Climate Change Countermeasures in Ports Toward Carbon Neutrality: Empirical Analysis and Potential New Countermeasures (Sustainable Development Goals Series) 303134393X, 9783031343933

This book presents an in-depth examination of the challenges facing the port industry in reducing CO2 emissions. Through

118 91 5MB

English Pages 134 [129] Year 2023

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Climate Change Countermeasures in Ports Toward Carbon Neutrality: Empirical Analysis and Potential New Countermeasures (Sustainable Development Goals Series)
 303134393X, 9783031343933

Table of contents :
Preface
Contents
Introduction to Climate Change Countermeasures in Ports
1 General Introduction to Climate Change Countermeasures in Ports
1.1 Introduction
1.2 Port-Related CO2 Emissions
1.3 Potential Climate Change Countermeasures in Ports Based on Previous Studies
1.4 Themes Covered in This Book Given the Current Climate Change Countermeasures at Ports
1.5 Conclusions
References
2 Relationship Between Port Governance and Climate Change Action
2.1 Introduction
2.2 Port Reform and Port Governance
2.3 Case Study: Japan
2.4 Stakeholders in Ports and Their Roles on Climate Change Actions
2.5 Conclusions
References
Countermeasures for Cargo-Handling Machinery
3 CO2 Reduction Effects by Electrification of Cargo-Handling Machinery
3.1 Introduction
3.2 Port Environmental Policies in Japan and Development of Environmental Countermeasures at Hakata Port
3.3 Overview of E-RTG and HSC
3.4 CO2 Emission Reductions by Introducing E-RTG and HSC
3.4.1 Changes in CO2 Emissions of the Entire CT Due to Climate Change Countermeasures
3.4.2 CO2 Reduction Effect of E-RTG and HSC
3.5 Conclusions
References
4 Economic Feasibility of Electrification of Cargo-Handling Machinery
4.1 Introduction
4.2 Governance Structure of Hakata Port
4.3 Cost–Benefit Analysis of E-RTG and HSC Installation
4.4 Possibility of Using Carbon Credits
4.5 Challenges from the Perspective of Economic Feasibility and the Role of Stakeholders in Promoting Countermeasures
4.6 Conclusion
References
Countermeasures in Reefer Container Areas
5 Simulation Model for Analysis of Energy Savings by Roof Shade Installations
5.1 Introduction
5.2 Design and Operational System of the RS
5.3 Methodology
5.4 Experimental Measurements of the Impact of the RSs
5.5 RS Impact Analysis Using Simulation
5.5.1 Thermal Factors
5.5.2 Governing Equation of the Simulation Model
5.5.3 Geometrical Model and Parameter Setting
5.5.4 Thermal Simulation
5.5.4.1 Effect of Dummy Containers
5.5.4.2 Heat Penetration of the Container Walls
5.5.4.3 Initial Temperature of the Container Walls
5.5.4.4 Thermal Buoyancy Between RCs
5.5.4.5 Monitoring Cell Locations
5.5.5 Verification of the Thermal Simulation Results
5.6 Conclusions
References
6 Energy-Saving Effects and Economic Feasibility of Roof Shade Installation
6.1 Introduction
6.2 Energy Savings Owing to the RS at Hakata Port
6.3 Seasonal and Regional Characteristics of Energy Savings
6.4 Economic Analysis of RS Installation
6.5 Discussion
6.6 Conclusion
Reference
New Possibilities for Climate Change Countermeasures in Ports
7 Carbon Containment and Creation of BCEs Through Beneficial Utilization of Dredged Soil Generated by Port Development Projects
7.1 Introduction
7.2 Carbon Storage Effect of Beneficial Utilization of Dredged Soil
7.3 Quantitative Effects of Organic Carbon Containment of Dredged Soils
7.4 Future Significance of Containment of Organic Carbon in the Dredged Soil and the Creation of BCE
7.5 Conclusion
References
8 Role of Ports in the Trade-Off Problem Between Circular Economy and Climate Change Action: Potential for Increased Use of Secondary Raw Materials in the Copper Industry as a Climate Change Countermeasure
8.1 Introduction
8.2 Methods
8.2.1 Scope of Analysis
8.2.2 Calculation Methods and Cases
8.2.3 Data Set
8.3 Results
8.4 Discussion
8.4.1 Potential Utilization of Carbon Credits
8.4.2 Feasibility of Carbon Neutrality in the Copper Industry Assuming Technological Innovation
8.4.3 Roles of Ports
8.5 Conclusion
References
Appendix_1

Citation preview

SDG: 13 Climate Action

Yoshihisa Sugimura

Climate Change Countermeasures in Ports Toward Carbon Neutrality Empirical Analysis and Potential New Countermeasures

Sustainable Development Goals Series

The Sustainable Development Goals Series is Springer Nature’s inaugural cross-imprint book series that addresses and supports the United Nations’ seventeen Sustainable Development Goals. The series fosters comprehensive research focused on these global targets and endeavours to address some of society’s greatest grand challenges. The SDGs are inherently multidisciplinary, and they bring people working across different fields together and working towards a common goal. In this spirit, the Sustainable Development Goals series is the first at Springer Nature to publish books under both the Springer and Palgrave Macmillan imprints, bringing the strengths of our imprints together. The Sustainable Development Goals Series is organized into eighteen subseries: one subseries based around each of the seventeen respective Sustainable Development Goals, and an eighteenth subseries, “Connecting the Goals,” which serves as a home for volumes addressing multiple goals or studying the SDGs as a whole. Each subseries is guided by an expert Subseries Advisor with years or decades of experience studying and addressing core components of their respective Goal. The SDG Series has a remit as broad as the SDGs themselves, and contributions are welcome from scientists, academics, policymakers, and researchers working in fields related to any of the seventeen goals. If you are interested in contributing a monograph or curated volume to the series, please contact the Publishers: Zachary Romano [Springer; zachary. [email protected]] and Rachael Ballard [Palgrave Macmillan; rachael. [email protected]].

Yoshihisa Sugimura

Climate Change Countermeasures in Ports Toward Carbon Neutrality Empirical Analysis and Potential New Countermeasures

123

Yoshihisa Sugimura Ministry of Land, Infrastructure, Transport and Tourism National Institute for Land and Infrastructure Management Yokosuka, Japan

ISSN 2523-3084 ISSN 2523-3092 (electronic) Sustainable Development Goals Series ISBN 978-3-031-34393-3 ISBN 978-3-031-34394-0 (eBook) https://doi.org/10.1007/978-3-031-34394-0 © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 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

Preface

Transportation is a major source of pollution, accounting for approximately 15% of the total global greenhouse gas emissions; specifically, the percentage contribution by mode is 10%, 1.8%, 1.6%, and 0.4% for road, air, shipping, and rail, respectively (IPCC, 2022). As hubs for passenger and cargo transportation, ports also use a substantial amount of energy. Ports consume energy in the form of electricity or fuel; hence, their fossil fuel consumption and, consequently, CO2 emissions, including the consumption of purchased external electricity, are high (Iris and Lam, 2019; Sifakis and Tsoutsos, 2021). There has been increasing pressure to address the negative externalities created by port operations, such as air pollution (Eriksen, 2018; Schrobback and Meath, 2020). Notably, as the world moves toward decarbonization, the contribution of ports to climate change is receiving increased attention (Alamoush et al., 2020). Subsequently, the concepts of green ports and sustainable ports have emerged (Davarzani et al., 2016; Bjerkan and Seter, 2019; Lim et al., 2019) Environmental considerations in port operations are now a significant factor in competitiveness, and inclusion of social and environmental considerations in their planning and management have become mandatory for ports to operate (Hales et al., 2016). Under an international framework, 55 major global ports voluntarily adopted the World Ports Climate Declaration (WPCD) in 2008 in response to a request from the International Association of Ports and Harbors (IAPH), thereby creating the World Ports Climate Initiative (WPCI). These ports have implemented various environmental measures according to the WPCD. Additionally, the WPCD was extended into the World Ports Sustainability Program in 2018. In the same year, the International Maritime Organization (IMO) pledged to halve greenhouse gas (GHG) emissions by 2050 compared to 2008 levels (IMO, 2018) and further adopted the invitation to promote port cooperation in 2019 (IMO, 2019). Within certain countries and regions, additional local regulatory climate change measures have also been adopted, such as EU Directive 2014/94 (European Commission, 2014). Although port-related CO2 emissions are categorized as emissions from ships, ports, or transport in the hinterland (Chen et al., 2013; Sornn-Friese et al., 2021), ports cannot only reduce CO2 emissions under these aspects but also implement actions to reduce emissions from shipping and inland transport (Alamoush et al., 2020). Ports, as nodes connecting maritime and inland transport, can play a role in optimizing energy use in the entire supply chain (Bjerkan and Seter, 2019). However, while some advanced ports have v

vi

introduced environmentally friendly measures, most have not yet established a strategy for reducing CO2 emissions; there is currently a large gap between countermeasures and long-term efforts to achieve emission reduction targets. This is partly due to the lack of research on empirical findings in ports and the gap between research and application that affects decision making in ports. The social situation regarding climate change countermeasures is rapidly changing. Not only are the NDCs established by each country under the Paris Agreement insufficient to meet their objectives, but global GHG emissions in 2030 indicate that global warming is likely to surpass 1.5°C by the end of the twenty-first century, necessitating a rapid response (IPCC, 2022). Therefore, many countries have decided to implement stronger climate change countermeasures to become carbon neutral (CN). In response, various industries, including the transportation sector, are setting aggressive targets, and encouraging technological innovations to this end. Additionally, CO2 emissions during transportation are becoming an imminent deciding factor for shippers to choose transportation modes (Qi et al., 2022). These rapid changes in social conditions have considerably reshaped the environment in which climate change countermeasures are implemented at ports. This book discusses climate change countermeasures in ports. It presents empirical studies on cargo-handling machinery and reefer containers, as they are the main sources of energy consumption of ports, as well as new possibilities for climate change countermeasures based on new perspectives for continuous efforts in the future. These empirical analyzes include demonstrating the effectiveness of countermeasures using actual data on fuel and electricity consumption at a Japanese port, presenting an evaluation method to determine whether or not to implement countermeasures, and demonstrating its validity by comparing it with experimental data. In addition, the feasibility of implementing countermeasures was demonstrated through economic analysis, and the contents were designed to be applied to port management practices. This book contributes to the decision-making process of port stakeholders who are struggling with the introduction of countermeasures in practice and contains useful information for researchers in port operations and climate change countermeasures. Specifically, this book answers the following questions: • • • • •

What kind of CO2 emissions are associated with ports? What types of countermeasures are being implemented? To whom are the countermeasures implemented? Who implements the countermeasures? Why does the status of countermeasures implemented vary by countries and ports? • Why are climate change countermeasures not widely implemented in ports? • Are there any countermeasures based on new perspectives toward carbon neutrality?

Preface

Preface

vii

The book is divided into four parts. Part I introduces Introduction to Climate Change Countermeasures in Ports, including the relationship between port governance and climate change countermeasures. In Part II, Countermeasures for Cargo-Handling Machinery and their economic feasibility analyses are introduced. Part III focuses on Countermeasures in Reefer Container Areas. Finally, Part IV provides two examples of New Possibilities for Climate Change Countermeasures in Ports, such as the beneficial utilization of dredged soil and role of ports in the trade-off problem between circular economy and climate change action. The book has several features. Climate change countermeasures at ports are so diverse that exhaustive introduction would take up too much space in a single volume. Therefore, the book is organized with an emphasis on empirical aspects, economic feasibility, and new possibilities. This is due to a lack of empirical studies and widely adopted countermeasures owing to economic constraints, as well as the need for long-term efforts in climate change countermeasures. Notably, the subject of the empirical analysis is Japanese ports because as a Japanese national, the author is able to provide detailed explanations on the topic. Although Japanese ports are not leading the world in terms of climate change countermeasures, Hakata Port is leading Japan’s climate change countermeasures, and the author has been provided with sufficient empirical data. In line with this, the terms used in the Japanese case studies are those commonly used in Japan. For example, while port administration organizations have many names, the most common of which are port authority (PA), port management department, and port bureau, they are generally translated into English as port management body (PMB) in Japan. Therefore, PMB is used in the context of Japanese ports. Despite Japanese ports being the subject of empirical studies herein, this book will be useful for people throughout the world, mainly those in the industry, as well as researchers and students in ports and logistics. Because climate change action is a hot issue in the port and logistics industry, PAs or terminal operators and many shipping or logistics companies relying on ports are trying to address it. This issue has piqued the interest of many researchers and university students in maritime economics, industrial engineering, and environmental engineering. Yokosuka, Japan

Yoshihisa Sugimura

References Alamoush AS, Ballini F, Ölçer AI (2020) Ports’ technical and operational measures to reduce greenhouse gas emission and improve energy efficiency: a review. Marine Pollut Bull 160 Bjerkan KY, Seter H (2019) Reviewing tools and technologies for sustainable ports: does research enable decision making in ports? Transp Res Part D Transp Environ 72:243– 260 Chen G, Govindan K, Golias MM (2013) Reducing truck emissions at container terminals in a low carbon economy: proposal of a queuing-based bi-objective model for optimizing truck arrival pattern. Transp Res Part E: Logist Transp Rev 55:3–22

viii Davarzani H, Fahimnia B, Bell M, Sarkis J (2016) Greening ports and maritime logistics: a review. Transp Res Part D Transp Environ 48:473–487 Eriksen TH (2018) Scales of environmental engagement in an industrial town: glocal perspectives from Gladstone, Queensland. Ethnos 83:423–439 European Commission (2014) Directive 2014/94/EU of the European Parliament and of the council of 22 October 2014 on the deployment of alternative fuels infrastructure Hales D, Lam JSL, Chang YT (2016) The balanced theory of port competitiveness. Transp J 55:168–189 IMO (2018) Initial IMO strategy on reduction of GHG emissions from ships. Resol MEPC 304(72) IMO (2019) Inivitation to member states to encourage voluntary cooperation between the port and shipping sectors to contribute to reducing GHG emissions from ships. Resol MEPC 323(74) IPCC (2022) Climate change 2022: mitigation of climate change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change Iris Ç, Lam JSL (2019) A review of energy efficiency in ports: operational strategies, technologies and energy management systems. Renew Sustain Energy Rev 112:170– 182 Lim S, Pettit S, Abouarghoub W, Beresford A (2019) Port sustainability and performance: a systematic literature review. Transp Res Part D Transp Environ 72:47–64 Qi Y, Harrod S, Psaraftis HN, Lang M (2022) Transport service selection and routing with carbon emissions and inventory costs consideration in the context of the belt and road initiative. Transp Res Part E Logist Transp Rev 159 Schrobback P, Meath C (2020) Corporate sustainability governance: insight from the Australian and New Zealand port industry. J Clean Prod 255 Sifakis N, Tsoutsos T (2021) Planning zero-emissions ports through the nearly zero energy port concept. J Clean Prod 286 Sornn-Friese H, Poulsen RT, Nowinsk AU, de Langen P (2021) What drives ports around the world to adopt air emissions abatement measures? Transp Res Part D Transp Environ 90

Preface

Contents

Part I

Introduction to Climate Change Countermeasures in Ports

1 General Introduction to Climate Change Countermeasures in Ports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Port-Related CO2 Emissions . . . . . . . . . . . . . . . . . . . . . . 1.3 Potential Climate Change Countermeasures in Ports Based on Previous Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Themes Covered in This Book Given the Current Climate Change Countermeasures at Ports . . . . . . . . . . . . . . . . . . 1.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 Relationship Between Port Governance and Climate Change Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Port Reform and Port Governance . . . . . . . . . . . . . 2.3 Case Study: Japan . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Stakeholders in Ports and Their Roles on Climate Change Actions . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Part II

.. .. ..

3 3 4

..

5

.. .. ..

7 10 11

. . . .

. . . .

15 15 15 17

....... ....... .......

20 24 25

. . . .

. . . .

. . . .

. . . .

. . . .

Countermeasures for Cargo-Handling Machinery

3 CO2 Reduction Effects by Electrification of Cargo-Handling Machinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Port Environmental Policies in Japan and Development of Environmental Countermeasures at Hakata Port . . . . . 3.3 Overview of E-RTG and HSC . . . . . . . . . . . . . . . . . . . . . 3.4 CO2 Emission Reductions by Introducing E-RTG and HSC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.1 Changes in CO2 Emissions of the Entire CT Due to Climate Change Countermeasures . . . . . . . . . . 3.4.2 CO2 Reduction Effect of E-RTG and HSC . . . . . .

.. ..

29 29

.. ..

30 32

..

35

.. ..

35 35 ix

x

Contents

3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Economic Feasibility of Electrification of Cargo-Handling Machinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Governance Structure of Hakata Port . . . . . . . . . . . . . . . . 4.3 Cost–Benefit Analysis of E-RTG and HSC Installation . . 4.4 Possibility of Using Carbon Credits . . . . . . . . . . . . . . . . . 4.5 Challenges from the Perspective of Economic Feasibility and the Role of Stakeholders in Promoting Countermeasures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Part III

38 39

. . . . .

41 41 42 43 46

.. .. ..

47 48 49

. . . . . . . . . . . . .

. . . . . . . . . . . . .

53 53 54 55 56 58 58 59 64 66 69 71 71

.. .. ..

73 73 73

. . . . .

76 77 79 81 82

. . . . .

Countermeasures in Reefer Container Areas

5 Simulation Model for Analysis of Energy Savings by Roof Shade Installations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Design and Operational System of the RS . . . . . . . . . . . . 5.3 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Experimental Measurements of the Impact of the RSs . . . 5.5 RS Impact Analysis Using Simulation . . . . . . . . . . . . . . . 5.5.1 Thermal Factors . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5.2 Governing Equation of the Simulation Model . . . 5.5.3 Geometrical Model and Parameter Setting . . . . . . 5.5.4 Thermal Simulation . . . . . . . . . . . . . . . . . . . . . . . 5.5.5 Verification of the Thermal Simulation Results . . 5.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 Energy-Saving Effects and Economic Feasibility of Roof Shade Installation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2 Energy Savings Owing to the RS at Hakata Port . . . . . . . 6.3 Seasonal and Regional Characteristics of Energy Savings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4 Economic Analysis of RS Installation . . . . . . . . . . . . . . . 6.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . .

Contents

xi

Part IV

New Possibilities for Climate Change Countermeasures in Ports

7 Carbon Containment and Creation of BCEs Through Beneficial Utilization of Dredged Soil Generated by Port Development Projects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2 Carbon Storage Effect of Beneficial Utilization of Dredged Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3 Quantitative Effects of Organic Carbon Containment of Dredged Soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 Future Significance of Containment of Organic Carbon in the Dredged Soil and the Creation of BCE . . . . . . . . . 7.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 Role of Ports in the Trade-Off Problem Between Circular Economy and Climate Change Action: Potential for Increased Use of Secondary Raw Materials in the Copper Industry as a Climate Change Countermeasure . . . . . . . . . . 8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.1 Scope of Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.2 Calculation Methods and Cases . . . . . . . . . . . . . . 8.2.3 Data Set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.4.1 Potential Utilization of Carbon Credits . . . . . . . . . 8.4.2 Feasibility of Carbon Neutrality in the Copper Industry Assuming Technological Innovation . . . . 8.4.3 Roles of Ports . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Appendix: Parameters Used for the Calculations . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.. ..

85 85

..

86

..

88

.. .. ..

88 93 94

. . . . . . . . .

. . . . . . . . .

97 97 98 98 101 103 106 108 108

. . . . .

. . . . .

110 111 111 112 120

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

Part I Introduction to Climate Change Countermeasures in Ports

Port-related CO2 emissions are identified as emissions from ships, ports, or transport in the hinterland. It is necessary to consider port activities as well as CO2 emissions from ships and vehicles entering and leaving the port. Previous studies have explored potential climate change countermeasures in ports, including those for vessels and vehicles entering and leaving the port, and some ports have been leading practical efforts. The level of efforts varies by country and port; however, recent studies indicate that countermeasures have been organized to a certain extent. Nevertheless, there is a gap between the current adoption of climate change countermeasures and ambitious emission reduction targets expressed by international organizations, necessitating long-term efforts. In Part I, the sources of port-related CO2 emissions are confirmed, and potential climate change countermeasures that could be adopted by ports are summarized from practical experience and previous studies. The background of the topics covered in this book is explained based on the characteristics of the current potential countermeasures. The relationship between port governance and climate change countermeasures and roles and responsibilities of stakeholders in this regard are also discussed, as the port governance of a country has a direct impact on the level of implementation of countermeasures.

1

General Introduction to Climate Change Countermeasures in Ports

1.1

Introduction

As hubs for logistics and passengers, ports consume a lot of energy; therefore, their CO2 emissions are high, which includes the consumption of purchased external electricity (Iris and Lam 2019; Martínez-Moya et al. 2019; Sifakis and Tsoutsos 2021). Increasing concerns over climate change have resulted in a greater focus on addressing the negative externalities created by port operations, such as air pollution (Eriksen 2018; Schrobback and Meath 2020; Alamoush et al. 2020). To address the challenges arising from CO2 emissions from ports, green ports have been introduced globally to reduce emissions and mitigate climate change (Davarzani et al. 2016). Simultaneously, ports need to manage and balance three bottom lines: economic, social, and environmental (Lim et al. 2019), and the concept of sustainable ports has emerged to address this (Bjerkan and Seter 2019). Consequently, environmental considerations in port operations are now an important element of competitiveness, and ports cannot operate without including social and environmental considerations in their planning and management (Hales et al. 2016). Port-related CO2 emissions are identified as emissions from ships, ports, or transport in the hinterland (Chen et al. 2013; Sornn-Friese et al. 2021), nevertheless, it is necessary to consider not only port activities but also CO2 emissions from ships and vehicles entering and leaving the port. However, because CO2 emissions reductions can be achieved through global supply chains and logistics efforts (Cariou et al. 2019; Xing et al. 2020), ports, as nodes connecting maritime and inland transportation, could play a role in optimizing the entire supply chain (Bjerkan and Seter 2019). Therefore, ports need to be able to apply strategies to reduce their CO2 emissions and implement measures to reduce emissions from maritime and inland transport (Alamoush et al. 2020). Several studies have previously investigated potential climate change countermeasures in ports, including those for vessels and vehicles entering and leaving the port, and some ports have been leading practical efforts. Although the level of effort by ports varies by country and port (Sornn-Friese and Poulsen 2016), latest studies indicate that countermeasures have been organized to a certain extent. However, there is a gap between the current adoption of climate change countermeasures and ambitious emission reduction targets, such as those expressed by the International Maritime Organization (IMO) and the International Association of Ports and Harbors (IAPH), necessitating longterm efforts are (Sornn-Friese et al. 2021).

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0_1

3

4

1

General Introduction to Climate Change Countermeasures in Ports

In this chapter, the sources of port-related CO2 emissions are confirmed, and potential climate change countermeasures that could be adopted by ports are summarized based on practical experience and previous studies. In addition, the background of the topics covered in this book is explained based on the characteristics of current potential countermeasures.

1.2

Port-Related CO2 Emissions

Port-related CO2 emissions, which can be identified as emissions from vessels, in ports, or from port hinterland transport, have been estimated in previous studies. For example, Styhre et al. (2017) and Tichavska et al. (2019) estimated emissions from ships in ports, Peng et al. (2018) and MartínezMoya et al. (2019) estimated emissions from port activities, and Aregall et al. (2018) and Wang et al. (2020) examined emissions from logistics networks in the port hinterland. The WPCI provides guidance on port carbon footprints (WPCI 2010), and Mamatok and Jin (2017) measured integrated port carbon footprints. In this chapter, the shares of port-related CO2 emissions from port activities such as cargo handling and from vessels and vehicles entering and departing from ports are confirmed. Because the shares vary depending on port characteristics, this chapter examines the CO2 emissions of Japanese ports as a whole from a macro perspective. The Ministry of Land, Infrastructure, Transport and Tourism (MLIT 2009) estimated the total CO2 emissions from 125 Japanese ports (88% of the total cargo handled at Japanese ports), which can be used as a basis for estimating CO2 emissions from Japanese ports. The share of each emission source is shown in Fig. 1.1, and the details and estimation methods are listed in Table 1.1. As shown in Fig. 1.1 and Table 1.1, emissions from ships and vehicles account for a considerable proportion of emissions compared to emissions from port activities, such as cargo-handling machinery and reefer containers. Notably, Gibbs et al. (2014) found that emissions from inland transport were twice as large as those from port activities and emissions from vessels at anchor were 10 times as large as those from port activities. Sakai and Watanabe (2009), who examined emissions at the top five container handling ports in Japan, found a similar trend, consistent with previous

Port 6.2%

Land transport 56.3%

Fig. 1.1 Percentage of each source in port-related CO2 emissions

Ship 37.4%

1.3 Potential Climate Change Countermeasures in Ports …

5

Table 1.1 Details of port-related CO2 emissions and estimation methods Share

Source

Items

1000s tCO2

Ship-to-shore cranes

39

0.3%

Yard cargo-handling machinery

347

2.7%

Reefer containers

137

1.1%

Administrative buildings

262

2.1%

Ports total

785

6.2%

Ships (at arrival and departure)

1320

10.4%

Ships at berth (MLIT 2009)  emissions at arrival and departure  emissions at berth (average of five ports) Sakai and Watanabe (2009)

Ships (at berth)

3413

27.0%

MLIT (2009)

Ships total

4733

37.4%

Sum of the above 2 items

Trucks (congestion at the gate)

91

Trucks (in-port drayage)

46

Trucks (hinterland transportation)

6986

Land transportation total

7123

Total (first perspective)

12,641

0.7%

Cargo-handling machinery in MLIT (2009) is allocated as a proportion of container cranes/yard cargo-handling machinery (average of five ports) (Sakai and Watanabe 2009) MLIT (2009) Sum of the above 4 items

MLIT (2009)

0.4% 55.3% 56.3% 100.0%

Sum of the above 3 items Sum of total

studies. The proportion of trucks (hinterland transportation) is particularly large in Fig. 1.1 and Table 1.1 because they include the entire hinterland transportation to the origin and destination of the cargo; however, the results would be closer to the trend reported in Gibbs et al. (2014) if they were limited to the suburbs of ports. Additionally, emissions from factories and other facilities located in ports were not included, but if they were, the emissions would be large enough to change the order of magnitude.

