7% of freight movement whilst only accounting for ~ 3% of transport sector energy use [6]. Urban and high-speed rail infrastructure has been expanding rapidly over the past decade laying the foundations for convenient low emission transport within and between cities [6]. Notwithstanding competition from other transport modes, rail travel share has remained steady. According to the International Energy Agency (IEA), the share of rail passenger kilometres travelled worldwide has remained at slightly less than 10% since 2000 despite rapid rail infrastructure expansion [6]. Furthermore, in 2019 rail services consumed ~ 0.6% of global oil and ~ 1.2% of global electricity use but © The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_4

51

52

4 Trains

were only responsible for ~ 0.3% of direct carbon dioxide (CO2 ) emissions [6]. This highlights that the modest energy use and emission outputs of rail travel remains a promising strategy to enhance energy security and reduce emissions. Researchers have claimed that rail travel can fulfil an individual’s desire for high mobility with low negative environmental, social, economic, and financial impacts [1]. Rail travel is considered a (mostly) inexpensive, safe and (in some contexts) profitable transport method [1], as well as being free from traffic congestion whilst offering a large carrying capacity [7]. Furthermore, long distance intercity, regional, or even international rail travel provides vital connections between populations and economies [8]. Urban rail is regarded as an ideal solution to reduce the restrictions of urban mobility due to the superior capacity, safety, reliability, and environmental performance of the service when compared to the private car. However, even though rail has become an alternative in many cities, its popularity has been limited due to the accessibility advantages of personal vehicles, despite growing levels of congestion and reduced parking availability acting as deterrents [9]. Although there can be capacity constraints that leave little scope for increasing future urban rail use, “push” and “pull” travel demand management (TDM) initiatives can be implemented by transport authorities to actively encourage rail use. However, in terms of environmental emissions, there is a concern that urban rail may lose its competitive edge if it does not reduce its energy usage whilst maintaining or enhancing its service quality and capacity [10]. To enhance mobility in urban areas ~ 200 cities worldwide have introduced metro systems, with their combined length exceeding ~ 32,000 km [6]. Since 2010 new metro systems have been introduced in 46 countries, 34 of which are in Asia (by the end of 2019, 37 cities in mainland China had opened metro systems with a total length of 5,180.6 km) [11]. Simultaneously, new light rail transit systems have been launched in 65 cities with 20 in Europe and the rest dispersed between North America, Asia and the Middle East and North Africa. The first light rail systems are also entering into operation in Saharan Africa in Ethiopia and Nigeria. The length of urban rail lines which comprises both metro and light rail has expanded by ~ 3.5% per year in the past decade and even faster growth would help mitigate CO2 emissions. This study however did not consider suburban heavy rail. Encouraging rail travel in modern society is challenging. For example, a railway journey is almost always in several parts i.e., with an individual leaving their home, the rail journey and onward travel from the railway station to the final destination [12]. However, most important railway stations are usually located centrally, therefore the requirement to interchange remains an important part of the railway journey [12]. Ensuring that the critical stages of travel run smoothly is essential to achieve a seamless travel experience and to ensure that rail remains a viable and attractive alternative option to personal vehicles [12]. Previous studies have highlighted that a major barrier to real uptake is the cost of rail [13]. For example, when price changes significantly, this can directly influence demand, causing passengers to transition to other modes of transport, often personal vehicles [13]. It is well established that if rail fares increase, passenger demand tends to decrease. For example, in the City of Zagreb, Croatia, free tickets introduced

4.1 Introduction

53

for pensioners, students and individuals holding social cards, led to an increase in rail transport demand [14]. Other factors including income or employment are also possible determinants of real travel demand [7]. The barriers to widespread rail uptake are further discussed in Chap. 5.

4.2 Conventionally Fuelled Trains Passenger rail is one of the least energy and CO2 intensive of all motorised transport modes, with ~ 25% of conventional passenger rail activity using conventionally fuelled, i.e., diesel, trains [6]. There is currently considerable research being done to aid the decarbonisation of rail; for example, most modern ETs use regenerative breaking to recover some of the kinetic energy lost during their frequent stops, including most underground metro networks and London’s Dockland Light Railway. Broader utilisation of this technology will help reduce overall energy usage by trains and it remains clear that many countries should phase out diesel use in favour of low emission alternatives. Furthermore, even though the GHG emissions from conventionally fuelled trains (CFTs) remains low, improvement to the current rail rolling stock will be required as a contribution towards meeting national level emission reduction targets. This will also contribute to the attractiveness of rail services, thus providing an alternative to personal vehicles and achieve a reduction in emissions. This phasing out process of stock replenishment should begin as early as possible because a train has a life expectancy of between 20 and 40 years. For example, some models such as the Train à Grande Vitesse (TGVs) in France and the UK 125 s have a retiring age of > 40. Therefore, new trains entering service in 2020 will still be in service by 2040 or 2060, long after France and the UK’s emission reduction targets.

4.3 Electric Trains Electric locomotives were originally demonstrated by Siemens in 1879, where power was supplied through a third rail between the two running rails [15]. Larger scale development of the electric rail system began parallel to the progression within electrical energy distribution systems and the development of electric machines towards the end of the nineteenth century, with the industrial production of electric locomotives beginning in the 1930s [16, 17]. Thus, electric rail should not be considered a new technology, but it has evolved over the years from DC third rail to high voltage overhead supply. Traction motors, signals and control equipment have also evolved from analogue to digital technology vastly improving efficiency. The emission levels associated with rail are dependent on how the electricity is generated. An electrified rail system works by distributing the electrical energy through a dedicated low or medium-voltage system (via an overhead conductor or

54

4 Trains

a third rail) to the train, which can operate without having a primary energy source on-board [17]. Electric trains are characterised by higher power-to-weight ratios than diesel trains and are also able to recover some kinetic energy through regenerative braking [18]. Currently ~ 75% of conventional passenger rail activity uses electric rail with virtually all urban (including metro and light rail) and high-speed rail networks using electric trains with electrification of conventional rail expected to continue at a rapid pace in future [6]. Many countries have already begun transitioning towards electric passenger rail including China, the EU, India, Japan, and Russia who together account for around 90% of the global passenger rail activity [6]. Although in some countries progress remains slow, new policies have been introduced to encourage the transition towards ETs. For example, in the UK, current policy states that diesel-only trains, which currently account for almost a third of passenger rail travel, will be phased out by 2040. However, only 40% of the rail network is electrified, therefore significant investment and planning is needed to ensure that there is a rapid and successful transition. This will require either overhead power lines for electrification or onboard energy storage from batteries or hydrogen fuel cells, however it is likely that the solution will be a combination of both. Current policy targets diesel trains alone, though a significant part of the train network is currently under hybrid power. To ensure de-carbonisation is taken as far as achievable, policy makers should also consider phasing out this power mode. Increasing the overall efficiency of electric rail will be critical to achieve energy savings and GHG emission reduction. Electric rail consists of a network of rails supplied by geographically distributed power supply substations. However, in some rural areas electrifying rail tracks may be too difficult; in this case using hydrogen (H2 ) generated by decarbonised electricity would allow a low emission alternative to CFTs [2]. Although they are currently not in use in the UK, there are plans for HTs technology to be used from as early as 2022 and it is relevant to note that hydrogen trams are already deployed in some Chinese cities [19].

4.4 Hydrogen Trains It is believed that the first hydrogen locomotive was developed, designed and initiated in the USA in 2002. This train used a polymer electrolyte membrane Fuel Cell (PEMFC) with metal hydrate batteries and storage [15, 20]. By 2012, five mining locomotives were introduced for commercial operation in South Africa. There has been a significant amount of research and development into hydrogen fuel cell (FC) powered transport since then with a large focus being placed on passenger vehicles [15]. In Japan, the JR East trialled a hydrogen rail car between 2006 and 2007 which had the maximum speed of 100 km/h [15]. In 2016, the Alstom Coradia iLint was manufactured and had a range of 1,000 km and a maximum speed of 140 km/h. This train emitted only steam and condensed water as a by-product as it combined various innovative elements including clean energy conversion, flexible

4.4 Hydrogen Trains

55

battery energy storage and management of traction power [15]. In 2017, a prototype of hydrogen light rail was introduced by the China Tangshan railway. This was powered by a pack of 150 kw fuel cells which were working in parallel with packages of battery and ultrareceptors [21]. A 12 kg tank of hydrogen was used to support 40 km of range in each of the 15-min trips. In addition, the line had four stations as well as a 100 kg capacity refuelling station [21], highlighting the commercially availability of hydrogen for rail. In June 2019, the HydroFlex was introduced in the UK which is a hybrid model intended to draw a proportion of its power from overhead lines or third rails, whilst the hydrogen FC can be used to supply electricity if overhead power is unavailable. Hydrogen train projects are being developed further in Germany, France, the Netherlands, and the UK [22]. Previous studies have discussed the technological feasibility of the introduction of hydrogen trains and have completed economic evaluations to develop a case for their introduction [2, 23–27]. Hydrogen rail is being proposed as a viable option for multiple reasons. Firstly, hydrogen can be produced from multiple energy sources which allows for reduced stress on specific renewable electricity technologies and its power systems can be technically acceptable and implemented in real traction systems [15]. Secondly, the power component can be accommodated within the space available on the train and permissible restrictions [15]. Therefore, taking this into consideration hydrogen rail is likely to be a good fit within the transport network. As this technology develops it remains likely it will be best utilised in rural areas where the traffic is low and does not warrant electrification and requires a larger range than a battery powered train for longer journeys with infrequent stops.

4.5 High-Speed Rail High-speed rail (HSR) has been developed worldwide over the past 60 years and has gained favour in many countries due to the advantages in speed, convenience, and safety [28], competing with short-haul flights. In 2020, more than 900 billion passenger kilometres were travelled by HSR compared to the 3,100 billion by conventional rail. Between 2008 and 2018, China’s HSR expanded 3.4 times with an annual growth of 13% and the introduction of 24,000 track kilometres, with two more highspeed rail corridors being added in 2019 [6]. This rapid rollout in China is one of the largest infrastructure projects in recent history and has resulted in HSR activity catching up with domestic aviation [6]. Similarly, the Indian Government is currently completing the land acquisition to construct an HSR corridor between Mumbai and Ahmedabad [6]. This will be the first of six HSR lines that the Indian government has planned totalling ~ 510 kms to build in the upcoming years to connect India’s largest cities [6]. HSR makes it an alternative mode of transport to flying (this is further discussed in Chap. 6). However, in different countries the speed of HSR travel varies. For example, in Japan, the Shinkansen HSR system operates at speeds of at least 200 km/h [29–31]. In order to achieve these higher speeds a different type of infrastructure, rolling stock,

56

4 Trains

signalling and operation is required [32]. The regular operating speed on most lines runs between 140 and 160 km/h, with a few rail corridors accommodating higher speeds of between 180 and 200 km/h [32]. Since the introduction of the Shinkansen in Japan, several other countries including Germany, Italy, France, and Taiwan have introduced a wave of HSR construction beginning in the 1980s [3]. China has started the construction of HSR in many regions and cities, covering 29,000 km as of 2018, accounting for more than two-thirds of the total mileage of HSR in the world [3]. Currently the maximum operating speed of China’s HSR is over 350 km/h. Even with the introduction of HSR, there remains criticism in some countries that will need to be addressed to ensure public acceptability and encouragement of low carbon rail. For example, the HSR 2 being built in England, UK, which is expected to decrease travel time between major cities including between Birmingham and London (~160 km) from ~ 81 to ~ 52 min. However, the cost to the consumer is expected to remain the same [33]. Although there are benefits including the convenience of getting to a destination in almost half the time, the number of individuals currently using rail remains low so it may be some time before the benefits are fully accounted [34]. In addition, hundreds of habitats and special wildlife sites are potentially under threat due to its construction. Although there are expected trade-offs when transitioning to low carbon electricity and transport, ensuring that there remains minimal impacts on natural capital and ecosystem services should also remain a key concern [35, 36].

4.6 Key Findings Low carbon rail offers an alternative to other transportation options, including domestic and in some cases international air travel, which emit higher levels of emissions per person per kilometre travelled (further discussed in Chap. 6). Although rail travel is considered a ‘green’ mode of transport, without improvements to the current rolling stock, as well as tracks and platforms, emissions could remain static which will result in a bigger challenge to reaching net zero emission reduction targets. This is particularly important to consider when current ruling stock has an average life expectancy of between 20 and 40 years. Numerous studies have highlighted that electric trains produce lower levels of emissions compared to CFTs and hydrogen trains. However, this remains dependent on the electricity generation mix. Furthermore, it remains likely that a combination of both HTs and ETs will be required to travel within urban and rural areas due to the different capabilities of these different transport options. Policymakers need to encourage a transition from personal vehicles towards low emission rail. However, this remains challenging with alternatives such as personal vehicles remaining the preferred option due to cost and convenience. The implementation of new technologies such as the HSR is hopeful to encourage travellers to choose rail however there may need to be additional incentives to encourage the necessary shift.

References

57

References 1. Hampaeyan Miandoab M, Ghezavati V, Mohammaditabar D (2020) Developing a simultaneous scheduling of passenger and freight trains for an inter-city railway considering optimization of carbon emissions and waiting times. J Clean Prod 248:119303 2. Logan KG, Nelson JD, McLellan BC, Hasting A (2020) Electric and hydrogen rail: potential contribution to net zero in the UK. Transp Res Part D Transp Environ 87:102523 3. Guo X, Sun W, Yao S, Zheng S (2020) Does high-speed railway reduce air pollution along highways? Evidence from China. Transp Res Part D Transp Environ 89:102607 4. Chen Y, Whalley A (2012) Green infrastructure: the effects of urban rail transit on air quality. Am Econ J Econ Policy 4:58–97 5. Guo S, Chen L (2019) Can urban rail transit systems alleviate air pollution? Empirical evidence from Beijing. Growth Change 50:130–144 6. IEA (2020) Rail. https://www.iea.org/reports/rail. Accessed 6 Oct 2021 7. Zamir Khan M, Naheed Khan F (2021) A dynamic analysis of rail travel demand in Pakistan. Case Stud Transp Policy 9:860–869 8. Sobieralski JB (2021) Energy consumption and emissions dynamics of U.S. domestic intercity air travel. Transp Res Part D Transp Environ 99:102993 9. Mulley C, Hensher DA, Cosgrove D (2017) Is rail cleaner and greener than bus? Transp Res Part D Transp Environ 51:14–28 10. González-Gil A, Palacin R, Batty P, Powell JP (2014) A systems approach to reduce urban rail energy consumption. Energy Convers Manag 80:509–524 11. Lin D, Nelson JD, Beecroft M, Cui J (2021) An overview of recent developments in China’s metro systems. Tunn Undergr Sp Technol 111:103783 12. Givoni M, Rietveld P (2007) The access journey to the railway station and its role in passengers’ satisfaction with rail travel. Transp Policy 14:357–365 13. Ntlatywa K (2019) Determinants of rail passenger transport usage: a case of Buffalo City Municipality. EIRP Proceedings, vol 14 14. Bubalo T, Rajsman M, Kukec T (2020) Dynamics of passenger demand and transport work in Croatian public road traffic system. Am J Traffic Transp Eng 5:34 15. Sun Y, Anwar M, Hassan NMS, Spiryagin M, Cole C (2021) A review of hydrogen technologies and engineering solutions for railway vehicle design and operations. Railw Eng Sci 29:212–232 16. Spiryagin M, Cole C, Sun YQ, McClanachan M, Spiryagin V, McSweeney T (2014) Design and simulation of rail vehicles. CRC Press 17. Cipek M, Pavkovi´c D, Kljai´c Z, Mlinari´c TJ (2019) Assessment of battery-hybrid diesel-electric locomotive fuel savings and emission reduction potentials based on a realistic mountainous rail route. Energy 173:1154–1171 18. Mandi´c M, Ugleši´c I, Milardi´c V, Filipovi´c-Grˇci´c B (2015) Application of regenerative braking on electrified railway lines in AC traction systems 25 kV, 50 Hz. In: 12th Symposium. HRO CIGRÉ 8 19. Staffell I, Scamman D, Velazquez Abad A, Balcombe P, Dodds PE, Ekins P, Shah N, Ward KR (2019) The role of hydrogen and fuel cells in the global energy system. Energy Environ Sci 12:463–491 20. Jones WD (2006) Hydrogen on track. IEEE Spectr 43:10–13 21. Haji Akhoundzadeh M, Panchal S, Samadani E, Raahemifar K, Fowler M, Fraser R (2021) Investigation and simulation of electric train utilizing hydrogen fuel cell and lithium-ion battery. Sustain Energy Technol Assessments 46:101234 22. Ruf Y, Zorn T, De Neve PA, Andrae P, Erofeeva S, Garrison F, Schwilling A (2019). Study on the use of fuel cells and hydrogen in the railway environment. https://doi.org/10.2881/495604 23. Marin GD, Naterer GF, Gabriel K (2010) Rail transportation by hydrogen vs. electrification— case study for Ontario Canada, I: propulsion and storage. Int J Hydrogen Energy 35:6084–6096 24. Marin GD, Naterer GF, Gabriel K (2010) Rail transportation by hydrogen vs. electrification— case study for Ontario, Canada, II: energy supply and distribution. Int J Hydrogen Energy 35:6097–6107

58

4 Trains

25. Hoffrichter A, Hillmansen S, Roberts C (2016) Conceptual propulsion system design for a hydrogen powered regional train. IET Electr Syst Transp 6:56–66 26. Haseli Y, Naterer GF, Dincer I (2008) Comparative assessment of greenhouse gas mitigation of hydrogen passenger trains. Int J Hydrogen Energy 33:1788–1796 27. Hoffrichter A, Miller AR, Hillmansen S, Roberts C (2012) Well-to-wheel analysis for electric, diesel and hydrogen traction for railways. Transp Res Part D Transp Environ 17:28–34 28. Khodaparastan M, Mohamed AA, Brandauer W (2019) Recuperation of regenerative braking energy in electric rail transit systems. IEEE Trans Intell Transp Syst 20:2831–2847 29. Jani´c M (2021) Estimation of direct energy consumption and CO2 emission by high-speed rail, transrapid maglev and hyperloop passenger transport systems. Int J Sustain Transp 15:696–717 30. Givoni M (2006) Development and impact of the modern high-speed train: a review. Transp Rev 26:593–611 31. Ji H, Yoshitsugu H, Peng J, Quan Y (2012) Economic effect of high-speed rail: empirical analysis of Shinkansen’s impact on industrial location. J Transp Eng 138:1551–1557 32. Baumeister S, Leung A (2021) The emissions reduction potential of substituting short-haul flights with non-high-speed rail (NHSR): the case of Finland. Case Stud Transp Policy 9:40–50 33. DfT (2016) High-speed two: from Crewe to Manchester, the West Midlands to Leeds and beyond. https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attach ment_data/file/897407/high-speed-two-crewe-manchester-west-midlands-leeds-document. pdf. Accessed 7 Oct 2021 34. Lalive R, Luechinger S, Schmutzler A (2018) Does expanding regional train service reduce air pollution? J Environ Econ Manage 92:744–764 35. The Wildlife Trusts (2021) High-Speed Rail (HS2). https://www.wildlifetrusts.org/hs2. Accessed 8 Oct 2021 36. Cornet Y, Dudley G, Banister D (2018) High-Speed Rail: Implications for carbon emissions and biodiversity. Case Stud Transp Policy 6:376–390

Chapter 5

Challenges of Implementing Electric and Hydrogen Public Transport

Abstract Although it remains clear that low carbon public transport will produce less emissions than conventionally fuelled alternatives, and much less when compared to personal vehicles per person per kilometre travelled, encouraging the use of public transport has multiple barriers that need to be overcome to increase widespread use. This includes a shift in travel behaviour for consumers in terms of public acceptance of new low carbon public transport, including encouraging public transport use at younger age so that individuals continue to use public transport as they get older. The main barrier remains the cost of the technology itself as well as the additional infrastructure when integrating low emission public transport. As technology advances it is expected that electric and hydrogen public transport costs will decrease, however hydrogen technology is still relatively new making this more challenging. Since travellers are cost-sensitive fares it is important that public transport remains competitive with personal vehicles. Furthermore, the practicalities of integrating new technologies need to be considered. For example, the introduction of electric and hydrogen in public transport will require additional infrastructure for charging which will have a geospatial impact on natural capital and ecosystem services. This is because electric transport will need to be charged more frequently than hydrogen transport due to the range of the vehicles. Furthermore, electrification of public transport will result in additional challenges for power generation and infrastructure development which will need to be addressed for widescale uptake. Therefore, electric charging infrastructure is more likely to be situated within city centres compared to hydrogen transport which can be situated in rural areas.