1.3

Potential Climate Change Countermeasures in Ports Based on Previous Studies

As indicated in the previous section, emissions from ships and vehicles account for a considerable share of emissions from port activities. However, ports are nodes connecting maritime and inland transport, and can thus play a role in optimizing the entire supply chain (Bjerkan and Seter 2019). Ports can reduce CO2 emissions from their activities and implement actions to reduce emissions from shipping and inland transport (Alamoush et al. 2020). There are two types of climate change counter measures in ports: those that are physically implemented in the port area by whoever is the main actor, and those that involve port management bodies (PMBs) or terminal operators (TOs). In this book, climate change countermeasures in ports are discussed in the latter sense. In this section, climate change countermeasures in ports that have been demonstrated in previous studies are reviewed, and potential countermeasures are summarized. While previous studies have focused directly on climate change countermeasures in ports, they have often been taken up as an element in the context of port greening and sustainability. Bergqvist and Monios (2019) pointed out that

6

1

General Introduction to Climate Change Countermeasures in Ports

sustainability in ports tends to focus on air pollution. Hua et al. (2020) noted a similar trend in green ports. Climate change was an important factor in previous studies addressing environmental issues in ports more broadly. This chapter reviews the literature to organize potential climate change countermeasures in ports. In addition to the climate change countermeasures themselves, this chapter also reviews previous studies on the implementation methods and environment in which the countermeasures are implemented. Sornn-Friese et al. (2021) pointed out that, PMBs can implement a wide range of countermeasures, and various studies have been conducted on emission countermeasures. In this regard, literature, which provides a comprehensive review of countermeasures, has already been developed, such as Bjerkan and Seter (2019), who present 26 climate change countermeasures as tools and technologies for sustainable ports, and Alamoush et al. (2020), who classified technical and operational countermeasures at ports to reduce GHG emissions and improve energy efficiency, including land transportation and at the ship-port interface, into 7 broad categories and 19 subcategories. Xing et al. (2020) reviewed countermeasures against CO2 emissions at ports from a ship-side perspective, organizing countermeasures in terms of port operations. Iris and Lam (2019) organized countermeasures at ports, with a focus on energy efficiency. Although not included in the literature review, Lam and Notteboom (2014) and Acciaro et al. (2014a) summarized environmental actions in ports, including climate countermeasures, in terms of PMBs. Additionally, the international organization IAPH has published a toolkit that establishes the planning process in addressing port-related air quality and climate change-related issues (IAPH 2020a, b). Moreover, PIANC has published reports on renewable energy and energy efficiency for ports (PIANC 2019a), as well as carbon management for ports and navigation infrastructure (PIANC 2019b). These publications can be used as references for this book. Regarding the implementation methods of countermeasures, Li et al. (2020) identified controltype policies, such as the establishment of emission control zones and emission limits, and incentive-type policies, such as the establishment of incentive and disciplinary systems, as the means adopted by governments. Lam and Notteboom (2014) pointed out that advanced ports tend to adopt an enforcement approach for environmental strategies. Concerning control-type policies, Geerlings and Van Duin (2011) argued that financial support from PMBs is necessary for TOs to replace cargo-handling machinery, while Lam and Notteboom (2014) stressed the importance of designing financial incentives for entities that develop voluntary practices to achieve their environmental goals. As Bergqvist and Egels-Zandén (2012) demonstrated, incentives for environmental considerations in inland transport, such as the use of differentiated port charges, could be an example. Implementing climate change measures in ports is quite costly (Woo et al. 2018), and even for onshore power supply, which is a typical countermeasure (Bjerkan and Seter 2019), Dai et al. (2019) and Zis (2019) argue that the biggest barriers for implementation are high initial costs. They further emphasized that even if investment in implementation is feasible only from a government standpoint that includes environmental benefits, it is difficult on a business basis, and the only way to achieve this is to mandate the use of onshore power supply systems. Regarding the implementation environment of countermeasures, Sornn-Friese and Poulsen (2016) discovered that the level of adoption of environmental initiatives varies by country and port, while Lam and Notteboom (2014) noted that geographical, economic, regulatory, and political backgrounds are important contexts for these differences. Although some countermeasures, such as onshore power supply, are becoming more widely implemented, Sornn-Friese et al. (2021) noted that there is a lack of a universally adopted countermeasure in the port industry and that there is a gap between the ambitious emission reduction targets expressed by the IMO and IAPH and current adoption of countermeasures, necessitating long-term effort. While ports act on the basis of national and regional regulations and rules (Alamoush et al. 2020), Poulsen et al. (2018) argued that climate change

1.4 Themes Covered in This Book Given the Current Climate Change …

7

countermeasures are complex and require voluntary efforts and broad cooperation among port stakeholders. Although ports have several stakeholders, Acciaro et al. (2014a) and Lam and Li (2019) argued that PMBs, in particular, need to take an active role to address environmental challenges. Sornn-Friese et al. (2021) empirically demonstrated that one of the key drivers of CO2 emission reductions is changes in port governance, which have a significant impact on environmental efforts (Munim et al. 2020). This is discussed in detail in the next chapter. Given the countermeasures organized in previous studies, especially in review articles, such as Bjerkan and Seter (2019) and Alamoush et al. (2020), the list of potential climate change countermeasures in ports can be summarized as shown in Table 1.2. In Table 1.2, countermeasures are divided into those related to port activities, vessels, and vehicles, based on the fact that their CO2 emissions are associated with ports, and are therefore, subject to countermeasures. Table 1.2 also shows the implementation status of each countermeasure in Japan and examples from previous studies that can be used as references when implementing countermeasures. This is because it is helpful to understand how difficult it is to realize each countermeasure by looking at Japan’s status.

1.4

Themes Covered in This Book Given the Current Climate Change Countermeasures at Ports

Although there are numerous potential climate change countermeasures for ports, as shown in Table 1.2, a globally adopted countermeasure in the port industry in practical terms is lacking (SornnFriese et al. 2021). For example, when observing the application status of climate change countermeasures in Japanese ports, as shown in Table 1.2, the degree of adoption and effectiveness of each countermeasure has not yet been high. Although the government has taken the initiative in the form of pilot projects and subsidies for onshore power supply facilities and energy-efficient cargo-handling machinery, the projects have not been fully adopted due to economic factors because neither national laws nor regulations at individual ports have been introduced. While advanced ports frequently cited in previous studies as developing environmentally friendly strategies include the Antwerp, Rotterdam, Los Angeles/Long Beach, Hamburg, and Singapore ports (Acciaro 2015; Aregall et al. 2018), Lam and Notteboom (2014) note that these advanced ports tend to adopt an enforcement approach. While the United States and European countries have tried to adopt market-based mechanisms, such as emissions trading schemes (Arimura et al. 2019), Japan's environmental policy is characterized by its reliance on voluntary approaches compared to the, which may account for its low effectiveness. Furthermore, port governance models underlie the different levels of environmental initiative adoption by ports in different countries and ports (Munim et al. 2020). This is discussed in detail in the next section. Another major reason for insufficient progress in climate change actions in ports is that countermeasures in ports are costly (Woo et al. 2018) and difficult to implement on a business basis because of barriers posed by high initial investments (Dai et al. 2019; Zis 2019). The challenges on the research side are also part of the background, as there are few previous studies based on empirical findings, and the gap between research and practice does not adequately support decision makers in ports (Bjerkan and Seter 2019). Therefore, it is important to analyze economic feasibility with empirical data, and Parts 2 and 3 of this book will address these gaps. Next, several characteristics emerged when examining the countermeasures in Table 1.2. First, the list of climate change countermeasures in ports to date consists mostly of emission source countermeasures, and barring a few exceptions, almost no studies mention sink countermeasures (IS2S 2013; Lam and Notteboom 2014; Bosman et al. 2018). Furthermore, although environmental impacts are occasionally mentioned in port development (Lum and Li 2019), they have not been recognized as

8

1

General Introduction to Climate Change Countermeasures in Ports

Table 1.2 List of potential climate change countermeasures at ports Port

Measures

Existing measures and status

Examples

Development of green port plan, etc.

Port environmental plan, carbon–neutral port (CNP) formation plan (government supported)

Acciaro et al. (2014), Schipper et al. (2017)

Monitoring GHG emissions, information measures

No legal or institutional framework

Puig et al. (2017), Cammin et al. (2020)

Inclusion of environmentally friendly actions in terminal leasing

Some examples, including environmental incentives in terminal leases

Berg and Langen (2014), Notteboom and Lam (2018)

Introduction of incentives (for ships and vehicles)

Voluntarily introduced at some ports

Chen et al. (2013), Aregall et al. (2018)

Collaboration between shipping, land transport, and PAs; optimization of the entire supply chain

Government subsidy system available (project to promote CO2 reduction measures in the logistics sector)

Cheon (2017), Aregall et al. (2018)

More efficient operation of cargohandling machines (including automation), environmentally friendly terminal layout

Government-led study on artificial intelligence terminals underway (with subsidies for rubber-tired gantry crane remoting)

He et al. (2015), He (2016)

Electrification of cargo-handling machinery via the use of fuel cells

Government subsidy system available (project to promote CO2 reduction measures in the logistics sector)

Martínez-Moya et al. (2019), Di Ilio et al. (2021)

Development of onshore power supply facilities

No full-scale introduction, although the government has conducted pilot projects at some ports

Zis et al. (2014) Winkel et al. (2016)

Renewable energy utilization

Promoting the introduction of offshore wind power generation, etc.

Lam and Notteboom (2014), Kang and Kim (2017)

Development of large-scale energy storage facilities

Government implementation of a pilot project on disaster response and other emergencies to reduce carbon emissions

Kotrikla et al. (2017), Papaio-annou et al. (2017)

Introduction of smart grid

No port-specific examples

Ihle et al. (2016), Yiğit et al. (2016)

Development of hydrogen stations

Under consideration at some ports (government supported)

Kang and Kim (2017), Bicer and Dincer (2018)

Installation of carbon-free power supplies in reefer containers and administration buildings

Voluntary introduction of roof shades in some ports

Wilmsmeier et al. (2012), van Duin et al. (2018)

Creation of blue carbon ecosystems, including within offshore wind power generation areas

Many areas have been developed under the Sea Blue projects, but their use as a measure against climate change is limited (the carbon offset system has only been introduced at some ports)

PIANC (2019b)

Beneficial utilization of dredged soil

Many areas have been developed under the Sea Blue and Ordinary Port Renovation projects, but they have not been assessed as climate change countermeasures

Development of green areas at ports

The Port Environment Improvement project has been applied to many areas, but they have not been assessed as climate change countermeasures

I2S2 (2013)

Development of import/export and distribution bases for next-generation energy

Liquefied natural gasbunkering bases are being promoted. Next-generation energy is under consideration at some ports

Acciaro (2015)

Development of import/export and delivery bases for recyclable resources

Exports are being promoted under the Recycle Port policy, but there are no equivalent to promoting imports

Poulsen et al. (2018),

(continued)

1.4 Themes Covered in This Book Given the Current Climate Change …

9

Table 1.2 (continued) Ship

Land transport

Measures

Existing measures and status

Examples

Utilization of onshore power supply facilities

No full-scale introduction, although the government has conducted pilot projects at some ports

Zis et al. (2014), Boile et al. (2016)

Implementation of efficient navigation within the port

Some examples of voluntary initiatives

Styhre et al. (2017), Linder (2018)

Introduction of environmentally friendly ships (zero-emission ships, etc.)

Government subsidies are available for R&D

Winnes et al. (2015), Styhre et al. (2017)

Reduction of traffic congestion at the gate, and improvement of the efficiency of intra-port storage transportation

Each port is dealing with the issue, but analyses as climate change countermeasures have just commenced

Chen et al. (2013), He et al. (2013)

Modal shift/split

Government subsidy system available

Aregall et al. (2018)

Introduction of environmentally friendly vehicles

Under consideration at some ports (government supported)

Poulsen et al. (2018), Di Ilio et al. (2021)

climate change countermeasures. Table 1.2 lists the creation of blue carbon ecosystems and the beneficial utilization of dredged soil as measures. Blue carbon refers to the carbon sequestered within marine ecosystems as a result of biological activities in the ocean, as defined by the United Nations Environment Programme (UNEP) in 2009; blue carbon has recently attracted attention as a mitigation measure for global warming (Lovelock and Duarte 2019). The ocean absorbs slightly more than half of the CO2 absorbed by organisms on Earth, and more than half of the carbon absorbed by the ocean is also absorbed by shallow-water areas known as blue carbon ecosystems (BCEs), such as mangrove forests, salt marshes, and seagrass beds (Nellemann et al. 2009). As many aspects of BCEs have not been scientifically verified (Lovelock and Duarte 2019; Macreadie et al. 2019), depending on their elucidation, future expansion as sinks can be expected. As Taljaard et al. (2021) argued for the incorporation of the natural environment as an integral component of port infrastructure systems, if BCEs, which are CO2 sinks, are created as part of port development, it is acceptable to categorize them as a climate change countermeasure in ports. PIANC (2019b) also mentions the use of coastal areas with vegetation and dredged soil as carbon sinks. Furthermore, the effective use of dredged soil associated with port development has the potential to be a new sink countermeasure. Although the effective use of dredged soil to create BCEs for CO2 absorption has already been shown (PIANC 2019b), the carbon contained in the dredged soil itself has not been focused on thus far. Because dredged soil contains a large amount of carbon, it has the potential to substantially contribute to CO2 absorption if it is contained in the ground for the creation of seaweed beds and tidal flats, backfilling of depressions, and land reclamation. The relationship between climate change countermeasures and the logistics function of ports has not received much attention. The use of next-generation energy and recyclable resources as climate change countermeasures (Poulsen et al. 2018) may improve feasibility by reducing costs through improved logistics efficiency, and the role of ports as transportation hubs should be significant. For example, the spread of low-carbon technologies (electric vehicles and low-carbon power generation technologies) could considerably increase the demand for copper, which would likely increase dependence on recyclable resources as raw materials. This would lead to the realization of a circular economy (CE), which is a common goal globally, along with climate change countermeasures. However, while transportation may have a significant impact on the cost and CO2 emissions of procuring recyclable resources with complex reverse logistics (Reddy et al. 2020), there are insufficient academic findings on the impact of the increased introduction of recyclable resources,

10

1

General Introduction to Climate Change Countermeasures in Ports

including logistics and climate change action, in which the realization of a CE may be conflicting goals in some cases. This book also focuses on the trade-off problem between CE and climate change countermeasures. Specifically the impact of increased introduction of recyclable resources as climate change countermeasures on CE needs to be verified, to determine whether it can be a feasible climate change countermeasure in ports. Two potential new climate change countermeasures for ports, the effective use of dredged soil and role of ports in the trade-off between the CE and climate change countermeasures, will be discussed in Part 4.

1.5

Conclusions

Because ports are hubs of logistics and passengers that consume a lot of energy and emit large amounts of CO2, climate change countermeasures at ports have been attracting attention as the world moves toward a decarbonized society. While port-related CO2 emissions include emissions from shipping, port activities, and port hinterland transport, ports, as nodes connecting maritime and inland transport, can play a role in optimizing the entire supply chain. Therefore, ports can apply strategies to reduce their CO2 emissions and implement measures to reduce the emissions of shipping and inland transport. Although several previous studies have investigated potential climate change countermeasures in ports, including those for ships and vehicles entering and departing the port, as well as ports leading practical efforts, there is a gap between the current adoption of climate change countermeasures and the ambitious emission reduction targets stated by international organizations, necessitating long-term efforts. In this chapter, the sources of port-related CO2 emissions were identified, and shares of inland transportation, vessels, and port activities were confirmed to be the largest, in that order. Next, potential climate change countermeasures that could be adopted by ports were summarized by dividing them into those for port activities, vessels, and vehicles, and the status of application of the countermeasures in Japan was presented. The level of climate change action varies depending on country and port; however, in practice, as is the case in Japan, climate change action has not been widely adopted. One of the major reasons for the insufficient progress of countermeasures is that they are costly and difficult to implement on a business basis, but there are few previous studies based on empirical findings, and the gap between research and practice does not adequately support decision makers in ports. For this reason, economic analysis, including empirical data, is important, and Parts 2 and 3 of this book will fill these gaps. Additionally, the gap between the adoption of climate change countermeasures and ambitious emission reduction targets indicates that new climate change countermeasures are needed, both practically and academically. In this regard, it is important to note that the list of climate change countermeasures in ports to date is characterized primarily by emission source countermeasures, but almost none by sink countermeasures. The relationship between climate change countermeasures and the logistics function of ports has not received much attention, but there is a possibility that the role of ports as transportation hubs could attract attention from a variety of perspectives toward a decarbonized society. This book focuses on these points and deals with the effective use of dredged soil and the role of ports in the trade-off between CE and climate change countermeasures in Part 4 as new possibilities for climate change countermeasures in ports.

References

11

References Acciaro M (2015) Corporate responsibility and value creation in the port sector. Int J Log Res Appl 18:291–311 Acciaro M, Ghiara H, Cusano MI (2014) Energy management in seaports: a new role for port authorities. Energy Policy 71:4–12 Alamoush AS, Ballini F, Ölçer AI (2020) Ports’ technical and operational measures to reduce greenhouse gas emission and improve energy efficiency: a review. Marine Pollut Bull 160 Aregall MG, Bergqvist R, Monios J (2018) A global review of the hinterland dimension of green port strategies. Transp Res Part D Transp Environ 59:23–34 Arimura TH, Kaneko S, Managi S, Shinkuma T, Yamamoto M, Yoshida Y (2019) Political economy of voluntary approaches: a lesson from environmental policies in Japan. Econ Anal Policy 64:41–53 Berg RVD, Langen PWD (2014) An exploratory analysis of the effects of modal split obligations in terminal concession contracts. Int J Shipp Trans Logist 6:571–592 Bergqvist R, Egels-Zandén N (2012) Green port dues—the case of hinterland transport. Res Transp Bus Manag 5:85–91 Bergqvist R, Monios J (2019) Green ports in theory and practice. Green Ports. Elsevier, pp 1–17 Bicer Y, Dincer I (2018) Clean fuel options with hydrogen for sea transportation: a life cycle approach. Int J Hydro Energy 43:1179–1193 Bjerkan KY, Seter H (2019) Reviewing tools and technologies for sustainable ports: does research enable decision making in ports? Transp Res Part D Transp Environ 72:243–260 Boile M, Theofanis S, Sdoukopoulos E, Plytas N (2016) Developing a port energy management plan issues, challenges, and prospects. Transp Res Record J Transp Res Board 2549:19–28 Bosman R, Loorbach D, Rotmans J, Raak RV (2018) Carbon lock-out: leading the fossil port of Rotterdam into transition. Sustainability 10 Cammin P, Yu J, Heilig L, Voß S (2020) Monitoring of air emissions in maritime ports. Transp Res Part D Transp Environ 87 Cariou P, Parola F, Notteboom T (2019) Towards low carbon global supply chains: a multi-trade analysis of CO2 emissions reductions in container shipping. Int J Prod Econ 208:17–28 Chen G, Govindan K, Golias MM (2013) Reducing truck emissions at container terminals in a low carbon economy: proposal of a queuing-based bi-objective model for optimizing truck arrival pattern. Transp Res Part E: Logist Transp Rev 55:3–22 Cheon S (2017) The economic-social performance relationships of ports: roles of stakeholders and organizational tension. Sustain Dev 25:50–62 Dai L, Hu H, Wang Z, Shi Y, Ding W (2019) An environmental and techno-economic analysis of shore side electricity. Transp Res Part D Transp Environ 75:223–235 Davarzani H, Fahimnia B, Bell M, Sarkis J (2016) Greening ports and maritime logistics: a review. Transp Res Part D Transp Environ 48:473–487 Di Ilio G, Di Giorgio P, Tribioli L, Bella G, Jannelli E (2021) Preliminary design of a fuel cell/battery hybrid powertrain for a heavy-duty yard truck for port logistics. Energy Convers Manage 243 Eriksen TH (2018) Scales of environmental engagement in an industrial town: glocal perspectives from Gladstone, Queensland. Ethnos 83:423–439 Geerlings H, Van Duin R (2011) A new method for assessing CO2-emissions from container terminals: a promising approach applied in Rotterdam. J Clean Prod 19:657–666 Gibbs D, Rigot-Muller P, Mangan J, Lalwani C (2014) The role of sea ports in end-to-end maritime transport chain emissions. Energy Policy 64:337–348 Gonzalez-Aregall M, Bergqvist R, Monios J (2018) A global review of the hinterland dimension of green port strategies. Transp Res Part d: Transp Environ 59:23–34 Hales D, Lam JSL, Chang YT (2016) The balanced theory of port competitiveness. Transp J 55:168–189 He J (2016) Berth allocation and quay crane assignment in a container terminal for the trade-off between time-saving and energy-saving. Adv Eng Inform 30:390–405 He J, Huang Y, Yan W, Wang S (2015) Integrated internal truck, yard crane and quay crane scheduling in a container terminal considering energy consumption. Expert Syst Appl 42:2464–2487 He J, Zhang W, Huang Y, Yan W (2013) A simulation optimization method for internal trucks sharing assignment among multiple container terminals. Adv Eng Inform 27:598–614 Hua C, Chen J, Wan Z, Xu L, Bai Y, Zheng T, Fei Y (2020) Evaluation and governance of green development practice of port: a sea port case of China. J Clean Prod 249 IAPH (2020a) IAPH tool box for greenhouse gasses. Int Assoc Ports Harb. http://iaphtoolbox.wpci.nl/DRAFT% 20IAPH%20GHG%20TOOL%20BOX%20dea.pdf IAPH (2020b) IAPH tool box for clean air programs. Int Assoc Ports Harb. http://iaphtoolbox.wpci.nl/DRAFT% 20IAPH%20TOOL%20BOX%20%20dea.pdf

12

1

General Introduction to Climate Change Countermeasures in Ports

Ihle N, Runge S, Meyer-Barlag C, Grundmeyer N, Appelrath HJ (2016) Software components for demand side integration at a container terminal. Comput Sci Res Dev 31:25–31 Iris Ç, Lam JSL (2019) A review of energy efficiency in ports: operational strategies, technologies and energy management systems. Renew Sustain Energy Rev 112:170–182 I2S2 (2013) Environmental initiatives at seaports worldwide: a snapshot of best practices. International Institute for Sustainable Seaports (I2S2). VA, USA. Kang D, Kim S (2017) Conceptual model development of sustainability practices: the case of port operations for collaboration and governance. Sustainability 9 Kotrikla AM, Lilas T, Nikitakos N (2017) Abatement of air pollution at an aegean island port utilizing shore side electricity and renewable energy. Mar Policy 75:238–248 Lam JSL, Li KX (2019) Green port marketing for sustainable growth and development. Transp Policy 84:73–81 Lam JSL, Notteboom T (2014) The greening of ports: a comparison of port management tools used by leading ports in Asia and Europe. Transp Rev 34(2):169–189 Li Y, Li J, Wei F, Huang Q (2020) CO2 emission performance evaluation of Chinese port enterprises: a modified metafrontier non-radial directional distance function approach. Transp Res Part D Transp Environ 89 Lim S, Pettit S, Abouarghoub W, Beresford A (2019) Port sustainability and performance: a systematic literature review. Transp Res Part D Transp Environ 72:47–64 Linder A (2018) Explaining shipping company participation in voluntary vessel emission reduction programs. Transp Res Part D Transp Environ 61:234–245 Lovelock CE, Duarte CM (2019) Dimensions of blue carbon and emerging perspectives. Biol Lett 15 Macreadie PI, Anton A, Raven JA, Beaumont N, Connolly RM, Friess DA, Kelleway JJ, Kennedy H, Kuwae T, Lavery PS, Lovelock CE, Smale DA, Apostolaki ET, Atwood TB, Baldock J, Bian-chi TS, Chmura GL, Eyre BD, Fourqurean JW, Hall-Spencer JM, Huxham M, Hendriks IE, Krause-Jensen D, Laffoley D, Luisetti T, Marbà N, Masque P, McGlathery KJ, Megonigal JP, Murdi-yarso D, Russell BD, Santos R, Serrano O, Silli-man BR, Watanabe K Duarte CM (2019) The future of blue carbon science. Nat Commun 10 Mamatok Y, Jin C (2017) An integrated framework for carbon footprinting at container seaports: the case study of a Chinese port. Marit Policy Manag 44:1–19 Martínez-Moya J, Vazquez-Paja B, Maldonado JAG (2019) Energy efficiency and CO2 emissions of port container terminal equipment: evidence from the port of Valencia. Energy Policy 131:312–319 Ministry of Land, Infrastructure, Transport and Tourism (2009) Guidelines for developing a greenhouse gas emission reduction plan at ports (draft). https://www.mlit.go.jp/common/000043030.pdf Munim ZH, Sornn-Friese H, Dushenko M (2020) Identifying the appropriate governance model for green port management: applying analytic network process and best-worst methods to ports in the Indian Ocean Rim. J Clean Prod 268 Nellemann C, Corcoran E, Durate CM, Valdes L, De Young C, Fonseca L, Grimsditch G (2009) Blue carbon—a rapid response assessment, United Nations environmental programme. GRID-Arendal. Brikeland Trykkeri AS, Birkeland Notteboom T, Lam JSL (2018) The greening of terminal concessions in seaports. Sustainability 10:3318 Papaioannou V, Pietrosanti S, Holderbaum W, Becerra VM, Mayer R (2017) Analysis of energy usage for RTG cranes. Energy 125:337–344 Peng Y, Li X, Wang W, Liu K, Li C (2018) A simulation-based research on carbon emission mitigation strategies for green container terminals. Ocean Eng 163:288–298 PIANC (2019a) Renewables and energy efficiency for maritime ports—MarCom WG Report n°159–2019a. The World Association for Waterborne Transport Infrastructure PIANC Maritime Navigation Commission. Brussels, Belguim PIANC (2019b) Carbon management for port and navigation infrastructure—EnviCom WG Report n° 188—2019b. The World Association for Waterborne Transport Infrastructure PIANC Maritime Navigation Commission. Brussels, Belguim Poulsen RT, Ponte S, Sornn-Friese H (2018) Environmental upgrading in global value chains: the potential and limitations of ports in the greening of maritime transport. Geoforum 89:83–95 Puig M, Michail A, Wooldridge C, Darbra RM (2017) Benchmark dynamics in the environmental performance of ports. Mar Pollut Bull 121:111–119 Reddy KN, Kumar A, Sarkis J, Tiwari MK (2020) Effect of carbon tax on reverse logistics network design. Comp Indust Eng 139 Sakai H, Watanabe Y (2009) Evaluation and reduction of carbon dioxide in container terminals. Transp Policy Stud Rev 9:15–21 Schrobback P, Meath C (2020) Corporate sustainability governance: insight from the Australian and New Zealand port industry. J Clean Prod 255 Schipper CA, Vreugdenhil H, de Jong MPC (2017) A sustainability assessment of ports and port-city plans: comparing ambitions with achievements. Transp Res Part D Transp Environ 57:84–111 Sifakis N, Tsoutsos T (2021) Planning zero-emissions ports through the nearly zero energy port concept. J Clean Prod 286

References

13

Sornn-Friese H, Poulsen RT (2016) Ports as sustainability hubs in global maritime supply chains: a content analysis of the external business communication of the worlds largest ports. Paper Presented at the International Association of Maritime Economist Sornn-Friese H, Poulsen RT, Nowinsk AU, de Langen P (2021) What drives ports around the world to adopt air emissions abatement measures? Transp Res Part D Transp Environ 90 Styhre L, Winnes H, Black J, Lee J, Le-Griffin H (2017) Greenhouse gas emissions from ships in ports—case studies in four continents. Transp Res Part D Transp Environ 54:212–224 Taljaard S, Slinger JH, Arabi S, Weerts SP, Vreugdenhil H (2021) The natural environment in port development: a ‘green handbrake’ or an equal partner? Ocean Coastal Manage 199 Tichavska M, Tovar B, Gritsenko KD, Johjansson L, Jakanen JP (2019) Air emissions from ships in ports: does regulation make a difference? Transp Policy 75:128–140 van Duin JHR, Geerlings H, Verbraeck A, Nafde T (2018) Cooling down: a simulation approach to reduce energy peaks of reefers at terminals. J Clean Prod 193:72–86 Wang L, Peng C, Shi W, Zhu M (2020) Carbon dioxide emissions from port container distribution: spatial characteristics and driving factors. Transp Res Part D Transp Environ 82 Wilmsmeier G, Froese J, Zotz AK (2012) Energy consumption and efficiency: emerging challenges from reefer trade in South American container terminals. FAL Bull 329:1–9 Winkel R, Weddige U, Johnsen D, Hoen V, Papaefthimiou S (2016) Shore side electricity in Europe: potential and environmental benefits. Energy Policy 88:584–593 Winnes H, Styhre L, Fridell E (2015) Reducing GHG emissions from ships in port areas. Res Transp Bus Manag 17:73–82 Woo JK, Moon DSH, Lam JSL (2018) The impact of environmental policy on ports and the associated economic opportunities. Transp Res Part A Policy Pract 110:234–242 WPCI, 2010. Carbon Footprinting for Ports: Guidance Document. World Port Climate Initiative. Xing H, Spence S, Chen H (2020) A comprehensive review on countermeasures for CO2 emissions from ships. Renew Sustain Energy Rev 134 Yiğit K, Kökkülünk G, Parlak A, Karakaş A (2016) Energy cost assessment of shoreside power supply considering the smart grid concept: a case study for a bulk carrier ship. Mar Policy 43:469–482 Zis TPV (2019) Prospects of cold ironing as an emissions reduction option. Transp Res Part A Policy Pract 119:82–85 Zis T, North RJ, Angeloudis P, Ochieng WY, Bell MGH (2014) Evaluation of cold ironing and speed reduction policies to reduce ship emissions near and at ports. Maritime Econ Logist 16:371–398

2

Relationship Between Port Governance and Climate Change Action

2.1

Introduction

As discussed in the previous chapter, although a wide range of climate change countermeasures can be applied to ports, currently, there is a lack of universally adopted countermeasures in the port industry in practical terms. Moreover, the ambitious targets set by organizations, such as the IMO, and the level at which countermeasures are currently implemented, are not aligned; thus, a long-term strategy is required (Sornn-Friese and Poulsen 2016). While the level of countermeasure implementation depends on geographical, economic, regulatory, and political contexts (Lam and Notteboom 2014), the intensity of the policies adopted by the government and port governance of the country directly influence the level of implementation (Li et al. 2020; Munim et al. 2020). The impact of policies can be easily understood from literature. For example, Lam and Notteboom (2014) pointed out that ports with environmentally advanced strategies tend to adopt a regulatory approach. However, the impact of port governance, combined with the diversity of its forms, can be very complex and fluctuating. Poulsen et al. (2018) noted that climate change action is complex, thereby requiring extensive cooperation among the numerous stakeholders of ports. Notably, Acciaro et al. (2014) and Lam and Li (2019) argue that port authorities (PAs) need to take an active role to effectively address environmental challenges, which is in the context of changes in port governance. Factors such as, the kind of port governance and corresponding regulations affect climate change action, governing body/authority responsible for the implementation of climate change countermeasures, and governing body/authority responsible for taking the initiative in promoting climate change countermeasures may vary depending on port governance. This chapter provides an overview of port reforms and the resulting changes in port governance followed by an explanation of the relationship between port governance and climate change countermeasures. As a case study, the chapter then discusses how port governance in Japan affects the climate change countermeasures of each port and which actors’ initiatives are needed to address the issue. The roles and responsibilities of the stakeholders are also discussed.

2.2

Port Reform and Port Governance

Against the background of environmental changes, such as trade globalization, technological innovation, and new public management philosophy (Brooks and Cul-linane 2007; Debrie et al. 2013; Ferrari et al. 2015), port reforms have been implemented in many countries since the late 1980s to © The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0_2

15

16

2

Relationship Between Port Governance and Climate Change Action

reduce costs, improve services (Chiu and Yen 2015), enhance efficiency and productivity (Zhang et al. 2019), and increase private sector involvement in the financing of infrastructure development (Parola et al. 2013). The World Bank (2007) identified the modernization of port administration and management, liberalization, commercialization, terminal corporatization, PA corporatization, and privatization as port reform options. Tull and Revely (2008) described three main types of port reforms: commercialization, corporatization, and privatization. Zhang et al. (2019) cited devolution as an institutional governance tool and categorized port reform options for this purpose as decentralization, commercialization, corporatization, and privatization. Thus, liberalization, decentralization, commercialization, corporatization, and privatization are the key words in port reform options, of which commercialization represents ports privately managed by a PA with discretion and responsibility; corporatization is where the port is legally turned into a private company while owned by the public and having public bodies, such as the government, as shareholders and thus, is a special form of commercialization (Zhang 2016). The definition and intended concept of the term “privatization” varies in literature; for example, in some literature liberalization, commercialization, corporatization, concession (including leasing and BOT), joint venture, and outsourcing (contracting) are categorized as “privatization to some extent” (Chen et al. 2017). Inoue (2013) reported that the main reason for the varying definitions of privatization is that the privatization of terminal operations is not distinguished from the privatization of the PA itself, while Zhang et al. (2018) argued that the concept of port governance is equivalent to similar concepts, such as devolution of authority and privatization. In this regard, if the introduction of operations from the private sector's perspective and the transfer of even a portion of PA functions to the private sector are defined as privatization in the broad sense, and the transfer of all PA assets and functions to the private sector is defined as privatization in the narrow sense (complete privatization), commercialization and corporatization are the introduction of operations from the private sector’s perspective by PAs themselves, and can be categorized as privatization in the broad sense. The typology of governance models that emerge under various port reforms has been summarized by Ferrari et al. (2015) and Parola et al. (2013), based on World Bank typology (private port, Landlord, Tool port, and public port in order of strongest private sector involvement). In the 1990s, the keywords for the first wave of port reforms were devolution, decentralization, and privatization. However, since 2000, reforms have evolved in their complexity and diversity and hence, can no longer be grouped under specific keywords, and a ports have moved away from a typical governance model (Brooks et al. 2017). Although the typology by the World Bank (2007) is widely accepted by both researchers and practitioners (Ferrari et al. 2015), it should be modified to fit regional characteristics and cultural backgrounds; this classification is also adopted in this book. The combinations of ownership and operations based on the classification are presented in Table 2.1. Although port reform originated in the United Kingdom (Baird 2013), the “full privatization” model common in the UK is not a standard model worldwide; contrarily, most of the large container ports worldwide follow a “landlord” model (Zhang 2016; Monios 2019). The landlord model facilitated the acceleration of privatization of terminal operations with the application of concessions (Notteboom 2007). Farrell (2012) classified the participating entities in operations as global terminal operators (GTOs), regional TOs, stevedores, shipping companies, logistics companies, construction companies, manufacturers, real estate companies, conglomerates, financial institutions, and public entities. Parola et al. (2013) categorized two types of entry strategies: (1) “direct public private partnership (PPP) entry,” which involves public–private negotiations, such as concessions and leases, and (2) “indirect PPP entry,” which involve shareholding transactions in existing PPPs; additionally, Parolaet al. (2013) empirically demonstrated that indirect PPPs have become mainstream since the early 2000s, although direct PPPs were dominant until the late 1990s, the earliest stage of port

2.3 Case Study: Japan

17

Table 2.1 Classification of port governance models Smaller < Private sector involvement > Greater Public port

Tool port

Landlord port

Private port

Regulation

Public

Public

Public

Private

Land

Public

Public

Public

Private

Infrastructure

Public

Public

Public

Private

Superstructure

Public

Public

Private

Private

Operation

Public

Private

Private

Private

Source Prepared by the author with reference to previous studies

privatization. Currently, GTOs have strengthened their market dominance through foreign investments via repeated mergers and acquisitions, and financial institutions have also entered the port operation industry (Notteboom and Rodrigue 2012; Parola et al. 2013; Monios 2019). Concessions in ports have also been discussed in previous studies, with Notteboom (2007) highlighting the procedures and clauses included in concession contracts under the landlord model and Theys et al. (2010) summarizing the discussion points at the pre-bid, bid, and post-bid stages. However, although concession contracts are effective as a strategic tool in the landlord model, there is no universal method for concession procedures (Ferrari et al. 2015). Farrell (2012) stated that procedures and leasing schemes are with region, country, and even within ports. There are two important factors in the relationship between port governance and climate change actions. One is the changing role of PAs. Because their devolution has been the most important aspect of port reform (Debrie et al. 2007), it has resulted in changes in the PAs themselves (Brooks et al. 2017). Historically, PAs originally held the responsibility of landlords, regulations, and operations as central or local government bodies (Brooks et al. 2017), but with the devolution of authority, decentralization, privatization following port reforms, and diversification of stakeholders, the community manager function has also been emphasized in addition to the traditional role of PAs (Verhoeven 2010). This has led to PAs playing an active role in environmental countermeasures. PAs are expected to be able to encourage the cooperation of many stakeholders and manage the entire process. Furthermore, concession contracts in the landlord model function as environmental governance (Van den Berg and De Langen 2014; Lam and Notteboom 2014; Notteboom and Lam 2018). Environmental clauses are included in 85% of all concession contracts between PAs and TOs (Notteboom and Verhoeven 2010). TOs win concession contracts based on environmental countermeasures, which become contractual obligations during operation. Thus, PAs are involved in management and supervision (Lam and Notteboom 2014). Therefore, the adoption of the landlord model is an important driver of climate change countermeasures in ports (Munim et al. 2020; Sornn-Friese et al. 2021), and concession contracts are an important policy tool to achieve environmental objectives (Martínez-Moya et al. 2019). This implies a significant role for PAs in environmental countermeasures in terms of concession contracts.

2.3

Case Study: Japan

Leading ports that have developed environmentally friendly strategies include Antwerp, Rotterdam, Los Angeles/Long Beach, Hamburg, and Singapore ports (Acciaro 2015; Aregall et al. 2018); notably, these ports have tended to adopt an enforcement approach with regard to implementation of climate change countermeasures (Lam and Notteboom 2014). Ports in the United States and European

18

2

Relationship Between Port Governance and Climate Change Action

countries have adopted regulatory approaches, such as emissions trading schemes (Arimura et al. 2019). However, a characteristic feature of Japan's environmental policy is that it relies on a voluntary approach. A bill that included the introduction of a cap-and-trade system in Japan was repealed in 2012 due to opposition from industrial stakeholders (Gokhale 2021) Therefore, although there exists a legal reporting system, a regulatory framework for corporate CO2 emissions in Japan is lacking. However, it has been demonstrated that CO2 emissions harm evaluations expressed by investors and rating agencies, and consequently corporate value (Safiullah et al. 2021); therefore many Japanese companies are under strong social pressure to achieve decarbonization by participating in international initiatives such as the Science Based Targets and Renewable Energy 100. Furthermore, in Japan, due to the restructuring of financial markets in 2022, the disclosure of Scope 1 and 2 emissions is mandatory, and that of Scope 3 emissions is recommended for companies listed on the highest rank, the prime market (Tokyo Stock Exchange 2022). These changes will likely accelerate the actions of shippers toward decarbonization. In this environment, Japan's port governance system is key to whether port stakeholders actively promote decarbonization activities. The actor responsible for and implementing climate change countermeasures and taking the initiative for the same in ports primarily depends on the port governance structure of the country. In this section, the relationship between port governance and climate change actions in ports is discussed using Japan as a case study. In Japan, the Ports and Harbors Act was enacted in 1950, soon after the end of World War II. However, because under the strong influence of the General Headquarters of the Allied Forces, the national government had supervisory authority over ports, and local governments were tasked with management (Sugimura 2020). The PA system, modeled after those in Europe and the United States, was assumed to be the main port management system, but it was not widely implemented; thus, port management has been conducted by local governments until now, which is a characteristic of the port governance system in Japan (Sugimura et al. 2022b) In this regard, Inoue (2018) noted that “it might be a surprise that ports of Japan, including the largest ones, continue to be managed under the general administration of local governments.” While containerization in Japan began in 1969, Foreign Trade Terminal Public Corporations were established at large ports in 1967 because there was a need to finance enormous expenses for the construction of container terminals. The container terminals developed by the corporations were leased to shipping lines and operated as dedicated terminals. In 1982, Foreign Trade Terminal Public Corporations were taken over and dissolved by port terminal public corporations under the concept of shifting the focus of port policy from development to operation because it was commonly believed that there would be no further increase in the size of container ships at that time. Over-Panamax ships were deployed in the 1990s, and most Asian countries responded quickly to these changes, but Japan failed to do so (Inoue 2018). The international competitiveness of Japanese container ports deteriorated. (Sugimura 2020). In response, hub and core international ports were designated by the central government in 1995, which determined the national layout of container ports. However, as local PMBs preferred routes to foreign ports such as Busan rather than domestic feeder transport, the concentration of cargo in international hub ports did not improve. As the international competitiveness of container ports in Japan continued to decline the Super Hub Port Policy was subsequently proposed in 2001. This policy aimed to develop mega-terminals and train a single domestic TO per port to achieve port services that would exceed those of major Asian ports. By this time, port reforms were underway worldwide and many countries had adopted the landlord model (Monios 2019), leading to the emergence of GTOs (Notteboom and Rodrigue 2012). Because countering global expansion by GTOs constituted the background of this policy, the fact that foreign GTOs have rarely entered the Japanese market in the wake of this policy is also a unique characteristic of Japanese port governance. Although three ports were designated as super hub ports through a public call for proposals in 2004, integrated management by mega TOs has rarely been achieved (Sugimura 2020).

2.3 Case Study: Japan

19

The International Container Strategy Port Policy was launched in 2009 to further reduce the number of ports in which to invest (Shinohara 2017). Under the motto “Selection and Concentration,” the maintenance and expansion of the trunk routes to Japan and recovery of transships in Busan and other ports were set as goals, to achieve integrated management by port operating companies with a private sector perspective. The port operating company system was applied to large-scale ports, where a port operating company with a private sector perspective operates a group of wharves by leasing them from the PMB, thereby providing subleases to TOs. This two-tiered structure, wherein the PMB leases all container terminals to a port operating company (POC) that subsequently leases the terminals to the TOs, is unique to Japan. Through a public call for proposals, two ports were selected, and strategic international ports were specified as those constituting the highest level in the revision of the 2011 Ports and Harbors Act. The central government invests in port management companies of strategic international ports. As described above, although port reform has been promoted in Japan, port governance has several characteristics compared to global trends, such as management by local governments, a unique POC system, and the absence of GTOs. The current port governance model lies somewhere between the landlord and tool models at ports, subject to the port operating company system. Moreover, there is an increasing trend toward private sector involvement. However, practically no competition exists in the selection process of POCs, and leases to TOs are non-competitive dedicated leases (Notteboom et al. 2017). Therefore, even in the case of private operations, the operator is not selected competitively. This governance structure has had a significant impact on implementation of climate change countermeasures. Non-corporatized PMBs do not have CSR or ESG perspectives as motivations for their environmental activities, and POCs, which are still strongly influenced by PMBs in their complex governance structures, have not yet fully developed their private sector perspective. Furthermore, the fact that local governments are PMBs may result in the ineffectiveness of the central government's national strategic policies, as local interests are prioritized and there is a lack of a business perspective in port operations (Sugimura et al. 2022b). Therefore, from the perspective of inter-port competition, it is unlikely that climate change countermeasures that would lead to higher port fees will be introduced on a port-by-port basis, regardless of the international trends. The absence of a GTO also negates the possibility of domestic TOs taking initiative in climate change countermeasures unlike the situation in other advanced countries with GTOs. As terminal leases to TOs have traditionally been in the form of dedicated leases rather than concession agreements (Sugimura 2020), it is unlikely that environmental initiatives at the expense of the TOs will be included in lease agreements where the PMB does not have strong bargaining power. Thus, the Japanese port governance model is not characterized by the voluntary and proactive promotion of climate change countermeasures at each port or container terminal (CT). Instead, the government's initiatives regarding climate change countermeasures are more important than that in other countries. PMBs have a significant influence on the decarbonization activities of POCs, TOs, and port users (Alamoush et al. 2020). Furthermore, previous studies have pointed out the increasing role of port managers (Acciaro et al. 2014; Lam and Li 2019). However, given the characteristics of Japan’s legislation or port governance regarding decarbonization, the probability of PMBs taking the lead in policy development and environmental investment is not high at the moment. As port management is a part of local government administration, the budget needs to be approved by the local council; however, our understanding of climate change countermeasures in ports is remains insufficient. For example, the financial resources for economic incentives for POCs, TOs, and port users and competitive relationship with neighboring ports for regulatory measures ahead of other ports are likely to be problematic. Ports in Japan are typically publicly owned and operated terminals (Farrell 2012), and

20

2

Relationship Between Port Governance and Climate Change Action

given the strength of public involvement, a regulatory approach, such as the legalization of climate change countermeasures, would make it easier to promote countermeasures from the perspective of PMBs because it can shift accountability to the central government. POCs, which are a characteristic of Japan, are decision makers regarding the introduction of environmental-friendly cargo-handling machinery and are the entities that lease terminals to TOs; therefore, they can be more active promoters than PMBs. If they adopt a corporate strategy that considers climate change countermeasures as the basic source of competitiveness, they may be able to actively introduce environmental-friendly cargo-handling equipment and incorporate environmentalfriendly clauses in lease contracts to TOs. In addition, as a stock company, it could also have a motive to promote climate change countermeasures from CSR and ESG perspectives. However, because the two-tier structure described above leaves the authority of the PMB over operations, they are still far from a private sector perspective, and do not have a global network such as corporatized PAs and GTOs overseas (Sugimura et al. 2022b). The key is to manage the port operations from the perspective of a full-fledged private sector but in the context of discretionary authority. Accordingly, the central government is also in a position to take the initiative because it is a shareholder in the POCs of strategic international ports and has the authority to consent to the managerial structure at hub international ports. Given Japan's port governance, central government initiatives are important from multiple perspectives for promoting climate change countermeasures in ports. Given the characteristics of Japan's environmental policy, although there are regulatory and incentive approaches (Li et al. 2020), the latter approach, which focuses on financial support, is more realistic for POCs and TOs in promoting countermeasures. It is important to clarify the government's stance by strongly promoting decarbonization port policies and providing subsidies for the introduction of countermeasures and financial support for the technological development of countermeasures to increase the feasibility of such policies. According to PMB, as port management is part of the local government administration, a regulatory approach by the central government would be easier to promote the countermeasures. Although POCs can be more proactive promoters than PMBs, it is necessary to establish a fullfledged management system from a private sector perspective with discretionary authority according to the purpose of the POC system, under the initiative of the central government. For climate change countermeasures to be actively promoted at each port, the central government needs to play a key role in reforming governance, such as the corporatization of PMBs and further strengthening the authority of POCs, so that internationally oriented port management professionals can be substantially involved in port operations. However, given the path dependency often discussed in port reform (Notteboom et al. 2013), a drastic change from the current situation is not realistic and would be a long-term challenge.

2.4

Stakeholders in Ports and Their Roles on Climate Change Actions

In the previous section, the relationship between Japan's port governance system and the promotion of climate change countermeasures has been discussed. In a country such as Japan, where a regulatory approach to environmental policy is not adopted and port governance does not function as environmental governance, the promotion of climate change countermeasures at ports is highly dependent on the voluntary efforts and cooperation of stakeholders. Climate change countermeasures in ports are complex and require stakeholder cooperation due to their role in port operations and also the impact that the countermeasures would have on them. To confirm this, Table 2.2 lists the stakeholders involved in some of the representative climate change countermeasures at the ports listed in Table 1.1 in Chap. 1.

2.4 Stakeholders in Ports and Their Roles on Climate …

21

Table 2.2 Typical climate change countermeasures and relevant stakeholders Countermeasure example

Type

Shipper

PA

Port operating company

Development of onshore power supply facilities

Hard

C

A

A

Reduction of traffic congestion at gate

Soft

C

Modal shift/split

Soft

C

A

A

Development of large-scale energy storage facilities

Hard

C

A

A

C

Introduction of smart grid

Hard

C

A

A

C

Renewable energy utilization

Hard

C

A

A

C

Development of hydrogen stations

Hard

C

A

A

C

Introduction of electrified or fuel cell cargo-handling machinery

Hard

C

A

A

Efficient operation of cargo-handling machines

Soft

C

Implementation of efficient navigation within the port

Soft

C

Introduction of environmental-friendly ships

Hard

C

Improvement of efficiency of intra-port drayage

Soft

C

Introduction of environmentally friendly vehicles

Hard

C

A

Terminal operator

Shipping company

Trucker

B A

B B

B

C

A A

B A

A

A

B A

A implementation entity; B supporting entity; C entity that may be affected by the use of the port in terms of cost

As shown in Table 2.2, many countermeasures involve multiple stakeholders, except for those that are voluntarily implemented independently under the goals of the respective industries, such as the introduction of environmentally friendly ships and vehicles by shipping companies and trucking companies. In particular, for hardware countermeasures, the owner of the facility is primarily responsible for implementation; however, the cost burden is a difficult problem to solve. For example, if onshore power supply facilities are installed, shippers, shipping companies, PMBs, and POCs can bear the cost of installation and an increase in energy charges. Similarly, the electrification of cargo-handling machinery could cost shippers in the form of higher port charges. If the cost is not transferred to port charges to avoid an impact on port competitiveness, facility owners could face financial difficulties, and if the cost is transferred to port charges, shippers and shipping companies could be discouraged from calling ports, resulting in the loss of port competitiveness. A previous study pointed out that a balance between CO2 emission reductions and fair competition is necessary because if environmental regulations are applied only to certain ports, the ports and shipping companies concerned will suffer damage in terms of reduced cargo volume and profits (Sheng et al. 2017). For example, in an interview with a PMB, the response was “it is difficult to mandate the use of land-based electric supply systems, which would increase costs for shipping companies, given the competition among ports.” PMBs, POCs, and TOs face the impact of climate change countermeasures on port competitiveness and their financial problems. Soft countermeasures that require stakeholders to bear the costs may also affect port competitiveness if introduced in a regulatory manner and may not be effective from the perspective of increased costs if left to the

22

2

Relationship Between Port Governance and Climate Change Action

Fig. 2.1 Relationships among stakeholders in Japanese ports

autonomy of the related stakeholders. The above points can be confirmed by examining the stakeholders and their relationships in Japanese CTs where the POC system has been introduced, as shown in Fig. 2.1. As shown in Fig. 2.1, the relationship among stakeholders is that shippers pay costs and demand transportation services and port services, while other stakeholders conduct corporate activities using their respective charges as revenues. Except for PMBs, they are private enterprises that conduct their business activities to maximize profits. Shippers select the most suitable shipping company for their supply chain, shipping companies select CTs that can handle more cargo, and shipping companies select PMBs and POCs. For PMBs and POCs, being selected by shipping lines indicates the port's competitiveness. Port selection is a complex process involving multiple criteria such as cost, time, reliability, and intermodal connectivity (Notteboom et al. 2017). However, in recent years, CO2 emissions during transportation have become another pressing objective factor for shippers when choosing a mode of transport (Qi et al. 2022), with cost maintaining its importance (Chang et al. 2008). Therefore, if climate change countermeasures increase port charges, the possibility of selecting that port will be reduced. However, if CO2 emissions become an important criterion for port selection, as shippers become more aware of the need for decarbonization, higher port fees may not be an issue. In terms of a port's competitiveness, determining the bearer for the expenses of climate change countermeasures is a challenging issue. Therefore, the action principles of each stakeholder regarding climate change actions are important. In Japan, the situation of typical stakeholders concerning decarbonization is as follows: • Shippers: Currently, there is no regulatory framework for corporate CO2 emissions in Japan, only a legal reporting system. However, many Japanese companies are under strong social pressure to decarbonize by participating in international initiatives. The disclosure of Scope 1 and 2 emissions is mandatory, and that of Scope 3 emissions is recommended for companies ranked highest on the prime market, due to the restructuring of the financial markets in 2022. Under these circumstances, as more companies commit to reducing supply chain emissions, CO2 emissions from maritime transportation and ports could become targets. • PMB: As indicated in previous studies, PMB is the most important actor in climate change action in ports, as it is responsible for developing decarbonization strategies and implementing public-

2.4 Stakeholders in Ports and Their Roles on Climate …





• •

23

side countermeasures. Furthermore, it is in a position to promote climate change countermeasures for other stakeholders through regulations and incentives, but it also has to consider the impact on the port's competitiveness. POC: Because the POC leases a group of CTs from the PMB and operates them, it has authority and responsibility similar to that of the PMB for decarbonization. As a private company, it is subject to a law-based reporting system; however, there are some cases wherein the electrification of cargo-handling machinery owned by the company is abandoned because of the cost of the initial investment. TO: The TO is the main CO2 emitter in the CT as its primary role of cargo operations require the use of cargo-handling machinery, which is considered to be the main source of emissions in CT. However, there is no industry-wide CO2 emission reduction target, and each company aims to reduce CO2 emissions through its initiatives. Nevertheless, because many companies are relatively small, they do not participate in international initiatives, nor do they have the financial strength to make aggressive investments to decarbonize their operations. Shipping companies: The shipping industry aims to meet the IMO's goals of reducing CO2 emissions by at least 40% by 2030 and GHG emissions by at least 50% by 2050 (both compared to the 2008 levels) (IMO 2018). Trucking companies: The trucking industry has set a goal of 31% reduction in CO2 emissions intensity in 2030 compared to the 2005 level, and aims to increase the proportion of electric vehicles to achieve this goal.

While all stakeholders have the scope to promote climate change countermeasures, it should be noted that, particularly with respect to countermeasures for port service providers (PMBs, POCs, and TOs), if there is a cost increase for stakeholders due to implementation of the countermeasure, without agreement on the sharing of roles in terms of implementation and cost, port calls may be discouraged, and the effectiveness of the countermeasures may be reduced. Therefore, the community manager function of the PMB is more important when implementing countermeasures. If shippers’ behavior change toward decarbonization precedes the implementation of the countermeasures, stakeholders on the supply side of port services may voluntarily adopt climate change countermeasures from the perspective of ensuring port competitiveness and allow them to burden the increased port charges to shippers. Otherwise, the financial constraints of the introducing entity would hinder the introduction of the countermeasures, except for economically feasible countermeasures that do not require shippers to bear increased charges. This may explain why climate change countermeasures are not currently widely adopted in Japanese ports. In this regard, the economic feasibility of countermeasures may be a factor that determines whether climate change countermeasures can be promoted in countries adopting port governance models that do not function as environmental governance unless behavioral change by shippers precedes them. The solution is either a regulatory approach by the government or a PMB, as adopted by developed ports, or incentives, including subsidies. However, the central government’s role is important from the viewpoint of ensuring fairness among ports, while the role of PMBs and POCs is important from the viewpoint of ensuring fairness among stakeholders within a port. The community manager function of the PMB is also important from the viewpoint of chronological alignment of the countermeasures taken by each stakeholder. For example, the onshore power supply requires the same timing of investment on both the land and ship sides, and the electrification of cargo-handling machinery and the use of renewable energy will have the same effect as FC, thus it is necessary to clarify the direction of the port to prevent double investment.