5.1 Introduction This chapter discusses the factors that influence the use of public transport and what can be done to actively encourage this transition away from personal vehicles. Section 5.1 discusses factors to encourage and influence consumers and the reasons why they travel. Section 5.2 highlights the public acceptability of new technologies including electric and hydrogen low emission public transport. Section 5.3 © The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_5

59

60

5 Challenges of Implementing Electric and Hydrogen Public Transport

discusses the technology and infrastructure challenges for electric and hydrogen public transport, including both monetary and non-monetary factors. Section 5.4 compares the cost of these new technologies and the associated infrastructures that would be required for wide scale implementation. Section 5.5 discusses the electricity demands for electric and hydrogen low emission public transport integration and the challenges that need to be addressed before widespread integration, such as energy storage. Section 5.6 discusses the environmental impact of low emission transport, including the impact on natural capital (NC) and ecosystem services (ES) from these transport alternatives. Section 5.7 highlights the key conclusions from this chapter.

5.2 Factors Influencing Public Transport Use Demand for public transport is mostly driven by an individual’s need to overcome the spatial mobility constraints of distance, time, and isolation from social (i.e., visiting family/friends, shopping etc.), economic (i.e., work) and public services (i.e., healthcare or school etc.) aspects of everyday life [1–5]. For a successful integration of public transport, the social sustainability of transport systems will need to be able to provide all members of society with equal access to opportunities and services [6]. Therefore, policymakers and planners need to consider the design of the network, routes, and frequency and how they meet the transport needs of the population [7, 8]. Scheduling a fleet of vehicles is an essential task within the planning process of public transport systems [9]. When constructing a schedule, a bus or train company’s primary objective is to determine the assignment of a set timetable to provide a service (i.e., the trip for transporting a passenger from a departure stop to an arrival stop or through deadhead trips without passengers to change location), at the minimal cost to the company taking into consideration the cost, maintenance, and fuel use [9]. This needs to ensure the trips are executable without time overlaps, each trip is covered, and a vehicle begins and ends at a specific depot [9]. Furthermore, through analysis of perceived satisfaction of public transport, four satisfaction key dimensions have been identified including: system, comfort, staff, and safety [10]. However, these satisfaction constructs were not cognitively perceived similarly across different cities due to dissimilarities in culture and tradition [10]. The service being provided to the consumer will need to ensure consumer retention as well as actively encouraging new users, primarily by shift from personal vehicles [11]. For example, a study of Montreal, Canada highlighted that public transport use decreased with age, but the individuals who used public transport in their youth tended to decrease their public transport mode share less than the individuals that relied on other modes of transport when they were young [12]. A study of Kyoto University students in Japan who regularly commute using personal vehicles were given a free one-month bus pass [13]. The results of this study demonstrated that the habits and frequency of bus use increased during the study period, with personal vehicle use decreasing even after the intervention period [13]. Therefore, encouraging

5.2 Factors Influencing Public Transport Use

61

public transport at a younger age suggests that even temporary structural changes may be important for encouraging mode shift towards public transport. Experience shows that the habits of transport usage can influence public transport uptake as they act at an unconscious level until it is ‘second nature’ [14–16]. For example, an individual’s previous experience with public transport has the potential to influence whether or not they considered using public transport after moving to a different residential area [17]. Furthermore, individuals who had previously lived in areas where public transport was poor or unavailable were more likely to commute by personal vehicle than individuals who moved to an urban area where public transport was of higher quality [17]. These studies highlight the need to ensure passenger retention, as a small change in ridership numbers could significantly impact the growth of the public transport market, and most strategies are designed to increase use as well as retain current passenger numbers [18, 19]. Furthermore, there has been a considerable consensus in literature that public transport remains a low status mode of transport [20–25]. Although bus use in some cities, such as London remains a popular method of transport [26], bus users are commonly described as being the most economically and socially disempowered members of society, with disproportionally higher numbers of older individuals, single mothers, recent immigrants, domestic service workers, individuals on low-income or individuals with disabilities using this transport method [20, 24, 25, 27–29]. This stigma related to lower class often leads buses to be regarded as a last resort transport mode [20, 24, 28, 29]. For example, in New Zealand buses are sometimes labelled ‘loser cruisers’, suggesting that only ‘losers’ use this transport method [30–32]. Therefore, until the stigma against buses has been addressed, encouraging bus use will remain challenging. In order for policymakers to develop comprehensive strategies to retain and encourage public transport use, it is necessary to identify which aspects of public transport influence satisfaction and loyalty for the consumer [33, 34]. In the context of public transport, satisfaction can be defined as a customer’s overall experience of a service compared to their pre-defined expectations [35], whereas loyalty is the customers intention to use the service in the future based on pervious experiences [36]. The range for bus satisfaction across European cities is very different exhibiting a higher percentage in Northern and Western Europe (i.e., Zurich (95%), Helsinki (89%), Rotterdam and Rennes (87%)) and low rates in central Eastern and Southern European countries (i.e., Palermo, Naples, Budapest) [37]. Poor bus satisfaction stems from poor performance including regularity, punctuality, speed, comfort, and design [37]. These are all contributing factors to the attractiveness of the bus compared to other modes and can reduce the loyalty of users. Although improvements have been made to address these factors, individuals are still far from changing their perception of buses [38]. Public transport has high operational and capital costs and the ability to recover this via the fare box revenue varies considerably. For example, in the USA, the average operating and capital costs of the ten largest bus systems are $0.85 and $0.16 per passenger mile, substantially higher than those of private vehicles at $0.11 and $0.14 per passenger mile [39]. Transitioning a fleet to electric and hydrogen transport brings higher upfront costs. Therefore, subsidies are frequently needed to

62

5 Challenges of Implementing Electric and Hydrogen Public Transport

encourage fleet transition whilst maintaining public transport use by keeping costs to the individual consumer down. One way of doing this is for government to lease vehicles to operators for a fixed period after which the bus company takes ownership; thus, avoiding procurement costs for the operator. Multimodal fare and ticketing systems have become common allowing individuals to travel on different modes of public transport within a city during the day [4, 40]. For example, in the São Paulo municipality in Brazil the Interligado System was introduced which optimised bus routes and services. This saw total public transport trips increased by 15% with passenger boarding growing by 49% between 2002 and 2006, however it did not mention how many times a passenger boarded as they may have taken multiple trips [41]. Similarly, in London, Transport for London introduced Oyster cards in 2003 which allows individuals to use multiple methods of public transport daily (i.e., buses, trains, and the underground) with a daily cost cap. These incentives make the experience convenient for the consumer and are likely to encourage multimodal public transport use throughout the day, especially if available via a smart phone app. Smartcard payment has been introduced across the world (e.g., in Sweden, Paris, Barcelona, China, and Australia). In addition, several travel demand management (TDM) initiatives have been introduced that provide advantages to public transport. For example, many countries have introduced bus lanes, priority signals, transit malls, bus gates etc. to provide priority. These bus prioritisation schemes can improve the public transport service and have been widely investigated within the literature [42–47]. These schemes have been used to alleviate urban congestion and allow buses to be a more efficient transport method. Additionally, some major cities, such as London in the UK, have introduced ultra-low emission zones to reduce the greenhouse gas (GHG) emissions within city centres. These zones ‘cap’ the gCO2 km−1 a private vehicle can emit. If the vehicle emits higher levels of emissions than this ‘cap’, the vehicle owner will need to pay a fine. If low emission buses, are already in use within these areas, this could encourage modal shift. Accessibility is generally defined as the ability to reach an opportunity (i.e., the socio-economic activities or services) by one or several transport modes within a given travel time or cost [48–50]. Bus use plays an important role providing accessibility, however, there remains a large gap with availability in rural compared to urban areas [51]. This inequality in terms of available transport options can have serious impacts on certain demographics, for example, the elderly, who may be reliant on public transport. This in turn can affect social inclusion, leading to social injustice [52]. In addition, bus services within more rural areas tend to be of lower quality than urban areas due to limited investments and subsidies [52]. This can lead to a ‘chicken and egg’ approach as policymakers are less likely to invest in the infrastructure for rural public transport if use is low. Furthermore, low-income individuals tend to suffer the most when it comes to bus use as they are the least able to commute longer distances when public transport is not available locally or in an affordable way [53–57]. Therefore, without further considerations, individuals in rural areas will prioritise personal vehicle use over public transport as this remains the most cost effective and convenient transport type.

5.3 Public Acceptance of New Technologies

63

Urban areas typically have higher populations and economic activity than rural areas. Furthermore, the closer proximity of an individual’s home to services and amenities (relative to that in rural areas) makes the provision of adequate transport services much easier than in rural areas [58]. Due to the cost and convenience, personal vehicles are therefore often favoured in rural areas. Given that ensuring an adequate public transport network within rural areas is not always achievable, encouraging a transition to personal electric vehicles (EVs) in these areas becomes more and more important. In rural areas, deployment of ‘ultra-rapid’ charging (150 kW) infrastructure will more likely be favoured to encourage the use of battery electric vehicles (BEVs) over conventionally fuelled personal vehicles as consumers typically want the conveniences of charging their vehicle for the same duration as fuelling a vehicle at a petrol station.

5.3 Public Acceptance of New Technologies When designing a public transport network, there are two different perspectives that need to be considered: the perception of quality the consumer has which is based on the users’ experiences, and the expected quality, which determines the users’ expectations of a public transport network [59]. Perceived service quality is one of the most important constructs in marketing literature and remains a significant variable that correlates to both customer satisfaction and value [60] since it indirectly measures how well a service delivery matches or exceeds customer expectations. Public acceptance of new technologies will be dependent on a myriad of spatial and temporal aspects that are measured and defined at different levels. Whilst ‘community’ acceptance describes an individual’s perspective at a local level, ‘sociopolitical’ acceptance refers to general attitudes towards energy technologies, typically measured through national opinion polls [61]. Therefore, the success of integrating electric and hydrogen public transport will not only be determined by the technical parameters discussed later in this chapter, but the acceptance of the end users [62]. For the successful integration of a new technology, the technology itself must be simple enough for the general public to understand the basics. For example, a lot of individuals today know how to drive a personal vehicle but may not directly know the principles of an internal combustion engine [62]. Therefore, when introducing a new technology, focusing primarily on entrepreneurs or environmental groups will not be enough for a successful integration. Introducing a media campaign alongside new technology integration will become a necessity if the public are to accept hydrogen public transport. However, as knowledge of hydrogen technologies improves, then public acceptance of hydrogen vehicles could improve [63–65]. However, it remains unknown whether individuals are aware of the transport type they are on as most buses for example in London are now electric and hydrogen.

64

5 Challenges of Implementing Electric and Hydrogen Public Transport

5.4 Technology and Infrastructure Challenges for Electric and Hydrogen Public Transport Electric and hydrogen public transport are subject to the same operational challenges as conventionally fuelled buses (CFBs). However, there are several additional challenges associated with electric and hydrogen buses that will need to be addressed before widescale introduction. The rate and manner in which low carbon transport infrastructures (e.g., charging stations, electricity generation etc.) are deployed will play an important role in determining the GHG emissions, energy demand and economic impact of the sector [66]. Research has highlighted that transport infrastructure and asset locations create an inertia on transport emissions, which is larger than the inertia of the vehicle fleet itself [67]. The environmental impact of climate policies on infrastructure investment remains ambiguous, with the long-terms implications of infrastructure often overlooked in research, therefore transport infrastructure remains at the intersection between climate and development objectives [67, 68]. Analysis of the physical infrastructure required for transport and buildings, highlighted that if per capita levels of infrastructure in Western countries was constructed globally, using current technologies, it would require using between 35 and 60% of the remaining carbon budget until 2050 to build a global infrastructure to keep average global temperature is to remain below 2 °C [69]. Furthermore, analysis has highlighted that at least 25 million kilometres of new roads are anticipated by 2050, which is the equivalent of a 60% increase in total road length in relation to 2010 levels [70]. Therefore, future emission projections from the transport sector should consider the ‘induced demand’ from the infrastructure and the emissions from the construction of the infrastructure itself [66]. This will be particularly important as mobility levels increase requiring new infrastructure, as well as upgrades to existing infrastructure [68]. Through careful infrastructure planning, a transition towards low carbon transport can be integrated in both developed and developing countries, which minimises the overall emissions through a whole systems approach [71, 72]. The scale and rate of this infrastructure transition is unprecedented, with the high material intensity of infrastructures likely to have a significant impact on material uses [73]. The materials required to build the infrastructure for low carbon technologies will require a wide use of materials than have not been used before, especially not at the scale required for demand [74]. For example, rare earth elements (i.e., neodymium [75]) and dysprosium for wind turbines, and tellurium and indium in solar panels [76], as well as other metals (i.e., cobalt, lithium, and platinum metals) will be required in large quantities [77]. Some of these required materials have been classified as ‘critical’ due to the resulting high risk of supply shortage, with policymakers driving academic research to identify potential critical materials [78, 79]. This demand for ‘critical’ materials can have negative implications for ES and NC as these two areas are linked. Therefore, if ES are used at an unsustainable rate, the stocks of NC which provided them may be depleted, negatively impacting the environment. As electric and hydrogen transport gain momentum, the issue of locating and securing

5.4 Technology and Infrastructure Challenges …

65

the availability, efficiency, and effectiveness of charging infrastructure, whilst not negatively impacting ES and NC, becomes a complex issue [80].

5.4.1 Electric Public Transport The design of an integrated EBs and electric train (ET) network is more complicated than a conventionally fuelled bus or train network due to the battery charging and range constraints. Because of this, optimisation approaches for installing electric infrastructure should be considered [81]. Vehicle range remains an issue for low carbon transport [82], as in general, diesel CFBs have a continuous range of ~ 300 km in urban areas [83], however the maximum driving range for most current EBs varies between 70 and 200 km [82]. This is between 25 and 65% less than diesel alternatives [84]. In addition, the batteries used in EBs have a low energy density, which means that, for a reasonable range, they have to be large, heavy, and expensive [85]. For example, an EB with a range of ~ 200 km requires ~ 150 kg of lithium-ion cells or more than 500 kg of lead acid batteries [85]. Similarly, ETs that need to be recharged have a shorter range, although some trolleybuses (as mentioned in Chap. 3) have overhead wires that allow them to be more flexible and not require frequent charging. This makes electric low carbon alternatives more difficult to operate as this additional weight will cause the bus or train to need to be charged regularly to ensure they can travel the required distances to meet the advertised service levels. This limited range in comparison to CFBs imposes significant limitations one the use of electric buses. Electric alternatives are often more suited to city centres where charging infrastructure would be more readily available. Currently there are three methods available to remedy the short-range limitation: battery swapping, wireless lane-based charging, and station-based charging [86, 87], all of which present challenges. Firstly, battery swapping allows operators to replace the depleted batteries with fully charged batteries instantly [88]. This has the potential to only cause a small delay on bus scheduling [89–91]. However, this method requires high construction costs and land use requirements. For example, in Qingdao, China an area of 5800 m2 was required to build one battery swapping station for EBs [88]. Although the range of electric transport is suited towards urban areas, this prohibits the wide scale roll out where land resource remain expensive [88]. Furthermore, operators will need to consider the minimum number of spare batteries to have charged and in stock and to ensure there is a schedule for recharging to ensure buses can be used [91–94]. Secondly, wireless lane-based charging utilises dynamic wireless power transfer to charge electric transport whilst it is in motion [95, 96]. Wireless lane-based charging has the potential to be introduced on rapid bus corridors and could reduce the size of batteries used on buses [87]. This technology has begun being piloted in Gumi City, South Korea and at Utah State University in the USA [88]. This technology is likely

66

5 Challenges of Implementing Electric and Hydrogen Public Transport

to have high construction costs associated as the wireless lane-based charging will need to be embedded underneath the roads [87]. Finally, station-based charging, where there are two different types: fast chargers and slow chargers. Fast chargers offer higher power infrastructure installed at terminals or intermediate bus stops. Buses can be fully charged in less than 10 minutes adding an operation of around 80 km [88], however their short-added range make them somewhat unsuitable for long distance bus routes when their increased cost is taken into consideration. Introducing fast-charging infrastructure has received a great deal of attention due to the convenience of integrating fast-charging system planning into existing bus lines [97]. With this type of infrastructure, operators will need to consider the scheduled time and type of charger to ensure costs remain low, the location of the charging stations and the minimum number of charging points available to ensure demands are met [82]. Slow chargers are usually located at bus terminals or depots with buses being charged once they have finished a whole day’s operation. Charging overnight also helps to reduce electricity costs as this is generally during off-peak time, therefore reducing the associated environmental impacts. Even with these three charging strategies taken into consideration, the energy consumption rates of electric transport can fluctuate depending on-road gradient, traffic congestion, whether air conditioning is on or off or the use of auxiliary systems [98]. For example, an EB with a range of 250 km can only run for 175 km if its air conditioning is on [88]. Therefore, the actual charging frequency remains uncertain and can influence the frequency of charging, and thus the overall environmental impact. Through modelling and simulation tools, policymakers can determine the best possible routes for the transport type in terms of terrain, weather, and operational conditions to limit the impact on the range [99–101]. This is important to consider as batteries require an effective heating and cooling system to avoid power degradation and ageing [102, 103]. Therefore, extreme weather conditions can be challenging for low emission buses [98, 104]. A lack of available charging infrastructure for public transport can also pose a challenge. For example, a smaller number of charging stations throughout a city can reduce the construction costs but will require a larger battery on the vehicle, therefore increasing the investment costs of a bus fleet [87]. On the other hand, a larger supply of charging stations has the potential to reduce battery size, with frequent recharging adding additional charging delays to the service [87]. This would result in a larger supply being required to guarantee service frequency. Therefore, these trade-offs guaranteeing the frequency of the service and charging needs will need to be considered for an optimal design plan [87]. Although there are pilots of low carbon charging infrastructure for buses and trains across the world, infrastructure is not currently widespread enough for the mass introduction of low carbon public transport. Practical issues for depots will also need to be considered (e.g., implications for design and location and access to the electricity grid). In terms of contracting, an attractive future proposition is the concept of supply chain procurement contracts which would bring together chassis and body builders, energy providers, battery pack and electric or hydrogen bus providers alongside the bus operator. Furthermore, in cities where hydrogen or electric buses are currently operating, there remains

5.4 Technology and Infrastructure Challenges …

67

conventionally fuelled alternatives. Therefore, the current infrastructure in these pilot areas needs to be further expanded.

5.4.2 Hydrogen Public Transport A fundamental stage towards the proliferation of hydrogen transport is the deployment of the hydrogen infrastructure, i.e., generation, storage, and delivery and how this is going to be able to meet transport demands [105, 106]. Hydrogen transport requires hydrogen filling stations, with the ideal location recommended to be on site at the bus or train depot [84]. This is because it will minimise costs and the GHG emissions of the hydrogen pathway. However, this will require large areas of land which is within urban areas and therefore most refuelling stations are likely to be situated in rural areas where land is cheaper. The ideal locations for refuelling HBs are based on the geographic location, relative size of the fuelling stations and hydrogen costs. Globally, there is very little commercial H2 refuelling infrastructure and it only exists in localised areas [107, 108]. Therefore, an operator wanting to purchase their own HB will be very restricted or even prohibited from doing so due to the lack of support infrastructure [109]. In some countries, there are additional challenges involved with the construction and placement of new hydrogen refuelling stations [110]. For example, according to Dutch law, hydrogen refuelling stations are required to go through a permitting process to become established [110]. This is because the main aim of the permit is to limit the risks to individuals and the environment associated with storing, dispensing, and generating hydrogen at the refuelling stations [110]. Although this helps to reduce risks, this does slow down the process of integrating new infrastructure. For example, as part of the permitting process, although there is currently no regulatory necessity to provide a Quantitative Risk Assessment (QRA) for hydrogen refuelling stations with a storage capacity below 5,000 kg in the Netherlands, permitting authorities have begun to request QRAs as part of the permit applications to assess potential safety risks associated [110]. Hydrogen generation plants are typically equipped with hydrogen energy storage systems in order to manage the hydrogen production and delivery based on the expected demand [111]. By storing hydrogen, this will reduce demand on the grids and excess hydrogen can be generated during off-peak time at night. However, this process is not commercially available yet and can have more damaging environmental impacts as the electricity required for hydrogen production will be generated as needed. As hydrogen generation is a two-step process, ensuring efficiency when generating hydrogen is important and losses can increase the cost of generation. This is because hydrogen is affected by the delivery method (i.e., through pipes and transport by trucks), the state of the hydrogen (i.e., whether it is a gas or liquid), and the demand [112, 113]. However, investment into pipeline infrastructure to transport hydrogen requires a huge infrastructure cost [114]. One solution to reduce costs would be to

68

5 Challenges of Implementing Electric and Hydrogen Public Transport

reuse old infrastructure. For example, to transport hydrogen produced from electricity from offshore wind turbines, natural gas pipelines could be used as this mitigates any additional emissions and reduces impact on both natural capital and ecosystem services although this requires special seals and valves due to the small molecule size. In the absence of pipelines, hydrogen is most commonly transported and delivered as a liquid [114]. This remains a feasibility study and has not been undertaken yet.

5.5 Costs of Electric and Hydrogen Public Transport For a successful transition to low emission public transport use cost remains a considerable barrier. When introducing new technologies, any upfront costs, infrastructure costs, maintenance costs, environmental costs etc. need to be considered when considering the feasibility of new technologies. New technologies are likely to have higher costs but over time are likely to decrease in costs, however a ‘chicken and egg’ scenario regarding their integration costs will need to be considered.

5.5.1 Electric Buses EBs currently range from between about U.S.$540,000 to $1,050,000, not including the infrastructure required which is determined by the bus type and battery size [115]. Although the costs of EBs are expected to decrease, the costs resulting from the need to replace the battery remain high. This is because the battery life is approximately half that of the bus [116] and are often only under warranty for the first six years of implementation [117]. For EBs, the costs and durability of lithium-ion batteries have a significant influence on the life cycle cost, for example, a 2020 battery pack could cost between $275 and $375/kWh [118, 119]. However, these cost levels may not be entirely valid for EBs due to their larger battery systems and different chemistries [99]. Furthermore, battery replacement costs are often not considered within the reported costs by agencies and can range between $60,000–$72,000 [120]. Infrastructure costs can vary depending on the type of EB and the charging method. For example, the cost of plug-in charging equipment can range between $19,000 to $50,000, however this does not take into consideration the installation costs which fall within a similar range ($5,000–$55,000) [115]. Furthermore, the payback period of charging infrastructure for EBs is still long and future research is required on the refinement of these cost calculations [121]. Additionally, the maintenance costs of EBs can depend on the availability of parts from the manufacturers and whether a bus is still under warranty. On average, these costs have been reported at $0.27 km−1 ($0.05–0.91 km−1 ), which is lower than $0.7 km−1 for diesel buses ($0.20–0.91 km−1 ) [115].

5.5 Costs of Electric and Hydrogen Public Transport

69

5.5.2 Electric Trains The energy and material costs for electrified trains and infrastructure represents a significant proportion of the transport costs [122], therefore many studies highlight that diesel alternatives will likely still be needed for passenger and freight transport [123]. This means that it is likely that the share of diesel fuelled alternatives such as biodiesel, methanol and hydrogen in transport will not be likely to exceed 10% by 2040 [124]. As ETs are a new technology to CFTs and are more likely to require additional infrastructure. Furthermore, Popovich et al. [125] highlighted that near-future train battery prices for ETs can achieve similarity with diesel-electric trains if environmental costs are included or if rail companies utilise wholesale electricity prices and achieve 40% use of fast-charging infrastructure. Furthermore, transitioning to ETs would not only reduce emissions in the USA could save the U.S. freight rail sector ~ $94 billion over a twenty-year period [125]. Even with a low cost of operation the electric infrastructure required ensures that ETs remain expensive to install. For example, ~ e1–2 million per kilometre, therefore highlighting it can remain it can be costly to operate and maintain.

5.5.3 Hydrogen Buses Like EBs, there are several ‘chicken and egg’ challenges that need to be addressed for widespread HB introduction, with cost remaining a major barrier. HBs are still considered a relatively new technology. Over the past several decades the cost of new HBs have decreased significantly, from $3.2 million in 2007 to $1.27 million in 2018 [126]. Despite decreasing upfront bus costs HBs remain significantly higher than EBs and CFBs, however the mass production of HBs is expected to mature and reduce the upfront costs. HBs have higher costs due to the significant expenses associated with hydrogen production, especially as electrolysis is not available on a commercial scale [127, 128]. Similarly, the cost of the refuelling infrastructure is expected to follow this trend but remain high. The construction cost of one hydrogen refuelling station is approximately $1.5–3 million, which is the equivalent of ten times higher than a gasoline station [129]. This therefore limits the number of hydrogen refuelling stations that governments can introduce due to the high installation costs.

5.5.4 Hydrogen Trains Although hydrogen trains (HTs) remain a newer technology and are not readily available on a commercial scale the true cost of this technology on a wide scale is in unknown on a wider scale. However, the costs of hydrogen generation for fuel

70

5 Challenges of Implementing Electric and Hydrogen Public Transport

cell HT operation are beginning to become more cost competitive. For example, in Brandenburg, Germany ~ 10.1 million train kilometres would be switched to fuel cell technology which would save ~ 9.5 million litres of fuel [130]. If these trains were fuelled using fuel cells, ~ 2198 tonnes of hydrogen annually would be required to fuel these trains [130]. This relates to ~ 6.40 e/kg, which also includes the cost of a hydrogen refuelling station, therefore making hydrogen available at a cost-effective rate for consumers [130]. In this scenario it is assumed that hydrogen was generated from electrolysis which is one of the more environmentally friendly methods of generation. Similarly, a study in Italy demonstrated financial indicators that showing that the levelised cost of hydrogen and total cost of ownership was ~ 8 e/kg and ~ 12 e/kg respectively and return on the investment of ~ 19% [131]. Therefore, confirming that there remains a high potential for hydrogen this technology within heavy duty transport, especially as costs are expected to reduce in the future [131, 132].

5.6 Electricity Demands for Low Carbon Transport Integration Understanding how electricity is generated will be important to assess the environmental impact of electric and hydrogen transport. This is because low carbon transport requires low carbon electricity, not generated from fossil fuels, which could negate the benefits of transitioning to electrical motive power. Furthermore, during this transition there will likely need to be an increase in renewable electricity generation as current electricity generation is unlikely to meet transport demands. Therefore, until enough renewable electricity technologies have been deployed, fossil fuels are still likely to be generating electricity. Several interim technologies have begun to be developed to ‘trap’ excess carbon dioxide emissions. Firstly, the introduction of carbon capture and storage (CCS) which refers to the chain of technologies that captures CO2 from the flue gas of power stations or other industrial resources. They are then managed, transported, and stored in deep underground geological formations [133, 134]. CCS has already been introduced in numerous countries across the world, with the International Energy Agency (IEA) stating that if the 2ºC pathway for energy is to be met, CCS will need to capture 6000 Mt CO2 yr−1 from transport by 2050 [135]. It is important to note that these processes can act as ‘interim technologies’ whilst the construction of additional renewable energy catches up to the energy demand. These technologies should not be considered a ‘solution’ to climate change as decarbonisation is still necessary if net zero targets are to be met. Therefore, CCS and direct air capture needs to be used in combination with other technologies to offset emissions that are very difficult to mitigate, like food production.

5.6 Electricity Demands for Low Carbon Transport Integration

71

5.6.1 Reducing Emissions and Demand on the Grid Ensuring that electricity is generated from renewable sources will help decrease the environmental impact of low emission transport. Diversification of a country’s energy mix from coal and oil towards renewables, natural gas and nuclear energy generation is more likely to ensure stable electricity generation. An efficient electricity network is also required to reduce grid losses. Hydrogen generation is a two-step process, requiring electricity, therefore electricity decarbonisation will be necessary. In addition, hydrogen generation is not 100% efficient and loses energy during the generation and transportation processes. Electrolysis is considered one of the most environmentally friendly methods of H2 generation, also known as ‘green hydrogen’, however it is not commercially available yet. When designing new technologies to meet energy demands, the spatial location of these energy sources should be a consideration. For example, when installing electricity and hydrogen electricity generation infrastructure and refuelling stations, site characteristics and utilisation feasibility should be considered. For example, new building designs could be compelled by building regulations to incorporate solar panels on their roofs, which would reduce land use conflicts elsewhere as solar requires significant land area and impacts ES and NC. This has the added benefit of localising energy production, if linked to a parking structure for example, or to reduce the network reliance of the buildings. Furthermore, tidal and wind generated electricity can be used to produce hydrogen locally to avoid transporting hydrogen long distances if grid connection in remote areas is not currently available (e.g., Orkney, Scotland). Utilisation of hydrogen energy locally can avoid restrictions that currently hinder the development of low emission transport. To reduce demand on the grid, ensuring adequate measures, such as the introduction of smart metres, has allowed a more co-ordinated timing of widespread charging. The introduction of smart metres will be particularly important for BEV users as it is likely that individuals will plug their vehicle in to charge as soon as they return home from work. Therefore, there is likely to be a peak demand at this time although individuals are not likely to use their vehicle until the next day; smart metres can ensure vehicles are charged overnight when energy demands are lower. For example, in the UK, BEV charging regulations are continually changing with the Automated and Electric Vehicles Act 2018 stating in Section 15 that infrastructure installed for the purposed of charging BEVs are to have ‘smart functionality’. This allows charging points to receive, understand and respond to signals sent by energy system participants (i.e., energy companies, National Grid etc.) to balance energy supply and demand. Therefore, operators will be required to modify their charging infrastructure to ensure ‘smart’ functionality. Smart charging can be used to minimise emissions, cost, and peak demand. As technology advances, smart charging will likely become the norm and marginal emissions for BEVs are likely to fall below average. These issues need to be addressed if nations are to meet their net zero targets.

72

5 Challenges of Implementing Electric and Hydrogen Public Transport

By shifting lower energy density generation types to areas unsuitable for higher value ES, this has the potential to minimise land use conflicts and minimise impacts on ES and NC optimising land use. Utilisation of hydrogen energy locally can avoid restrictions that currently hinder the development of low emission transport. Therefore, future planning for sustainable cities should ensure both compact and energy efficient designs as well as allowing a maintainable and liveable landscape [136]. In addition, encouraging greater use of public transport will also help to alleviate this demand as there will be less vehicles needing to be charged.

5.6.2 Energy Storage Technologies Energy storage through battery farms or hydrogen storage will be needed to ensure that cleaner energy, generated by nuclear or renewable sources during non-peak times, can be best utilised and reduce reliance on fossil fuels. This is often a critique of renewable energy, with this current intermittency issue dealt with using dispatchable gas powered electricity to balance the load or by paying for the temporary shutdown of windfarms, neither of which is a sustainable route in the future. Whilst this strategy is possible at the larger grid scale, the demands on the network as it transitions will further be enabled by possible financial incentives for individuals to charge their vehicles during certain times of the day to make system wide planning more attainable. This means that future energy networks require a dual axis approach, with encouragement of BEVs supported by appropriate grid management approaches to manage the peak demand curves which will overall decrease BEV emission impact. In addition, there is the potential for BEVs to be used to help cope with the broader energy network demands if the network infrastructure is developed with this consideration in mind. For hydrogen transport to be a successful and viable transport option long-term, the storage of hydrogen will be necessary. To store hydrogen from the location of the off-site production to the utilisation sites, it must be pressurised and delivered either as a compressed gas or liquefied [137]. This process also requires electricity to convert the hydrogen to either compressed gas or liquid form and then converted back at a later stage. It is therefore very important to take a whole system perspective when considering hydrogen storage. Keeping hydrogen in a low-pressure gaseous form is the most preferable method in terms of efficiency, however vehicles are not able to store enough hydrogen using this method [138]. Therefore, compressing hydrogen into its liquid form or in its gaseous form will be necessary. Converting hydrogen into its liquid form reduces its efficiency although it is sold in this capacity [138]. Technology has not advanced enough for the liquid hydrogen form to be used on-board vehicles [138]. Therefore, hydrogen storage technology should be considered from the energy production to the energy use. This process is not considered particularly energy efficient as significant hydrogen is lost during the generation and transporting stages.

5.7 Environmental Implications

73

5.7 Environmental Implications Considering the demands of increased electricity requirements together with transport infrastructure requirements within city and regional planning, there are likely significant environmental impacts associated with the integration of low carbon transport. The implications of these impacts are important to consider when structuring policy approaches for implementing additional infrastructure that will be needed for low carbon transport to meet net zero. Future planning for sustainable cities should ensure both compact and energy efficient designs as well as allowing a maintainable and liveable landscape [135]. Considering the demands of increased energy requirements together with transport infrastructure requirements within city and regional planning, should lead to a whole system design approach. For example, incorporating solar PV on the roof areas within city design regulations and refuelling stations for electric and hydrogen transport by considering the site characteristics and utilisation feasibility. This can reduce the impact on NC and ES and minimise land use conflicts as few, if any, other services can utilise roof areas. This has the added benefit of localising energy production for either electricity or hydrogen transport, if linked to a parking structure for example, or to reduce the network reliance of the buildings. Moreover, when designing a city, the’15 (or 20) minute’ city structure should be further considered as this allows for a more practical layout and ensures that a city doesn’t centre around personal vehicles [139]. This encourages active travel which in turn should aid in the reduction of GHG emissions. Furthermore, rural areas have the space to develop renewable electricity generation so that whilst public transport may be more difficult to implement in rural areas, the supply of local low carbon energy as electricity or via hydrogen production, may be more easily implemented and have less impact on ES and NC. Additionally, this has the benefit of reducing local area land use ES conflicts in densely populated areas. To meet the needs of hydrogen transport, significantly more investment is required into renewable energy than for electric transport. To reduce the environmental impact on both ES and NC in terms of land area required to meet hydrogen demands, emphasis on hydrogen generation from steam methane reforming with carbon capture and storage will be required until hydrogen generated from electrolysis is available on commercial scale. If energy storage technologies are developed this is likely to have a reduced impact as less electricity generation stations will need to be constructed. The area requirements for offshore and onshore wind and solar panels are important to consider in terms of impact on ES (disruption to hydrological process or scenic spots) and therefore on NC.

74

5 Challenges of Implementing Electric and Hydrogen Public Transport

5.8 Key Findings Although it has been proven that low carbon public transport will emit lower emission levels than private vehicles, individual attitudes will need to be shifted and new technologies accepted by the public. By encouraging the use of public transport at a young age, individuals are more likely to continue using public transport throughout their lives. Therefore, ensuring that public transport meets the needs of the individual by being convenient, frequent, and cost competitive will be important. For widespread integration of low carbon public transport, several barriers will need to be addressed. On an individual level, this will include socio-demographic barriers, focusing on whether the vehicles meet the needs of the individual such as the frequency and reliability of the network. Additionally, economic barriers, including the cost to the operator of critical infrastructure and the cost to the individual traveller at the point of use need to be overcome. Furthermore, as it is likely that a combination of both electric and hydrogen public transport will be required to meet net zero, hydrogen public transport is likely to be situated in rural areas. This is because hydrogen fuelled public transport tends to have a larger range than electric transport, therefore requiring less charging and travelling longer distances. In addition, hydrogen refuelling stations are suited to rural areas due to the large land areas required as hydrogen is generated on site. To ensure widescale connectivity, governments and policymakers will need to take a whole systems approach to ensure that the charging infrastructures introduced will accommodate the transport types as well as have limited impact on NC and ES. Further focus will need to be placed on how the electricity generated to ‘fuel’ these low carbon alternatives is made to ensure minimal environmental impact.

References 1. Ruiz M, Segui-Pons JM, Mateu-LLadó J (2017) Improving bus service levels and social equity through bus frequency modelling. J Transp Geogr 58:220–233 2. Joseph L, Neven A, Martens K, Kweka O, Wets G, Janssens D (2021) Exploring changes in mobility experiences and perceptions after implementation of the bus rapid transit system in Dar es Salaam. Case Stud Transp Policy. https://doi.org/10.1016/j.cstp.2021.04.012 3. Lucas K, Porter G (2016) Mobilities and livelihoods in urban development contexts: introduction. J Transp Geogr 55:129–131 4. Chowdhury S, Hadas Y, Gonzalez VA, Schot B (2018) Public transport users’ and policy makers’ perceptions of integrated public transport systems. Transp Policy 61:75–83 5. Pojani D, Stead D (2015) Sustainable urban transport in the developing world: beyond megacities. Sustainability 7:7784–7805 6. Casas I (2007) Social exclusion and the disabled: an accessibility approach. Prof Geogr 59:463–477 7. Ruano-Daza E, Cobos C, Torres-Jimenez J, Mendoza M, Paz A (2018) A multiobjective bilevel approach based on global-best harmony search for defining optimal routes and frequencies for bus rapid transit systems. Appl Soft Comput 67:567–583 8. Mulley C, Nelson J (2021) Public transport network planning. In: International encyclopedia of transportation, Elsevier, pp 338–394

References

75

9. Olsen N, Kliewer N, Wolbeck L (2020) A study on flow decomposition methods for scheduling of electric buses in public transport based on aggregated time–space network models. Cent Eur J Oper Res. https://doi.org/10.1007/s10100-020-00705-6 10. Fellesson M, Friman M (2012) Perceived satisfaction with public transport service in Nine European Cities. J Transp Res Forum. https://doi.org/10.5399/osu/jtrf.47.3.2126 11. Allen J, Muñoz JC, de Dios OJ (2018) Modelling service-specific and global transit satisfaction under travel and user heterogeneity. Transp Res Part A Policy Pract 113:509–528 12. Grimsrud M, El-Geneidy A (2014) Transit to eternal youth: lifecycle and generational trends in Greater Montreal public transport mode share. Transportation (Amst) 41:1–19 13. Fujii S, Kitamura R (2003) What does a one-month free bus ticket do to habitual drivers? An experimental analysis of habit and attitude change. Transportation (Amst) 30:81–95 14. Hitchings R (2012) People can talk about their practices. Area 44:61–67 15. Setten G (2020) Habitus 16. Shove E, Pantzar M, Watson M (2012) The dynamics of social practice: everyday life and how it changes. Sage 17. Weinberger R, Goetzke F (2011) Drivers of auto ownership: the role of past experience and peer pressure. Auto Motives 18. Barlow S, Clark H (2006) Findings of customer surveys conducted for Sacramento Regional Transit 19. Perk V, Flynn J, Volinski J (2008) Transit ridership, reliability, and retention 20. Fitt H (2018) Habitus and the loser cruiser: how low status deters bus use in a geographically limited field. J Transp Geogr 70:228–233 21. Beirão G, Sarsfield Cabral JA (2007) Understanding attitudes towards public transport and private car: a qualitative study. Transp Policy 14:478–489 22. Ellaway A, Macintyre S, Hiscock R, Kearns A (2003) In the driving seat: psychosocial benefits from private motor vehicle transport compared to public transport. Transp Res Part F Traffic Psychol Behav 6:217–231 23. Goodman A, Jones A, Roberts H, Steinbach R, Green J (2014) “We can all just get on a bus and go”: rethinking Independent mobility in the context of the universal provision of free bus travel to young Londoners. Mobilities 9:275–293 24. Green J, Jones A, Roberts H (2014) More than A to B: The role of free bus travel for the mobility and wellbeing of older citizens in London. Ageing Soc 34:472–494 25. Griskevicius V, Tybur JM, Van den Bergh B (2010) Going green to be seen: status, reputation, and conspicuous conservation. J Pers Soc Psychol 98:392–404 26. Steinbach R, Green J, Datta J, Edwards P (2011) Cycling and the city: a case study of how gendered, ethnic and class identities can shape healthy transport choices. Soc Sci Med 72:1123–1130 27. Guiver JW (2007) Modal talk: discourse analysis of how people talk about bus and car travel. Transp Res Part A Policy Pract 41:233–248 28. ANGROSINO MV (1994) On the bus with Vonnie Lee: explorations in life history and metaphor. J Contemp Ethnogr 23:14–28 29. Sheller M, Urry J (2000) The city and the car. Int J Urban Reg Res 24:737–757 30. Fitt H (2017) Do social meanings matter? How and how much social meanings influence everyday transport practices? N Z Geog 73:181–191 31. Moore T (2010) Perspectives on sustainable transport. Sizing Up City Urban Form Transp. New Zeal 32. Fitt H (2019) Beyond the urban commute: why we should be powering up for electric vehicle holidays. Lincoln Plan Rev 10:17–24 33. Friman M, Fellesson M (2009) Service supply and customer satisfaction in public transportation: the quality paradox. J Public Transp 12:4 34. Fornell C, Johnson MD, Anderson EW, Cha J, Bryant BE (1996) The American customer satisfaction index: nature, purpose, and findings. J Mark 60:7–18 35. Morfoulaki M, Tyrinopoulos Y, Aifadopoulou G (2010) Estimation of satisfied customers in public transport systems: a new methodological approach. J Transp Res Forum 46:63–72