24

2.5

2

Relationship Between Port Governance and Climate Change Action

Conclusions

The lack of a widely adopted climate change countermeasure in the port industry, indicates the need for long-term efforts by all stakeholders to achieve decarbonization targets. While differences in the level of implementation are directly influenced by the intensity of government policies and port governance of the country, the impact of port governance can be complex because of the diversity of its forms. In this chapter, the relationship between port governance and climate change countermeasures was discussed based on a case study of how port governance in Japan affects climate change countermeasures in each port, and what kinds of initiatives are needed based on this relationship. An important point in the relationship between port governance and climate change countermeasures is the emphasis on the community manager function of PAs, as stakeholders have become more diverse due to devolution, decentralization, and privatization of ports following port reforms. Additionally, the concession contracts in the landlord model also function as environmental governance. Therefore, the adoption of the landlord model is an important driver of climate change in ports, and concession contracts are very important as a policy tool to achieve environmental objectives. Notably, this implies that PAs play a significant role in environmental countermeasures from the perspective of concession contracts. Although port reform has been promoted in Japan, several characteristics of port governance, such as management by local governments, a unique POC system, and the absence of GTOs, are in contrast to global trends. There is practically no competition in the selection process of POCs, and leases to TOs are also non-competitive dedicated leases; therefore, even in the case of private operations, the operator is not competitively selected. These characteristics of Japan's port governance model make it difficult for climate change countermeasures to be actively promoted at individual ports and CTs, and it has become clear that central government initiatives are even more important than in other countries. In a country such as Japan, where a regulatory approach to environmental policy is not adopted and port governance does not function as environmental governance, the promotion of climate change countermeasures at ports is highly dependent on the voluntary efforts and cooperation of stakeholders. Particularly, with respect to countermeasures for port service providers (PMBs, POCs, and TOs), if there is a cost increase for stakeholders due to implementing countermeasures, without consensus on the sharing of roles in terms of implementation and cost, port calls may be discouraged, and the effectiveness of the countermeasures may be reduced. Therefore, the community manager function of the PMB is particularly significant when implementing countermeasures. If shippers’ behavior change toward decarbonization precedes the implementation of the countermeasures, stakeholders on the supply side of port services may voluntarily adopt climate change countermeasures from the perspective of ensuring port competitiveness and allow them to place the burden of the increased port charges on shippers. Otherwise, the financial constraints of the introducing entity would hinder the introduction of the countermeasures, except for economically feasible countermeasures that do not require shippers to bear increased charges. In this regard, the economic feasibility of countermeasures may be a factor that determines whether climate change countermeasures can be promoted in countries adopting port governance models that do not function as environmental governance unless behavioral change by shippers precedes them. Although this chapter focuses on Japan as a case study, the relationship between port governance and climate change countermeasures requires similar considerations for each country, as port governance systems differ from country to country. Thus, the approach used in this study can be used as a reference.

References

25

References Acciaro M (2015) Corporate responsibility and value creation in the port sector. Int J Log Res Appl 18:291–311 Acciaro M, Ghiara H, Cusano MI (2014) Energy management in seaports: a new role for port authorities. Energy Policy 71:4–12 Alamoush AS, Ballini F, Ölçer AI (2020) Ports’ technical and operational measures to reduce greenhouse gas emission and improve energy efficiency: a review. Marine Poll Bull 160 Aregall MG, Bergqvist R, Monios J (2018) A global review of the hinterland dimension of green port strategies. Transp Res Part D Trans Environ 59:23–34 Arimura TH, Kaneko S, Managi S, Shinkuma T, Yamamoto M, Yoshida Y (2019) Political economy of voluntary approaches: a lesson from environmental policies in Japan. Econ Anal Policy 64:41–53 Baird AJ (2013) Acquisition of UK port equity funds. Res Transp Bus Manag 8:158–165 Brooks MR, Cullinane KPB (2007) Devolution, port performance and port governance. Elsevier, Oxford Brooks MR, Cullinane KPB, Pallis AA (2017) Revisiting port governance and port reform: a multi-country examination. Res Transp Bus Manag 22:1–10 Chang Y-T, Lee S-Y, Tongzon JL (2008) Port selection factors by shipping lines: different perspectives between trunk liners and feeder service providers. Mar Policy 32:877–885 Chen PSL, Pateman H, Sakalayen Q (2017) The latest trend in Australian port privatization: drivers, processes and impacts. Res Transp Bus Manag 22:201–213 Chiu R-H, Yen DC (2015) Application of organizational life cycle theory for port reform initiatives in Taiwan. Res Transp Manage 14:14–24 Debrie J, Gouvernal E, Slack B (2007) Port devolution revised: the case of regional ports and the role of lower tier governments. J Transp Geogr 15:455–464 Debrie J, Lavaud-Letilleul V, Parola F (2013) Shaping port governance: the territorial trajectories of reform. J Transp Geogr 27:56–65 Farrell S (2012) The ownership and management structure of container terminal concession. Marit Policy Manag 39:7–26 Ferrari C, Parola F, Tei A (2015) Governance models and port concession in Europe: commonalities, critical issues and policy perspectives. Transp Policy 41: 60–67 Gokhale H (2021) Japan’s carbon tax policy: limitations and policy suggestions. Curr Res Environ Sustain 3 IMO (2018) Initial IMO strategy on reduction of GHG emissions from ships. Resol MEPC 304(72) Inoue S (2013) Port management challenges in the supply chain age: strategic effectiveness and framework of port corporatization. Transp Policy Stud Rev 15:19–30 Inoue S (2018) Realities and challenges of port alliance in Japan -ports of Kobe and Osaka. Res Transp Bus Manag 26:45–55 Lam JSL, Li KX (2019) Green port marketing for sustainable growth and development. Transp Policy 84:73–81 Lam JSL, Notteboom T (2014) The greening of ports: a comparison of port management tools used by leading ports in Asia and Europe. Transp Rev 34(2):169–189 Li Y, Li J, Wei F, Huang Q (2020) CO2 emission performance evaluation of Chinese port enterprises: a modified metafrontier non-radial directional distance function approach. Transp Res Part D Transp Environ 89 Martínez-Moya J, Vazquez-Paja B, Maldonado JAG (2019) Energy efficiency and CO2 emissions of port container terminal equipment: evidence from the port of Valencia. Energy Policy 131:312–319 Ministry of the Environment (2021) List of companies involved in TCFD, SBT, and RE100 (as of December 31, 2021) Monios J (2019) Polycentric port governance. Transp Policy 83:26–36 Munim ZH, Sornn-Friese H, Dushenko M (2020) Identifying the appropriate governance model for green port management: applying analytic network process and best-worst methods to ports in the Indian Ocean Rim. J Clean Prod 268 Notteboom T (2007) Concession agreement as port governance tools. Res Transp Econ 17:437–455 Notteboom T, De Langen P, Jacobs W (2013) Institutional plasticity and path dependence in seaports: interactions between institutions, port governance reforms and port authority routines. J Transp Geogr 27:26–35 Notteboom T, Lam JSL (2018) The greening of terminal concessions in seaports. Sustainability 10:3318 Notteboom TE, Parola F, Satta G, Pallis AA (2017) The relationship between port choice and terminal involvement of alliance members in container shipping. J Transp Geogr 64:158–173 Notteboom T, Rodrigue JP (2012) The corporate geography of global container terminal operators. Marit Policy Manag 39:249–279 Notteboom T, Verhoeven P (2010) The awarding of seaport terminals to private operators: European practices and policy implications. Europ Transp 45:83–101

26

2

Relationship Between Port Governance and Climate Change Action

Parola F, Notteboom T, Satta G, Rodrigue JP (2013) Analysis of factors underlying foreign entry of terminal operators in container ports. J Transp Geogr 33:72–84 Poulsen RT, Ponte S, Sornn-Friese H (2018) Environmental upgrading in global value chains: the potential and limitations of ports in the greening of maritime transport. Geoforum 89:83–95 Qi Y, Harrod S, Psaraftis HN, Lang M (2022) Transport service selection and routing with carbon emissions and inventory costs consideration in the context of the belt and road initiative. Transp Res Part E Logist Transp Rev 159 Safiullah M, Kabir MN, Miah DM (2021) Carbon emissions and credit ratings. Energy Econ 100 Sheng D, Li Z-C, Fu X, Gillen D (2017) Modeling the effects of unilateral and uniform emission regulations under shipping company and port competition. Transp Res Part E Logist Transp Rev 101:99–114 Sornn-Friese H, Poulsen RT (2016) Ports as sustainability hubs in global maritime supply chains: a content analysis of the external business communication of the worlds largest ports. Paper Presented at the International Association of Maritime Economist Sornn-Friese H, Poulsen RT, Nowinsk AU, de Langen P (2021) What drives ports around the world to adopt air emissions abatement measures? Transp Res Part D Transp Environ 90 Sugimura Y (2020) Public-private partnerships in Japan's cruise terminal operations. Res Transp Business Manage Sugimura Y, Okada T, Kuwae T, Mito Y, Naito R, Nakagawa Y (2022a) New possibilities for climate change countermeasures in ports: organic carbon containment and creation of blue carbon ecosystems through beneficial utilization of dredged soil. Marine Policy 141 Sugimura Y, Wakashima H, Liang Z, Shibasaki R (2022b) Logistics strategy simulation of second-ranked ports on the basis of Japan’s port reforms: a case study of Hakata Port. Marit Pol Manage Theys C, Notteboom T, Pallis AA, DeLangen PW (2010) The economics behind the awarding of terminal in seaport: towards a research agenda. Res Transp Econ 27:37–50 Tokyo Stock Exchange (2022) Compliance status with the corporate governance code (as of July 14, 2022). https://www.jpx. co.jp/equities/listing/cg/tvdivq0000008jdy-att/nlsgeu000006jzbl.pdf Tull M, Revely J (2008) Introduction: port reform in the Asia-Pacific Region. In: Tull M, Reveley J (Eds) Port privatisation: the Asia-Pacific experience. Edward Elgar, Cheltenham Van den Berg R, De Langen PW (2014) An exploratory analysis of the effects of modal split obligations in terminal concession contracts. Int J Shipp Transp Log 6:571–592 Verhoeven P (2010) A review of port authority functions: toward a renaissance? Marit Policy Manag 37:247–270 World Bank (2007) Port reform toolkit second edition module 3 alternative port management structures and ownership models. https://ppiaf.org/sites/ppiaf.org/files/documents/toolkits/Portoolkit/Toolkit/pdf/modules/03_TOOLKIT_ Module3.pdf Zhang J (2016) Quasi-landlord port financing in China: features, practice and a contract theory analysis. Transp Res Part A 89:73–88 Zhang Q, Geerlings H, El Makhloufi A, Chen S (2018) Who governs and what is governed in port governance: a review study. Transp Policy 64:51–60 Zhang Q, Zhen S, Geerlings H, El Makhloufi A (2019) Port governance revisited: how to govern and for what purpose? Transp Policy 77:46–57

Part II Countermeasures for Cargo-Handling Machinery

Cargo-handling machinery RCs have been identified as the main sources of CO2 emissions in CTs. Technological measures such as electrification, hybridization, and use of fuel cells, as well as measures to improve the efficiency of cargo-handling operations, including automation, have been researched as climate change countermeasures for cargo-handling machinery. However, these countermeasures have not been widely adopted in practical terms. High initial investments, expensive countermeasures, and challenges in implantation on a business basis are a major reason for insufficient progress in climate change actions in ports. Additionally, the lack of previous studies based on empirical findings, and the gap between research and practice does not provide adequate support to decision makers for ports. Part II introduces empirical CO2 emission reduction effects of the countermeasures using Hakata Port, which is leading Japan’s climate change countermeasures by electrification of tire-type transfer cranes and hybridization of straddle carriers, as a case study. Economic feasibility analysis of the implementation of these cargo-handling machinery at Hakata Port is also presented.

3

CO2 Reduction Effects by Electrification of Cargo-Handling Machinery

3.1

Introduction

Port-related CO2 emissions are primarily attributed to vessels, port activities, or port hinterland transport (Chen et al. 2013; Sornn-Friese et al. 2021); however, the substantial number of hinterland transport vehicles and vessels traveling into ports also account for a large share of the CO2 emissions from vehicles. However, CO2 emissions from vehicles and ships are likely to decrease in the future as society moves toward decarbonization. Under the Paris Agreement, anthropogenic GHG emissions must be reduced rapidly and and by a significant proportion to limit the increase in the global average temperature to less than 1.5 °C above pre-industrial levels (IPCC 2021). However, the NDCs set by each country are insufficient to achieve the target, and rapid action is also required, as global GHG emissions in 2030 indicate that global warming is likely to exceed 1.5 °C during the twenty-first century (IPCC 2022). Consequently, Japan and other countries have decided to implement stronger climate change actions to meet the targets and become CN. In response, the shipping and vehicle industries as well as individual companies are aiming to accelerate their efforts to reduce CO2 emissions, including technological innovation. Considering this, reducing CO2 emissions from port activities should be the target of port stakeholders, which can be directly achieved through their efforts. In addition, CO2 emissions are likely to become an important factor in port competitiveness. Selecting the transport mode involves several decisions, especially by logistics service providers and shippers; particularly, shippers can specify requirements that constrain modal choice (Holguín-Veras et al. 2021), and CO2 emissions during transportation have become a major deciding factor for shippers when selecting transportation modes (Qi et al. 2022). For example, Apple Inc., one of the world’s largest companies, places heavy weightage on carbon emissions when selecting a transportation service (Apple 2022). Hence, ports with low CO2 emissions are more likely to be selected by shippers. The behavioral change of shippers toward carbon neutrality has become an important environmental change for reducing CO2 emissions at ports. Regarding CO2 emissions in CT, cargo-handling machinery and reefer containers have been identified as the main sources of emissions in previous studies (Iris and Lam 2019; Wang et al. 2020; Xing et al. 2020; Di Ilio et al. 2021). For example, according to Martínez-Moya et al. (2019), the breakdown of electricity consumption at the Noatum CT in the port of Valencia is 33.9% and 47.9% for quay cranes and reefer containers, respectively; the breakdown of fuel consumption for RTGs and yard trucks is 63.2% and 32.6%, respectively; and the breakdown of CO2 emissions for RTGs, yard trucks, reefer containers, and quay cranes is 44.9%, 23.2%, 13.8%, and 9.8%, respectively. © The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0_3

29

30

3

CO2 Reduction Effects by Electrification of Cargo-Handling Machinery

Technological measures, such as electrification and hybridization (Martínez-Moya et al. 2019; PIANC 2019; Sifakis and Tsoutsos 2021), the use of fuel cells (Iris and Lam 2019; Di Ilio et al. 2021), and measures to improve the efficiency of cargo-handling operations, including automation (He et al. 2015; Sha et al. 2017) have been explored to reduce emissions from cargo-handling machinery. However, CO2 emissions from CTs have rarely been estimated (Martínez-Moya et al. 2019), and the conversion to environmental-friendly cargo-handling machinery requires time owing to technical and economic challenges (Zhong et al. 2019). The current port governance environment does not allow for practical identification of CO2 emissions from port cargo-handling machinery and implementation of corresponding countermeasures; this can primarily be attributed to the lack of studies based on empirical findings and the gap between research and practice, which does not adequately support decision makers in ports (Bjerkan and Seter 2019). This chapter introduces the CO2 emission reduction effects of the countermeasures using Hakata Port, which is leading Japan’s climate change countermeasures by electrification of tire-type transfer cranes (RTG) and hybridization of straddle carriers (S/C), as a case study.

3.2

Port Environmental Policies in Japan and Development of Environmental Countermeasures at Hakata Port

In this chapter, the reason why Hakata Port was selected as a case study is first explained to introduce the effect of electrification of RTGs and hybridization of S/C using actual data from Hakata Port in Japan. Hakata Port began operating a CT in 1982 and is currently the sixth largest port in Japan in terms of container handling volume, handling approximately one million TEUs. It is an emerging port, characterized by the flexibility of its policies (Sugimura et al. 2022). For example, the port community system was introduced in 1994 to improve the efficiency of yard operations, and the Hakata Port Logistics IT System was introduced in 2000 to address the chronic gate congestion that occurred in the late 1990s, thereby eliminating the congestion problem. Hakata Port is also one of the first ports in Japan to adopt the perspective of the private sector in the operation of CTs. Hakata Port Terminal Co., Ltd., established as a third-sector company, received a long-term lease for port facilities from Fukuoka City under the Special Area for Structural Reform Act in 2004. However, the greatest feature of this port is its environmental impact. To understand how the environmental initiatives of Hakata Port are leading in Japan, an overview of the history of port policy in Japan is presented. Table 3.1 shows the history of the port environmental policy in Japan. Although environmental considerations have been given to ports since the early 1970s in Japan, until the early 2000s, the port environmental policy was to develop green areas and seashores, reduce water pollution, and form export bases for recyclable resources. The Ports and Harbors Act of 1999 stipulated that environmental conservation should be considered during port development. The first port environmental policy to explicitly address climate change came into effect following the Kyoto Protocol, which initially focused on emission sources such as onshore power supply and the introduction of energy-efficient cargo-handling machinery. However, following the UNEP report in 2009, the focus shifted toward the use of BCEs as sinks. Most recently, carbon-free ports (CFP) (MLIT 2018) and carbon–neutral port (CNP) policies (MLIT 2021) have been adopted to combat climate change. The government’s long-term port policy, PORT2020, aims to create the world's first CFP by controlling CO2 emission source measures via the introduction of offshore wind power generation, low-carbon transportation machinery, onshore power supply systems, and CO2 sink measures such as the utilization of BCEs (e.g., seagrass meadows). CNP policies were introduced in response to the Japanese government’s pledge of carbon neutrality by 2050 that aims for the

3.2 Port Environmental Policies in Japan …

31

Table 3.1 History of Japanese port environmental policies Year

Port environmental policies, etc.

1973

Amendment of the ports and harbors act: start of the port environment improvement project (federally subsidized green areas, beaches, and waste disposal facilities)

1985

Long-term port policy: ports for the 21st century

1988

Commenced development of biologically symbiotic port structures (port facilities with tidal flat and reef functions)

1994

Port environmental policy: new environmental measures for ports, toward the establishment of environmentally friendly ports (eco-ports); establishment of the Eco-port Model Project (focused on the advanced development of port environmental infrastructure)

1997

Adoption of the Kyoto Protocol

1999

Revision of the ports and harbors act (consideration of environmental preservation during port development)

2002

Recycle port policy (port as a base for reverse logistics network)

2005

Greening port administration: basic directions for future port environmental policy [council report] (conservation, restoration and creation of natural environment, formation of circular society, measures against global warming)

2006

Commencement of pilot tests of ship idling stop system (onshore power supply system)

2008

Commencement of pilot tests of biologically symbiotic seawalls (port facilities with tidal flat and reef functions)

2009

Ways of port policy for climate change caused by global warming [council report] (onshore power supply, energy-saving cargo-handling machinery, etc.)

2015

Adoption of the Paris agreement

2018

Long-term port policy: PORT2020, Carbon-free port policy

2020

Carbon–neutral port policy

realization of a wide-scale decarbonized society in Japan by utilizing alternative energy sources (e.g., hydrogen) through the mass import, storage, and utilization of next-generation energy at ports, which serve as nodes and industrial centers for international logistics, as well as through the upgradation of port functions and accumulation of waterfront industries in consideration of decarbonization. Under such policies, the central government has taken the initiative in the form of pilot projects and subsidies for the introduction of onshore power supply facilities and energy-efficient cargohandling equipment (MLIT 2009). The Hakata Port, which is characterized by flexible policy development, was a suitable target for such projects. However, except for the launch of pilot projects and subsidies, a full-scale introduction is yet to be achieved because laws and regulations have not been introduced at each port. Even when overlooking climate change countermeasures at Japanese ports, the degree of introduction and effectiveness of each policy menu is not high at this moment. For example, only three ports in Japan have implemented electrification of RTGs. Therefore, the number of ports wherein empirical evaluations of countermeasure effectiveness can be conducted is limited. Hakata Port aimed to realize an eco-terminal even before the CFP and CNP policies were adopted (Hiyoshi 2012), and has introduced electric RTG (E-RTG), hybrid S/C (HSC), and roof shades (RSs) in the reefer area since 2010. The port is an environmentally advanced port that has been leading the way in environmental policies among Japanese ports, winning the International Association of Ports and Harbors (IAPH)’s Port Environmental Gold Award in 2013 (Hakata Port Terminal Co., Ltd. 2013).

32

3.3

3

CO2 Reduction Effects by Electrification of Cargo-Handling Machinery

Overview of E-RTG and HSC

Hakata Port has two CTs, Kashii Park Port (KPPCT) and Island City (HICCT), as shown in Fig. 3.1, and both CTs utilize different cargo-handling methods, S/C, and transfer cranes (Table 3.2). Six HSCs were introduced in the KPPCT, and all RTGs in the HICCT were electrified and implemented under central government initiatives. Specifically, the electrification of RTGs (4 new units and conversion of 13 units) and introduction of HSC (1 unit) were conducted in 2010 and 2011 as pilot projects by the MLIT, and all necessary expenses, including infrastructure development for E-RTGs, were covered by the central government. Subsequent installations of E-RTGs (four units) and HSCs (five units) were subsidized by the Ministry of the Environment (1/3 of the national expense ratio). Figure 3.2 demonstrates the appearance of the E-RTG and Fig. 3.3 provides an overview. Electrification refers to changing the power source from a diesel engine generator to an onshore power supply system, which requires the conversion of the RTG itself as well as the construction of power supply poles called BUS-BARs and electrical facilities to connect the RTG to the power supply as infrastructure support. Figure 3.4 shows the arrangement of BUS-BAR. Diesel engines are fueled by diesel oil, and even if the power source is electric power, Scope 2 emissions remain, but the difference in CO2 emission coefficients makes it possible to substantially reduce CO2 emissions using E-RTG. Additionally, other benefits, such as improved terminal efficiency owing to reduced refueling and idling time, reduced engine maintenance costs owing to reduced engine operating time, and an

Fig. 3.1 Overview of Hakata port

3.3 Overview of E-RTG and HSC

33

Table 3.2 Overview of the CTs at Hakata port Terminal

KPPCT

HICCT

Quay length

300 m  2

330 m

350 m

Quay depth

−13 m

−14 m

−15 m

Area

22 ha

36 ha

Storage capacity

8973 TEU

22,208 TEU

Cargo-handling machine

Gantry crane 4, S/C 17

Gantry crane 5, RTG 22

Fig. 3.2 E-RTG at Hakata port

CO2

CO2

Fig. 3.3 Overview of E-RTG

34

3

CO2 Reduction Effects by Electrification of Cargo-Handling Machinery

T S H Q

G

P

F

O

E

N

D

M

C

L

B

J

A

Fig. 3.4 Layout of BUS-BAR and electrical facilities

Fig. 3.5 Appearance of HSC

improved work environment owing to the elimination of engine noise, are expected. Note that only when moving between lanes where contact with the BUS-BAR is not possible is electric power not available; hence, the conversion type uses diesel fuel, while the fully electrified type uses a secondary battery (lithium-ion battery) as its power source. Figure 3.5 depicts the appearance of the HSC and Fig. 3.6 shows a diagrammatic outline. The hybrid system comprises a diesel-fueled generator and a lithium-ion battery, and the regenerative energy acquired during operation is stored in a storage battery. When the load on the generator is large, such as when the HSC starts or when the spreader is being hoisted, the storage battery is discharged to reduce the fuel consumption and support the motor output. This reduces fuel consumption, thereby enabling the reduction of CO2 emissions.

3.4 CO2 Emission Reductions by Introducing …

35

Fig. 3.6 Overview of HSC

3.4

CO2 Emission Reductions by Introducing E-RTG and HSC

3.4.1 Changes in CO2 Emissions of the Entire CT Due to Climate Change Countermeasures Hakata Port has accumulated data on the amount of electricity and fuel used by each emission source in CTs. Figure 3.7 shows a comparison of CO2 emissions between FY2009 (before the implementation of climate change countermeasures) and FY2020 (latest available data), calculated based on the data and CO2 emission factors (Ministry of the Environment 2020; Ministry of the Environment 2021). Table 3.3 lists the number of cargo-handling machines and cargo-handling volume, which are the assumptions for CO2 emissions as well as CO2 emissions per TEU. Compared to FY2009, while the number of cargo-handling machines and cargo-handling volume increased significantly, the overall CO2 emissions decreased, with a 36.5% decrease in CO2 emissions per TEU. The value of 8.26 kg-CO2/TEU for Hakata Port is smaller than the value of 12 kg CO2/TEU for Tokyo Port (Tokyo Port Terminal Co., Ltd. 2020). In terms of emissions by source, both the absolute amount and percentage of RTGs decreased significantly, and the S/C also decreased. However, gantry cranes and prime movers have increased as container handling volume increased. Although the outcomes of the countermeasures were clear, CO2 emissions increased per the increase in cargo volume in targets where no countermeasures were implemented, and CO2 emissions also increased in reefer areas where no countermeasures were implemented in earnest.

3.4.2 CO2 Reduction Effect of E-RTG and HSC Of the 22 RTGs in the HICCT, 9 were fully electrified, and 13 units were converted from diesel engine generators to onshore power supply systems. Because the current energy consumption is the value after electrification, the effect of installing E-RTGs in the entire CT is calculated using the following equation and the values listed in Table 3.4. EFERTG ¼ 1 

with EERTG without EERTG

ð3:1Þ

36

3

CO2 Reduction Effects by Electrification of Cargo-Handling Machinery

8,000 7,000

355 (4.7%) 438 (5.8%)

347 (5.0%) 402 (5.8%)

Light mast

862 (12.5%)

Office building

681 (9.9%)

Prime mover

513 (6.8%) 6,000 5,000

611 (8.1%)

1,667 (22.1%)

Quay crane

4,000 1,900 (27.5%) 3,000

Reefer container

1,683 (22.3%) 712 (10.3%)

RTG

2,000 S/C 1,000

2,286 (30.3%)

1,992 (28.9%)

2009

2020

0 Fig. 3.7 Comparison of CO2 emissions before and after implementation of climate change countermeasures

Table 3.3 Comparison of conditions before and after implementation of climate change countermeasures FY2009

FY2020

Increase or decrease (%)

Gantry crane (unit)

9

9

0

RTG (unit)

14

22

+57.1

S/C (unit)

17

17

Prime mover (unit)

24

30

0

Container handling volume (TEU)

581,061

835,375

CO2 emissions (t-CO2)

7553

6987

−8.7

CO2 emissions per TEU (kg-CO2/TEU)

13.00

8.26

−36.5

+25.0 +43.8

f with e EERTG ¼ ef e  CERTG þ ef f  CERTG

ð3:2Þ

without EERTG ¼ ef f  uhRTG  N h

ð3:3Þ

Of the 17 S/C in the KPPCT, 6 units were of the hybrid type, and 11 units were of the normal type. Because the current energy consumption is the value after the introduction of HSC, the effect of the introduction of HSC on the entire CT is calculated using the following equation and the values listed in Table 3.5.