76

5 Challenges of Implementing Electric and Hydrogen Public Transport

36. Morpace International Transit Cooperative (1999) A handbook for measuring customer satisfaction and service quality. Transportation Research Board 37. Corazza MV, Guida U, Musso A, Tozzi M (2016) A European vision for more environmentally friendly buses. Transp Res Part D Transp Environ 45:48–63 38. Tozzi M, Guida U, Knote T (2014) 3iBS: the intelligent, innovative integrated bus systems. In: Transp. Res. Arena 5th Conf. Transp. Solut. from Res. to Deployment European Comm. Eur. Dir. Roads Eur. Road Transp. Res. Advis. Counc. WATERBORNETP European Rail Res 39. Sun R, Chen Y, Dubey A, Pugliese P (2021) Hybrid electric buses fuel consumption prediction based on real-world driving data. Transp Res Part D Transp Environ 91:102637 40. Luk J, Olszewski P (2003) Integrated public transport in Singapore and Hong Kong 41. Hidalgo D (2009) Citywide transit integration in a large city: the interligado system of São Paulo, Brazil. Transp Res Rec 19–27 42. Ben-Dor G, Ben-Elia E, Benenson I (2018) Assessing the impacts of dedicated bus lanes on urban traffic congestion and modal split with an agent-based model. Procedia Comput Sci 130:824–829 43. Janos M, Furth PG (2002) Bus priority with highly interruptible traffic signal control: simulation of San Juan’s Avenida Ponce de Leon. Transp Res Rec 1811:157–165 44. Levinson HS, St. Jacques KR (1998) Bus lane capacity revisited. Transp Res Rec 1618:189– 199 45. Eichler M, Daganzo CF (2006) Bus lanes with intermittent priority: strategy formulae and an evaluation. Transp Res Part B Methodol 40:731–744 46. Viegas J, Lu B (2004) The intermittent Bus Lane signals setting within an area. Transp Res Part C Emerg Technol 12:453–469 47. Wu J, Hounsell N (1998) Bus priority using pre-signals. Transp Res Part A Policy Pract 32:563–583 48. Gutiérrez J (2001) Location, economic potential and daily accessibility: an analysis of the accessibility impact of the high-speed line Madrid–Barcelona–French border. J Transp Geogr 9:229–242 49. Geurs KT, van Wee B (2004) Accessibility evaluation of land-use and transport strategies: review and research directions. J Transp Geogr 12:127–140 50. Bertolini L, le Clercq F, Kapoen L (2005) Sustainable accessibility: a conceptual framework to integrate transport and land use plan-making. Two test-applications in the Netherlands and a reflection on the way forward. Transp Policy 12:207–220 51. González-González E, Nogués S (2019) Long-term differential effects of transport infrastructure investment in rural areas. Transp Res Part A Policy Pract 125:234–247 52. Tran Y, Yamamoto T, Sato H, Miwa T, Morikawa T (2020) Attitude toward physical activity as a determinant of bus use intention: a case study in Asuke, Japan. IATSS Res 44:293–299 53. Diaz Olvera L, Plat D, Pochet P (2013) The puzzle of mobility and access to the city in Sub-Saharan Africa. J Transp Geogr 32:56–64 54. Diaz Olvera L, Plat D, Pochet P (2015) Assessment of mobility inequalities and income data collection. Methodological issues and a case study (Douala, Cameroon). J Transp Geogr 46:180–188 55. Grengs J (2010) Job accessibility and the modal mismatch in Detroit. J Transp Geogr 18:42–54 56. Oviedo Hernandez D, Titheridge H (2016) Mobilities of the periphery: informality, access and social exclusion in the urban fringe in Colombia. J Transp Geogr 55:152–164 57. Shay E, Combs TS, Findley D, Kolosna C, Madeley M, Salvesen D (2016) Identifying transportation disadvantage: mixed-methods analysis combining GIS mapping with qualitative data. Transp Policy 48:129–138 58. Mounce R, Beecroft M, Nelson JD (2020) On the role of frameworks and smart mobility in addressing the rural mobility problem. Res Transp Econ 83:100956 59. Bordagaray M, dell’Olio L, Ibeas A, Cecín P (2014) Modelling user perception of bus transit quality considering user and service heterogeneity. Transp A Transp Sci 10:705–721 60. Laroche M, Ueltschy LC, Abe S, Cleveland M, Yannopoulos PP (2004) Service quality perceptions and customer satisfaction: evaluating the role of culture. J Int Mark 12:58–85

References

77

61. Wüstenhagen R, Wolsink M, Bürer MJ (2007) Social acceptance of renewable energy innovation: an introduction to the concept. Energy Policy 35:2683–2691 62. Kovaˇc A, Paranos M, Marciuš D (2021) Hydrogen in energy transition: a review. Int J Hydrogen Energy 46:10016–10035 63. Lipman TE, Elke M, Lidicker J (2018) Hydrogen fuel cell electric vehicle performance and user-response assessment: Results of an extended driver study. Int J Hydrogen Energy 43:12442–12454 64. O’Garra T, Mourato S, Pearson P (2005) Analysing awareness and acceptability of hydrogen vehicles: a London case study. Int J Hydrogen Energy 30:649–659 65. Yetano Roche M, Mourato S, Fischedick M, Pietzner K, Viebahn P (2010) Public attitudes towards and demand for hydrogen and fuel cell vehicles: a review of the evidence and methodological implications. Energy Policy 38:5301–5310 66. Ó Broin E, Guivarch C (2017) Transport infrastructure costs in low carbon pathways. Transp Res Part D Transp Environ 55:389–403 67. Guivarch C, Hallegatte S (2011) Existing infrastructure and the 2°C target. Clim Change 109:801–805 68. Fisch-Romito V, Guivarch C (2019) Transportation infrastructures in a low carbon world: an evaluation of investment needs and their determinants. Transp Res Part D Transp Environ 72:203–219 69. Müller DB, Liu G, Løvik AN, Modaresi R, Pauliuk S, Steinhoff FS, Brattebø H (2013) Carbon emissions of infrastructure development. Environ Sci Technol 47:11739–11746 70. Laurance WF, Clements GR, Sloan S et al (2014) A global strategy for road building. Nature 513:229–232 71. Henao A, Piatkowski D, Luckey KS, Nordback K, Marshall WE, Krizek KJ (2015) Sustainable transportation infrastructure investments and mode share changes: a 20-year background of Boulder, Colorado. Transp Policy 37:64–71 72. Tiwari G, Jain D, Ramachandra Rao K (2016) Impact of public transport and non-motorized transport infrastructure on travel mode shares, energy, emissions and safety: case of Indian cities. Transp Res Part D Transp Environ 44:277–291 73. Fishman T, Schandl H, Tanikawa H (2016) Stochastic analysis and forecasts of the patterns of speed, acceleration, and levels of material stock accumulation in society. Environ Sci Technol 50:3729–3737 74. Greenfield A, Graedel TE (2013) The omnivorous diet of modern technology. Resour Conserv Recycl 74:1–7 75. Du X, Graedel TE (2011) Global rare earth in-use stocks in NdFeB Permanent Magnets. J Ind Ecol 15:836–843 76. Helbig C, Bradshaw AM, Kolotzek C, Thorenz A, Tuma A (2016) Supply risks associated with CdTe and CIGS thin-film photovoltaics. Appl Energy 178:422–433 77. Busch J, Dawson D, Roelich K (2017) Closing the low carbon material loop using a dynamic whole system approach. J Clean Prod 149:751–761 78. Erdmann L, Graedel TE (2011) Criticality of non-fuel minerals: a review of major approaches and analyses. Environ Sci Technol 45:7620–7630 79. Roelich K, Dawson DA, Purnell P, Knoeri C, Revell R, Busch J, Steinberger JK (2014) Assessing the dynamic material criticality of infrastructure transitions: a case of low carbon electricity. Appl Energy 123:378–386 80. Xylia M, Leduc S, Patrizio P, Kraxner F, Silveira S (2017) Locating charging infrastructure for electric buses in Stockholm. Transp Res Part C Emerg Technol 78:183–200 81. Alwesabi Y, Liu Z, Kwon S, Wang Y (2021) A novel integration of scheduling and dynamic wireless charging planning models of battery electric buses. Energy 230:120806 82. Wang Y, Huang Y, Xu J, Barclay N (2017) Optimal recharging scheduling for urban electric buses: a case study in Davis. Transp Res Part E Logist Transp Rev 100:115–132 83. Soylu S (2015) Development of PN emission factors for the real world urban driving conditions of a hybrid city bus. Appl Energy 138:488–495

78

5 Challenges of Implementing Electric and Hydrogen Public Transport

84. Mahmoud M, Garnett R, Ferguson M, Kanaroglou P (2016) Electric buses: a review of alternative powertrains. Renew Sustain Energy Rev 62:673–684 85. Correa G, Muñoz P, Falaguerra T, Rodriguez CRR (2017) Performance comparison of conventional, hybrid, hydrogen and electric urban buses using well-to-wheel analysis. Energy 141:537–549 86. Li J-Q (2016) Battery-electric transit bus developments and operations: a review. Int J Sustain Transp 10:157–169 87. Chen Z, Yin Y, Song Z (2018) A cost-competitiveness analysis of charging infrastructure for electric bus operations. Transp Res Part C Emerg Technol 93:351–366 88. An K (2020) Battery electric bus infrastructure planning under demand uncertainty. Transp Res Part C Emerg Technol 111:572–587 89. An K, Jing W, Kim I (2020) Battery-swapping facility planning for electric buses with local charging systems. Int J Sustain Transp 14:489–502 90. You P, Low SH, Tushar W, Geng G, Yuen C, Yang Z, Sun Y (2018) Scheduling of EV battery swapping—Part I: centralized solution. IEEE Trans Control Netw Syst 5:1887–1897 91. You P, Yang Z, Zhang Y, Low SH, Sun Y (2016) Optimal charging schedule for a battery switching station serving electric buses. IEEE Trans Power Syst 31:3473–3483 92. Choi W, Kim J (2014) Electrification of public transportation: battery swappable smart electric bus with battery swapping station. In: 2014 IEEE Conference on Expo Transp. Electrif. Asia-Pacific (ITEC Asia-Pacific). pp 1–8 93. Dai Q, Cai T, Duan S, Zhao F (2014) Stochastic modeling and forecasting of load demand for electric bus battery-swap station. IEEE Trans Power Deliv 29:1909–1917 94. Kang Q, Wang J, Zhou M, Ammari AC (2016) Centralized charging strategy and scheduling algorithm for electric vehicles under a battery swapping scenario. IEEE Trans Intell Transp Syst 17:659–669 95. Chen Z, Liu W, Yin Y (2017) Deployment of stationary and dynamic charging infrastructure for electric vehicles along traffic corridors. Transp Res Part C Emerg Technol 77:185–206 96. Riemann R, Wang DZW, Busch F (2015) Optimal location of wireless charging facilities for electric vehicles: flow-capturing location model with stochastic user equilibrium. Transp Res Part C Emerg Technol 58:1–12 97. Ding H, Hu Z, Song Y (2015) Value of the energy storage system in an electric bus fastcharging station. Appl Energy 157:630–639 98. Prohaska R, Kelly K, Eudy L (2016) Fast charge battery electric transit bus in-use fleet evaluation. In: 2016 IEEE transportation electrification conference & expo, pp 1–6 99. Lajunen A (2018) Lifecycle costs and charging requirements of electric buses with different charging methods. J Clean Prod 172:56–67 100. Gao Z, Lin Z, LaClair TJ, Liu C, Li J-M, Birky AK, Ward J (2017) Battery capacity and recharging needs for electric buses in city transit service. Energy 122:588–600 101. Ranta M, Karvonen V, Potter JJ, Pasonen R, Pursiheimo E, Halmeaho T, Ponomarev P, Pihlatie M (2016) Method including power grid model and route simulation to aid planning and operation of an electric bus fleet. In: 2016 IEEE vehicle power and propulsion conference, pp 1–5 102. Lajunen A, Kalttonen A (2015) Investigation of thermal energy losses in the powertrain of an electric city bus. In: 2015 IEEE transportation electrification conference & Expo, pp 1–6 103. Jaguemont J, Boulon L, Dubé Y (2016) A comprehensive review of lithium-ion batteries used in hybrid and electric vehicles at cold temperatures. Appl Energy 164:99–114 104. Shirazi Y, Carr E, Knapp L (2015) A cost-benefit analysis of alternatively fueled buses with special considerations for V2G technology. Energy Policy 87:591–603 105. Farag HEZ, Al-Obaidi A, Khani H, El-Taweel N, El-Saadany EF, Zeineldin HH (2020) Optimal operation management of distributed and centralized electrolysis-based hydrogen generation and storage systems. Electr Power Syst Res 187:106476 106. Shabani MJ, Moghaddas-Tafreshi SM (2020) Fully-decentralized coordination for simultaneous hydrogen, power, and heat interaction in a multi-carrier-energy system considering private ownership. Electr Power Syst Res 180:106099

References

79

107. Thomas CE (2009) Fuel cell and battery electric vehicles compared. Int J Hydrogen Energy 34:6005–6020 108. Weeda M, de Wilde H, Wurster L--R, Bünger U, Schaap S-G, Wallmark C (2014) Towards a comprehensive hydrogen infrastructure for fuel cell electric cars in view of EU GHG reduction targets. Hydrog. Infrastruct. Transp. (HIT), Eur. Union Brussels, Belgium 109. Hardman S, Steinberger-Wilckens R, van der Horst D (2013) Disruptive innovations: the case for hydrogen fuel cells and battery electric vehicles. Int J Hydrogen Energy 38:15438–15451 110. Honselaar M, Pasaoglu G, Martens A (2018) Hydrogen refuelling stations in the Netherlands: an intercomparison of quantitative risk assessments used for permitting. Int J Hydrogen Energy 43:12278–12294 111. Xiao W, Cheng Y, Lee W-J, Chen V, Charoensri S (2011) Hydrogen filling station design for fuel cell vehicles. IEEE Trans Ind Appl 47:245–251 112. Ahluwalia RK, Peng JK, Roh HS, Hua TQ, Houchins C, James BD (2018) Supercritical cryocompressed hydrogen storage for fuel cell electric buses. Int J Hydrogen Energy 43:10215– 10231 113. Langford BC, Cherry C (2012) Transitioning a bus transit fleet to hydrogen fuel: a case study of Knoxville Area Transit. Int J Hydrogen Energy 37:2635–2643 114. Reuß M, Grube T, Robinius M, Preuster P, Wasserscheid P, Stolten D (2017) Seasonal storage and alternative carriers: a flexible hydrogen supply chain model. Appl Energy 200:290–302 115. Deliali A, Chhan D, Oliver J, Sayess R, Pollitt KJG, Christofa E (2021) Transitioning to zero emission bus fleets: state of practice of implementations in the United States. Transp Rev 41:164–191 116. Brecher A et al (2012) Transit bus applications of lithium ion batteries: Progress and Prospects 117. Hanlin J, Reddaway D, Lane J (2018) Battery electric buses—state of the practice. TCRP Synth. Transit Pract 118. Lajunen A (2014) Energy consumption and cost-benefit analysis of hybrid and electric city buses. Transp Res Part C Emerg Technol 38:1–15 119. Pihlatie M, Kukkonen S, Halmeaho T, Karvonen V, Nylund N-O (2014) Fully electric city buses—The viable option. In: 2014 IEEE international electric vehicle conference, pp 1–8 120. California Air Resources Board (ARB) (2016) Advanced clean transit battery cost for heavy duty electric vehicles 121. Wang N, Li Y, Liu Y (2014) Economic evaluation of electric bus charging infrastructure. In: 17th IEEE international conference on intelligent transportation systems, pp 2799–2804 122. Federici M, Ulgiati S, Basosi R (2008) A thermodynamic, environmental and material flow analysis of the Italian highway and railway transport systems. Energy 33:760–775 123. Cipek M, Pavkovi´c D, Kljai´c Z, Mlinari´c TJ (2019) Assessment of battery-hybrid dieselelectric locomotive fuel savings and emission reduction potentials based on a realistic mountainous rail route. Energy 173:1154–1171 124. Kalghatgi G (2018) Is it really the end of internal combustion engines and petroleum in transport? Appl Energy 225:965–974 125. Popovich ND, Rajagopal D, Tasar E, Phadke A (2021) Economic, environmental and gridresilience benefits of converting diesel trains to battery-electric. Nat Energy 6:1017–1025 126. Eudy L, Post M, Eudy L, Post M (2018) Fuel cell buses in U.S. Transit fleets: current status 2018 fuel cell buses in U. S. Transit fleet: current status 2018. 127. Bubna P, Brunner D, Gangloff JJ, Advani SG, Prasad AK (2010) Analysis, operation and maintenance of a fuel cell/battery series-hybrid bus for urban transit applications. J Power Sources 195:3939–3949 128. Mohammedi M, Kraa O, Becherif M, Aboubou A, Ayad MY, Bahri M (2014) Fuzzy logic and passivity-based controller applied to electric vehicle using fuel cell and supercapacitors hybrid source. Energy Procedia 50:619–626 129. Kim H, Eom M, Kim B-I (2020) Development of strategic hydrogen refueling station deployment plan for Korea. Int J Hydrogen Energy 45:19900–19911 130. Herwartz S, Pagenkopf J, Streuling C (2021) Sector coupling potential of wind-based hydrogen production and fuel cell train operation in regional rail transport in Berlin and Brandenburg. Int J Hydrogen Energy 46:29597–29615

80

5 Challenges of Implementing Electric and Hydrogen Public Transport

131. Piraino F, Genovese M, Fragiacomo P (2021) Towards a new mobility concept for regional trains and hydrogen infrastructure. Energy Convers Manag 228:113650 132. Zhang C, Greenblatt JB, Wei M, Eichman J, Saxena S, Muratori M, Guerra OJ (2020) Flexible grid-based electrolysis hydrogen production for fuel cell vehicles reduces costs and greenhouse gas emissions. Appl Energy 278:115651 133. Bellamy R, Lezaun J, Palmer J (2019) Perceptions of bioenergy with carbon capture and storage in different policy scenarios. Nat Commun 10:743 134. de Queiroz Fernandes Araújo O, de Medeiros JL (2017) Carbon capture and storage technologies: present scenario and drivers of innovation. Curr Opin Chem Eng 17:22–34 135. Haszeldine RS, Flude S, Johnson G, Scott V (2018) Negative emissions technologies and carbon capture and storage to achieve the Paris Agreement commitments. Philos Trans R Soc A Math Phys Eng Sci 376:20160447 136. Kalantari Z, Khoshkar S, Falk H, Cvetkovic V, Mörtberg U (2017) Accessibility of waterrelated cultural ecosystem services through public transport—a model for planning support in the Stockholm region. Sustain. https://doi.org/10.3390/su9030346 137. Reuß M, Grube T, Robinius M, Stolten D (2019) A hydrogen supply chain with spatial resolution: comparative analysis of infrastructure technologies in Germany. Appl Energy 247:438–453 138. Mori D, Hirose K (2009) Recent challenges of hydrogen storage technologies for fuel cell vehicles. Int J Hydrogen Energy 34:4569–4574 139. Moreno C, Allam Z, Chabaud D, Gall C, Pratlong F (2021) Introducing the “15-Minute City”: sustainability, resilience and place identity in future post-pandemic cities. Smart Cities 4:93–111

Chapter 6

Low Carbon Public Transport and the Competition with Aviation

Abstract Transitioning away from aviation to reduce emissions remains challenging, especially due to the demographics of current aviation users. This is because although only 11% of the global population have travelled by air, frequent flyers from more economically developed countries contribute significant levels of emissions through international and national travel. Although there are already policies in place to reduce emissions including the introduction of voluntary offsetting and the emission trading schemes, these do not appear efficient enough to reduce emissions if net zero targets are to be met. Therefore, introducing more drastic measures such as banning short-haul flights is more likely to see alternative transport modes such as high-speed rail (HSR) becoming a more viable and environmentally friendly option. However, ensuring the alternative transport options are in place before making these changes will also be necessary to ensure this transition towards low carbon alternative transport options.

6.1 Introduction Although road transportation remains the largest greenhouse gas (GHG) emitting transport type, aviation is rapidly expanding and is one of the most energy intensive forms of transport [1, 2]. Global aviation fuel use and carbon dioxide (CO2 ) emissions have increased in the past four decades with large growth occurring in Asia and other developing countries due to the rapid expansion of civil aviation. Although this pattern of growth is expected to be maintained, there remains uncertainty within this expectation due to the slowdown in aviation operations in 2020 and 2021 due to the nove coronavirus (COVID-19), or SARS-CoV-2, pandemic [3]. Aviation is currently responsible for ~10.6% of emissions within the transport sector, which is growing at an exponential rate of 5% annually, doubling in size every twenty years [4, 5]. Air travel is considered the most environmentally damaging form of transport in respect to climate change as emissions from aviation are considered more harmful than those from road transport [2]. For example, in 2018, global aviation has been estimated to account for ~2.4% of anthropogenic emissions of CO2 © The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_6

81

82

6 Low Carbon Public Transport and the Competition with Aviation

including land use changes [6]. Although new generations of aircraft have a significantly lower fuel consumption than previous generations, aviation’s contribution to global emissions has been projected to rise to 22% by 2050 [7]. Therefore, new solutions to reduce emissions from aviation need to be considered as this sector is expected to increase. Decarbonising the aviation sector has received significant interest by airlines, airplane manufacturers and policy makers [8]. However, electrifying aviation may only be achieved for short distances (i.e., a short-haul flight is typically less than 3,000 miles or has a flight time of less than 6 hours) in the medium term, with liquid fuels, such as biofuels, likely to dominate the industry beyond 2030 for long haul flights [9, 10]. However, this remains challenging as due to the current safety standards and compatibility within the aircraft fleets only sustainable aviation fuels with excellent performance within jet engines are approved [11, 12]. Reducing aviation use in favour of low emission alternatives where possible, such as high-speed rail (HSR), would allow a decrease in GHG emissions. This chapter will discuss current aviation policy and which countries are the largest GHG emitters (Sect. 6.3), policies to reduce GHG emissions from aviation including: taxes, emission trading schemes and incentives to encourage low carbon public transport alternatives (Sect. 6.4), before drawing key conclusions for aviation (Sect. 6.5).

6.2 Aviation Emission Policy Both the 1997 Kyoto Protocol and the 2015 Paris Agreement considered averaged per capita emissions for Annex 1 countries (i.e., countries that were considered industrialised countries and were members of the OECD (Organisation for Economic Co-operation and Development) in 1992), with the expectation that high-emitting countries (on a per capita basis) would make a ‘fair and ambitious’ contribution to emission reductions [2]. However, this has not been the case as there remains no specific policy to encourage countries to reduce their aviation emissions. Both the Kyoto Protocol and the Paris Agreement focus on the reduction of CO2 emissions and omit aviation contribution to radiative forcing from short-lived emissions, such as nitrous oxides (NOx ) or through contrails or clouds (H2 O) [6]. Although these non-CO2 emissions are not directly comparable with GHG emissions, they do contribute to global warming and are expected to remain relevant in the short- and medium-term future [13]. This has led studies to conclude that emissions from aviation are warming the climate at approximately three times the rate associated with CO2 emissions from aviation alone [6]. Furthermore, the Carbon Offsetting and Reduction Scheme for International Aviation (CORSIA), initiated by the International Civil Aviation Organization (ICAO), introduced a global market-based measure programme to complement the global carbon reduction target [14]. During the first phase between 2021 and 2026, all participating nations must reduce CO2 emissions relative to the average baseline emission for 2019 and 2020, with any countries exceeding the upper limit purchasing

6.2 Aviation Emission Policy

83

an allowance [14, 15]. In the second phase, from 2027 to 2035, all ICAO member states, excluding developing countries and countries with low CO2 emission levels, must also participate in this scheme [14].

6.2.1 Responsibility for Aviation Emissions Various studies have highlighted that high GHG emitters are mostly found within the highly affluent countries [16–19]. For example, the top 10% of GHG emitting countries in the world account for 45% of the global CO2 eq, whilst the bottom 50% of GHG emitters contribute to 13% of emissions [20, 21]. Furthermore, in 2020 the top five emitting countries in the world were the USA, Luxembourg, Singapore, Saudi Arabia, and Canada where population per capita emitted more than 200 tCO2 e per person per year [20]. The world average emitters were Tanzania, Mongolia, Germany, China, and France, where individuals emitted an average of 7 tCO2 e per person per year [20]. The bottom five emitting countries were in Africa and Latin America countries (Honduras, Mozambique, Rwanda, Malawi, and Zambia) with emissions ten to twenty times below the continental average and approximately fifty times below the world average [20]. Whilst this indicates very significant differences in per capita emissions between countries, frequent movement, and in particular, access to private transport is one of the key contributors to carbon-intensive consumption [22]. National surveys have established that air travellers are disproportionately wealthy [16–18, 23, 24]. By comparing passenger numbers to population and wealth on a population of ~7,594,000,000 in 2020, the number of flights averaged over the population was the equivalent of ~0.03 per person per year in low-income countries, ~0.15 in lower middle-income countries, ~0.49 in upper middle-income countries, and ~2.02 in high income countries [2]. These results indicate that the theoretical maximum share of the population that have used aviation is ~1.63% in low-income countries, ~7.51% in lower middle-income countries, ~24.72% in upper middleincome countries, and ~100% in high income countries [2]. High income countries reached 100%, as within these countries there is the potential that each individual in the population could have participated in at least one trip [2]. These results do not consider that there is a significant share of the population in every country that does not fly. Therefore, when discussing methods to reduce emissions, taking into consideration who is mostly responsible for these emissions is an important factor. Additionally, further considerations need to be made to the ‘super emitters’. This is because these individuals are the 10% of the most frequent flyers, emitting more than half of global CO2 emissions from commercial air travel, in addition to private aircraft users who emit up to 7,500 t CO2 per year [2]. These ‘super emitters’ may contribute to global warming at a rate 225,000 times higher than the globe’s poorest individuals (0.1 tCO2 per person per year) [6]. Therefore, focusing on policies to target the ‘super emitters’ should help to dramatically decrease aviation emissions, however putting a monetary value, i.e., introducing a carbon offset fee, may not be

84

6 Low Carbon Public Transport and the Competition with Aviation

enough. Although the need to fly for international and domestic business trips and conferences may be less frequent as a result of COVID-19, due to online business meeting and conferences, this may not be enough to deter travel for super emitters for work completed in person (further discussed in Chap. 8).

6.3 Policies to Reduce Emissions from Aviation To actively encourage a transition away from aviation, Governments and policymakers have begun to introduce new policies to push and encourage more sustainable transport options. For example, replacing short-haul flights whilst making long distance terrestrial public transport a more favourable option or introducing alternative taxes to actively encourage lower carbon transport types.

6.3.1 Taxes Tax is a useful way to reduce air passenger transport fossil fuel consumption and carbon emissions [25, 26]. As the aviation sector has relatively thin profit margins, air travel companies can be influenced object to certain policies, including a carbon tax [15]. Thin margins will be further exacerbated by a reduction in demand resulting in an increase in prices for the consumer which results in aviation becoming more exclusive and only used in extremes. Therefore, some airlines will not have over capacity especially after situations like COVID-19 which means that some airlines will not survive (see Chap. 8). Almost all countries have taxes on fossil fuels for road transport, which remains an effective method for decreasing GHG emission as well as financing the public sectors [27]. However, few countries have a tax on jet fuel, with Norway and Japan being the exceptions. For example, Japan has introduced the Aviation Fuel Tax of Japan and studies have projected this will create a reduction in the CO2 emissions from planes [26]. Therefore, the introduction of a fuel tax may result in a decrease in air travel, however after a 30% reduction in aviation fuel tax in 2011, the jet fuel demand increased by 10% highlighting an increased travel demand. However, in some countries, including in the EU, a tax on jet fuel for international aviation is not permitted under current international agreements [28]. Under the EU Energy Tax Directive (2003/96/EC, Article 14.1), EU Member States are prohibited from imposing a general tax on jet fuel for all international aviation [28]. To get around this, the UK Government announced in 2010 that they wanted to replace the tax per passenger with a tax per plane [28]. This was based on the weight of the aircraft, which would give airlines stronger incentives to fly with full aircraft, reducing the emissions produced per person [28]. Despite this, under article 14.2 of the same directive, it is possible to introduce a tax on jet fuel between two or more countries if

6.3 Policies to Reduce Emissions from Aviation

85

it is agreed upon in their bilateral Aviation Service Agreement, and if the countries have reserved the rights to tax jet fuel in relation to the ICAO resolution. An alternative to fuel tax is to tax the tickets purchased, especially to reflect the demographics of the individual’s purchasing the tickets [28]. The UK introduced an Air Passenger Duty in 1994, however this was not popular with consumers as it could double the cost of flights. Many other countries introduced similar taxes for both domestic and international air travel, including Germany, Sweden, France, Norway, Austria, and South Africa [28, 29]. Similarly, it is possible to tax distancebased air passengers as there are no restrictions on international aviation. However, for countries within the EU, the taxes must be the same for all flights [29]. Currently, the Netherlands, Norway, the USA, and Sweden have introduced policies to encourage bio-jet fuel production [30]. For example, Norway, has introduced a quota obligation for biofuels for aviation which means that from 2020 onwards, at least 0.5% of all fuel sold must be advanced biofuels for both domestic and international aviation [28]. Similarly, the Swedish Government announced in 2020 its intention to introduce a GHG emission mandate for aviation fuel [30]. Their reduction level is expected to be 0.8% in 2021 and gradually increase to 27% by 2027, with most savings expected from sustainable aviation fuel [30].

6.3.2 Emission Trading Scheme (ETS) Where it is not possible to meet emission reduction targets, some countries have introduced an Emission Trading Scheme (ETS) which allows countries to mitigate their GHG emissions with flexibility, cost savings and effectiveness within the transport sector [31, 32]. An ETS scheme works through a ‘cap and trade’ system as a policy instrument for pollution quantity control coupled with a defined, tradable unit. For example, it provides a country with a permit to emit a pre-determined level of a pollutant for a pre-determined duration of time [33]. These schemes have been used within the EU and are an important measure for Governments from regional to international level to achieve and address their GHG emission reduction targets [33]. The 2008/101/EC decree, enacted in November 2008, brought the international airline business into the EU ETS [34]. This caused great controversy across the world as from January 1st, 2012, each international flight taking-off and landing in EU was to be given an emission permit [35]. However, considering this controversy, the EU suspended the emission taxes of non-EU airlines in 2012. However, the ETS defines limits for carbon from intra-European flights with more than 80% allocated to the airlines free of charge, therefore the ETS has little to no impact on the pollution abatements costs of ETS [34]. The failure of EU ETS in the aviation sector is attributed to the apprehensions of the developing economies that such policies would levy additional monetary burden on their developing aviation markets [15]. In mid-2020, the European Commission published their Roadmap for the legislative initiative aimed at amending the EU ETS regarding aviation [30]. This will serve

86

6 Low Carbon Public Transport and the Competition with Aviation

to implement the CORSIA that is consistent with the EU’s 2030 climate objectives [30]. During the period 2021–2035, and based on expected participation, the scheme is estimated to offset around 80% of the emissions above 2020 levels [30]. This proposal will be part of the broader European Green Deal [30].

6.3.3 Phasing Out Short Haul Flights Short-haul flights produce the highest level of GHG emissions per kilometre, although longer flights emit higher level of emissions in absolute terms [36]. Some studies suggest short-haul flights produce more than twice as much CO2 emissions per kilometre than long haul flights [37]. This is due to the energy intensive take off and climb phase, which is distributed over a shorter flight time for short-haul flights compared to medium and longer-haul flights. Therefore short-haul flights remain the least efficient and could be replaced the most easily by other modes of transport. Furthermore, in urban areas the average distance between passenger stations is short for example between one to two miles [38]. Therefore, trains accelerate rapidly to their maximum speed bracket (e.g., 80–100 km/h) and decrease shortly afterwards to prepare for their next stop [38]. The typical average acceleration and deceleration rates are between 1.1 m/s2 and −1.3 m /s2 respectively [38]. This in turn makes rail an ideal alternative when travelling between cities with minimal numbers of stops. Among the plethora of environmental measures to mitigate the negative impacts of air travel, a transition towards greener modalities includes replacing short haul flights with high-speed rail (HSR) [39–43]. Research has highlighted that HSR can have dramatic effects on air travel demand, for example on the Paris-Nantes route, the introduction of the TGV network decreased air traffic by 30% [44]. Furthermore, the HSR link between Madrid and Seville shifted the air/rail passenger split from 67–33 to 16–84%, demonstrating that if implemented, HSR could be competitive with aviation [45]. This is because HSR has the potential to compete with aviation on distances up to 200 km, and in some cases up to 2,000 km [46–50]. The urgent need for airline aid and recovery packages because of the COVID19 pandemic (discussed further in Chap. 8) has provided an opportunity for policymakers to transition towards these low carbon alternatives. Several initiatives have been recently undertaken in different European countries. For example, French lawmakers have banned short-haul internal flights in France, where a train journey of less than two and a half hours could be provided as an alternative [43]. Similarly in Austria, the Austrian Government has introduced clauses into their state aid package that will result in Austrian Airlines halving their emission levels from domestic flights by 2030 by ending domestic routes already served by an HSR connection in less than 180 min [43]. As discussed in Chap. 4, if HSR is powered by renewably generated electricity, there is the potential to emit significantly less GHGs emissions per passengerkilometre than air transport [51]. For example, a Series N700 ‘Nozomi’ Shinkansen train in Japan travelling between Tokyo and Osaka consumed nearly one-eighth of

6.3 Policies to Reduce Emissions from Aviation

87

the energy consumed by a B777-200 airplane per seat. This means that the train emitted about one-twelfth of the CO2 of air transport per seat [51]. Therefore, from an environmental emissions perspective, HSR remains the more appealing transport type to decision-makers. Although HSR can provide benefits in comparison to short-haul flights, such as similar travel time and lower emissions, it also requires significant investments in time and infrastructure [36, 52], which may have a negative impact on climate change and biodiversity [53]. However, a full life cycle analysis would need to be considered to ensure a more realistic reflection on the environmental impact.

6.3.4 Airport Surface Access Strategies Recognising that how air passengers, staff and visitors access airports is a relevant component of the air transport system, airport operators have given increasing attention to the development of surface access strategies. Ideally, an airport should be accessible by a variety of modes and travellers should be provided with information to allow them to make sustainable travel choices. Budd and Ison (2021) note that London Heathrow airport’s central terminal area is the busiest bus and coach station in the UK and the airport is also connected to the London Underground and the Heathrow Express rail services [54]. As part of their Airport Masterplan Newcastle International Airport in the northeast of England has developed a surface access strategy that ‘seeks to get passengers and staff to the Airport efficiently and sustainably and with sensitivity to our community’ [55]. The plan also details how future growth can be sustainable in terms of noise levels, environmental impact, and energy consumption through encouraging the use of a metro link and a heavy rail connection.

6.4 Conclusions Reducing emissions from aviation will remain challenging. Due to the demographics of aviation users, the current proposed policies already in place, including voluntary offsetting and the ETS, appear not to be sufficient to counterbalance the growing emissions and passenger numbers from aviation. Therefore, introducing polices such as banning short-haul flights will likely see a higher reduction in users when flying is no longer an option. Before introducing these policies, ensuring a strong terrestrial public transport network is in place will be essential to ensure individuals do not transition to their own personal vehicles as an alternative r. A key example of this is in Japan and France which using the bullet trains and the TGV respectively have been able to encourage a smooth transition towards low carbon public transport for inter-city travel.

88

6 Low Carbon Public Transport and the Competition with Aviation

References 1. IPCC (2014) Climate change 2014 synthesis report summary chapter for policymakers. https:// doi.org/10.1017/CBO9781107415324 2. Gössling S, Humpe A (2020) The global scale, distribution and growth of aviation: implications for climate change. Glob Environ Chang 65:102194 3. Le Quéré C, Jackson RB, Jones MW et al (2020) Temporary reduction in daily global CO2 emissions during the COVID-19 forced confinement. Nat Clim Chang 10:647–653 4. Cohen SA, Higham JES (2011) Eyes wide shut? UK consumer perceptions on aviation climate impacts and travel decisions to New Zealand. Curr Issues Tour 14:323–335 5. Dubois G, Ceron JP (2006) Tourism/leisure greenhouse gas emissions forecasts for 2050: factors for change in France. J Sustain Tour 14:172–191 6. Lee DS, Fahey DW, Skowron A et al (2021) The contribution of global aviation to anthropogenic climate forcing for 2000 to 2018. Atmos Environ 244:117834 7. Cames M, Graichen J, Siemons A, Cook V (2015) Emission reduction targets for international aviation and shipping. Policy Dep A Econ Sci Policy Eur Parliam B-1047 Brussels 8. EASA (2016) European aviation environmental report 9. Gegg P, Budd L, Ison S (2014) The market development of aviation biofuel: Drivers and constraints. J Air Transp Manag 39:34–40 10. Deane JP, Pye S (2018) Europe’s ambition for biofuels in aviation—a strategic review of challenges and opportunities. Energy Strateg Rev 20:1–5 11. Gaspar RMP, Sousa JMM (2016) Impact of alternative fuels on the operational and environmental performance of a small turbofan engine. Energy Convers Manag 130:81–90 12. Why ESK, Ong HC, Lee HV, Gan YY, Chen W-H, Chong CT (2019) Renewable aviation fuel by advanced hydroprocessing of biomass: challenges and perspective. Energy Convers Manag 199:112015 13. Bock L, Burkhardt U (2019) Contrail cirrus radiative forcing for future air traffic. Atmos Chem Phys 19:8163–8174 14. Kito M, Nagashima F, Kagawa S, Nansai K (2020) Drivers of CO2 emissions in international aviation: the case of Japan. Environ Res Lett 15:104036 15. Rathore H, Jakhar SK (2021) Differential carbon tax policy in aviation: one stone that kills two birds? J Clean Prod 296:126479 16. Büchs M, Schnepf SV (2013) Who emits most? Associations between socio-economic factors and UK households’ home energy, transport, indirect and total CO2 emissions. Ecol Econ 90:114–123 17. Gill B, Moeller S (2018) GHG emissions and the rural-urban divide. A carbon footprint analysis based on the German official income and expenditure survey. Ecol Econ 145:160–169 18. Irfany MI, Klasen S (2016) Inequality in emissions: evidence from Indonesian household. Environ Econ Policy Stud 18:459–483 19. Ummel K (2014) Who pollutes? A household-level database of America’s greenhouse gas footprint. Cent Glob Dev Work Pap 20. Piketty T, Chancel L (2015) Carbon and inequality: from Kyoto to Paris. Trends Glob Inequal Carbon Emiss Prospect An Equitable Adapt Fund Paris Sch Econ 21. Hubacek K, Baiocchi G, Feng K, Patwardhan A (2017) Poverty eradication in a carbon constrained world. Nat Commun 8:912 22. Beaverstock J V, Faulconbridge JR (2014) Wealth segmentation and the mobilities of the super-rich. Elit mobilities 40–61 23. Carlsson-Kanyama A, Lindén A-L (1999) Travel patterns and environmental effects now and in the future: implications of differences in energy consumption among socio-economic groups. Ecol Econ 30:405–417 24. Banister D (2018) Inequality in transport 25. Qiu R, Xu J, Xie H, Zeng Z, Lv C (2020) Carbon tax incentive policy towards air passenger transport carbon emissions reduction. Transp Res Part D Transp Environ 85:102441

References

89

26. González R, Hosoda EB (2016) Environmental impact of aircraft emissions and aviation fuel tax in Japan. J Air Transp Manag 57:234–240 27. Sterner T, Coria J (2013) Policy instruments for environmental and natural resource management. Routledge 28. Larsson J, Elofsson A, Sterner T, Åkerman J (2019) International and national climate policies for aviation: a review. Clim Policy 19:787–799 29. Faber JF, Huigen T (2018) A study on aviation ticket taxes. CE Delft 30. Panoutsou C, Germer S, Karka P, Papadokostantakis S, Kroyan Y, Wojcieszyk M, Maniatis K, Marchand P, Landalv I (2021) Advanced biofuels to decarbonise European transport by 2030: markets, challenges, and policies that impact their successful market uptake. Energy Strateg Rev 34:100633 31. Jiang J, Xie D, Ye B, Shen B, Chen Z (2016) Research on China’s cap-and-trade carbon emission trading scheme: overview and outlook. Appl Energy 178:902–917 32. Tang L, Wang H, Li L, Yang K, Mi Z (2020) Quantitative models in emission trading system research: a literature review. Renew Sustain Energy Rev 132:110052 33. Zelljadt E, Mehling M (2021) Emission trading systems in transportation. In: Brewer T (ed) Transportation air pollutants: black carbon and other emissions. Springer International Publishing, Cham, pp 87–96 34. Cui Q, Li Y, Wei Y-M (2017) Exploring the impacts of EU ETS on the pollution abatement costs of European airlines: an application of network environmental production function. Transp Policy 60:131–142 35. Anger A, Köhler J (2010) Including aviation emissions in the EU ETS: much ado about nothing? A review. Transp Policy 17:38–46 36. Baumeister S, Leung A (2021) The emissions reduction potential of substituting short-haul flights with non-high-speed rail (NHSR): the case of Finland. Case Stud Transp Policy 9:40–50 37. Grimme W, Jung M (2018) Towards more sustainability?—The development of aviation emissions from Germany between 1995 and 2016. In: Proceedings of 22nd Air Transp Res Soc World Conf 38. Khodaparastan M, Mohamed AA, Brandauer W (2019) Recuperation of regenerative braking energy in electric rail transit systems. IEEE Trans Intell Transp Syst 20:2831–2847 39. Danapour M, Nickkar A, Jeihani M, Khaksar H (2018) Competition between high-speed rail and air transport in Iran: the case of Tehran-Isfahan. Case Stud Transp Policy 6:456–461 40. D’Alfonso T, Jiang C, Bracaglia V (2016) Air transport and high-speed rail competition: environmental implications and mitigation strategies. Transp Res Part A Policy Pract 92:261– 276 41. Robertson S (2016) The potential mitigation of CO2 emissions via modal substitution of highspeed rail for short-haul air travel from a life cycle perspective—an Australian case study. Transp Res Part D Transp Environ 46:365–380 42. Zhang A, Wan Y, Yang H (2019) Impacts of high-speed rail on airlines, airports, and regional economies: a survey of recent research. Transp Policy 81:A1–A19 43. Avogadro N, Cattaneo M, Paleari S, Redondi R (2021) Replacing short-medium haul intraEuropean flights with high-speed rail: Impact on CO2 emissions and regional accessibility. Transp Policy 114:25–39 44. Dobruszkes F (2011) High-speed rail and air transport competition in Western Europe: a supplyoriented perspective. Transp Policy 18:870–879 45. Barrón I, Campos J, Gagnepain P, Nash C, Ulied A, Vickerman R (2009) Economic analysis of high-speed rail in Europe. Spain 46. Button K (2012) Is there any economic justification for high-speed railways in the United States? J Transp Geogr 22 47. Chen Z (2017) Impacts of high-speed rail on domestic air transportation in China. J Transp Geogr 62:184–196 48. Dalla Chiara B, De Franco D, Coviello N, Pastrone D (2017) Comparative specific energy consumption between air transport and high-speed rail transport: a practical assessment. Transp Res Part D Transp Environ 52:227–243

90

6 Low Carbon Public Transport and the Competition with Aviation

49. Sun X, Zhang Y, Wandelt S (2017) Air transport versus high-speed rail: an overview and research agenda. J Adv Transp 2017:8426926 50. Prussi M, Lonza L (2018) Passenger aviation and high-speed rail: a comparison of emissions profiles on selected European routes. J Adv Transp 2018:6205714 51. Yu K, Strauss J, Liu S, Li H, Kuang X, Wu J (2021) Effects of railway speed on aviation demand and CO2 emissions in China. Transp Res Part D Transp Environ 94:102772 52. Bukovac S, Douglas I (2019) The potential impact of high-speed rail development on Australian aviation. J Air Transp Manag 78:164–174 53. Cornet Y, Dudley G, Banister D (2018) High-speed rail: implications for carbon emissions and biodiversity. Case Stud Transp Policy 6:376–390 54. Budd L, Ison S (2021) Public transport integration. In: Mulley C, Nelson JD, Ison S (eds) Handbook of public transport. Routledge, Abingdon, pp 72–81 55. Newcastle International (2021) Masterplan 2035 Summary. https://www.newcastleairport.com/ about-your-airport/masterplan/masterplan-2035-summary/. Accessed 10 Oct 2021

Chapter 7

Freight

Abstract This chapter discusses fuel alternatives for shipping and road freight as a method to reduce greenhouse gas (GHG) emissions and meet emission targets. Transitioning away from diesel fuelled transport will be an important step, as this is currently the main transport fuel for both transport types and an important barrier to overcome. However, most alternative fuels for both shipping and road transport including hydrogen, ammonia, and methanol, which are new technologies, face numerous challenges including infrastructure and storage, and have not yet been implemented on a large scale. Other alternatives including unmanned aerial vehicles are also discussed briefly, however as this is such a new technology, regulations and legislation for their implementation is not yet comprehensive enough to ensure its viability. Without significant changes in how freight is moved, emission reduction objectives will not be met as current projections foresee a dramatic increase in emission levels from freight transport by 2050.