3.4 CO2 Emission Reductions by Introducing …

37

Table 3.4 Parameters used to calculate the effect of E-RTG Item

Explanation

EFERTG

Ratio of CO2 emission reduction by electrification (%)

with EERTG

CO2 emissions with electrification (t-CO2/year)

without EERTG

CO2 emissions without electrification (t-CO2/year)

Item

Explanation

e CERTG

Value

Source

Power consumption after electrification (kWh/year)

1,343,092

Actual results for FY2020

96,778

Actual results for FY2020

CO2 intensity of electricity (t-CO2/kWh)

0.000344

Ministry of the Environment (2021)

ef f

CO2 intensity of diesel oil (t-CO2/kL)

2.58

Ministry of the Environment (2021)

uh

Diesel oil consumption per handling before electrification (kL/time)

1.093

Actual results for FY2009

Nh

Number of handlings (times/year)

1,034,604

Actual results for FY2020

f CERTG

ef

*

e

Diesel oil consumption after electrification (kL/year)

*

Diesel oil is used to move between lanes, even after electrification

Table 3.5 Parameters used to calculate the effect of HSC Item

Explanation

EFHSC

Ratio of CO2 emission reduction by hybridization (%)

with EHSC

CO2 emissions with hybridization (t-CO2/year)

without EHSC

CO2 emissions without hybridization (t-CO2/year)

rHSC

Reduction rate of power consumption per handling by hybridization (%)

Item

Explanation

Value

Source

f CHSC

Diesel oil consumption of HSC (kL/year)

374,818

Actual results for FY2020

f CNSC

Diesel oil consumption of normal type S/C (kL/year)

505,124

Actual results for FY2020

uhHSC

Diesel oil consumption per handling of HSC (kL/time)

1.24

Actual results for FY2019*

Diesel oil consumption per handling of normal type S/C (kL/time)

1.75

Actual results for FY2019*

uhNSC *

Although two normal S/C units were renewed and have been in full operation since FY2020, the data for FY2019 before the renewal is used because renewed S/Cs are high-performance units and the effect of introducing HSCs is reduced

EFHSC ¼ 1 

with EHSC without EHSC

ð3:4Þ

  f f with EHSC ¼ ef f CHSC þ CNSC without EHSC ¼ ef f

f CHSC f þ CNSC 1  rHSC

ð3:5Þ ! ð3:6Þ

38

3

CO2 Reduction Effects by Electrification of Cargo-Handling Machinery

Table 3.6 CO2 emission reductions of E-RTG and HSC E-RTG

HSC

with EERTG

712 t-CO2/year

without EERTG

2920 t-CO2/year

without with  EERTG EERTG

2208 t-CO2/year

EFERTG

76%

CO2 emission reduction per unit

100 t-CO2/year

with EHSC

1961 t-CO2/year

without EHSC without EHSC

2295 t-CO2/year 

with EHSC

334 t-CO2/year

rHSC

29%

EFHSC

12%

CO2 emission reduction per unit

56 CO2/year

rHSC ¼ 1 

uhHSC uhNSC

ð3:7Þ

Table 3.6 lists the calculation results. Because all RTGs were electrified, the overall reduction in CTs was as large as 76%, while the reduction in HSCs was lower than that of E-RTGs at 29% per unit, and the reduction in CTs was 12% when the ratio of the number of units installed was considered. Two normal type S/C units were renewed in the middle of 2019, but the performance improved after the renewal, so that rHSC decreased to 25% when using the FY2020 results as uhHSC and uhNSC .

3.5

Conclusions

Hakata Port has led to climate change countermeasures at ports in Japan by introducing E-RTG, HSC, and RS. The changes in overall CT CO2 emissions due to the countermeasures, contribution of ERTG and HSC to those changes, and quantitative effects of the introduction of E-RTG and HSC, were empirically verified using Hakata Port as a case study. At Hakata Port, climate change countermeasures have been implemented since 2010, and CO2 emissions have decreased while container handling volume has increased significantly. In particular, the emission reduction per TEU is 36.5% for the entire CT. The most significant contribution to this reduction is the electrification of RTGs, which has a CO2 emission reduction effect of 76%; additionally, HSCs have a per unit CO2 emission reduction effect of 29%, which is smaller than that of E-RTGs. However, the CO2 emission reduction effect of the E-RTG installation is highly dependent on the CO2 intensity of electricity. For example, if a country has a high CO2 intensity of electricity, the apparent effect of introducing E-RTGs will be greatly reduced. Conversely, as some previous studies have referred to the use of renewable energy for electricity used in ports (Kang and Kim 2017; Alamoush et al. 2020), CO2 emissions can be greatly reduced by decreasing the CO2 intensity. Accordingly, it would be fairer to evaluate the effect of electrification of cargo-handling machinery in terms of its energy-saving effect than in terms of its CO2 reduction effect, which varies greatly depending on the level of the CO2 intensity of electricity. Currently, electrification and hybridization are considered the most efficient technical measures for cargo-handling machinery in terms of energy savings (Sifakis and Tsoutsos 2021). If the CO2 intensity of electricity becomes zero, CO2 emissions

References

39

will be zero regardless of how inefficient the operation is as long as it is electrified. However, from the viewpoint of energy savings, it is also necessary to improve operational efficiency. Because inefficient cargo-handling operations in container yards lengthen the turnaround time of ships and hinterland transport vehicles and cause congestion in ports (Feng et al. 2022), efficient deployment and scheduling of cargo-handling machinery is one of the most important challenges in terminal operations (Yu et al. 2019), cargo-handling efficiency is also important in that it can reduce CO2 emissions from ships and vehicles staying in ports (Styhre et al. 2017). Operational efficiency is often associated with the use of smart technologies and automation, for example in terms of minimizing rehandling (Martín-Soberón et al. 2014; Bjerkan and Seter 2019; Sifakis and Tsoutsos 2021), and the relationship between digital transformation and decarbonization at ports will further deepen in the near future. As mentioned in the introduction of this chapter, there are few previous studies based on empirical findings on climate change countermeasures in ports, and there exists a gap between research and practice that hinders decision-making. The main academic significance of this chapter, which empirically analyses the effects of the introduction of E-RTG and HSC using actual data from Hakata Port, lies in the fact that it addresses both these limitations.

References Alamoush AS, Ballini F, Ölçer AI (2020) Ports’ technical and operational measures to reduce greenhouse gas emission and improve energy efficiency: a review. Marine Poll Bull 160 Apple (2022) Environmental progress report Bjerkan KY, Seter H (2019) Reviewing tools and technologies for sustainable ports: does research enable decision making in ports? Transp Res Part D Transp Environ 72:243–260 Chen G, Govindan K, Golias MM (2013) Reducing truck emissions at container terminals in a low carbon economy: proposal of a queuing-based bi-objective model for optimizing truck arrival pattern. Transp Res Part E Logist Transp Rev 55:3–22 Di Ilio G, Di Giorgio P, Tribioli L, Bella G, Jannelli E (2021) Preliminary design of a fuel cell/battery hybrid powertrain for a heavy-duty yard truck for port logistics. Energy Conver Manage 243 Feng Y, Song DP, Li D (2022) Smart stacking for im-port containers using customer information at automated container terminals. Eur J Oper Res 301:502–522 Hakata Port Terminal Co., Ltd. (2013) Press release, gold award at IAPH (the international association of ports and harbors), Los Angeles general meeting He J, Huang Y, Yan W, Wang S (2015) Integrated internal truck, yard crane and quay crane scheduling in a container terminal considering energy consumption. Expert Syst Appl 42:2464–2487 Hiyoshi K (2012) Aiming for the best ECO terminal in Japan. J Japan Inst Marine Eng 47:54–58 Holguín-Veras J, Kalahasthi L, Campbell S, González-Calderón CA, Wang XC (2021) Freight mode choice: results from a nationwide qualitative and quantitative research effort. Transp Res Part A Policy Pract 143:78–120 IPCC (2021) Climate change 2021: the physical science basis. contribution of working group I to the sixth assessment report of the intergovernmental panel on climate change. IPCC (2022) Climate change 2022: mitigation of climate change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change Iris Ç, Lam JSL (2019) A review of energy efficiency in ports: operational strategies, technologies and energy management systems. Renew Sustain Energy Rev 112:170–182 Kang D, Kim S (2017) Conceptual model development of sustainability practices: the case of port operations for collaboration and governance. Sustainability 9 Martínez-Moya J, Vazquez-Paja B, Maldonado JAG (2019) Energy efficiency and CO2 emissions of port container terminal equipment: evidence from the Port of Valencia. Energy Policy 131:312–319 Martín-Soberón AM, Monfort A, Sapiña R, Mon-terde N, Calduch D (2014) Automation in port container terminals. Proc Soc Behav Sci 19:195–204 Ministry of Land, Infrastructure, Transport and Tourism (2009) FY2009 ports and harbors bureau related budget summary Ministry of Land, Infrastructure, Transport and Tourism (2018) Medium- and long-term policy for ports “PORT 2030” Ministry of Land, Infrastructure, Transport and Tourism (2021) Development manual for “carbon neutral port (CNP) formation plan”

40

3

CO2 Reduction Effects by Electrification of Cargo-Handling Machinery

Ministry of the Environment (2020) List of calculation methods and emission factors in calculation, reporting, and publication systems Ministry of the Environment (2021) Emission factors by electric utility (for calculation of greenhouse gas emissions by specific emitters) Nellemann C, Corcoran E, Durate CM, Valdes L, De Young C, Fonseca L, Grimsditch G (2009) Blue carbon—a rapid response assessment, United Nations environmental programme. GRID-Arendal. Brikeland Trykkeri AS, Birkeland PIANC (2019) Renewables and energy efficiency for maritime ports—MarCom WG report n°159–2019. The World Association for Waterborne Transport Infrastructure PIANC Maritime Navigation Commission. Brussels, Belguim Qi Y, Harrod S, Psaraftis HN, Lang M (2022) Transport service selection and routing with carbon emissions and inventory costs consideration in the context of the Belt and Road Initiative. Transp Res Part E Logist Transp Rev 159 Sha M, Zhang T, Lan Y, Zhou X, Qin T, Yu D, Chen K (2017) Scheduling optimization of yard cranes with minimal energy consumption at container terminals. Comput Ind Eng 113:704–713 Sifakis N, Tsoutsos T (2021) Planning zero-emissions ports through the nearly zero energy port concept. J Clean Prod 286 Sornn-Friese H, Poulsen RT, Nowinsk AU, de Langen P (2021) What drives ports around the world to adopt air emissions abatement measures? Transp Res Part D Transp Environ 90 Styhre L, Winnes H, Black J, Lee J, Le-Griffin H (2017) Greenhouse gas emissions from ships in ports—case studies in four continents. Transp Res Part D Transp Environ 54:212–224 Sugimura Y, Wakashima H, Liang Z, Shibasaki R (2022) Logistics strategy simulation of second-ranked ports on the basis of Japan’s port reforms: a case study of Hakata port. Marit Policy Manage Tokyo Port Terminal Co., Ltd. (2020) Tokyo port terminal Co., Ltd. environmental efforts Wang L, Peng C, Shi W, Zhu M (2020) Carbon dioxide emissions from port container distribution: Spatial characteristics and driving factors. Transp Res Part D Transp Environ 82 Xing H, Spence S, Chen H (2020) A comprehensive review on countermeasures for CO2 emissions from ships. Renew Sustain Energy Rev 134 Yu D, Li D, Sha M, Zhang D (2019) Carbon-efficient deployment of electric rubber-tyred gantry cranes in container terminals with workload uncertainty. Eur J Oper Res 275:552–569 Zhong H, Hu Z, Yip TL (2019) Carbon emissions re-duction in China’s container terminals: optimal strategy formulation and the influence of carbon emissions trading. J Clean Prod 219:518–530

4

Economic Feasibility of Electrification of Cargo-Handling Machinery

4.1

Introduction

In the previous chapter, the CO2 emission reduction effects of the introduction of E-RTG and HSC at Hakata Port in Japan were quantitatively analyzed. It was found that E-RTG and HSC reduced CO2 emissions by 76% and 29% per unit, respectively. The overall CO2 emissions of CTs decreased despite a large increase in the number of cargo-handling machines and cargo-handling volume, compared to before the introduction of the countermeasures, resulting in a 36.5% reduction in CO2 emissions per TEU. Although the CO2 emission reduction effect can be confirmed, when two S/C units were renewed at Hakata Port in 2019, the normal type with higher CO2 emissions was purchased instead of the hybrid type. Why is this happening even though HSC has certain decarbonization benefits, and Hakata Port Terminal Co., Ltd. (HPT) itself is aware of this? Implementing climate change countermeasures in ports is costly (Woo et al. 2018), and the high initial investment hinders the implementation of countermeasures at a business scale (Dai et al. 2019; Zis 2019). Attempts to recover the cost of implementation by adding it to port charges may reduce a port's competitiveness. A previous study pointed out that a balance between emissions reduction and fair competition is necessary because if environmental measures are applied selectively, the associated ports and shipping companies would suffer damage in terms of reduced cargo volume and profits (Sheng et al. 2017). Conversely, shippers may prefer ports with low CO2 emissions (Qi et al. 2022), in which case economic feasibility may be secured through an increase in cargo-handling volume. Port stakeholders are faced with the decision-making problem of whether to adopt climate change countermeasures in their ports, considering their impact on the economic feasibility and competitiveness of the port in question. As mentioned in the previous section, the introduction of E-RTGs and HSCs at Hakata Port was conducted ahead of other ports in Japan in the form of pilot projects by the central government. Considering the purpose of the pilot projects, other ports that recognize the effectiveness of the projects should have followed them and expanded their use nationwide; however, only three ports in Japan have completed the electrification of RTGs. The financial burden and lack of empirical studies regarding the effectiveness of countermeasures in ports may be the cause of this situation. Therefore, in this chapter, an empirical economic feasibility analysis of the implementation of E-RTG and HSC at Hakata Port is presented. First, based on the governance structure of Hakata Port, the entities responsible for and potentially affected by the introduction of E-RTG and HSC are identified, following which, the economic feasibility from the perspective of individual stakeholders, that is, the benefits and costs

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0_4

41

42

4

Economic Feasibility of Electrification of Cargo-Handling Machinery

attributed to each stakeholder, is discussed. Subsequently, strategies to promote the introduction of countermeasures are discussed. These considerations will also clarify the significance of pilot projects.

4.2

Governance Structure of Hakata Port

Before analyzing the economic feasibility of introducing E-RTG and HSC, it is necessary to confirm the entities responsible for and affected by the changes, that is, the owner, operator, and user of the facilities, as a port governance structure of Hakata Port. Although the Ports and Harbors Act of Japan was enacted in 1950, the national government retained supervisory authority over ports, and local governments were tasked with management (Sugimura 2020). The PA system, modeled after those in Europe and the United States, was assumed to be the main port management system, but it has not been widely implemented. Thus far, port management has been performed by local governments. However, as local governments do not have the financial and technical capacity to construct large port facilities, this responsibility has been directly assumed by the national government, which is managed by PMBs per a management contract under the Ports and Harbors Act. As port reforms have been promoted worldwide, certain policies related to port governance structures have been developed in Japan, where private sector participation in port operations has been promoted. Under the Strategic International Container Port Policy, a port operating company system is currently applied to the top-ranked strategic international ports and second-ranked hub international ports; the governance structure of the ports under this system lies somewhere between the landlord and tool models (Sugimura et al. 2022). In this system, a POC, a joint-stock company from a private sector perspective, leases a group of wharves from a PMB and operates them as a single entity. However, the POC is not in charge of cargo-handling operations itself, but rather leases them to terminal operators (TOs) on a dedicated basis for operation in strategic international ports, allowing them to use the facilities in hub international ports. Hakata Port, a hub international port, also follows the POC system; and Fig. 4.1 depicts the governance structure and Table 4.1 shows the ownership/operation structure of each facility. Figure 4.2 shows the relationships among the stakeholders of Hakata Port CT. At Hakata Port, the yard cargo-handling machinery is owned by HPT, and TOs conduct cargo-handling operations while paying a fee to HPT for the use of cargo-handling machinery. Electricity and fuel costs for E-RTGs and HSCs are not included in the usage charges paid to the HPT at Hakata Port and are borne directly by the TOs. This practice affects the attribution of benefits. Central government Management contract (quay)

Regulation, Landlord

Fukuoka City (Port management body) Long-term lease (quay, yard)

Operation

Shareholder

Lease fee 51%

Fukuoka City

49%

Local companies (including stevedores)

Hakata Port Terminal (Port operating company) License agreement (yard, crane)

Usage fee

Terminal operators (6 stevedores)

Fig. 4.1 Governance structure of Hakata port

4.3 Cost–Benefit Analysis of E-RTG and HSC Installation

43

The responsibility for formulating port plans and policies remains with the PMB, which is a shareholder of the HPT; therefore, the PA operation function is not completely separated from the HPT, as in the usual landlord model (Sugimura et al. 2022). The overall strategy for climate change countermeasures is developed by the PMB, while the HPT is responsible for individual countermeasures in the CTs. The HPT, together with the PMB, is the entity that is responsible for maintaining port competitiveness because the increase in the number of port calls and cargo-handling volume will increase revenues through quay fees and cargo-handling machinery fees. The two-tiered structure of PMB-POC-TO is unique and creates a complex structure of port governance.

4.3

Cost–Benefit Analysis of E-RTG and HSC Installation

This section presents an economic evaluation of the introduction of E-RTGs and HSCs. HPT, as the owner of E-RTGs and HSCs, would normally be the primary decision maker, but the actual introduction of E-RTGs and HSCs was conducted under the initiative of the central government. Specifically, in 2010 and 2011, the MLIT conducted pilot projects to electrify the RTGs (4 new units were introduced and 13 existing units were retrofitted) and HSC (1 unit); all necessary expenses, including infrastructure development for E-RTGs, were borne by the government. Subsequent installations of E-RTGs (four units) and HSCs (five units) were subsidized by the Ministry of the Environment (1/3 of the national expense ratio). E-RTGs require an initial investment for the infrastructure and reinvestment for the replacement of lithium-ion, in addition to the higher cost of the main body. However, the energy source is electricity, which is cheaper than traditional fuel. The HSC is more expensive than a normal carrier and requires the replacement of lithium-ion batteries as a reinvestment, but the consumption of fuel is reduced. In the economic evaluation of the introduction of E-RTGs and HSCs, the difference in benefits and costs compared to normal types should be considered, where benefits are the reduction in energy costs, and costs are the initial investment, reinvestment, and maintenance costs. In the case of E-RTGs, infrastructure development is required, in addition to the RTG itself; therefore, this should also be accounted for. Economic analysis was performed using the following equation, and the values are shown in Table 4.2. All data used were actual data from Hakata Port or the values used in the management of Hakata Port. TBi ¼

Ti X t¼1

TCiwithout ¼ ICiMB þ ICiINF þ

RCit

ð4:1Þ

Ti X MCt þ RIM t þ RII t i

i¼1

TCiwith ¼ ICiMB þ ICiINF  GSi þ

i

i

ð1 þ rÞt

ð4:2Þ

Ti X MC t þ RIM t þ RII t i

i¼1

NPViwithout ¼ TBi  TCiwithout

i

ð1 þ r Þt

i

ð4:3Þ ð4:4Þ

44

4

Economic Feasibility of Electrification of Cargo-Handling Machinery

CBRwithout ¼ i

TBi TCiwithout

ð4:5Þ

NPViwith ¼ TBi  TCiwith CBRwith ¼ i

ð4:6Þ

TBi TCiwith

ð4:7Þ

The results of the analysis, including the attribution of each benefit and cost, are shown in Tables 4.3 and 4.4, and the final results are shown in Table 4.5. At Hakata Port, HPT owns RTGs and S/C, and TOs use them while paying usage fees, electricity, and fuel costs. Therefore, as shown in Tables 4.3 and 4.4, the energy cost reduction benefits of ERTGs and HSCs are not attributed to HPT and are unprofitable from HPT's perspective of HPT. However, because the two-tier structure of PMB-POC-TO is characteristic of Japanese port governance, the results in Table 4.5 are obtained if the standard model of HPT and TO combined (or the case where the energy cost reduction is attributed to HPT) is assumed. Neither RTG nor HSC can be implemented on a business basis, but it can be implemented with a subsidy of 1/3 of the initial cost. At the beginning of this chapter, a case was introduced in which two S/C units were renewed at Hakata Port in 2019; normal type units with higher CO2 emissions were purchased instead of hybrid type units. The results of the economic analysis suggest the reasons for this decision. In the past, government subsidies made it possible to purchase the more expensive hybrid type at practically twoTable 4.1 Ownership and operation form of facilities at Hakata port Quay Central government

Quay crane

Owner

Owner

Yard crane

Owner

Fukuoka city (PMB) Hakata Port Terminal Co. (port operating company)

Yard

Operator

Terminal operators

Fig. 4.2 Relationships among stakeholders at Hakata port

Operator

Operator

Owner, Operator

User

User

User

4.3 Cost–Benefit Analysis of E-RTG and HSC Installation

45

Table 4.2 Parameters used in economic evaluation Item

Explanation

i

Type of machinery (E-RTG, HSC)

t

Year t after start of service

r

Social discount rate (%) (use 4%)

Ti

Useful life of machinery i (years)

TBi

Total benefit of installing machinery i (yen)

TCiwithout

Total cost of installing machinery i without subsidy (yen)

TCiwith RCit

Total cost of installing machinery i with subsidy (yen)

ICiMB

Difference in the initial cost of the main body of machinery i from the normal type (yen)

ICiINF MCit RIMit RIIit

Infrastructure costs to install machinery i (yen)

GSi

Government subsidies for the installation of machinery i (yen)

NPViwithout

Net present value of installing machinery i without subsidy (yen)

CBRwithout i

Cost–benefit ratio of installing machinery i without subsidy

NPViwith

Net present value of installing machinery i with subsidy (yen)

CBRwith i

Reduction in energy costs in year t when machinery i is installed (yen)

Increase in maintenance costs in year t when machinery i is installed (yen) Increase in reinvestment cost of the main body in year t when machinery i is installed (yen) Increase in infrastructure reinvestment costs in year t when machinery i is installed (yen)

Cost–benefit ratio of installing machinery i with subsidy

Item

E-RTG

HSC

Ti

25 years

25 years

RCit

2298 thousand yen

1511 thousand yen

ICiMB

35,000 (200,000–165,000) thousand yen

30,000 (140,000–110,000) thousand yen

ICiINF MCit RIMit

38,000 thousand yen

0 yen

0 yen

0 yen

−4000 thousand yen (t ¼ 5; 15), 11,000 thousand yen (t ¼ 10; 20)

24,000 thousand yen (t ¼ 10; 20)

RIIit

5970 thousand yen (t ¼ 10; 20)

0 yen

GSi

1/3 of the initial cost

1/3 of the initial cost

thirds of the price, that is, less expensive than the normal type. However, because the subsidy is no longer applied, the hybrid type is no longer inexpensive, making it difficult to introduce it on a business basis from an economic perspective. Similar cases have occurred at other ports in Japan, and currently, it is difficult to select environmentally friendly cargo-handling machinery from an economic standpoint.

46

4

Economic Feasibility of Electrification of Cargo-Handling Machinery

Table 4.3 Benefit incidence table of E-RTG installation (per unit) (in thousand yen) Government

HPT

TO

Total

Initial investment in the main body

−35,000

−35,000

Initial investment in infrastructure

−47,479*

−44,083

Reinvestment in the main body

−14,000

−14,000

Reinvestment in infrastructure

−11,942

Energy cost Subsidy

−82,493

82,493

Total

−82,493

−25,928

*

−11,942 57,460

57,460

57,460

−50,960

0

Additionally, there was an initial infrastructure investment of 6083 thousand yen by the PMB

Table 4.4 Benefit incidence table of HSC installation (per unit) (in thousand yen) Government

HPT

TO

Initial investment in the main body

−30,000

Reinvestment in the main body

−27,167

Energy cost Subsidy

−46,667

46,667

Total

−46,667

−10,500

Total −30,000 −27,167

23,598

23,598

23,598

−33,568

0

Table 4.5 Summary of cost–benefit analysis (NPV in thousand yen) Without subsidy NPV

With subsidy CBR

NPV

CBR

E-RTG

−50,960

0.53

31,533

2.22

HSC

−33,568

0.41

13,098

2.25

4.4

Possibility of Using Carbon Credits

As mentioned in the previous section, subsidies would increase the possibility of E-RTGs and HSCs, but it is also necessary to consider financing methods that do not rely on subsidies. Tolliver et al. (2020) argued that various financing mechanisms that leverage both public and private resources are crucial in capitalizing on the Paris Agreement and the NDC initiative, while Sun and Yang (2021) identified cap-and-trade and carbon taxes as two key policies. In fact, approximately 20% of global emissions are covered by carbon pricing in the form of carbon taxes and trading systems (Hepburn et al. 2020). In Japan, carbon taxes, emissions trading (total volume control), government-operated credit trading, and private credit trading have been identified as economic approaches to decarbonization, and the government is aiming to create a carbon market and consider regulatory approaches, such as setting emission allowances (Ministry of Economy, Trade and Industry 2021). Therefore, this section discusses the possibility of introducing E-RTG and HSC by utilizing carbon credit.

4.5 Challenges from the Perspective of Economic Feasibility …

47

Table 4.6 Possibility of utilizing carbon credits E-RTG

HSC

CO2 emission reductions

100.4 t-CO2/unit/year

55.7 t-CO2/unit/year

Income from credits

156,000 yen/year

86,000 yen/year

Present value (25 years)

2.43 million yen

0.35 million yen

32,397 yen

38,565 yen

Credit price *

*

Price required to make a CBR of 1

In Japan, carbon credits include J-Credits, Tradable Green Certificates, and Non-Fossil Fuel Certificates, which can be used for legal reporting and international initiatives such as CDP, RE100, and SBT Among Japanese carbon credits, the highest Non-Fossil Certificate is about 13,700 yen/tCO2 (Japan Electric Power Exchange 2021), and the cheapest J-Credit (energy-saving type) is about 1550 yen/t-CO2 (J-Credit System Office 2022). Table 4.6 shows the impact of converting CO2 reductions by introducing E-RTGs and HSCs into credits, setting the cheapest credit price at 1550 yen/t CO2, and adding income from the sale of credits. The present value of credit income assuming the current J-Credit price cannot compensate for the NPV without subsidies (Table 4.5, and conversely, the credit price required to secure CBR = 1 is in the 30,000-yen range, making it difficult to introduce E-RTGs and HSCs on a business basis using the carbon credit system if the current credit price is assumed. However, credit prices vary with various variables, such as project type (Vanderklift et al. 2019), and the recognition of value depends on the motives of the buyer, such as pricing differently for buyers interested only in offsetting CO2 emissions than for those interested in the source and location of the offset (Hamrick and Goldstein 2016). Some buyers choose credits that support their supply chains while others choose ones that support the regions in which they operate (Vanderklift et al. 2019). Given the high price of J-Blue credits related to BCEs traded in the voluntary market in Japan (Kuwae et al. 2022), the possibility of business-based adoption through the use of carbon credits remains. The perspective of maritime domain relations may also motivate the purchase of high-value credits, as 90% of investors are interested in investing in sustainable blue economy projects, of which one-third consider this field to be important (Credit Suisse 2020).