7.1 Emissions from Freight This chapter focuses on alternatives to the current freight shipping and road freight trends. The International Transport Forum’s (ITF) International Freight Model foresees an increase of trade-related freight transport emissions by a factor of 3.9 to 2050, which would see an increase in CO2 emissions of ~2,108Mt in 2010 to ~8,132Mt by 2050 [1]. This almost fourfold increase in emission levels would undermine any attempt at meeting Paris Agreement targets, therefore adjusting current freight transport practices and policies to ensure they are aligned with climate mitigation is essential. Urban freight transport has a large impact on cities, not only on traffic and congestion but also on-air pollution and the use of urban space. Previous research has attempted to minimise this impact [2–5]. Section 7.2 discusses emissions from freight shipping and considers alternative fuels including hydrogen, ammonia, and methane. In this context, there are several different types of ships operating within the international shipping market, however their classifications are less formal. Cargo ships, or freighters, have been designed to © The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_7

91

92

7 Freight

transport bulk cargo, e.g., shipping containers, bulk commodities, oil, personal vehicles etc. which results in their capabilities and size being matched to their intended markets or intended travel routes [6]. Section 7.3 discusses the emissions produced and considers alternative fuels for freight trucks. In this context, freight trucks are often defined by cargo weight or axle loading, with different regimes used in different regions including the USA, Europe, China, and Japan. Furthermore, the smallest truck class is often ~3.5 tonnes. Section 7.4 discusses alternatives for both freight shipping and trucks using unmanned aerial vehicles (UAVs), also known as a drone which is unmanned and controlled by a ground-based controller. Although this has potential to reduce traffic congestion, there remain several challenges that will need to be addressed before implementation. Section 7.5 discusses the key conclusions from this chapter. This chapter does not discuss freight trains as most of the conclusions and alternatives considered for freight trains, i.e., electric and hydrogen alternatives, drawn in Chap. 4 for passenger rail are relevant in this context. This is because improvements/extensions of the electrification of passenger services will increase opportunities for electrified rail freight.

7.2 Freight Shipping Emissions within the shipping industry have been recognised as an important source of air pollution in port cities and inland river regions and are responsible for 2.4% of total global emissions [7, 8]. This is of particular importance due to the large amounts of different air pollutants including sulphur oxides (SOx ), nitrogen oxides (NOx ) and particulate matters (PM) from both operating and berthing in ports [9–11]. Furthermore, large ship engines tend to burn heavy oil bunker fuel when at sea which has a very high sulphur content. Under EU regulations, these ships have to switch to diesel (lower sulphur) in European and UK waters and ports. As the increasingly strict emission controls of road vehicle policy and regulations is introduced, focus needs to be placed within the shipping industry as international maritime trade is likely to expand in the long-term [7, 12, 13]. The International Maritime Organization (IMO), the main governing body for this sector, has set a target of a 50% reduction in emissions by 2050 compared to 2008 levels [8]. The IMO’s Greenhouse Gas study released in 2014 highlighted that for global international shipping the CO2 estimate decreased from 2.8% in 2007 to 2.2% in 2012 [14]. Although this is a decrease in the level of emissions, total shipping emissions is still expected to increase between now and 2050 as expansions in freight movements occur. Therefore, emissions are expected to rise if technological and operational improvements are not made [14]. Although these improvements are expected to yield significant energy savings, business as usual scenarios of maritime CO2 emissions up to 2050 project an increase of between 50 and 250% [14]. Estimates within the EU suggest that by 2050, freight shipping will produce 17% of global GHG emissions if regulatory measures are not

7.2 Freight Shipping

93

introduced. Similarly, East Asia was responsible for ~16% of global shipping CO2 emissions in 2013, increasing from between 4 and 7% in 2002–2005 [15]. Reducing emissions from the shipping industry is particularly important, with several studies highlighting the negative health impacts that this can cause. For example, particulate related emissions were annually responsible for ~60,000 cardiopulmonary and lung cancer related deaths [16]. Individuals most affected by this lived in coastal regions on major trade routes in Europe, East Asia, and South Asia [16]. Similarly, another studied highlighted that emissions from shipping in East Asia resulted in adverse health impacts with between 14,500 and 37,000 premature deaths per year [15]. To reduce emissions, utilising detailed emission inventories from the shipping industry is crucial to better understand the total emissions and to project total emissions forward [17]. Earlier inventories commonly used a ‘top-down’ approach to calculate emissions based on ship fuel consumption, however this method was latterly considered less accurate when it comes to a regional scale [18]. More recently a ‘bottom-up’ method, using position tracking data from the Automatic Identification System (AIS), has been introduced which is believed to have a higher level of precision and reduce uncertainties within the shipping industry. This method calculates emissions based on high temporal resolution ship navigation data such as speed, location, routing, and duration etc. via AIS satellite [18–23]. This is particularly useful with any ship larger than 300 tonnes globally being required to report their position within a few certain intervals [17]. This has resulted in the availability of new data which has made it possible to refine different methods that can significantly improve the quality of ‘bottom-up’ ship emission inventories [17]. This approach has now been widely accepted in Europe, America, Australia, the Arctic, as well as other countries and regions [24–26]. There are several possible methods for reducing the emissions within the shipping industry. Examples include slow steaming i.e., reducing the speed of the ship to save fuel or by introducing energy saving technologies such as using waste heat from the engines for electricity generation, heating, and air conditioning, or rerouting for milder weather conditions. Although these measures will help reduce emissions, they will not be significant enough to reduce to achieve net zero. Transitioning away from heavy fuel oil (HFO) towards more sustainable fuel types for propulsion has been identified as the most effective method to reduce emissions, with capabilities of up to 14% reduction [27]. Several alternative methods for the shipping industry have been considered as they are expected to provide lower levels of emissions.

7.2.1 Alternatives to Conventionally Fuelled Shipping Currently, the most common fuels used within the shipping industry are HFO, marine diesel oil and liquefied natural gas (LNG), with HFO being the most economical for

94

7 Freight

long range shipping [8]. However, the shipping industry faces many unique challenges and pathways towards a net zero emission reduction, especially as ships typically have a relatively long lifespan (over 20 years). This means that solutions to decrease emissions will need to be introduced within the next 5–10 years if net zero by 2050 is to be met. Due to the power demand of long-range ships, successful solutions that have already been introduced for other transport methods i.e., batteries for electric cars, may not be a viable alternative within the shipping industry [8]. Therefore, early identification of future solutions will be crucial to guide investment and policy, saving both money and time.

7.2.1.1

Hydrogen as a Fuel Type

Hydrogen as a fuel source is currently receiving significant attention as a potential energy carrier by many countries as it has no carbon which means it is an efficient fuel with no CO2 emissions, except for any related to its production [28]. Hydrogen can be cleanly generated by means of an electrolyser technology or recovered from the industrial hydrogen-rich waste streams using various separating techniques [29–32]. Therefore, in its pure form, hydrogen has potential to be a zero-emission transport option within the shipping industry. This is because, the only by-product of approach of a Proton Exchange Membrane (PEM) fuel cell is water. However, several challenges need to be overcome before widespread integration. Firstly, the scalability. Currently, a few small hydrogen powered ships have been developed and introduced with small energy demands, e.g., the Hamburg Ferry [8]. However, this would need to be scaled up to meet the energy demand of larger vessels. As most current ships use an internal combustion engine (ICE), there is potential to retrofit them with hydrogen, however significant modifications would be required due to the different burning rates of hydrogen and the current fuels used. New ships would have to be designed with fuel cells and electrical propulsion motors. Furthermore, generating hydrogen is also problematic. This is because less than 1% of hydrogen is readily available as a naturally occurring gas, with many methods to produce hydrogen using fossil fuels and emitting emissions [33, 34]. Electrolysis is a method of producing hydrogen using water and electricity and could be considered zero emission. Although technology has developed significantly in recent years, this method does not allow hydrogen to be produced at the scale required to supply the shipping fleet. Some ships can support on-board electrolysis which could eliminate this issue as it would ease concerns of the security of the supply and less infrastructure would be required. Furthermore, if the ship (either a ship with sails or through rotating cylinders) itself could produce electricity on-board (i.e., via wind or solar), this could help in production. However, this is still a relatively new concept and would also need to overcome several challenges to be feasible. Secondly, health and safety would need to be a major consideration. For example, in the EU all fuels must comply with EU regulations, such as keeping containers in well-aired locations and away from ignition sources [8]. This is because

7.2 Freight Shipping

95

hydrogen has a high flammability range between 4 and 77% in air [35], and therefore if not handled correctly could be considered explosive [36]. Thirdly, storage. As hydrogen has a critical temperature of ~33 K (−240 °C), hydrogen is gaseous at the ambient temperature [8, 37]. Furthermore, the gravitational energy density is high, whilst the volumetric energy density is low. Therefore, unless the latter can be increased, the volume required to meet energy demand would be far too vast. To reduce this gap, storing hydrogen within a high-pressure container (with 700 bars typically considered the maximum practical pressure), the energy density can be increased. However, this does require additional infrastructure to maintain the pressure, including complex structural considerations, as well as using H2 energy density at 700 bar a tank to power a Panamax ship ~15,000 high pressure turning for 3–4 weeks would require a significant proportion the ships volume [8, 38]. This would mean the transportation of goods would be significantly more difficult. Storing hydrogen as a liquid could result in the energy density being increased further, however this requires a constant temperature of between 13.8 and 33.2 K which would result in a significant energy cost [8]. When storing hydrogen at this state, there is a level of ‘boil off’ where liquids begin to evaporate. This will require additional stages to manage, for example, directing the ‘boil off’ to the engine to be used for propulsion, which has been demonstrated to be efficient in liquified natural gas tankers (however there is less control of the rate at which the fuel is consumed). An alternative is re-liquefying the gas, which has more control over the additional size of infrastructure, however there is an energy cost to do this. Another alternative storage method to increase energy density is by absorbing hydrogen into metals via hydrogen bonding, also known as ‘metal hydrides’ [8]. This process has potential to increase the weight requirements. Finally, transportation of hydrogen. As highlighted above there are many factors when considering the best way to store hydrogen and this is important to consider within the transportation of the hydrogen phase. Currently, hydrogen can be transported in liquid form in trucks, large containers, and ships and in gaseous form via pipelines or trucks [39].

7.2.1.2

Ammonia as a Fuel Type

Recently, the idea of utilising ammonia as a maritime fuel has been becoming more popular as ammonia has no carbon content and would produce new carbon emissions at the point of use [8, 40], however, the release of NOx is possible due to the combustion of ammonia. Similarly to hydrogen, ammonia as a fuel source has not been utilised on such a large scale and will face many challenges before integration. As ammonia is more commonly used within fertilisers, it does have a pre-existing global supply chain, which means that there are currently pre-existing global safety protocols in place. This means that increasing the scalability of this fuel source may be a more viable option, however current methods to produce ammonia mainly use fossil fuels, however Norway produces a lot of ammonia and fertiliser using its hydropower.

96

7 Freight

By fitting post-combustion devices, such as catalytic converters, NOx emissions can be reduced, however this would be at the expense of a high energy cost and size requirement [8, 41]. A common issue with an NOx treatment process is called an ‘ammonia slip’ where certain levels of ammonia can pass through the system, causing the release of both NOx and N2 O. Ammonia is not flammable in air but there is still potential for secondary ignition fuel is required for combustion such as natural gas or hydrogen [8, 42] . Therefore, a blend of these gases could be feasible for combustion, although this would still produce NOx and would be less efficient than fuel cells [8]. A major challenge that needs to be considered when transitioning towards ammonia as a fuel source is ammonia’s high toxicity levels, with relatively small levels of exposure required for the loss of consciousness [42, 43]. Therefore, if a large volume of ammonia is stored both onshore (i.e., at different ships docks) and offshore, a further increase to existing safety protocols will need to likely occur, increasing capital expenditure and storage requirements, i.e., an additional layer of casing. Ammonia is also a corrosive material meaning the consideration of the storage materials would be required so that they do not degrade [8]. Some researchers have indicated that due to the methods to store ammonia and high hydrogen content, it could be considered a carrier of hydrogen to allow for ‘ammonia cracking’ which is the process of converting ammonia back into hydrogen. This process has been considered relatively efficient but does require high temperatures to prevent (or reduce as far as possible) the release of NOx and N2 O into the environment. Furthermore, it has been debated whether the energy costs of converting ammonia to hydrogen would be more or less than the costs of cooling or compressing into hydrogen [44]. It is important to note that whilst ammonia is not the only component that remains capable of ammonia cracking, it is the only component able to do this with a zero-carbon content [45]. Ammonia cracking is also a complex process requiring temperatures ~1000 °C [46], however some researchers believe it may be possible to perform this process at lower temperatures [46–48]. These studies have not been tested on large scale maritime shipping but could be used as fuel cells specifically for this purpose.

7.2.1.3

Methanol as a Fuel Type

Methanol has also been considered as a potential fuel type due to the similar carbon content to methane. For example, Stena successfully retrofitted a ferry, powered by methanol recycled from residual steel gases also known as ‘blue methanol’, which operated between Gothenburg, Sweden and Kiel, Germany. This is because there are several properties that are unique to methanol making it a viable alternative, for example, it can be used to feed a fuel cell directly, which although it would produce CO2 emissions, would be significantly easier to capture and store [49]. In terms of combustion, CO2 emissions at the point of use would be considered like LNG, with no nitrogen or sulphur content in methanol, meaning no SOx emissions, however NOx could be emitted due to the nitrogen content in air, which would be ~60% of HFO and significantly lower than ammonia.

7.2 Freight Shipping

97

Methanol has a high toxicity to humans; therefore, the IMO has suggested that methanol storage would require more monitoring systems than current fuel. This in turn may increase the cost and engineering challenges associated with methanol. Furthermore, methanol has a much lower boiling point than hydrogen and ammonia at ~65 C therefore storage, especially in its liquid form, is significantly easier and the refuelling time would be much quicker. However, the primary method for methanol extraction does require much warmer temperatures of between ~200-300 C, therefore due to the energy cost of conversion, a methanol fuel cell remains the most efficient method of propulsion from methanol [50]. Methanol remains one of the largest globally produced chemicals at ~85 mt per year using one of three main feed stocks: natural gas or coal, biomass, or agricultural waste [51]. The most common method is using fossil fuels; therefore, the process remains very energy intensive and can lead to high levels of associated emissions. One solution considered was to produce methanol via a supply of CO2 , hydrogen, and renewable electricity, however this process is not considered efficient due to the large thermodynamic penalties and has a significantly lower yield than methanol produced from syngas [52].

7.3 Freight Trucks Road freight represents ~7% of global energy related CO2 emissions in comparison to international shipping which represented ~2.6% of global emissions in 2015 [6, 53]. Furthermore, current truck fleets produce a high share of road traffic emissions, although the total numbers of freight trucks are relatively low [54]. For example, ~3.8 millions heavy duty trucks have been sold globally between 2011 and 2018, with most being sold for commercial use and fitted with diesel engines [54]. Similar to the freight shipping industry the main strategy to reduce emissions from freight trucks is through alternative fuels, especially in substitution of diesel oil which has one of the highest emission factors within the available fuels for heavy duty transport [55]. Furthermore, many countries have designated emission reduction standards for heavy duty trucks, including the EU, India, and the USA to ensure emissions decrease to meet net zero emission reduction targets [56]. Utilising hydrogen to substitute for diesel is a feasible way to reduce greenhouse gas emissions and pollutant emissions whilst reducing regional economic expenditure [57]. As discussed above and in previous chapters, hydrogen can be utilised as a zero emission and compared to diesel engines, which have an efficiency of less than 40% [58], the working efficiency of hydrogen fuel cells is usually ~60% [59], with the maximum of ~80% [60]. Furthermore, as discussed in Chap. 3, electric and hydrogen fuelled road transport may act as suitable alternatives. A study focusing on Latin American countries which analysed electric truck adoption in Argentina, Brazil, Chile, Colombia, and Uruguay, highlighted that similar to personal electric vehicle use, a key barrier was upfront costs (compared to the current price of diesel trucks) [61]. Also comparable to personal

98

7 Freight

electric vehicles, several key barriers would need to be addressed before widespread adoption including, the provision of charging infrastructure. Furthermore, although electric trucks would be cheaper than hydrogen alternatives, hydrogen trucks are likely to have a larger range and therefore able to travel further with less stops. As hydrogen fuelling stations are more likely to be situated in more sparsely populated areas due to the large space required for their infrastructure, hydrogen trucks may be a viable alternative. However, similar challenges will need to be addressed including upfront costs and availability of charging infrastructure etc. Furthermore, for trucks, both ammonia and methanol could be used, however similar constraints to their application in shipping may need to be addressed.

7.4 Alternatives to Land-Based Freight Movements As technology advances, new transportation technologies have been introduced to aid consumers requirements, save time and money, and allow a better delivery service whilst ensuring a company’s sales are not affected [62]. For example, in Thailand, transportation was the second largest source of CO2 emissions, and although traffic congestion was low in rural areas, the quality of the road infrastructure posed a challenge to delivery services [63]. Therefore, alternatives such as UAVs or drones have been introduced by a number of companies including Amazon, DHL, Google, UPS etc. [64]. This method has potential to be used for multiple services when online shopping, including for goods and supplies, food deliveries, health and emergency medicines, security etc. [65]. Furthermore, over 100 robotaxis have been introduced in China (in Shanghai, Shenzhen, Wuhan, and other cities) by the company AutoX [65]. Also, some studies have suggested that this online shopping method is one of the most environmentally friendly transport options [64] if the electricity used to charge them is from low emissions sources. However, further research is needed into the environmental effects of the materials required to get a true understanding of the environmental implications. Furthermore, the lack of range will limit drones to compete for short or remote distances and they would likely compete with electric light vans. Currently, drones have some limitations regarding regulation and legislation and a limited range in operations due to the battery size and energy consumption. Like electric vehicles, drones will need their batteries recharged to keep running between services [66]. Furthermore, configuration of the drone’s network, customers, and the pick-up point of the delivery all play a substantial role, however, are not usually controllable. For example, if a drone was required to pick up food from different restaurants through a delivery service and drop them off at different locations, each restaurant would need to have their own fleet of drones [66]. Furthermore, other challenges including the identification of drones and their purpose of flight from afar using wireless electronic means, provision of safety features on drones in case of a crash, flying over private property or the security of the drone (as they could be hijacked), are all concerns that need considered before widespread implementation.

7.5 Key Findings

99

7.5 Key Findings As discussed in this chapter, reduction of emissions from freight shipping and trucks will be essential to meet climate change targets as these transport types, although not high in numbers emit significant levels of emissions. As examined, the main method to reduce emissions from freight are by using alternative fuels, as diesel powered engines have historically been the dominant choice for both heavy duty trucks and shipping. Although alternative fuels have been discussed here, significant work is still needed to better understand the viability of these alternatives as many of these are not yet on a large scale. As these technologies have such a long-life expectancy, the decisions regarding alternative fuel sources, including the installation of new or modified infrastructure will need to occur soon for a higher success rate. The adopted technology will also be influenced by changes in motive power for passenger transport and its refuelling infrastructure. As technology advances, other alternatives such as UAVs have begun to be introduced into the network. However, as this also remains a relatively new technology additional teething problems will need to be worked out in terms of configuration etc. to allow for successful implementation. Furthermore, implementing alternative fuelled transport types or UAVs will require a collaborative effort as many countries do not have these alternative fuel sources available or within proximity. Therefore, supporting refuelling infrastructure close to points of interchange such as ports or within rural areas will need to be co-ordinated.

References 1. ITF-OECD (2015) The Carbon Footprint of Global Trade. Int Transp Forum Glob dialogue better Transp 12 2. Ezquerro S, Alonso B, Moura JL (2021) Sharing bus stops: an efficient use of public spaces. J Clean Prod 294:126282 3. Bandeira RAM, D’Agosto MA, Ribeiro SK, Bandeira APF, Goes GV (2018) A fuzzy multicriteria model for evaluating sustainable urban freight transportation operations. J Clean Prod 184:727–739 4. Ezquerro S, Romero JP, Moura JL, Benavente J, Ibeas Á (2018) Minimizing the impact of large freight vehicles in the city: a multicriteria vision for route planning and type of vehicles. J Adv Transp 2018:1732091 5. Hai D, Xu J, Duan Z, Chen C (2020) Effects of underground logistics system on urban freight traffic: a case study in Shanghai, China. J Clean Prod 260:121019 6. Speirs J, Balcombe P, Blomerus P et al (2020) Natural gas fuel and greenhouse gas emissions in trucks and ships. Prog Energy 2:12002 7. Chen D, Wang X, Nelson P, Li Y, Zhao N, Zhao Y, Lang J, Zhou Y, Guo X (2017) Ship emission inventory and its impact on the PM2.5 air pollution in Qingdao Port, North China. Atmos Environ 166:351–361 8. McKinlay CJ, Turnock SR, Hudson DA (2021) Route to zero emission shipping: hydrogen, ammonia or methanol? Int J Hydrogen Energy 46:28282–28297

100

7 Freight

9. Zhu M, Li KX, Lin K-C, Shi W, Yang J (2020) How can shipowners comply with the 2020 global sulphur limit economically? Transp Res Part D Transp Environ 79:102234 10. Zhu M, Li KX, Shi W, Lam JSL (2017) Incentive policy for reduction of emission from ships: a case study of China. Mar Policy 86:253–258 11. Okada A (2019) Benefit, cost, and size of an emission control area: a simulation approach for spatial relationships. Marit Policy Manag 46:565–584 12. Eyring V, Köhler HW, Lauer A, Lemper B (2005) Emissions from international shipping: 2. Impact of future technologies on scenarios until 2050. J Geophys Res Atmos. https://doi.org/ 10.1029/2004JD005620 13. Vutukuru S, Dabdub D (2008) Modeling the effects of ship emissions on coastal air quality: a case study of southern California. Atmos Environ 42:3751–3764 14. Gritsenko D (2017) Regulating GHG emissions from shipping: local, global, or polycentric approach? Mar Policy 84:130–133 15. Liu H, Fu M, Jin X, Shang Y, Shindell D, Faluvegi G, Shindell C, He K (2016) Health and climate impacts of ocean-going vessels in East Asia. Nat Clim Chang 6:1037–1041 16. Corbett JJ, Winebrake JJ, Green EH, Kasibhatla P, Eyring V, Lauer A (2007) Mortality from ship emissions: a global assessment. Environ Sci Technol 41:8512–8518 17. Johansson L, Jalkanen J-P, Kukkonen J (2017) Global assessment of shipping emissions in 2015 on a high spatial and temporal resolution. Atmos Environ 167:403–415 18. Miola A, Ciuffo B (2011) Estimating air emissions from ships: meta-analysis of modelling approaches and available data sources. Atmos Environ 45:2242–2251 19. Eyring V, Isaksen ISA, Berntsen T, Collins WJ, Corbett JJ, Endresen O, Grainger RG, Moldanova J, Schlager H, Stevenson DS (2010) Transport impacts on atmosphere and climate: shipping. Atmos Environ 44:4735–4771 20. Olesen HR, Miljøundersøgelser D (2009) Ship emissions and air pollution in Denmark: Present situation and future scenarios. Danish Environmental Protection Agency 21. Perez HM, Chang R, Billings R, Kosub TL (2009) AIS data use in marine vessel emission estimation. In: 18th Annu Int Emiss Invent Conf e17 22. Saxe H, Larsen T (2004) Air pollution from ships in three Danish ports. Atmos Environ 38:4057–4067 23. Tzannatos E (2010) Ship emissions and their externalities for the port of Piraeus–Greece. Atmos Environ 44:400–407 24. Goldsworthy L, Goldsworthy B (2015) Modelling of ship engine exhaust emissions in ports and extensive coastal waters based on terrestrial AIS data—an Australian case study. Environ Model Softw 63:45–60 25. Jalkanen J-P, Brink A, Kalli J, Pettersson H, Kukkonen J, Stipa T (2009) A modelling system for the exhaust emissions of marine traffic and its application in the Baltic Sea area. Atmos Chem Phys 9:9209–9223 26. Winther M, Christensen JH, Plejdrup MS, Ravn ES, Eriksson ÓF, Kristensen HO (2014) Emission inventories for ships in the arctic based on satellite sampled AIS data. Atmos Environ 91:1–14 27. Buckingham JE (2020) Future fuels for commercial shipping 28. Gopalakrishnan B, Khanna N, Das D (2019) Chapter 4—Dark-fermentative biohydrogen production. In: Pandey A, Mohan SV, Chang J-S, Hallenbeck PC, Larroche C (eds) Biohydrogen, 2nd edn. Elsevier, Amsterdam, pp 79–122 29. Ortiz-Imedio R, Caglayan DG, Ortiz A, Heinrichs H, Robinius M, Stolten D, Ortiz I (2021) Power-to-ships: future electricity and hydrogen demands for shipping on the Atlantic coast of Europe in 2050. Energy 228:120660 30. Hanley ES, Deane JP, Gallachóir BPÓ (2018) The role of hydrogen in low carbon energy futures—a review of existing perspectives. Renew Sustain Energy Rev 82:3027–3045 31. Carmo M, Fritz DL, Mergel J, Stolten D (2013) A comprehensive review on PEM water electrolysis. Int J Hydrogen Energy 38:4901–4934 32. Yáñez M, Ortiz A, Brunaud B, Grossmann IE, Ortiz I (2018) Contribution of upcycling surplus hydrogen to design a sustainable supply chain: the case study of Northern Spain. Appl Energy 231:777–787

References

101

33. Schlapbach L, Züttel A (2010) Hydrogen-storage materials for mobile applications. Mater Sustain Energy 265–270 34. Holladay JD, Hu J, King DL, Wang Y (2009) An overview of hydrogen production technologies. Catal Today 139:244–260 35. Goldmann A, Sauter W, Oettinger M, Kluge T, Schröder U, Seume JR, Friedrichs J, Dinkelacker F (2018) A study on electrofuels in aviation. Energies. https://doi.org/10.3390/en11020392 36. Mao X, Ying R, Yuan Y, Li F, Shen B (2021) Simulation and analysis of hydrogen leakage and explosion behaviors in various compartments on a hydrogen fuel cell ship. Int J Hydrogen Energy 46:6857–6872 37. Züttel A (2003) Materials for hydrogen storage. Mater Today 6(9):24–33 38. Rodrigue J-P, Ashar A (2016) Transshipment hubs in the New Panamax Era: The role of the Caribbean. J Transp Geogr 51:270–279 39. Sherif SA, Barbir F, Veziroglu TN (2005) Towards a hydrogen economy. Electr J 18:62–76 40. Zamfirescu C, Dincer I (2008) Using ammonia as a sustainable fuel. J Power Sources 185(1):459–465 41. Hussein NA, Valera-Medina A, Alsaegh AS (2019) Ammonia-hydrogen combustion in a swirl burner with reduction of NOx emissions. Energy Procedia 158:2305–2310 42. Klerke A, Christensen CH, Nørskov JK, Vegge T, (2008) Ammonia for hydrogen storage: challenges and opportunities. J Mater Chem 18(20):2304–2310 43. Little DJ, Smith III MR, Hamann TW (2015) Electrolysis of liquid ammonia for hydrogen generation. Energy & Environ Sci 8(9):2775–2781 44. Comotti M, Frigo S (2015) Hydrogen generation system for ammonia–hydrogen fuelled internal combustion engines. Int J Hydrogen Energy 40:10673–10686 45. Lan R, Irvine JTS, Tao S (2012) Ammonia and related chemicals as potential indirect hydrogen storage materials. Int J Hydrogen Energy 37:1482–1494 46. Lamb KE, Viano DM, Langley MJ, Hla SS, Dolan MD (2018) High-purity H2 produced from NH3 via a ruthenium-based decomposition catalyst and vanadium-based membrane. Ind Eng Chem Res 57:7811–7816 47. Zhang Z, Liguori S, Fuerst TF, Way JD, Wolden CA (2019) Efficient ammonia decomposition in a catalytic membrane reactor to enable hydrogen storage and utilization. ACS Sustain Chem Eng 7:5975–5985 48. Jo YS, Cha J, Lee CH, Jeong H, Yoon CW, Nam SW, Han J (2018) A viable membrane reactor option for sustainable hydrogen production from ammonia. J Power Sources 400:518–526 49. Joghee P, Malik JN, Pylypenko S, O’Hayre R (2015) A review on direct methanol fuel cells—in the perspective of energy and sustainability. MRS Energy Sustain 2 50. Mekhilef S, Saidur R, Safari A (2012) Comparative study of different fuel cell technologies. Renew Sustain Energy Rev 16:981–989 51. Sheldon D (2017) Methanol production-a technical history. Johnson Matthey Technol Rev 61:172–182 52. Boot-Handford ME, Abanades JC, Anthony EJ et al (2014) Carbon capture and storage update. Energy Environ Sci 7:130–189 53. Olmer N, Comer B, Roy B, Mao X, Rutherford D (2017) Greenhouse gas emissions from global shipping, 2013–2015 Detailed Methodology. Int Counc Clean Transp Washington, DC, USA, pp 1–38 54. Inkinen T, Hämäläinen E (2020) Reviewing truck logistics: solutions for achieving low emission road freight transport. Sustainability. https://doi.org/10.3390/su12176714 55. Rodrigues Teixeira AC, Borges RR, Machado PG, Mouette D, Dutra Ribeiro FN (2020) PM emissions from heavy duty trucks and their impacts on human health. Atmos Environ 241:117814 56. Mahesh S, Ramadurai G, Nagendra SMS (2019) On-board measurement of emissions from freight trucks in urban arterials: Effect of operating conditions, emission standards, and truck size. Atmos Environ 212:75–82 57. Lao J, Song H, Wang C, Zhou Y, Wang J (2021) Reducing atmospheric pollutant and greenhouse gas emissions of heavy duty trucks by substituting diesel with hydrogen in Beijing-TianjinHebei-Shandong region, China. Int J Hydrogen Energy 46:18137–18152

102

7 Freight

58. Tauzia X, Maiboom A (2013) Experimental study of an automotive Diesel engine efficiency when running under stoichiometric conditions. Appl Energy 105:116–124 59. Lohse-Busch H, Stutenberg K, Duoba M, Liu X, Elgowainy A, Wang M, Wallner T, Richard B, Christenson M (2020) Automotive fuel cell stack and system efficiency and fuel consumption based on vehicle testing on a chassis dynamometer at minus 18 °C to positive 35 °C temperatures. Int J Hydrogen Energy 45:861–872 60. Haseli Y (2018) Maximum conversion efficiency of hydrogen fuel cells. Int J Hydrogen Energy 43:9015–9021 61. Tanco M, Cat L, Garat S (2019) A break-even analysis for battery electric trucks in Latin America. J Clean Prod 228:1354–1367 62. Shrivastava K (2013) An approach of shopping in 21st century: online shopping. SIJ Trans Comput Sci Eng Appl (CSEA) 1:133–135 63. Mangmeechai A (2016) An economic input-output life cycle assessment of food transportation in Thailand. Int J Environ Stud 73:778–790 64. Koiwanit J (2018) Analysis of environmental impacts of drone delivery on an online shopping system. Adv Clim Chang Res 9:201–207 65. Loke SW, Rakotonirainy A (2021) Automated vehicles, urban robots and drones: three elements of the automated city. Autom City Internet Things Ubiquitous Artif Intell. Springer International Publishing, Cham, pp 69–108 66. Pinto R, Zambetti M, Lagorio A, Pirola F (2020) A network design model for a meal delivery service using drones. Int J Logist Res Appl 23:354–374

Chapter 8

Low Carbon Transport for a Modern Working Environment

Abstract Due to the impact of unexpected circumstances such as COVID-19, several adaptations have been made to the current working environment which makes it more conducive for less travel. For example, because of the pandemic more individuals are working from home and do not need to travel daily to and from a workplace. In addition, some workplace practices such as the ‘10-day fortnight’ have been introduced. Furthermore, due to virtual meetings and conferences there is less demand for international and domestic business flights. Although this potentially means less daily travel, this also has negative implications as individuals are more likely to choose a method of transport that’s convenient and cost effective. This often means using a personal vehicle. Public transport has been negatively impacted because of the pandemic and will require a significant behavioural change to recover and consolidate its position as a viable alternative to the personal vehicles.

8.1 Coronavirus and Emissions The global coronavirus (COVID-19), or SARS-CoV-2, has created unforeseen circumstances for individual travel and travel behaviour [1], with a sudden reduction of both greenhouse gas (GHG) emissions and air pollutants [2]. Previous studies have highlighted that human mobility and interaction patterns may directly contribute to the spread of infectious diseases during pandemics, therefore travel is often restricted [3–8]. A study in the USA found a strong correlation between reduced mobility behaviour and decreased COVID-19 case growth rates. This study indicated that behavioural changes were able to be seen days to weeks before movement restrictions were implemented. Their study highlighted how individuals desired to avoid the pandemic [9]. This has been further demonstrated during the Spanish flu in 1918 [10, 11]. To reduce the spread of COVID-19, various control and preventative measures have been introduced and legislated by governments from around March 2020. This has seen a reduction in mobility across the world [12]. Strategies labelled ‘social distancing’ were introduced to help mitigate transmission as it is spread through © The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_8

103

104

8 Low Carbon Transport for a Modern Working Environment

respiratory droplets and require a certain proximity to other individuals [13]. This included closures of places of learning with remote or online learning introduced, working from home, closure of non-essential shops and restaurants, restrictions on public gatherings, suspending public transport and restricting international travel. These policies saw a rapid decline in road traffic, air travel and public transport ridership [14, 15]. These policies to contain the virus and restrict mobility have reduced transport and energy demand across the world. By early April 2020, daily global carbon dioxide (CO2 ) emissions have decreased by 17% (11–25% for ± 1σ) compared with the mean 2019 levels, with just under half from changes in road transport [16]. At their peak, emissions in individual countries decreased by an average of 26% [16]. However, this decline has since been reversed. GHG emissions have been forecast to rise once the COVID-19 pandemic has finished, paving a ‘rebound’ affect similar to the 2008 financial crisis [17]. In this instance, COVID-19 pandemic has caused a 1% decrease in CO2 emissions, with emissions expected to increase by 5% once the pandemic is over [17]. Government actions and economic incentives post-COVID-19 will likely influence the CO2 emission path for decades through a ‘new normal’ [16]. However, this decrease in emission levels is likely to remain temporary as they do not reflect the structural changes within the economic, transport or energy systems. As governments aim to address the pandemic and mitigate economic consequences of closing society for months to try and slow spread of the virus, discussions are being held about what the ‘new normal’ will look like and how, when, and under what conditions future transport and mobility will operate.

8.2 The ‘New Working Normal’ With one of the primary reasons to travel the commute to work or education, lockdown restrictions significantly impacted individual travel behaviour. For example, in Spain, mobility to workplaces decreased by 80% compared with pre-COVID-19 trends. Furthermore, Spain reported the lowest vehicle miles travelled in Europe, with only 12% of the pre-COVID-19 miles travelled during the second week in April 2020 [18]. Similarly, in Greece, traffic volume decreased by more than 80% during the most serious period of the pandemic [19]. This highlights that the need to travel for work is no longer so widely needed as individuals, where possible, have begun working from home (WFH). A national survey in Australia found that in the early days of the pandemic the overall number of people WFH at least one day a week increased from 30 to 60% and the number WFH for five days a week increased from 7 to 30% [20]. This has been seen on a global scale during COVID-19 with online meetings as opposed to in person or in the office recommended as a method to reduce the spread of the virus [21]. This highlights a decreased need and reliance on transport and post-COVID-19 could become the ‘new normal’ way of working for many.

8.2 The ‘New Working Normal’

105

As a result of the pandemic, there have been multiple technological advances in the way individuals can WFH. This method of communication is likely to reduce the need to travel for work daily. Information and communication technologies also offers more flexibility for work and living arrangements however they can lead to fatigue and negative emotions due to the lack of interaction between co-workers [22, 23]. Furthermore, as workplaces have begun to open as cases of COVID-19 decrease, many workplaces are offering their staff the opportunity to continue to work from home, full time, if not a few days a week. Prior to COVID-19 some workplaces were offering the ‘nine-day fortnight’ which allowed individuals to work longer hours on nine days and get the tenth day as a day off. Although these schemes would mean individuals are travelling into work less days a week, it does mean they are likely to choose the most ‘convenient’ method of transport for the trip they are making. For example, if the individual has taken public transport five days a week as they have purchased a weekly bus pass for example, it may not make financial sense if the cost to drive is perceived to be lower, and they are unlikely to consider the impact on the environment. Alternatively, families who have previously owned multiple cars may no longer need them if they are able to co-ordinate working from home and going into an office. This may encourage individuals to downsize to just one vehicle or no vehicle at all. Many cities have now introduced car sharing schemes (e.g., Co-Wheels in the UK and Enterprise in the USA), which offer the opportunity to access a personal vehicle if an individual requires one urgently. However, this is not often a popular option, particularly if it is a family with young children, as the individual who takes on the family caring role often desires the freedom and flexibility of their own vehicle in case of emergencies.

8.3 Private Vehicles and Public Transport Public transport usage saw the largest decrease in ridership, with decreased commuter demand during the pandemic period. Although there remains great uncertainty on the challenges facing public transport, it is possible to forecast the short and medium terms implications of the COVID-19 pandemic for public transport [24]. Public transport can play an influential role in the transmission of a virus, particularly in highly populated cities [25]. For example, world passenger demands typically decreased by between 80 and 95% in the earlier stages of lockdowns [26]. Although public transport uptake is beginning to increase again as restrictions are eased, the number of passengers reflects about half of pre-pandemic level [26]. For example, in Poland, one study highlighted that one out of four Polish commuters will refrain from riding public transport in the future [27]. Similarly, in Canada, most commuters intended to increase their use of car or bike but decrease their share of other travel modes (especially the subway, bus, and taxi) after the stay-at-home orders were lifted [28]. This was mainly driven by the perceived additional risks of health safety,

106

8 Low Carbon Transport for a Modern Working Environment

peace of mind and travel experience [28]. Therefore, public transport operators must implement measures to minimise the risk of the virus spreading on-board as well as during the waiting time [24]. A study by Basu and Ferreira [29] reports that one in five of zero-car households in metropolitan Boston agreed that COVID-19 had enhanced their intention to purchase a car. This highlights the concern that many of those lost to public transport as a result of the virus will never return [29]. This includes implementing policies such as maintaining social distancing and mandatory facemasks when using public transport; however, it is unknown to what extent individuals follow these rules [30, 31]. For example, it has been reported that not all passengers have been using facemasks when using public transport [32]. In addition to the risk of infection [33, 34], uncertainty regarding the frequency of the service may also place a key role in public transport uptake [29]. As public transport is important for everyday travel, especially to allow essential workers to commute, it cannot be completely shut down during situations such as COVID-19. However, in Europe, the number of passengers using public transport declined by 80% in some cities as commuter demand decreased and transport operations reduced their services [14]. With a reduced number of passengers, there lies temptation of public transport operators to increase the fares and to supress the number of services to reduce the effects of the decreased public transport demands. However, this is likely to have significant negative impacts, including financial strain, on working class and poorer neighbourhoods reliant on public transport [35]. As a result, many individuals have begun changing their travel pattern to less crowded and more flexible transport types, such as private vehicles or active travel to reduce the risk of catching the virus [36, 37]. This is likely to have significant environmental ramifications which may therefore highlight the need for a rapid transition towards low carbon private vehicles until the confidence for public transport increases [3].