4.5

Challenges from the Perspective of Economic Feasibility and the Role of Stakeholders in Promoting Countermeasures

As described in Chap. 2, the Japanese port governance model is not characterized by the active promotion of climate change countermeasures at individual ports and CTs. Rather, the central government’s initiative is more important than in other countries. Therefore, the government's initiative to introduce environmentally friendly cargo-handling machinery as pilot projects, as in the case of Hakata Port, was a fitting approach. Additionally, as revealed in this chapter, the fact that it is difficult to introduce the measures on a business basis from an economic perspective and that public support is required, should also be considered when evaluating the pilot projects. In the absence of economic feasibility, it is difficult to adopt climate change countermeasures without public support. There are two approaches to promote climate change countermeasures: regulatory type and voluntary type. In the case of regulatory type, the government's initiative is important because it is not realistic for PMB to promote the countermeasures on a port-by-port basis

48

4

Economic Feasibility of Electrification of Cargo-Handling Machinery

from the perspective of the impact on port competitiveness, and the countermeasures must be promoted on a country-wide basis. In the case of the voluntary type, as cost burden is an issue, unless the importance of CO2 emissions increases in the CT selection criteria of shippers and shipping companies, it is difficult to impose the cost of introduction on port charges; therefore, facility owners, such as POCs and TOs, would have to bear the cost in the case of hard facilities. Incentives such as government or PMB subsidies are needed, especially when initial costs are a bottleneck and introducing the measures is challenging. However, if increased awareness about decarbonization among shippers precedes the implementation of countermeasures, then the increase in port charges is less likely to affect competitiveness, and shippers may be willing to bear the costs. Because the possibility that CTs with climate change countermeasures are selected will increase, soft measures are likely to be promoted under a scheme led by the PMBs or POCs, while hard measures are likely to be promoted under a scheme in which the owners of the facilities concerned are responsible for implementing them, requiring shippers to bear the costs. If it takes time to change the awareness on the shippers’ side, the realistic way forward is to adopt a regulatory approach on the PMBs and POCs’ sides through government initiatives to ensure a fair competitive environment among ports, or to encourage stakeholders to invest in decarbonization, especially through financial incentives. As described above, for countermeasures on facilities that are unlikely to be economically feasible, the owners of such facilities should play a role in introducing the countermeasures; however, the cost burden would be a major issue to port competitiveness, and multiple scenarios can be envisioned. If increasing awareness of decarbonization among shippers precedes the implementation, then the introduction of climate change countermeasures by stakeholders on the port service supply side could be promoted voluntarily from the perspective of ensuring port competitiveness, and shippers could be made to bear increased port charges. Else, it is unlikely that countermeasures affecting port competitiveness through increased port charges will be introduced. In this case, regulatory approaches and incentives from the government and PMBs are required. However, the government’s role is important from the perspective of ensuring fairness among ports, and the role of PMBs and POCs is important from the perspective of ensuring fairness among stakeholders within a port. To move toward decarbonization in ports while facing the difficult issue of economic feasibility, it would be important for stakeholders to cooperate in each countermeasure. Moreover, utilization of councils, wherein stakeholders participate, to avoid the issue of the cost burden and the role of PMB as a core member of such councils would be important, in addition to the role of the government in ensuring fairness among ports.

4.6

Conclusion

Climate change countermeasures in ports are costly (Woo et al. 2018), and high initial investment is often a barrier to their implementation on a business scale (Dai et al. 2019; Zis 2019). Attempts to recover the cost of implementation by adding it to port charges may reduce a port's competitiveness. Thus, port stakeholders face the decision-making problem of implementing climate change countermeasures in their ports, while maintaining the economic feasibility and competitiveness of the port. In this chapter, an empirical economic feasibility analysis of the implementation of E-RTG and HSC at Hakata Port was presented. First, based on the governance structure of Hakata Port, the entities responsible for and potentially affected by the introduction of E-RTG and HSC were identified, and then the economic feasibility from the perspective of individual stakeholders was discussed.

References

49

At Hakata Port, the yard cargo-handling machinery is owned by HPT, the POC of Hakata Port. TOs conduct cargo-handling operations while paying a fee to HPT for the use of cargo-handling machinery. Electricity and fuel costs for E-RTGs and HSCs are not included in the usage charges paid to the HPT at Hakata Port and are borne directly by the TOs. Therefore, the energy costreduction benefits of E-RTGs and HSCs are not attributed to HPT and are unprofitable from HPT's perspective of HPT. While the two-tier structure of PMB-POC-TO is a characteristic of Japanese port governance, it was revealed that even if the standard model of HPT and TO is assumed, neither E-RTG nor HSC can be implemented on a business basis, but can be implemented with a subsidy of 1/3 of the initial cost as a result of economic feasibility analysis. Although the CO2 emission reduction effect can be confirmed, when two S/C units were renewed at Hakata Port in 2019, the normal type with higher CO2 emissions was purchased instead of the hybrid type. The results of the economic feasibility analysis suggested that this was because government subsidies made it possible to previously purchase the more expensive hybrid type at practically two-thirds of the price. However, because the subsidy is no longer applied, the high cost of the hybrid type makes it economically unfeasible to introduce it on a business basis. For countermeasures on facilities that are unlikely to be economically feasible, the owners of such facilities should take the initiative to introduce the countermeasures; however, the cost burden would be a major issue in relation to port competitiveness. If increasing awareness of decarbonization among shippers does not precede the implementation of countermeasures, it is unlikely that changes affecting port competitiveness through increased port charges would be introduced. In this case, regulatory approaches and incentives from the government and PMBs are required. However, the government’s role is important from the perspective of ensuring fairness among ports, and the role of PMBs and POCs is important from the perspective of ensuring fairness among stakeholders within a port.

References Credit Suisse (2020) Investors and the blue economy Dai L, Hu H, Wang Z, Shi Y, Ding W (2019) An environmental and techno-economic analysis of shore side electricity. Transp Res Part D Transp Environ 75:223–235 Hamrick K, Goldstein A (2016) Raising ambition: state of the voluntary carbon markets 2016. Forest Trends’ Ecosyst Market Hepburn C, Stern N, Stiglitz JE (2020) “Carbon pricing” special issue in the European economic review. Europ Econ Rev 127 Japan Electric Power Exchange (2021) FY2021 non-fossil value trading market notification of transaction results. http:// www.jepx.org/market/nonfossil.html J-Credit System Office (2022) J-credit system. https://japancredit.go.jp/data/pdf/credit_001.pdf Kuwae T, Watanabe A, Yoshihara S, Suehiro F, Sugimura Y (2022) Implementation of blue carbon offset crediting for seagrass meadows, macroalgal beds, and macroalgae farming in Japan. Marine Policy Ministry of Economy, Trade and Industry (2021) Carbon pricing for growth—credit transactions, etc. https://www.meti. go.jp/shingikai/energy_environment/carbon_neutral_jitsugen/pdf/003_04_00.pdf Qi Y, Harrod S, Psaraftis HN, Lang M (2022) Transport service selection and routing with carbon emissions and inventory costs consideration in the context of the belt and road initiative. Transp Res Part E Logist Transp Rev 159 Sheng D, Li Z-C, Fu X, Gillen D (2017) Modeling the effects of unilateral and uniform emission regulations under shipping company and port competition. Transp Res Part E Logist Transp Rev 101:99–114 Sugimura Y (2020) Public-private partnerships in Japan's cruise terminal operations. Res Transp Business Manage Sugimura Y, Wakashima H, Liang Z, Shibasaki R (2022) Logistics strategy simulation of second-ranked ports on the basis of Japan’s port reforms: a case study of Hakata Port. Marit Policy Manage Sun H, Yang J (2021) Optimal decisions for competitive manufacturers under carbon tax and cap-and-trade policies. Comp Indust Eng 156 Tolliver C, Keeley AR, Managi S (2020) Drivers of green bond market growth: the importance of nationally determined contributions to the Paris agreement and implications for sustainability. J Clean Prod 244

50

4

Economic Feasibility of Electrification of Cargo-Handling Machinery

Vanderklift AM, Marcos-Martinez R, Butler AJR, Coleman M, Lawrence A, Prislan H, Steven A, Thomas S (2019) Constraints and opportunities for market-based and protection of blue carbon ecosystems. Marine Policy 107 Woo JK, Moon DSH, Lam JSL (2018) The impact of environmental policy on ports and the associated economic opportunities. Transp Res Part A Policy Pract 110:234–242 Zis TPV (2019) Prospects of cold ironing as an emissions reduction option. Transp Res Part A Policy Pract 119:82–85

Part III Countermeasures in Reefer Container Areas

Although reefer containers (RCs) are the main sources of energy consumption in port activities, the energy efficiency of RCs is mostly disregarded in strategy formulation for most ports and terminals, and energy-saving measures for RCs have rarely been the subject of research. Studies that could support decision-making are needed to overcome the challenge of being the main energy consumption source. Part III focuses on energy consumption reduction and decarbonization measures in RCs. Specifically, a simulation method is proposed to estimate the amount of energy required for cooling that predicts the surface temperature of container walls using computational fluid dynamics (CFD) to calculate the benefits of installing RSs in RC areas. The proposed simulation model is used to calculate the actual energy-saving effects at Hakata Port in Japan. This part also examines the economic feasibility of installing RSs using the amount of energy savings. This is because even if there are CO2 emission reduction and energy-saving effects, it may be difficult to introduce depending on economic feasibility, or conversely, it may lead to the consideration of issues and strategies for the introduction. The content of this part is very valuable in terms of showing the quantitative effects of climate change countermeasures in the RC areas. Moreover, because it is an empirical analysis, it would provide an important basis for ports worldwide to consider the implementation of RSs.

5

Simulation Model for Analysis of Energy Savings by Roof Shade Installations

5.1

Introduction

The main sources of energy consumption in port activities are cargo-handling machinery and reefer containers (RCs) (Acciaro and Wilmsmeier 2015; Ilio et al. 2021; Iris and Lam 2019; van Duin et al. 2018; Wang et al. 2020). While technological improvements to cargo-handling machinery, including practical energy-saving measures (Martínez-Moya et al. 2019), have been researched under countermeasures, the energy efficiency of RCs has been disregarded in strategy formulation for most ports and terminals, and only a few ports are certified according to the ISO 50001 energy efficiency standard (Acciaro and Wilmsmeier 2015). Although previous studies have shown that reefer areas are responsible for approximately 20–45% of the energy consumption at CTs (Acciaro and Wilmsmeier 2015; Iris and Lam 2019; van Duin et al. 2018), energy-saving measures have rarely been the focus of research (Iris and Lam 2019). As Bjerkan and Seter (2019) pointed out, studies based on the empirical findings at ports are rare, and the gap between the research and application affects decision-making; empirical studies demonstrating the effectiveness of introducing measures in RC areas and studies that could support decision-making are needed to overcome the current challenges in optimizing energy use. This chapter focuses on energy consumption reduction and decarbonization measures in RCs. Specifically, a simulation method is proposed to estimate the amount of energy required for cooling. The method predicts the surface temperature of container walls using CFD to calculate the benefits of installing roof shades (RSs) in RC areas. Among the few previous studies on RCs, Rijsenbrij and Wieschemann (2011) reported that covered areas prevent container heating. Werner (2014) suggested covering the roof of the reefer area with solar panels, which reduced the electrical energy required for cooling owing to container shading. To ensure efficient terminal operation, peak shaving, which is used to reduce the peak energy consumption of a port, can be implemented in RC operations (Iris and Lam 2019). RC areas require a variable amount of electricity, depending on the season and time of day. In addition, the intermittent distribution of electricity between reefer batches and limited allowance for electricity consumption have been reported to reduce the peak energy demand by 62.8 and 7.2%, respectively (van Duin et al. 2018). However, RSs have rarely been the subject of research, and the details of energy-saving mechanisms, CO2 emission reduction effects, and economic feasibility for their introduction have not been studied. The effects of introducing RSs for RCs are empirically summarized in this chapter. Additionally, a versatile simulation method for preliminary evaluation, which can be used as a basis for evaluating the feasibility of introducing RSs based on economic analysis, is presented. If the energy savings and © The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0_5

53

54

5

Simulation Model for Analysis of Energy Savings …

CO2 reduction effects of installing RSs can be estimated through simulation, it would be a powerful tool for the preliminary evaluation of installing RSs in RC areas and for supporting decision-making. Therefore, the results of this chapter can support decision makers, such as PAs and TOs, when installing RSs, thereby contributing to worldwide energy savings and decarbonization in ports.

5.2

Design and Operational System of the RS

The major cause of the increase in energy consumption by RCs is the heat penetration from strong solar insolation. Therefore, reducing the heat penetration of the walls of RCs from the surroundings and solar insolation in summer can reduce electrical energy consumption. During operation, the RCs are exposed to direct sunlight. On sunny days, direct insolation causes an increase in the surface temperature of the container. Consequently, heat penetrates the container, necessitating the removal of excess heat through a refrigeration system. The energy required to remove the heat is equal to the energy consumed by the load factor of the refrigeration system. A RS was installed at the Hakata Island City Container Terminal (HICCT), one of two CTs at Hakata Harbor in Japan, to protect RCs from direct sunlight exposure. The fundamental design of the RS is similar to that of an ordinary awning used in buildings, and the innovative aspect of this technology is its design as a retractable roof for container storage areas. A retractable roof was installed above the yard along the frame structures bonded with modular work platforms. The roof was made of fabric canvas and fitted onto frame structures. Furthermore, the frame structures were divided into six sets of roof folds linked to each other. The roof could be retracted by mechanical shifting on the horizontal track. This shifting of the frame structures was driven by an electric winch that was controlled by terminal operating systems. Under normal conditions, the roof is usually closed and only opened upon receiving work instructions from the terminal operating system. The HICCT introduced retractable RSs into two slots of its RC storage yard. Figure 5.1 shows the functional design of the installed RS. The operation of the shade was integrated with the terminal operating system used in the CT, as shown in Fig. 5.2. When an incoming container truck is checked at the gate, the container data are sent to the terminal operating system, which automatically arranges the cargo-handling machines and determines the container position. Subsequently, the terminal operating system sends work instructions to the appropriate cargo-handling machines, and the roof automatically opens before the arrival of these machines.

Fig. 5.1 Design of the RSs installation at the RC storage yard

5.3 Methodology

55 Terminal operation systems

Work instruction to RTG

Entrance information

Truck arrives at the gate

Instruction to open the roof shade

Fig. 5.2 Operational mechanism of the installed RS

5.3

Methodology

A simulation method is proposed to estimate the amount of energy required for cooling, which predicts the surface temperature of container walls using CFD. A phenomenon may be determined by several small factors, which is known as a self-assembly type phenomenon, and is difficult to analyze, unlike problems in which the main controlling factor is clear (Nicolis and Prigogine 1989). However, in this study, a self-assembly type thermo-fluid analysis model was introduced as a simulation method. First, the thermal factors, including the small factors related to energy consumption, were assumed as follows: • • • •

Solar radiation from the sun. Heat conduction on the container walls and through the road. Heat convection on the surrounding container surfaces. Thermal radiation among the solid surfaces.

Based on the above assumptions, the finite volume method (FVM), which is a discretization method, was used for the numerical analysis by applying the governing equations related to the thermo-fluid analysis. However, because the model was self-assembly type, the numerical analysis results did not easily match the experimental results. Therefore, the following method was implemented to solve the numerical analysis problem: • The concept of a dummy container was introduced to control the heat penetration into the container storage area using the RS. • The thermal response was confirmed to be gradual by observing heat penetrating the wall and recognizing the time factor as an issue in the numerical analysis. • The initial value was defined based on its effect on RCs when exposed to solar radiation. • The influence of the thermal buoyancy between the RCs was examined owing to the large difference between the surface temperatures of the sides of the upper and lower RCs. • The surface temperatures of the RCs measured in the experiment and those referred to in the numerical calculations were redefined on the mesh after careful examination. The proposed approach improves the accuracy of new numerical analysis methods; however, experimental results are essential for refining the simulation model. The next section describes the

56

5

Simulation Model for Analysis of Energy Savings …

various experimental measurements required to improve the simulation model. The development of a simulation model based on the above is described.

5.4

Experimental Measurements of the Impact of the RSs

Two types of experiments were conducted at the HICCT to verify the energy saving effect of RS installation and evaluate the thermal characteristics of the containers. Details of the experimental locations and orientations are listed in Table 5.1 and Fig. 5.3. The orientation of the RC complies with the following rules: the reefer unit faces west, the door faces east, one sidewall faces south, and one sidewall faces north. The 40-ft-long high-cube RCs were used as the measurement objects. Six RCs were prepared using the same manufacturer and production year. The design and specifications of the containers are Table 5.1 Location of the measurement experiment

Place

Hakata island city container terminal

Degree of longitude

130.40° (130°24′04.24″E)

Degree of latitude

33.65° (33°39′31.15″N)

Altitude above sea-level

2.5 m

Time zone

Universal Time Coordinated (UTC) + 9 h

Orientation of RC

334.185° from North (Wall Azimuth)

Fig. 5.3 CT layout of Hakata Island City Container Terminal (HICCT)

5.4 Experimental Measurements of the Impact of the RSs

57

Fig. 5.4 Design and specification of a high-cube RC

shown in Fig. 5.4. The containers were 12.1 m long, 2.4 m wide, and 2.8 m high. The structures comprised a ceiling wall, sidewalls, floor, and corner metal-casting foundations. The interior space of the RC mainly served as the cargo hold, with the floor equipped with a T-grating to circulate air from the refrigeration system attached at the end of the spaces. The container walls comprised three layers, aluminum, polyurethane, and stainless steel, with thicknesses of 0.8, 90, and 0.9 mm, respectively, and thermal conductivities of 204, 0.03, and 16 W/mK, respectively. Container design data available from the manufacturer show a minimum interior temperature of − 0°F (−18°C) and a maximum exterior temperature of 100°F (38°C). The overall heat-transfer rate is 7,556 BTU/h (1,904 kcal/h) with a U-Factor of 75.0 BTU/h °F (34 kcal/h °C). The exterior surface comprises flat aluminum panels, and all the surfaces are painted white. The devices and sensors used in this experiment included pyranometers, power meters, thermocouples, and weather stations. Figure 5.5 depicts the measurement devices and arrangement of the sensor locations. The pyranometers used to measure the solar radiation were placed on top of the building facility of the RC storage yard, and five units were set on the horizontal and vertical planes in all cardinal directions. Thermocouples were installed inside and outside the container surfaces at 20 points. Five sensors were attached to the middle of the inner surfaces (floor, sidewalls, ceiling, and center). In addition, 15 sensors were attached to the outside at the middle points of each wall, including those of the fan and compressor. The power meter employed to measure energy consumption was set at the power-plug station near the measurement object. Weather stations used to measure wind speed and direction were placed in the building offices. The measurement data from all the devices were recorded every minute under various weather conditions. Two experiments were conducted using six RCs over a continuous period. The experimental procedure considers the location and position of the RCs. The location of RCs is related to their stack arrangement in the storage yard, as shown in Fig. 5.6. To indicate the location of the containers, a naming convention was assigned, starting with the bay number, followed by the row and tier numbers. The detailed procedure for each experiment is described in Table 5.2 and Fig. 5.7.

58

5

a. Data loggers (CADAC 3)

Simulation Model for Analysis of Energy Savings …

b. Pyranometer set-up (EKO ML-020VM)

e. Arrangement of thermocouples c. Power meter (HIOKI PW3169)

d. Weather station (Davis Vantage Vue)

Fig. 5.5 Measurement devices and arrangement of sensor locations

Fig. 5.6 Stack arrangement at the RC storage yard

The experimental measurements are shown in Figs. 5.8, 5.9, 5.10, 5.11. The implications of the experimental measurements are presented in Table 5.3.

5.5

RS Impact Analysis Using Simulation

5.5.1 Thermal Factors Thermal simulation was conducted to predict the surface temperature of container walls using CFD by considering the physical properties and environmental conditions. The thermodynamic interaction between the container surfaces and environment was modeled using heat-transfer processes. The physical model of the heat-transfer processes considered in this simulation can be describes as:

5.5 RS Impact Analysis Using Simulation

59

Table 5.2 Detailed procedure for each experiment Performance test

Experimentation one (EXP 1)

Experimentation two (EXP 2)

Objectives

Preconditioning test to obtain benchmark parameter of each RC Compare one container with RS and without RS

Verify the effect of RS installation on the RC by directly comparing three tier RCs with and without RS

Evaluate the thermal characteristics among RCs by three tiers RCs placing side by side at the RC storage yard

Period

July 2011, July—October 2012

July—October 2013, July—October 2014

July—October 2015

Procedures

1. Preparation of RC; cleaning of the cargo hold, installation of measurement devices 2. Placing RCs on the empty space without any possibility of a shadow effect from another object 3. Test a refrigeration system by considering the cooling load of the refrigeration units 4. Set the cooling temperature of the RC to the constant condition at 0°C 5. Monitor the fluctuation of the energy consumption and temperature changes until constant ranges in 1-day measurement with a clear-sky condition

1. Arrange six RCs into two groups; one group consists of three tiers of RCs 2. Move the first group to the reefer storage yard without RS. Move the second group to reefer storage with RS 3. The position starts from the south side (R1), continues to the middle side (R4), and then moves to the north side (R8) 4. Data was acquired and saved by considering the weather factors. The main factor is the strong intensity of solar radiation, which requires a day with clear-sky conditions. The duration time for each step is usually 2–3 days

1. Arrange six RCs into two pairs of three tier RCs, placing them side by side at the RC storage yard 2. The first location of experimentation is at the reefer storage yard with the installation of RS 3. The position starts from the south rows (R1 and R2), continues to the middle rows (R4 and R5), and then moves to the north rows (R7 and R8) 4. Data was acquired and saved by considering the weather factors 5. After completion for the condition with RS, the same procedure is conducted for condition without RS

• Solar radiation from the sun transfers heat through the domain air in the simulation model. • Heat conduction on the container walls transfers heat through solid objects in the simulation model. • Heat convection on the container surfaces transfers heat between the solid surface and the fluid motion of the surroundings. • Thermal radiation on solid surfaces is related to the surface radiation emitted by other objects. The simulation object comprised an RS, container structures, and road as the storage bed. Additionally, the simulation considered the heat interactions between each part impacted by insolation (Shinoda and Budiyanto 2016). The effects of heat on individual parts is shown in Fig. 5.12.

5.5.2 Governing Equation of the Simulation Model The simulation model was based on partial differential equations governing flow and convective heat transfer. The heat-transfer process in the RC was simulated via CFD using PHOENICS, which is a thermo-fluid analysis software package that numerically calculates and visualizes the flow conditions of gases and liquids, such as changes in flow velocity, pressure, and temperature (CHAM 2010). The governing equations in the field of fluid flow based on the incompressibility and heat-transfer processes comprise the generalized transport equations, as shown in Eqs. (5.1), (5.2), (5.3) (Patankar 1980):

60

5

Simulation Model for Analysis of Energy Savings …

(a) Measurement experiment 1

(b) Measurement experiment 2 Fig. 5.7 Illustration of measurement experiments

Solar radiation without RS Fig. 5.8 Intensity of solar radiation with and without RSs

Solar radiation with RS on the top of tier 3

5.5 RS Impact Analysis Using Simulation

Temperature profiles without RS Fig. 5.9 Comparison of temperature profiles with and without RSs

Fig. 5.10 Power consumption with and without an RS

61

Temperature profiles with an RS

62

5

Simulation Model for Analysis of Energy Savings …

Fig. 5.11 Thermal characteristics on surface walls side by side Table 5.3 Implications of the experimental measurements Item

Implications of experimental measurements

Solar insolation

• In the absence of an RS, the maximum solar radiation was measured on the ceiling and west surfaces. The highest solar radiation was observed on the ceiling surface at 12:00 PM, whereas that on the west surface occurred at 3:30 PM. The east and south surfaces received approximately the same maximum solar radiation at around 08:30 AM and 10:30 AM, respectively • A fluctuation of high solar intensity occurred in the afternoon supposedly owing to sunlight penetrating through the gap in the folding structure of the RS

Temperatures of the container walls

• The surface temperature increase was apparent at the top tier when the RS was absent. All surface temperatures except that of the bottom increased and reached a peak. The ceiling surface exhibited high temperatures for almost the entire day from 8:00 AM to 6:00 PM and these temperatures were higher than those of the other surfaces. The bottom surface exhibited the lowest temperature, which was approximately the same as the air temperature • Surface temperatures decreased at all container tiers when the RS was used. At the top tier, the surface temperature reduction of the ceiling was significant. Each surface mainly followed the intensity trend of the solar radiations on the RS. Temperature fluctuations occurred several times owing to the high intensity of solar radiation entering through a gap in the roof frame. The temperature of the bottom surface did not decrease substantially with the introduction of an RS as the average temperature of the surface was equal to the air temperature

Power consumption

• Power consumption was classified as either high or low range. The power consumption of the RCs at the top and middle tier without an RS remained in the high range; however, the power consumption of the rest of the containers remained in the low range. Therefore, the RS was found to reduce energy consumption • The middle tier exhibited the highest reduction in power consumption owing to advantageous thermal protection from the upper and lower positions

Thermal interaction between containers

• Thermal interaction between containers can be observed in the gap between two rows. Thermal stratification was evident from the top to bottom surfaces. The surfaces facing each other tended to experience thermal stratification from the top to bottom tiers • Thermal stratification occurred due to the thermal buoyancy of the air in the narrow space; warm air tends to rise under the influence of buoyant forces, thereby resulting in a positive vertical temperature gradient between the floor and ceiling

5.5 RS Impact Analysis Using Simulation

63

Sun

Fig. 5.12 Heat-affected parts considered in the simulation model. RS) RS: This object is exposed to direct solar radiation; the heat increases significantly on a sunny day; CW1) Ceiling wall: This is affected by the emitted radiation from the RS and heat conduction from the walls; SW1) South side wall: This is affected by the heat conduction from other walls and radiation emitted from other surfaces.; SW2) North side wall: This is protected from solar radiation and faces the other container in the vertical direction.; SW3) Another container surface: This part contributes to the heat radiation on the north side.; BW1) Base wall: This wall faces the ceiling wall and is protected from solar radiation.; FP1) Foot part among containers: Heat is transferred between containers through this part.; CW1) Ceiling wall: Heat on this part is affected by the heat conduction from the other walls.; BW2) Base wall: This wall faces the road surface; the heat on this part is affected by heat conduction.; FP2) Foot part connected to the road: Heat transfers from the road to the container through this part.; RB1) Road base: Heat on this part is affected by direct solar radiation and heat transfer from the containers.; RB2) Road base under the container: This part receives heat from the road exposed to the sun

(a) Continuity equation rv ¼ 0

ð5:1Þ

  @v þ v  rv ¼  rp þ lr2 v þ fB q @t

ð5:2Þ

 @T þ v  rT dV ¼ @t

ð5:3Þ

(b) Momentum equation

(c) Energy equation Z

V

 qCp

Z

V

kr2 TdV þ

Z

ðqR þ qS ÞdS

S

where t represents the time, v is the velocity vector, q is the density, p is the pressure, l is the coefficient of viscosity, k is the thermal conductivity, Cp is the specific heat capacity, and T is the local

64

5

Simulation Model for Analysis of Energy Savings …

temperature. Furthermore, natural convection was modeled using the Boussinesq approximation, which uses a constant-density fluid model, but applies a local body gravitational force throughout the fluid, which is a linear function of the volumetric thermal expansion coefficient b and the local temperature difference. A buoyancy source was added to the momentum equation, as shown in Eqs. (5.4) and (5.5) (Spalding 2013): fB ¼  Dqg ¼  qref bðT  Tref Þg   qR ¼ uR eR er TR4  T 4

ð5:4Þ ð5:5Þ

where qref and Tref denote the density and temperature, respectively, at the boundary walls, and g is the acceleration due to gravity. The internal energy from the thermal radiation ðqR Þ was considered at the surface of the objective cell in the FVM within the immersed solid (IMMERSOL) per unit volume, as expressed in Eq. (5.5). Furthermore, uR is the radiation shape factor from the viewpoint of the radiation material; eR is the emissivity of the radiation material; e is the emissivity of the objective solid body; TR is the radiation temperature of the fluid; and r is the Stefan–Boltzmann constant (5.67  10−8 W/m2K4). When the solid areas of the model, such as the container walls, RS, or road, are considered, the second term on the right-hand side of Eq. (5.3) is neglected. Thus, the equation becomes that of heat conduction, except for the fluid near the surface of the solid body. In addition, solar radiation energy ðqS Þ was considered in this model. A Dirichlet boundary condition was used in the surrounding container to specify a uniform inlet velocity. Additionally, a no-slip boundary condition was assumed for all container walls. The inlet air temperature varied with time, and the walls were assumed adiabatic. The k − e turbulence model was used to close the set of governing equations.

5.5.3 Geometrical Model and Parameter Setting The geometrical model of the CFD simulation considered a two-dimensional analysis of the heat transfer in the middle section of the RC. The analysis was performed to reduce the computational time required for the CFD model. The middle section was assumed to be the most representative model for heat transfer through the container walls because a large amount of heat penetrated through the sides and ceiling owing to their large surface areas. Therefore, the doors and reefer sides were neglected in this model. Figure 5.13 shows the middle section of the RC and geometrical domain of the simulation model. The entire geometrical model comprised the RCs, RS, and a container storage yard. The overall dimensions of the geometric model were 704 m2 with widths and heights of 32 and 22 m, respectively. The geometrical model was the calculation domain of the simulation model, which comprised ambient air, container structures, cool air inside the containers, and the road as ground. Furthermore, the container structure contained three structural layers: external frame, insulation, and internal frame. The dimensions and positions of each object were considered in the measurement data. Furthermore, the parameters applied to each object domain of the simulation model included the material properties, radiation properties, and the initial temperature of the object. The details of the parameter settings for each part are summarized in Tables 5.4 and 5.5. The domain of each object was transformed into a finite volume through body-fitted mesh generation in the Cartesian framework. Subsequently, the solution domain was subdivided into several control volumes on a mono-block mesh using the conventional staggered-grid approach. The total mesh for the entire geometrical model contained 3,50,200 cells divided between each region of

5.5 RS Impact Analysis Using Simulation

65

Fig. 5.13 Geometrical model of the simulation Table 5.4 Parameter settings for air properties

Table 5.5 Parameter settings for material properties

Density (kg/m3)

1.22

Velocity (m/s)

0.3

Specific heat (kJ/kgK)

1.006

Initial temperature (°C)

23.39

Material properties

Thermal conductivity (W/mK)

Radiation properties (absorption and emission)

Aluminum

204

0.6

Polyurethane

0.03



Stainless steel

16



RS

0.84

0.5

Road

1.13

0.6

66

5

Simulation Model for Analysis of Energy Savings …

domain objects. In the narrow spaces in the domain, such as between the containers, a power-mesh ratio was introduced in the near wall to satisfy the calculation in the thin-layer region. The simulation model was set to a transient calculation for 12 h and was divided into 12 steps. The calculation began at 6:00 AM and continued until 6:00 PM. Furthermore, the total number of iterations was 100, with a global convergence criterion of 0.01. The boundary condition was treated as the source of quantities such as mass, momentum, and energy. Additionally, over the entire domain, the air velocity and radiative heat transfer were set as the boundary conditions and applied to the simulation model. An immersed solid model of radiative heat transfer has been used to handle surface-to-surface radiation (Spalding 2013).

5.5.4 Thermal Simulation In this section, the traits considered in the thermal simulation are discussed. Thermal analysis was conducted to explain the physical mechanism of heat-transfer problems and its effects on the temperature changes of various surfaces, particularly container walls. The heat transfer problems for this simulation model included the effect of dummy containers, heat penetration through the container walls, the initial temperature of the container walls, thermal buoyancy between the RCs, and monitoring cell locations.

5.5.4.1 Effect of Dummy Containers Dummy containers are artificial objects placed beside observed objects, and these containers serve as radiative walls to protect the object from direct exposure to sunlight. This geometrical technique was employed in the simulation model to handle the initial point of the surface temperature, which is expected to increase owing to solar radiation. This technique predicts that the initial points of the surface temperature are influenced by shading effects from other objects. Furthermore, dummy containers are beneficial for simulating a case with an unknown initial point. Figure 5.14 depicts the effect of dummy walls on the initial point of the ceiling surface temperature. In this case, three tiers of

Fig. 5.14 Effect of dummy containers on the simulation results

5.5 RS Impact Analysis Using Simulation

67

containers were placed on the south and north sides. The initial point of the surface temperature in the morning shifted from 06:00 AM to 07:30 AM, owing to the sunlight shaded by the dummy containers on the south side. The observed object was normally exposed to direct solar radiation when the solar altitude exceeded the height of the dummy container. Consequently, the surface temperature reached a peak at 12:00 PM, which was similar to that achieved without dummy walls at the same time. Following the attainment of the peak point, the temperature of the dummy containers decreased faster than normal because of the shade provided by the dummies on the north side. The temperature decrease continued until the afternoon, when the temperature trend was lower than that without dummy containers. The simulation technique suggested that future directions for the analysis should be related to the temperature fluctuations at certain measurement times.

5.5.4.2 Heat Penetration of the Container Walls The heat energy transferred from the surroundings to the cargo space of the RC is referred to as heat penetration, which is the physical mechanism of conductive heat transfer through container walls owing to the thermal invasion from solar radiation. Figure 5.15 shows the heat penetration of the container walls obtained from the simulation results. This simulation considered solar radiation under clear-sky conditions, with the air temperature inside the container maintained at 0 °C. The heat energy from solar radiation began to penetrate the walls at 07:00 AM. Furthermore, the penetration continued to increase gradually, reaching a maximum at 14:00. Consequently, the effect of heat penetration caused the inner surface temperature of the container walls to increase by approximately 5 °C at maximum penetration. Thereafter, the penetration gradually decreased until the balance condition was reached at 22:00. Heat penetrated the outer surfaces at a maximum temperature of 35 °C at noon. The temperature profile around the surface that interacted with the ambient air decreased exponentially until the ambient air temperature was reached. The temperature profile was influenced by the fluid motion of the air layer on the surface, wherein the changes in the convective heat-transfer coefficient followed the velocity of the air. When the wind was at a sufficient speed, the heat-transfer coefficient increased, thereby decreasing the surface temperature and resulting in reduced heat penetration through the walls. The temperature profile of the insulation layer decreased linearly with the temperature of the interior surfaces, which in the simulation was set to 0 °C. Under actual conditions, the decreased temperature profile did not reach 0 °C, owing to the heat penetration and temperature inside the container. Excessive heat penetration from the outside changes the temperature profile and causes the interior temperature to increase. The temperature profile of the insulation layer represents the ability of the material to inhibit heat penetration and is important for improving insulation performance.

Fig. 5.15 Heat penetration through the container walls

68

5

Simulation Model for Analysis of Energy Savings …

Fig. 5.16 Initial temperatures of the container walls

5.5.4.3 Initial Temperature of the Container Walls The ability of the RC to maintain cool air inside the cargo area can be assessed based on the thermal resistance of each wall. The container walls were important for reducing the heat penetration from external factors and comprised an insulation layer that was manufactured by following a standard design with the same thermal resistance as that of an RC. Under actual conditions, the thermal performance of the container walls is not fixed, even for RCs with the same production year and manufacturer. Further analyses related to these factors were performed using the initial temperature of the container walls at the saturation point. The surface temperatures were the lowest at the saturation point. Figure 5.16 depicts the initial temperatures on the ceiling and bottom surfaces of the RCs under the RS condition. The surface temperature on the container walls began to saturate at 23:00. Each wall reached its saturation point at different temperatures and times. The ceiling reached the saturation point at 24°C at approximately 06:00 AM. On average, the other surfaces reached saturation at 23°C at approximately 07:00 AM. The temperature at the saturation point represents the initial temperature at which heat begins to penetrate the walls. In the simulation analysis, the saturation point and initial temperature were important parameters for analyzing the surface temperature trend. The initial temperature affected the peak surface temperature, whereas the saturation point affected the curve shift of the surface temperature trend. Furthermore, high initial temperatures caused the peak surface temperature to be higher than expected, whereas low initial temperatures resulted in a lower peak. In addition, the early saturation points caused the peak to be reached earlier than expected, whereas the late saturation points caused the peak to occur later. 5.5.4.4 Thermal Buoyancy Between RCs Thermal buoyancy (also referred to as thermal stratification) is the temperature gradient that occurs on container surfaces owing to the natural convection flow. Natural convection within the air space causes warm air to rise to the ceiling because warm air is lighter than the surrounding cool air. Figure 5.17 illustrates the thermal buoyancy between the container tiers. The temperature profiles on the north and south faces were plotted together to explain this phenomenon, and the surface temperatures between the two exhibited the same trend for each tier. Thermal buoyancy was observed between the lower and upper tiers. The vertical spacing between each container tier was 3 m, the total height from the ground to the upper tier was 6.4 m. Further, the temperature difference from each tier was approximately 2.5°C; therefore, the total temperature difference between the lower and upper tiers was 10°C. These results are consistent with those of previous studies (Saïd et al. 1996; Walker

5.5 RS Impact Analysis Using Simulation

69

Fig. 5.17 Thermal buoyancy between container tiers

2006). Thermal buoyancy appeared clearly in the simulation and was consistent with the measurement results. In this case, thermal buoyancy was set in the buoyancy-driven case using the Boussinesq equation. Moreover, gravitational acceleration and surface velocity were essential parameters for the analysis.

5.5.4.5 Monitoring Cell Locations In the FVM, the location of the observed object monitoring point must be considered. The surface temperatures that provided different spot values in the different cells were considered as the monitoring points in this study. This circumstance was chosen because of the physical phenomena in the medium of the monitoring cells. In the case of the surface temperature, the location of the monitoring cells must be initially determined because the measurement result is obtained by sticking to the surface. The number of cells is an important factor affecting the observed object by influencing both the solid and air temperatures. The thinner cells in the medium near the solid wall were better at representing the surface temperature. Figure 5.18 shows the temperature profiles at different cell locations. The maximum temperatures obtained for the solid medium and air were different. The solid exhibited thermal conductivity and provided high temperature when the heat source was transferred. Subsequently, the high temperature of the solid gradually decreased to the air temperature, and suitable monitoring was performed on the layer of air near the wall.

5.5.5 Verification of the Thermal Simulation Results The simulation model was verified by comparing the simulation results with the experimental results, considering the important factors in the thermal simulation. The surface temperatures at the center of each wall were compared during the transient period from 06:00 to 18:00. The container used for verification is located at T3 in Fig. 5.19. Thereafter, a comparison was performed with and without RS under the clear-sky condition. Consequently, the surface temperatures from the simulation results were verified using data from August 15, 2015, for the condition with RS, and August 27, 2015, for the condition without RS.

70

5

Simulation Model for Analysis of Energy Savings …

Fig. 5.18 Temperature profile at difference monitoring cells

Basis of calculations

Roof Shade

(EC34)

R8 R7 R5 R5 R4 R3 R2 R1

(a) The condiƟon without roof shade

T3

T3

T2

T2

T1

T1 (North) R8 R7 R6 R5 R4 R3 R2 R1 (South)

(b) The condiƟon with roof shade

Fig. 5.19 Verification of the simulation result using the experimental measurements

Figure 5.20a shows the simulation and measurement results for the condition with an RS, which confirms that the temperatures at all locations obtained from the simulation were consistent with the measurements. The temperature was highest on the ceiling surface, followed by the south, north, and bottom surfaces. The simulation results fluctuated primarily on the southern lines because the altitude of the sun was lower than the height of the RS, thereby resulting in direct solar radiation exposure on the surface. Figure 5.20b shows the results without an RS, and confirms the consistency between the simulation and measurements. The ceiling surface was at its highest temperature, with its peak occurring at noon. Temperature fluctuations of the measurements were evident, particularly in the afternoon, and resulted in slight differences between the simulation and measurement values. Moreover, in the simulation, the temperature escalation over time tended to produce smooth lines owing to the time and calculation steps. A 1 h step interval was used in this study. The high southsurface temperature was followed by a peak in the morning. Thereafter, the temperatures of the north and bottom surfaces subsequently increased and reached a peak in the afternoon. As described above, the effectiveness of the proposed simulation model is confirmed.

References

71

(a) With RS

(b) Without RS

Fig. 5.20 Verification of the simulation result by comparing it with the experimental measurements

5.6

Conclusions

Although RCs are the primary energy consumers in port activities, the energy efficiency of RCs is mostly disregarded in strategy formulation for most ports and terminals, and energy-saving measures for RCs have rarely been the subject of research. This chapter focuses on energy consumption reduction and decarbonization measures in RCs, and a simulation method is proposed to estimate the amount of energy required for cooling that predicts the surface temperature of container walls using CFD to calculate the benefits of installing RSs in RC areas. A self-assembly type thermo-fluid analysis model was introduced as a simulation method. This type of model is determined by several small factors, unlike problems in which the main controlling factor is clear. Thermal factors, including small factors related to energy consumption, such as solar radiation from the sun, heat conduction on container walls and through roads, heat convection on surrounding container surfaces, and thermal radiation among solid surfaces, were considered. Then, the FVM, which is a discretization method, was used for numerical analysis by applying the governing equations related to the thermo-fluid analysis. A comparison with experimental results using RSs installed at Hakata Port in Japan confirmed the validity of the simulation model. The next chapter describes how this simulation model was used to calculate the actual energy savings and decarbonization benefits. Additionally, the economic feasibility of the installation is discussed based on installation costs.

References Acciaro M, Wilmsmeier G (2015) Energy efficiency in maritime logistics chains. Res Transp Bus Manag 17:1–7 Bjerkan KY, Seter H (2019) Reviewing tools and technologies for sustainable ports: does research enable decision making in ports? Transp Res Part D Transp Environ 72:243–260 CHAM (2010) PHOENICS-VR reference guide. Retrived from https://www.cham.co.uk/phoenics/d_polis/d_docs/ tr326/tr326top.htm Ilio DG, Giorgio PD, Tribioli L, Bella G, Jannelli E (2021) Preliminary design of a fuel cell/battery hybrid powertrain for a heavy-duty yard truck for port logistics. Energ Convers Manag 243:114423 Iris Ç, Lam JSL (2019) A review of energy efficiency in ports: operational strategies, technologies and energy management systems. Renew Sustain Energy Rev 112:170–182

72

5

Simulation Model for Analysis of Energy Savings …

Martínez-Moya J, Vazquez-Paja B, Maldonado JAG (2019) Energy efficiency and CO2 emissions of port container terminal equipment: evidence from the Port of Valencia. Energ Policy 131:312–319 Nicolis G, Prigogine I (1989) Exploring complexity. W.H.Freeman and Co Ltd. Patankar S (1980) Series in computational methods in mechanics and thermal sciences. Numerical heat transfer and fluid flow Rijsenbrij JC, Wieschemann A (2011) Sustainable container terminals: a design approach. Springer, New York Saïd MNA, MacDonald RA, Durrant GC (1996) Measurement of thermal stratification in large single-cell buildings. Energ Build 24(2):105–115 Shinoda T, Budiyanto MA (2016) Energy saving effect of roof shade for reefer container in marine container terminal. J Jpn Inst Navig 134:103–113 Spalding B (2013) Chapter one–trends, tricks, and try-ons in CFD/CHT. Adv Heat Transf, 1st edn., 45:1–78. Elsevier Inc. van Duin J, Geerlings H, Verbraeck A, Nafde T (2018) Cooling down: a simulation approach to reduce energy peaks of reefers at terminals. J Clean Prod 193:72–86 Walker E (2006) Methodology for the evaluation of natural ventilation in buildings using a reduced-scale air model. Ph.Dthesis, Massachusetts Institute of Technology Wang L, Peng C, Shi W, Zhu M (2020) Carbon dioxide emissions from port container distribution: spatial characteristics and driving factors. Transp Res Part D Transp Environ 82:102318 Werner B (2014) Reduction of the CO2 footprints of container terminals by photovoltaics, green efforts project. Technical report

6

Energy-Saving Effects and Economic Feasibility of Roof Shade Installation

6.1

Introduction

In this chapter, first, the proposed simulation model is used to calculate the actual energy-saving effects at Hakata Port. Although the simulation model proposed in the previous chapter was developed to evaluate the effects of the RS installed at HICCT and used detailed experimental measurements in summer as input values, such data are not usually available. If equivalent results can be obtained, even if the input values are general data that are easily available, the simulation model can be applied to other seasons and ports with different surrounding environments, such as temperature and solar radiation, and the versatility of the simulation model will be enhanced. This chapter verifies this hypothesis and confirms the seasonal and regional characteristics of the energy-saving effects of RS through simulations at other ports. This chapter examines the CO2 emission reduction effects and economic feasibility using the amount of energy saved. This is because, as described in Chap. 4, even CO2 emission reduction and energy-saving effects are confirmed, it may be difficult to implement depending on economic feasibility, or conversely, it may lead to the consideration of issues and strategies for the introduction. In this chapter, actual data, such as installation costs and electricity charges at Hakata Port, are used for economic evaluation. The content of this chapter is very valuable in terms of showing the quantitative effects of climate change countermeasures in the RC areas, and additionally, because it is an empirical analysis, it would serve as a useful guide for ports worldwide to consider the implementation of RSs.

6.2

Energy Savings Owing to the RS at Hakata Port

The energy-saving effect of the RS was estimated using the simulation model developed in the previous section. The heat penetration through the container walls is equal to the energy consumed by the refrigeration unit to maintain a set temperature. The heat penetration was calculated from the heat transferred across the exterior surface to the inside surface, thereby indicating that heat penetration can be obtained from the temperature difference between the exterior and interior surfaces of the container walls. The heat penetration of the container wall was calculated using Eqs. (6.1), (6.2) and (6.3) (Tanaka et al. 2006).

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0_6

73

6 Energy-Saving Effects and Economic Feasibility …

74

qwall ¼ K  A  ðDT Þ K ¼

1 a

þ

1 n P xi i¼1

ki

þ

1 a0

qtotal ¼ qCeiling þ qSouth þ qNorth þ qBottom

ð6:1Þ ð6:2Þ

ð6:3Þ

Equation 6.1 expresses the thermal transmission and is used to calculate the heat penetration across an RC wall. Parameter A in Eq. (6.1) denotes the area of the wall surface (m2), while ΔT represents the temperature difference between the interior and exterior surfaces of the RC (°C). Furthermore, K represents the overall heat-transfer coefficient of the RC wall, which is calculated using Eq. (6.2). In Eq. (6.2); k represents the thermal conductivity of each material; x represents the thickness of the insulation material; and ao and ai represent the convection heat-transfer coefficients for the exterior and interior air layers, respectively (23 and 9 W/mK, respectively). Therefore, the overall heattransfer coefficient of the RC wall was estimated to be 0.44 W/m2K. In Eq. (6.3), qCeiling , qSouth , qNorth , and qBottom represent the heat from the ceiling, south, north, and bottom walls, respectively. Furthermore, the heat penetration was calculated for each RC with and without RS. Using the change in temperature over time obtained from the simulation, the sum of the heat penetration through each wall of the RC, which is the amount of electricity required to maintain the temperature inside the RC, can be obtained using Eq. (6.3). In this section, the evaluation of energy savings was conducted at the location of the RC storage yard at HICCT. The middle columns represent the impact of the RS, as the surface of the RC was fully shaded when shade was present, and the energy savings of the middle column were estimated. Thereafter, the energy savings were evaluated using the ratio of energy consumption of the RCs. This ratio uses the energy consumption at the top tier of the middle column without RS (EC34 in Fig. 6.1), which is the basis of energy savings. The benchmark for the energy consumption ratio is expressed by Eq. (6.4): REti ¼

ECti  EC34  100 EC34

ð6:4Þ

where REti represents the ratio of the energy consumption of the RC in the objective location and ECti represents the objective energy consumption. The representative position is the RC at the top tier of the middle column in the condition without the RS. The total ratio of the energy consumption (TRE) in the middle column was calculated using Eq. (6.5): TRE ¼

3 X REt4 t¼1

3

ð6:5Þ

As described in the previous chapter, an experiment to verify the energy-saving effects of installing the RSs was conducted using six RCs with two rows of three tiers in parallel. Owing to the limitations of the experimental facility, the dates of the experiments with (August 15, 2015) and without RSs (August 27, 2015) were different. Therefore, the effectiveness of the simulation was examined by combining experimental and simulation results. Table 6.1 lists the estimated energy savings obtained from experiments and simulations.

6.2 Energy Savings Owing to the RS at Hakata Port

75

Basis of calculations

Roof Shade

(EC34) T3

T3

T2

T2

T1

T1 (North) R8 R7 R6 R5 R4 R3 R2 R1 (South)

R8 R7 R5 R5 R4 R3 R2 R1

(b) Condition with roof shade

(a) Condition without roof shade Fig. 6.1 RC location as basis for energy savings calculations Table 6.1 Estimated energy savings Measurements

Simulation 1

Simulation 2

Simulation 3

Without RS

August 27, 2015

August 27, 2015

August 15, 2015

August 15, 2015

With RS

August 15, 2015

August 15, 2015

August 15, 2015

August 15, 2015

Item Without RS

T3

qtotal

RE

qtotal

RE

qtotal

RE

qtotal

RE

2,094

0

2,068

0

2,179

0

2,157

0

T2

1,770

− 15.46

1,780

− 13.94

1,808

− 16.99

1,838

− 14.75

T1

1,749

− 16.46

1,773

− 14.28

1,789

− 17.89

1,786

− 17.17

1,816

− 13.29

− 10.64

Average of column With RS

T3

− 9.41 1,785

− 13.72

− 11.63 1,785

− 18.08

− 10.64 1,662

− 22.91

T2

1,620

− 22.64

1,587

− 23.29

1,587

− 27.16

1,630

− 24.40

T1

1,425

− 31.94

1,481

− 28.42

1,481

− 33.03

1,606

− 25.52

Average of column

− 22.62

− 21.81

− 25.76

− 24.28

Difference in TRE

− 11.98

− 12.40

− 14.13

− 13.64

The difference in TRE was 11.98% in the measurement column. As shown in column Simulation 1, the heat penetration values obtained using thermal simulations based on solar radiation on different days were similar. For example, the simulated value obtained (T3) in the Simulation 1 column without the RSs was 2,068 W and measured value was 2,094 W. The other positions and the case with the RSs showed a similar trend. The difference in TRE is 12.40%, which is close to the energy savings shown in the measurement column. This result confirmed the validity of the simulation model. According to Simulation 2, which estimated the heat flow using thermal simulations based on solar radiation with and without RSs on the same day, the difference in TRE was 14.13%, which is the final energy-saving effect of the RS installation based on the simulation. In Simulation 3, the heat penetration was simulated using data from the Japan Meteorological Agency (JMA) for solar radiation, temperature, and wind speed on August 15, 2015, to increase versatility. Unlike in Simulation 2, in the case of RSs, the heat penetrations are averaged, and the differences in quantity and stacking are less likely to be apparent. However, the TRE is − 25.76% for Simulation 2 versus − 24.28% for Simulation 3, which is comparable. The total difference in the heat penetration was 13.64%, which was not substantially different from that in Simulation 2. Therefore, the JMA data can be used to estimate the effect of RSs.

6 Energy-Saving Effects and Economic Feasibility …

76

6.3

Seasonal and Regional Characteristics of Energy Savings

Based on the results, the simulation model for calculating the energy-saving effects of RS, using JMA data as input values, is versatile and can be applied to other seasons and regions. In this section, the seasonal and regional characteristics of the energy-saving effects are verified using the JMA data. Specifically, simulations were conducted for the six ports shown in Fig. 6.2, which are considered to have different climatic characteristics in Japan, using JMA data for daily temperatures and solar radiation for sunny days closest to the middle of January, April, July, and October. First, to confirm the differences in the input data for the six ports, the highest values of temperature and solar radiation for each month are shown in Figs. 6.3 and 6.4. The time at which the highest values are recorded differs from port to port. The temperature is higher in July, October, April, and January, and the insolation is higher in July, April, October, and January, in that order, with higher latitudes presenting higher temperatures and insolation. Figure 6.5 shows the difference in TRE for each month at each port. Seasonally, January had the largest difference in TRE, followed by April and October, which were at the same level, and July had the smallest difference. Muroran Port had the largest difference in TRE by region, which generally decreases as latitude decreases, and Naha Port had the smallest difference. Thus, the energy-saving effect of RS is larger in winter than in summer, and larger in cold regions than in warm regions. The difference in results from intuitive expectations are because the main factor in summer and warmer regions is temperature. The effect of solar radiation is relatively small, reflecting the fact that the change in RC surface temperature is a self-assembling phenomenon.