8.4 Active Travel During the lockdown period, active travel has been encouraged for exercise, which has the added benefit of reducing the pressure on transport systems and the road network [38]. Furthermore, many streets are being closed to cars and public transport to encourage active travel with ‘pop up’ cycling lanes being installed to allow individuals to exercise and have access to amenities [39]. In addition, these cycle lanes have involved ‘taking back’ road space (e.g., suspending parking spaces, or closing-off/narrowing redundant lanes) to create wider footpaths and/or temporary cycle lanes [40]. These measures have been introduced as policymakers have become more receptive of cycling as a method of active travel as it combines time-savings as well as exercise [41]. Early leadership came from the Global South, with Bogota amongst the first to expand its existing cycle network to alleviate pressure on public transport [39].

8.4 Active Travel

107

Furthermore, governments, such as France, consider the idea of a’15-min city’ where residents can access all their needs within 15 min of active travel i.e., either through walking or cycling [42]. With all amenities within a short distance of each other, this reduces the need and reliance of private vehicles, decreasing the overall transport emissions and has been effective in Paris as they have preserved local and small specialised shops, local pharmacies, doctors surgeries and schools.

8.5 Aviation On an international basis, as international events transition online, reducing aviation for work purposes is likely. Aviation is one of the sectors impacted the most by COVID-19, with passenger air traffic falling by 90% within Europe [14]. It also remains extremely demanding and expensive for airline companies to provide effective measures to ensure social distancing, whilst also ensuring the planes are maintained to high cleanliness standards [17]. To reduce emission levels, it is often argued that planes should not take off without a full passenger log, therefore the limited economic resilience of aviation will likely be the key factor in the decreased share for the future of the sector [43]. The pandemic is likely to heavily influence aviation for years to come, with unclear effects on airline companies and aircraft manufacturers.

8.6 Future Considerations Mitigation measures must continue to take into consideration unexpected circumstances such as the COVID-19 pandemic and what influence this may have on emissions. Although road transport emissions have decreased due to Government lockdowns, once restrictions have fully lifted individuals are more likely to travel by personal vehicle than from public transport as a method to reduce exposure. The shift towards low carbon transport will need to be implemented as soon as possible to ensure road transport emissions do not rapidly increase. Therefore, this needs to be considered when estimating future transport predictions as the total distance travelled annually per car may increase, with the distance travelled on public transport remaining the same but with less daily journeys.

References 1. Tiikkaja H, Viri R (2021) The effects of COVID-19 epidemic on public transport ridership and frequencies. A case study from Tampere, Finland. Transp Res Interdiscip Perspect 10:100348

108

8 Low Carbon Transport for a Modern Working Environment

2. Forster PM, Forster HI, Evans MJ et al (2020) Current and future global climate impacts resulting from COVID-19. Nat Clim Chang 10:913–919 3. Abdullah M, Dias C, Muley D, Shahin M (2020) Exploring the impacts of COVID-19 on travel behavior and mode preferences. Transp Res Interdiscip Perspect 8:100255 4. Peixoto PS, Marcondes D, Peixoto C, Oliva SM (2020) Modeling future spread of infections via mobile geolocation data and population dynamics. An application to COVID-19 in Brazil. PLoS ONE 15:1–23 5. Yan QL, Tang SY, Xiao YN (2018) Impact of individual behaviour change on the spread of emerging infectious diseases. Stat Med 37:948–969 6. Funk S, Salathé M, Jansen VAA (2010) Modelling the influence of human behaviour on the spread of infectious diseases: a review. J R Soc Interface 7:1247–1256 7. Muley D, Shahin M, Dias C, Abdullah M (2020) Role of transport during outbreak of infectious diseases: evidence from the past. Sustainability. https://doi.org/10.3390/su12187367 8. Cooley P, Brown S, Cajka J et al (2011) The role of subway travel in an influenza epidemic: a New York city simulation. J Urban Heal 88:982–995 9. Badr HS, Du H, Marshall M, Dong E, Squire MM, Gardner LM (2020) Association between mobility patterns and COVID-19 transmission in the USA: a mathematical modelling study. Lancet Infect Dis 20:1247–1254 10. Trilla A, Trilla G, Daer C (2008) The 1918 “Spanish Flu” in Spain. Clin Infect Dis 47:668–673 11. Ammon CE (2002) Spanish flu epidemic in 1918 in Geneva, Switzerland. Eurosurveillance 7:190–192 12. Pepe E, Bajardi P, Gauvin L, Privitera F, Lake B, Cattuto C, Tizzoni M (2020) COVID-19 outbreak response, a dataset to assess mobility changes in Italy following national lockdown. Sci Data 7:230 13. Wilder-Smith A, Freedman DO (2020) Isolation, quarantine, social distancing and community containment: pivotal role for old-style public health measures in the novel coronavirus (2019nCoV) outbreak. J Travel Med 27:1–4 14. Budd L, Ison S (2020) Responsible transport: a post-COVID agenda for transport policy and practice. Transp Res Interdiscip Perspect 6:100151 15. Vingilis E, Beirness D, Boase P et al (2020) Coronavirus disease 2019: what could be the effects on-road safety? Accid Anal Prev 144:105687 16. Le Quéré C, Jackson RB, Jones MW et al (2020) Temporary reduction in daily global CO2 emissions during the COVID-19 forced confinement. Nat Clim Chang 10:647–653 17. Tardivo A, Zanuy AC, Martín CS COVID-19 Impact on Transport: A Paper from the Railways’ Systems Research Perspective. Transp Res Rec 0:0361198121990674 18. Awad-Núñez S, Julio R, Gomez J, Moya-Gómez B, González JS (2021) Post-COVID-19 travel behaviour patterns: impact on the willingness to pay of users of public transport and shared mobility services in Spain. Eur Transp Res Rev 13:20 19. Katrakazas C, Michelaraki E, Sekadakis M, Yannis G (2020) A descriptive analysis of the effect of the COVID-19 pandemic on driving behavior and road safety. Transp Res Interdiscip Perspect 7:100186 20. Beck MJ, Hensher DA (2020) Insights into the impact of COVID-19 on household travel and activities in Australia—the early days under restrictions. Transp Policy 96:76–93 21. Zhang J, Hayashi Y, Frank LD (2021) COVID-19 and transport: Findings from a world-wide expert survey. Transp Policy 103:68–85 22. Waizenegger L, McKenna B, Cai W, Bendz T (2020) An affordance perspective of team collaboration and enforced working from home during COVID-19. Eur J Inf Syst 29:429–442 23. Matusik SF, Mickel AE (2011) Embracing or embattled by converged mobile devices? Users’ experiences with a contemporary connectivity technology. Hum Relations 64:1001–1030 24. Gutiérrez A, Miravet D, Domènech A (2020) COVID-19 and urban public transport services: emerging challenges and research agenda. Cities Heal 1–4 25. Ghosh A, Nundy S, Ghosh S, Mallick TK (2020) Study of COVID-19 pandemic in London (UK) from urban context. Cities 106:102928

References

109

26. Vickerman R (2021) Will Covid-19 put the public back in public transport? A UK perspective. Transp Policy 103:95–102 27. Przybylowski A, Stelmak S, Suchanek M (2021) Mobility behaviour in view of the impact of the COVID-19 pandemic—public transport users in Gdansk Case Study. Sustainability. https:// doi.org/10.3390/su13010364 28. Labonté-LeMoyne É, Chen S-L, Coursaris CK, Sénécal S, Léger P-M (2020) The unintended consequences of COVID-19 mitigation measures on mass transit and car use. Sustainability. https://doi.org/10.3390/su12239892 29. Basu R, Ferreira J (2021) Sustainable mobility in auto-dominated Metro Boston: challenges and opportunities post-COVID-19. Transp Policy 103:197–210 30. Cartenì A, Di Francesco L, Martino M (2020) How mobility habits influenced the spread of the COVID-19 pandemic: results from the Italian case study. Sci Total Environ 741:140489 31. Chu DK, Akl EA, Duda S et al (2020) Physical distancing, face masks, and eye protection to prevent person-to-person transmission of SARS-CoV-2 and COVID-19: a systematic review and meta-analysis. Lancet 395:1973–1987 32. Dzisi EKJ, Dei OA (2020) Adherence to social distancing and wearing of masks within public transportation during the COVID 19 pandemic. Transp Res Interdiscip Perspect 7:100191 33. Hotle S, Murray-Tuite P, Singh K (2020) Influenza risk perception and travel-related health protection behavior in the U.S.: Insights for the aftermath of the COVID-19 outbreak. Transp Res Interdiscip Perspect 5:100127 34. Shah AUM, Safri SNA, Thevadas R, Noordin NK, Rahman AA, Sekawi Z, Ideris A, Sultan MTH (2020) COVID-19 outbreak in Malaysia: actions taken by the Malaysian government. Int J Infect Dis 97:108–116 35. Hensher DA, Wei E, Beck M, Balbontin C (2021) The impact of COVID-19 on cost outlays for car and public transport commuting—the case of the Greater Sydney Metropolitan Area after three months of restrictions. Transp Policy 101:71–80 36. Koehl A (2020) Urban transport and COVID-19: challenges and prospects in low- and middleincome countries. Cities Heal 0:1–6 37. Laverty AA, Millett C, Majeed A, Vamos EP (2020) COVID-19 presents opportunities and threats to transport and health. J R Soc Med 113:251–254 38. Ali N, Abdullah M, Javid MA (2021) Accessibility-based approach: shaping travel needs in pandemic situation for planners’ perspectives. Eng J 25:15–22 39. Nurse A, Dunning R (2020) Is COVID-19 a turning point for active travel in cities? Cities Heal 0:1–3 40. Dunning RJ, Nurse A (2020) The surprising availability of cycling and walking infrastructure through COVID-19. Town Plan Rev 41. Steinbach R, Green J, Datta J, Edwards P (2011) Cycling and the city: a case study of how gendered, ethnic and class identities can shape healthy transport choices. Soc Sci Med 72:1123– 1130 42. Moreno C, Allam Z, Chabaud D, Gall C, Pratlong F (2021) Introducing the “15-minute city”: sustainability, resilience and place identity in future post-pandemic cities. Smart Cities 4:93– 111 43. Gössling S (2020) Risks, resilience, and pathways to sustainable aviation: a COVID-19 perspective. J Air Transp Manag 89:101933

Chapter 9

Policy Recommendations

Abstract This brief highlights that although electric and hydrogen transport are considered ‘zero emission’ at their point of use, their true environmental impact is determined by the electricity used to ‘fuel’ these vehicles. Therefore, as there is a global effort to meet Paris Agreement targets, priority needs to be placed on decarbonising public transport and encouraging sustainable and long-term travel. Furthermore, it is likely that a combination of both electric and hydrogen public transport options will need to be incorporated into the transport network to ensure range requirements are met. Although there are trade-offs and environmental impacts of incorporating low carbon transport, without significant changes net zero targets are not going to be met. By learning from the successes and failures of other countries who are at the forefront of technology, there is the opportunity for a collaborative advancement and reduction in greenhouse gas emissions.

9.1 Environmental Impact This brief highlights that the environmental impact of transitioning to low carbon transport personal vehicles will not be enough to meet climate change targets due to the holistic approach required and the influences of other sectors i.e., energy generation for electric and hydrogen transport. Hence, there is a strong need to transition towards low carbon public transport. Although transport brings several advantages to society, both from a personal perspective and economically, these benefits can result in undesirable outcomes including climate change, air pollution, water pollution, congestion, accidents, and noise etc. [1]. Furthermore, transport has the potential to harm human health through the particles emitted, however as these emissions are generated by transport, and impose a cost on other economic factors, they are recognised as externalities or external costs and should also be considered. As low carbon transport does not emit tailpipe emissions, impact on human health can improve. Through this transition towards low carbon transport, it remains inevitable there will be an impact on the environment through additional energy generation, and © The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 K. G. Logan et al., Transportation in a Net Zero World: Transitioning Towards Low Carbon Public Transport, Green Energy and Technology, https://doi.org/10.1007/978-3-030-96674-4_9

111

112

9 Policy Recommendations

the required infrastructure (generation source, power distribution, charging infrastructure, new and improved rails lines and from the vehicles themselves). However, even with this additional energy generation, the environmental impact of electric and hydrogen transport, particularly from low carbon public transport, remains significantly lower than that of conventionally fuelled alternatives. Reduced reliance on fossil fuels removes the environmental risks that are associated with extraction; incidents such as oil spills often have decades long consequences for the surrounding ecosystem. Further downstream, impacts of fossil fuel usage such as ocean acidification and global temperature rises causing sea ice melt are all factors that impact the environment and therefore have a value from a natural capital (NC) and ecosystem service (ES) viewpoint. This is not to say that all renewable generation methods do not have any quantifiably negative effect on the local environment, however the impacts are generally small when compared to fossil fuels [2, 3]. Although beyond the scope of this study, further research should account for where the vehicle is constructed and how it got to the user with promotion of local manufacturing and assembly a priority. Furthermore, renewable energy sources and charging stations should be installed near to where vehicles are likely to be recharged. This will reduce transmission and distribution losses, reducing the additional energy that may need to be generated. This means that whilst public transport may be more difficult to implement in rural areas, the supply of local clean energy as fuel, either directly as electricity or via hydrogen production, may be more easily implemented and have a reduced impact on ES and NC. As technology continues to improve, the total number of onshore or offshore wind turbines or solar panels will be less as they can generate more energy and will require less land and sea area, reducing the impact on ES and NC. Attention will also need to be given to avoiding stranded assets in the event that public transport embraces electric but then switches to hydrogen. Finally, guaranteeing that electric and hydrogen transport vehicles can be charged at varying times throughout the day and night may also be necessary to ensure there is reduced impact on the grid. By ensuring low emission vehicles are charged during off-peak hours, primarily during the night, there is likely to be a reduction in the increased peak generation capacity required and this will allow for a smoother transition towards low carbon transport [4]. This will also allow for a decreased reliance on fossil fuels during the transition towards low carbon energy generation. One solution to limit this is to integrate ‘packages’ which allow consumers to either purchase an electric vehicle (EV) battery outright or to pay a mileage fee and return the battery after use. This would allow companies or governments to manage charging times and reduce the potential impact on the national electric grid networks [5].

9.2 Recommendations Within the Transport Sector For all transport types, a more concrete set of policies should be implemented in detail at national level regarding how countries will successfully transition towards low carbon transport. Although this has been documented for personal vehicles in

9.2 Recommendations Within the Transport Sector

113

several publications across different countries, there remains a lot of ambiguity in terms of the infrastructures required to meet demands and actively encourage low emission transport adoption (i.e., through the maintenance of grants and potential reforms to vehicle tax) [6]. Whilst low emission transport cannot yet be considered widespread, the early adopters with financial means and desire to adopt the new technology have already done so. Even by bringing the ban of new conventionally fuelled vehicles (CFVs) (and vans and hybrids) sales forward, there is nothing in place that prevents or disincentivises individuals purchasing CFVs up until the final year of sale. This is where the lessons learnt within China could be applied, where new consumers in the personal transport market are financially encouraged to take up low emission options over conventional fuelled vehicles. For net zero to be successfully met, a transition towards low carbon public transport, away from personal vehicles is necessary. This is because emissions produced per person per kilometre travelled would be significantly less for EBs and HBs and trains over CFVs and EVs. Although this transition towards electric and hydrogen public transport has already begun in some major cities, there is limited information regarding how the rest of the world will transition and what types of public transport will be implemented and where. Furthermore, as highlighted in Chap. 5 there are significant barriers for the implementation of electric and hydrogen public transport because of both monetary and non-monetary challenges as well as public acceptance of new technologies. To successfully allow a transition towards sustainable public transport, cost to the consumer should be reassessed to ensure competitiveness for both bus and train networks. As cost remains a major driver in personal transport choice, this may also reduce the number of trips made from planes, particularly domestic flights. With the introduction of fuel duty and carbon taxes for both domestic and international flights, the carbon dioxide (CO2 ) emissions from air travel can be reduced. Introducing fuel tax has already been piloted in Japan under the Aviation Fuel Tax of Japan which resulted in a reduction in the CO2 emissions from planes [7]. The cost and convenience of public transport, compared to private alternatives, needs to remain competitive to encourage use which could also be considered for a carbon tax. Whilst analysis clearly indicates that a transition towards electric and hydrogen public transport is necessary, the environmental benefits of this transition away from conventionally fuelled transport will be negated if the energy sector does not ensure rapid decarbonisation of their electricity generation mix. As low carbon transport is integrated into the network, additional energy will need to be generated with the available information on how and where this is going to occur currently uncertain with detailed plans only in place in some sectors. This requires joined up policies, planning and legislation. Consideration of emissions from the air and shipping industry will be necessary to reduce overall transport emissions if the net zero targets are to be met. As mentioned, Japan has already successfully introduced a carbon tax for passengers choosing to fly, therefore further research into the implications of this travel demand management (TDM) measure within domestic and international travel could be done to determine

114

9 Policy Recommendations

how emissions will be affected. This process may need further analysis to gauge a better understanding of how to introduce TDM initiatives to encourage low carbon travel options. With the introduction of high-speed rail, analysing future commuter behaviour may be necessary to ensure it is fully utilised, though implementation will need to be network wide to likely have the maximum effect of encouraging people to switch to public transport. As technology advances there are options including hydrogen or ammonia powered container ships which could reduce emissions within the international shipping industry. However, it would depend on how the hydrogen or ammonia is generated to determine the environmental benefits. Furthermore, reviews regarding which countries will be responsible for emissions produced will also be necessary, i.e., with reference to the country the ship/plane is departing from or where they are travelling to. Furthermore, considerations and plans should be implemented, including mitigation measures which address unexpected circumstances such as the COVID-19 pandemic and what influence this may have on emissions, should be developed. Although road transport emissions have decreased due to lockdowns, once restrictions have lifted individuals are more likely to travel by personal vehicle than from public transport as a method to reduce exposure to the virus. The shift towards low carbon transport will need to be implemented as soon as possible to ensure road transport emissions do not rapidly increase. Therefore, this needs to be considered when estimating future transport predictions as the total distance travelled annually per car may increase, with the distance travelled by public transport remaining the same but with less daily journeys.

9.3 Conclusions This brief highlights that there remain several key challenges for transport to meet the Paris Agreement if immediate focus is not placed on decarbonising the energy sector and encouraging low carbon transport in parallel. During the transition towards low carbon transport, additional energy will need to be generated to meet demand, which will require additional land area and infrastructure, creating trade-offs that need minimised as far as possible. Although fossil fuels are expected to be phased out soon, fossil fuel power stations will still be within the generation mix, with new technologies such as carbon capture and storage and direct air capture necessary in the interim as additional larger capacity from renewables and nuclear energy is implemented. Learning from other countries who are at the forefront of technology will allow further advancement and reduce their emissions to meet national and global targets as part of the Paris Agreement. This will require two key lessons to be implemented. Firstly, to better understand how to implement a high-capacity low carbon electric or hydrogen public transport network. Secondly, ensuring peak energy demands are met with decarbonised energy generation. Without these, there will remain a struggle to meet national and international targets.

9.3 Conclusions

115

For a successful low carbon transport network, a mixture of electric and hydrogen transport options will be required for all transport types. This will reduce ‘range anxiety’ and ensure the needs of all individuals are met, i.e., rural areas are more likely to require hydrogen transport to ensure sufficient distance can be travelled. Within urban areas, individuals generally travel shorter distances and can use electric vehicles as these have a shorter, though still sufficient, range to cope with consumer needs, assuming charging facilities are available. Early integration of low carbon transport will achieve the greatest reduction in GHG emissions and cumulative emissions to stay in line with the Intergovernmental Panel on Climate Change’s carbon budgets. Fast and decisive action will be required to meet net zero emission targets from transport. Finally, although low carbon transport is going to impact NC and ES, this impact will be significantly less than what is likely to occur if the current levels of use of conventionally fuelled transport continue.

References 1. Mostert M, Caris A, Limbourg S (2017) Road and intermodal transport performance: the impact of operational costs and air pollution external costs. Res Transp Bus Manag 23:75–85 2. Borja Á, Galparsoro I, Irigoien X et al (2011) Implementation of the European Marine strategy framework directive: a methodological approach for the assessment of environmental status, from the Basque Country (Bay of Biscay). Mar Pollut Bull. https://doi.org/10.1016/j.marpolbul. 2011.03.031 3. De Dominicis M, Wolf J, O’Hara Murray R (2018) Comparative effects of climate change and tidal stream energy extraction in a Shelf Sea. J Geophys Res Ocean 123:5041–5067 4. Brown S, Pyke D, Steenhof P (2010) Electric vehicles: the role and importance of standards in an emerging market. Energy Policy 38:3797–3806 5. Andersen PH, Mathews JA, Rask M (2009) Integrating private transport into renewable energy policy: the strategy of creating intelligent recharging grids for electric vehicles. Energy Policy 37:2481–2486 6. Brand C, Anable J (2019) ‘Disruption’ and ‘continuity’ in transport energy systems: the case of the ban on new conventional fossil fuel vehicles. In: ECEEE 2019 Summer Study. Giens, near Hyeres, France, pp 1–12 7. González R, Hosoda EB (2016) Environmental impact of aircraft emissions and aviation fuel tax in Japan. J Air Transp Manag 57:234–240