Fig. 6.2 Ports selected for analysis

6.4 Economic Analysis of RS Installation

40

Muroran

77

Niigata

Sendai Shiogama

Osaka

Hakata

Naha

35 30 25 20 15 10 5 0 January

April

July

October

Fig. 6.3 Highest monthly input value (temperature) at each port

1200

Muroran

Niigata

Sendai Shiogama

Osaka

Hakata

Naha

1000 800 600 400 200 0 January

April

July

October

Fig. 6.4 Highest monthly input value (solar radiation) at each port

6.4

Economic Analysis of RS Installation

In this section, the impact of energy savings from the installation of RSs on electricity charges and CO2 emissions is verified. The electricity charges for the entire HICCT, where the RSs are installed, are used to verify the impact on electricity charges and CO2 emissions, taking into account seasonal characteristics. First, the amount of energy saved by installing RSs can be calculated using the following Eqs. (6.6) and (6.7):

6 Energy-Saving Effects and Economic Feasibility …

78

30

Muroran

Niigata

Sendai Shiogama

Osaka

Hakata

Naha

25 20 15 10 5 0 January

April

July

October

Fig. 6.5 Difference in TRE by season and port based on simulation

ES ¼

12 X

esi ¼

i¼1

12 X

ðPi  TREi  Si Þ

ð6:6Þ

i¼1

RES ¼

ES 12 P i¼1

ð6:7Þ

TPi

where ES is the annual energy saved (kWh), esi is the energy saved in month i (kWh), Pi is the power consumption of RC in month i (kWh), TREi is the energy-saving rate in month i (%), and Si is the percentage of daylight hours in month i (%). The electricity charges consist of the basic charge, which is determined by the contract power calculated based on the power consumed over 30 min, and the usage charge, which is the unit price per 1 kWh and is collected only if electricity is used. In the case of HICCT, the unit price of the usage charge differs according to the time zone and season as follows: Peak (summer months from July to September, 13:00 to 16:00):15.32 yen/kWh, summer daytime (off-peak hours from 8:00 to 22:00 in summer):13.13 yen/kWh, daytime (off-peak hours from 8:00 to 22:00):12.28 yen/kWh, and nighttime (off-peak and daytime):8.59 yen/kWh. The reduction in electricity charges through energy saving can be calculated using the following Eqs. (6.8) and (6.9): CB ¼

 12  X Px Cib   TREx þ esi  Ui TPx i¼1 RCB ¼

CB 12 P i¼1

ð6:8Þ ð6:9Þ

ðCib þ Ciu Þ

where CB is the electricity charge reduction (yen); Cib is the basic charge in month i (yen); Px , TPx , and TREx are the overall electricity consumption (kWh) in month x, when the electricity consumption is maximum after installation of RS, the electricity consumption in the RC area (kWh), and

6.5 Discussion

79

energy-saving rate (%), respectively; Ui is the weighted average electricity charge (yen/kWh) considering the charges for each daylight hour in month i; RCB is the reduction rate of electricity charges (%); and Ciu is the electricity charge (yen) in month i. CO2 emission reductions can be calculated using the following Eq. (6.10). RE ¼

12 X

esi  ef

ð6:10Þ

i¼1

where RE is the CO2 emission reduction (t-CO2) and ef is the CO2 intensity of electricity (t-CO2/ kWh). As only the TREi of the representative month of each season was calculated in the previous section, TREi from December to February (winter) was defined as TRE1 , from March to May (spring) as TRE4 , from June to August (summer) as TRE7 , and from September to November (autumn) as TRE10 ; Si was used for each three-month period rather than monthly. Similarly, when divided into four periods, the period of maximum power consumption after RS installation is from June to August; thus, TRE7 was defined as TREx , with Srx being the power consumption rate of the RC area from June to August. Note that the RS system is not currently in operation at the HICCT; thus, the obtained power consumption and electricity charges for the HICCT are for the case when there is no RS. The results are shown in Table 6.2. Although the power consumption of the RS was larger in summer, the energy-saving rate was the lowest in summer. Furthermore, because the ratio of daylight hours also has an effect on the energy savings, there is no significant difference in the amount of energy saved by the RS, except in winter, but the order of the amount is spring, autumn, summer, and winter. The reductions in electricity charges and CO2 emissions also follow the same order, but the reduction in basic electricity charges during the summer, when electricity consumption is at its maximum after the installation of the RS, is applied throughout the year. Thus, the results indicate that the installation of the RS leads to an annual reduction of 4.1%, or 4.54 million yen in electricity charges and a reduction of 1.3%, or 49.9 t-CO2, in CO2 emissions. Regarding the reduction in electricity charges due to installation of RSs, the impact of the reduction in basic electricity charges through peak shaving is larger. The power consumption of CTs does not change significantly seasonally except for RCs, but because RCs have a large share to begin with and further increase in summer, the energy-saving rate in summer is the most important in terms of the impact on electricity cost reductions. Accordingly, because the energy-saving rate is the smallest in summer by season, the impact on peak shaving is also relatively small; conversely, if the energy-saving rate in summer is large, the effect of the reduction in basic rates will be enhanced throughout the year.

6.5

Discussion

From an energy-saving perspective, because no other practical countermeasures have been sufficiently demonstrated academically, the installation of the RS is one of the few RC-related energysaving countermeasures worth promoting as a policy measure, even though the energy-saving is only 1.9% in the case of HICCT. However, the method of promotion depends on the economic feasibility of the installation. The RSs at Hakata Port were installed as part of a national pilot project and operated by Hakata Port Terminal Co., Ltd. (HPT), the POC at Hakata Port. The installation cost was 260 million yen, and construction lasted from February 2010 to January 2011, with the government bearing 100% of

6 Energy-Saving Effects and Economic Feasibility …

80

Table 6.2 Energy saving, electricity charges reduction, and CO2 reduction effects of HICCT Item

DecemberFebruary

March– May

June– August

September– November

Annual

Source

Overall power consumption of HICCT (a) (kWh)

1,554,389

1,700,084

1,841,513

1,687,980

6,783,966

Actual results in 2020

Of which, power consumption of RC (b) (kWh)

707,622

897,515

1,055,265

887,274

3,547,676

Actual results in 2020

Ratio (%)

45.5

52.8

57.3

52.6

52.3

b/a

Energy-saving rate by RS (c) (%)

20.34

14.66

12.96

16.01

15.54

Fig. 6.5

Ratio of daylight hours (d) (%)

16.3

28.4

24.0

24.5

23.3

JMA data

Energy saved by RS (e) (kWh)

23,461

37,367

32,823

34,803

128,454

bcd

Ratio (%)

1.5

2.2

1.8

2.1

1.9

e/a

Overall electricity charges of HICCT (f) (yen)

25,937,937

27,194,943

30,040,350

27,751,893

110,925,123

Actual results in 2020

Of which, basic charge (g) (yen)

10,105,286

10,105,286

10,105,286

10,105,286

40,421,146

Actual results in 2020

Of which, usage charge (h) (yen)

15,832,651

17,089,657

19,935,064

17,646,606

70,503,978

Actual results in 2020

Weighted average electricity charges (i) (yen/kWh)

12.00

11.67

12.24

12.18

12.01

*1

Reduction of basic charge by RS (j) (yen)

750,712

750,712

750,712

750,712

3,002,849

e  g  b/a

Ratio (%)

7.4

7.4

7.4

7.4

7.4

j/g

Reduction of usage charge by RS (k) (yen)

274,593

437,507

402,770

423,751

1,538,621

ei

Ratio (%)

1.7

2.6

2.0

2.4

2.2

k/h

Electricity charge reduction by RS (l) (yen)

1,025,305

1,188,220

1,153,482

1,174,463

4,541,470

j+k

Ratio (%)

4.0

4.4

3.8

4.2

4.1

l/f

CO2 emissions from entire HICCT (m) (t-CO2)

858.5

919.1

954.2

891.9

3,623.6

Actual results in 2020

CO2 emission reduction by RS (n) (t-CO2)

8.6

13.6

12.0

12.7

49.9

e  CO2 intensity

Ratio (%)

1.0

1.5

1.3

1.4

1.3

n/m

*1 (Electricity charge for each month and each time period during the period x daylight hours during the concerned month and time period)/Total daylight hours during the period

6.6 Conclusion

81

the cost. The reduction in electricity charges is an economic benefit for POC, and an evaluation was conducted to determine whether the installation is feasible on a business basis to the initial investment. The reduction in electricity charges is 4.54 million yen in the case of Hakata Port HICCT, which handles 35,000 TEU RCs, and the initial investment that can be recovered in 12 years of service life is 42.62 million yen, assuming a social discount rate of 4% and no maintenance costs. The estimated cost of this measure for HICCT, if RSs are installed in the entire RC area is 2 billion yen (an initial investment of 260 million yen for the installation of two bays, which corresponds to 13% of the RC area); thus, it would be difficult to introduce this measure in HICCT on a business basis unless the cost is approximately 1/50th of that, or 2.78 million yen per bay. According to an interview with the person in charge of the installation at the time, the RS was designed to be massive in consideration of safety and other factor; it is possible to simplify the structure, which would reduce the cost. As previously described, the RS at Hakata Port was installed with the government bearing 100% of the cost. The fact that HPT has not installed any additional RSs at its own expense since the pilot project indicates that it would be difficult to install RSs on a business basis. In addition to clarifying the existence of energy-saving effects and the difficulty of installing RS on a business basis, the fact that either the development of inexpensive structures through research or public support is required for the introduction of RS should be taken as an outcome of the pilot project. In future, it will be necessary to support technological development to maximize energy-saving effects and reduce initial costs through effective structures and materials, and develop a supportive financial policy when economic feasibility cannot be guaranteed even with such support. The CO2 emission reduction effect is quite limited at 1.3%, even at the Port of Hakata ICCT, where the electrification of cargo-handling machinery has already progressed. This value is expected to be even lower for CTs in other ports where electrification has not progressed because the CO2 emissions of RCs account for a smaller share of the total. However, the CO2 emissions of RCs are currently highly dependent on the CO2 intensity of the purchased electricity, and the use of renewable energy sources for electricity is more effective in terms of CO2 emission reductions. If a structure covering the RS with solar panels becomes feasible, a synergistic effect between shading and on-site power generation will be achieved. This highlights the necessity of policy support for technological development.

6.6

Conclusion

Although the reefer area has a high energy consumption and is a major source of CO2 emissions in CTs, the energy efficiency of these containers has mostly been disregarded in terminal strategies, and empirical studies on RCs are scarce. In the previous chapter, a simulation method was developed to calculate the effect of RS installation as an energy-saving measure, and the effectiveness of this method was confirmed by comparing the simulation results with experimental data from Hakata Port. In this chapter, the energy-saving effect of introducing roof shades using the simulation results was quantitatively estimated, and it was also confirmed that JMA data could be used to estimate the effect of RSs. To verify the seasonal and regional characteristics of the energy-saving effects of RS installation, simulations were conducted for six ports, which are considered to have different climatic characteristics in Japan, using JMA data for daily temperatures and solar radiation in January, April, July, and October. The simulation results indicated that the energy-saving effect of RS is larger in winter than in summer, and larger in cold regions than in warm regions. This deviation in results from expectations can be attributed to the fact that the main factor in summer and warmer regions is temperature, and the effect of solar radiation is relatively small, indicating that the change in RC surface temperature is a self-assembling phenomenon.

82

6 Energy-Saving Effects and Economic Feasibility …

The reduction in electricity consumption charges was demonstrated through an economic evaluation. The results revealed that introducing the system from a business perspective is challenging. This is because the installation cost cannot be recovered during the operational period based on the money saved by the reduction in power consumption alone. The RSs at Hakata Port were installed as part of a government national pilot project and operated by the HPT, the POC of Hakata Port. The reduction in electricity charges was 4.54 million yen in the case of Hakata Port HICCT, which handles 35,000 TEU RCs. The initial investment that can be recovered in 12 years of service life is 42.62 million yen. The estimated cost for introducing RSs in the entire RC area of the HICCT is 2 billion yen, making it economically unfeasible to introduce this measure on a business basis. A supporting evidence to this result is the fact that HPT has not installed any additional RSs at its own expense. However, because no other practical countermeasures have been sufficiently demonstrated academically, the installation of RSs is one of the few RC-related energy-saving countermeasures worth promoting, even though the energy savings are limited. In the future, it will be necessary to support technological development to maximize energy-saving effects and reduce initial costs through effective structures and materials; additionally, a suitable financial policy should be developed in case economic feasibility cannot be ensured with technical and logistical support. The major contribution of the proposed simulation model and analysis of energy-saving effects is that it provides a powerful tool for the preliminary evaluation of roof shade installation at ports. Many climate change countermeasures other than roof shading can be adopted by ports, and the selection and combination of these measures is difficult. Further research, such as the simulation method introduced in this study, is required to support decision makers in determining practical applications.

Reference Tanaka S, Iwata T, Tsuchiya T, Akimoto T, Terao M, Takeda H (2006) Architectural environment engineering. Inoue Shoin

Part IV New Possibilities for Climate Change Countermeasures in Ports

A long-term effort is required to address the gap between the current adoption of climate change countermeasures in ports and ambitious emission reduction targets expressed by international organizations. In particular, many countries have decided to implement stronger climate change countermeasures to become CN. This is because not only are the NDCs set by each country under the Paris Agreement insufficient to achieve their targets, but global GHG emissions in 2030 indicate that global warming is likely to exceed 1.5 °C by the end of the 21st century, necessitating a rapid response. In addition to the need for long-term efforts, proposals for new countermeasures in ports backed by practical and academic support are urgently needed. Part IV presents two examples of new potential climate change countermeasures in ports. One is the effective use of dredged soil generated by port development projects, and the other is the role of ports as an import base for recyclable resources. Intuitively, it is difficult to connect either of these themes with climate change countermeasures. However, even these seemingly unrelated countermeasures will be necessary to achieve CNPs in the near future. Additionally, it is natural that the possibilities of climate change countermeasures that can be implemented in ports can be greatly expanded, considering the various functions of ports as infrastructure.

7

Carbon Containment and Creation of BCEs Through Beneficial Utilization of Dredged Soil Generated by Port Development Projects

7.1

Introduction

The preceding chapters discussed climate change countermeasures related to cargo-handling machinery and RCs. The current level of implementation of climate change countermeasures is not conducive to the realization of the ambitious emission reduction targets expressed by international organizations (Sornn-Friese et al. 2021). Thus, scientific and practical proposals for new countermeasures are required. In this regard, the list of climate change countermeasures in ports to date mostly pertain to emission source countermeasures, and very few studies have presented sink countermeasures (Bosman et al. 2018; IS2S 2013; Lam and Notteboom 2014). However, ports can be important players in carbon capture and storage (CCS) (Acciaro et al. 2014), and leading practices for CCS in ports are beginning to emerge. For example, the Port of Rotterdam is working with the Port of Antwerp as a CO2 transport hub and offshore storage project, which is expected to be the first large-scale CCS project in an EU member state (Wright 2022). Similarly, the Port of Copenhagen-Malmö announced its participation in a project to store and distribute the captured CO2 to ships that will transport it to storage in old oil fields (Tank News International 2020); additionally, the Port of Corpus Christi announced its participation in a project to sequester CO2 into a saline aquifer (Port of Corpus Christi 2022). However, in these projects, although the ports are involved as CCS fields and CO2 transport bases, the main actors are the industries in which they are located, not the ports themselves. This chapter deals with carbon storage that is proactively conducted as a port activity. Furthermore, although environmental impacts are sometimes mentioned in port development (Lam and Li 2019), they have not been recognized as climate change countermeasures. Potential climate change countermeasures in ports were previously discussed in Chap. 1. The creation of BCEs and beneficial utilization of dredged soil are the focus of this chapter. Blue carbon refers to the carbon sequestered within marine ecosystems as a result of biological activity in the ocean, as defined by the UNEP in 2009. It has recently attracted attention as a mitigation measure for global warming (Lovelock and Duarte 2019). The ocean absorbs slightly more than half of the CO2 absorbed by organisms on Earth, and more than half of the carbon absorbed by the ocean is also absorbed by shallow-water areas, such as mangrove forests, salt marshes, and seagrass beds (Nellemann et al. 2009); these areas have been categorized as BCEs. As several aspects of BCEs have not been scientifically verified (Lovelock and Duarte 2019; Macreadie et al. 2019), depending on their elucidation, future expansion as sinks can be expected. As Taljaard et al. (2021) argued for the incorporation of the natural environment as an integral component of port infrastructure systems, if BCEs, which are CO2 sinks, are created as part of port development, it is © The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0_7

85

86

7 Carbon Containment and Creation of BCEs …

acceptable to call them a climate change countermeasure in ports. PIANC (2019) reports the use of coastal areas with vegetation and dredged soil as carbon sinks. Furthermore, the effective use of dredged soil associated with port development has the potential to be a new sink countermeasure. Although the effective use of dredged soil to create BCEs for CO2 absorption has already been demonstrated (PIANC 2019), the carbon contained in the dredged soil itself has not been focused on thus far. Because dredged soil contains a large amount of carbon, it has the potential to substantially contribute to CO2 absorption if it is contained in the ground for the creation of seaweed beds and tidal flats, backfilling of depressions, and land reclamation. Government and PMBs in Japan have been working on port projects, such as the creation of eelgrass beds and the development of biologically symbiotic port structures. Although these projects were not initially intended as climate change countermeasures, they represent the creation of BCEs that can serve as carbon sinks. While most countermeasures previously discussed have addressed “emission sources” in “port operations”, the containment of organic carbon in dredged soil and the generation of BCE are novel in that they are “sink countermeasures” implemented during the “port development” process. Accordingly, this chapter examines the possibility of such climate change countermeasures in ports based on Japanese experience and provides a more reliable basis for policymaking concerning ports.

7.2

Carbon Storage Effect of Beneficial Utilization of Dredged Soil

The benefits of dredged soil for land development and wetland restoration have been acknowledged globally (Taljaard et al. 2021). In Japan, from the perspective of the biological growth environment and water quality purification, the development of tidal flats, seagrass meadows, and macroalgal beds, which are BCEs, and biosymbiotic port structures that serve as a foundation for the growth of BCEs, have been promoted as a port environmental policy from a relatively early stage. Between 1979 and 2018, port projects created approximately 80 BCEs (MLIT 2020). Dredged soil has been widely employed in the creation of tidal flats, seagrass meadows, and macroalgal beds because of its high organic matter and water content, which is used as a foundation material for habitats. Approximately 30–50 cm of sand is placed on top of the dredged soil to create tidal flats, seagrass meadows, and macroalgal beds with healthy habitats. During the high economic growth and rapid development of coastal areas from the 1950s to the 1970s, many borrow pits were left across the coastal areas of Japanese ports because the soil and sand had been collected and used as reclamation material for port construction. Recently, these borrow pits have been backfilled to alleviate the frequent occurrence of hypoxia in waterbodies (Kusuyama et al. 2017). After backfilling, it is possible to form healthy marine habitats by covering the surface with sand; thus, eutrophic dredged soil can be used in this process. Recently, the utilization of BCE as a climate change countermeasure has been included in Japanese port environmental policies, which refer to the beneficial use of dredged soil as a foundation material for BCEs. However, the main focus is on BCE as a sink and not on the containment of carbon in dredged soil. This chapter’s perspective is novel in this respect. As a basis for the discussion in this chapter, the amount of carbon in dredged soil and the status of effective utilization of dredged soil are shown in Tables 7.1. Of the current dredged soil volumes in Japan, 13% are used for beach and tidal flats and 6% for backfilling borrow pits (Table 7.1). Permanently storing this amount of carbon in the ground can help reduce GHG emissions from ports. For the containment effect of carbon in the dredged soil, two distinct concepts are used, as shown in Fig. 7.1. According to the first interpretation, dredging is the removal of the stored carbon. When the removed dredged soil is unloaded, the organic matter in the dredged soil decomposes under aerobic conditions and CO2 is released. Conversely, removal of

7.2 Carbon Storage Effect of Beneficial Utilization of Dredged Soil

87

Table 7.1 Amount of carbon in dredged soil and status of effective utilization of the dredged soil Parameters

Value

Source

Annual amount of dredged soil (status)

159,500,000 m3

Kuwae et al. (2019)

Carbon content of dredged soil

44.7 tCO2 per 1000 m3

Density = 1.5 gcm−3, soil particle density = 2.6 gcm−3, total organic carbon (TOC) = 1.5% (Naito et al. 2008)

Carbon content in the annual amount of dredged soil

712,965 tCO2

Multiplication of the values of the previous two parameters

Utilization rate for backfilling borrow pits

5.9%

Ministry of Finance (2018)

Utilization rate for seagrass meadows, macroalgal beds, and tidal flats

12.9%

Utilization rate for land development

53.2%

(a) Interpretation 1 CO2

Storage on land

Dredging

Seagrass meadows Macroalgal beds

Containment of dredged soil

Extraction

Sedimentation of organic matter Difference in carbon residual is emissions reduction

Carbon storage

(b) Interpretation 2 Seagrass meadows Macroalgal beds

Dredging

Containment of dredged soil Carbon storage Sedimentation of organic matter

Fig. 7.1 Two interpretations regarding carbon in dredged soil

88

7 Carbon Containment and Creation of BCEs …

dredged soil can reduce the decomposition of organic matter. The difference in CO2 emissions between the two is the reduction in CO2 emissions due to dredged soil. As per the second interpretation, because the sedimented organic matter in a port is periodically removed through maintenance dredging of the navigation channel, it should not be regarded as a carbon storage resource; however, only the amount of organic matter contained after dredging is regarded as storage. Utilization of dredged soil can be a potential climate change countermeasure in both interpretations. In particular, tidal flats, seagrass meadows, and macroalgal beds created with dredged soil can be compatible with shallow-water BCE, simultaneously providing a beneficial use of resources and a novel, nature-based solution to climate change (Raymond et al. 2017). In this regard, moving forward, the selection method for dredged soil utilization should include carbon balance as one of the indicators.

7.3

Quantitative Effects of Organic Carbon Containment of Dredged Soils

In this section, the potential for carbon sequestration by BCE as a sink measure and the effective use of dredged soil containment are examined. Specifically, estimations of the contribution to achieving carbon neutrality when these sink measures are considered are compared to port-related CO2 emissions. The CO2 emissions from Japanese ports are listed in Table 1.1, and the amount of CO2 sequestration by BCE was estimated as 1.32 million t CO2yr−1 in Japan (Kuwae et al. 2019), The port component can be estimated using the ratio of ports along the coastline of Japan. Based on the two perspectives described in the previous section, estimates of carbon stock in the bottom soil of ports, CO2 emissions by dredging, and reduction of emissions by effective utilization of dredged soil, which are necessary for the first perspective, and carbon stock through the utilization of dredged soil, which is necessary for the second perspective, are calculated. It is also necessary to consider the amount of CO2 emissions associated with dredging and seaweed bed development. The residual rate of carbon in each effective utilization method of dredged soil is shown in Table 7.2 and the calculation method and results are shown in Table 7.3. To evaluate these effects, the results, including CO2 emissions at ports and carbon storage by BCE at present, are shown in Fig. 7.2; subsequently, the CO2 emissions and storage of all ports in Japan were compared. Although storage by BCE and the beneficial utilization of dredged soil resulted in an effect of 378,000 tCO2yr−1, this amount is quite small compared to the amount of emissions. Notably, the amount of CO2 emissions reduction (399,000 tCO2yr−1) from dredging projects exceeded that of storage or emissions reduction (255,000 tCO2). Thus, it does not represent a practical storage measure unless devices are fabricated for both beneficial utilization and construction methods. At present, the beneficial utilization of dredged soil wisas smaller than that of CO2 emissions from port activities (728,000 and 796,000 tCO2, respectively).

7.4

Future Significance of Containment of Organic Carbon in the Dredged Soil and the Creation of BCE

The results in the previous section suggest that drastic reforms are required to realize carbon neutrality in ports because of the high CO2 emissions compared to the storage effect of the BCE and organic carbon containment. In this section, the future significance of organic carbon containment and creation of BCE through the effective use of dredged soil as a foundation material for BCE as a new climate change countermeasure for the realization of CNPs will be discussed through scenario

7.4 Future Significance of Containment of Organic Carbon in the Dredged Soil and the Creation of BCE

89

Table 7.2 Preconditions for calculation of carbon storage by effective utilization of dredged soil Items

Value (%)

Source

Utilization rate for backfilling deep excavation sites

5.9

Ministry of Finance (2018)

Utilization rate for seaweed beds and tidal flats

12.9

Utilization rate for land development

53.2

Residual rate for backfilling deep excavation sites

90

Residual rate for seaweed beds and tidal flats

80

Residual rate for land development

30

Author’s setting is based on Nakagawa (1998) and Okada et al. (2016)

Table 7.3 CO2 emissions, storage, and emissions reductions at Japanese ports No.

Methods

1,000s tCO2

Source

1

Ship-to-shore cranes

39

Table 1.1

2

Yard cargo-handling machinery

347

3

Reefer containers

137

4

Administrative buildings

262

5

Dredging projects (Construction machinery etc.)

399

Annual amount of dredged soil  TOC emission unit (0.025 tCO2m−3) (Hayashi 2011)

6

Dredging projects (Emissions of stocked carbon)

713

Table 7.1

7

Ships (at arrival and departure)

1,320

Table 1.1

8

Ships (at berth)

3,413

9

Trucks (congestion at the gate)

91

10

Trucks (in-port drayage)

46

11

Trucks (hinterland transportation)

6,986

12

Carbon storage in sediment soil within ports

− 781

Port area (634,000 ha)  TOC sedimentation unit (1.23 tCO2ha−1) (Kubo 2015)

13

Emissions reductions or increased carbon storage through the utilization of dredged soil

− 225

Annual amount of dredged soil (Table 7.1)  carbon content in dredged soil (Table 7.1)  utilization rate  residual rate (Table 7.3)

14

BCE

− 162

The current sink of 1.32 million tons (Kuwae et al. 2019)  port shoreline extension ratio (12%)

15 16

Total (first perspective)

12,523

Sum of No. 1–14

Port activities only

728

Sum of No. 1–6, 12–14

Total (second perspective)

12,652

Sum of No. 1–5, 7–11, 13, 14

Port activities only

796

Sum of No. 1–5, 13, 14

analysis. Specifically, the amount of CO2 generated and stored (or reduced emissions) by the port in 2030 and 2050 were estimated according to the two scenarios. Scenario 1 was the case where the effective utilization ratio of dredged soil for BCE foundation materials increased, while the emission source countermeasures that have been confirmed to be effective in previous studies and leading cases were implemented. Scenario 2 assumes that an absolute conversion to next-generation energy is the

7 Carbon Containment and Creation of BCEs …

90

Storage by BCE Carbon stock in ports Reduction of carbon emissions (or carbon storage) by beneficial utilization of dredged soil Emissions from trucks Emissions from ships Emissions from port activities Emissions from dredging projects Emissions of carbon stored in ports 15000

12000

9000

6000

3000

0

-3000

Interpretation 1

Interpretation 2

Fig. 7.2 CO2 emissions and storage of all ports in Japan

target, volume of dredging projects increases, and ratio of effective utilization for BCE foundation materials further increases. In both scenarios, we assumed that construction innovations were developed and ratio of carbon residuals was improved from that mentioned in the previous subsection. The details of each scenario are shown in Table 7.4, the preconditions for the calculation of emissions are shown in Table 7.5, and preconditions for the beneficial utilization of dredged soil are shown in Table 7.6. The calculation results are shown in Fig. 7.3. The results can be interpreted as follows. Carbon neutrality in ports is realized only by 2050 in Scenario 2, where emission source countermeasures are significantly advanced, suggesting that conversion to next-generation energy is indispensable. Because more dredged soil is generated in Scenario 2, the effective use of dredged soil could have a great effect as a sink countermeasure. If CO2 emissions can be controlled by improving the construction process, a certain net absorption effect can be obtained (CO2 emissions from dredging projects are less than the effects of the beneficial utilization of dredged soil in 2030 and 2050 in both scenarios). The beneficial utilization of dredged soil as a BCE foundation material plays a role in achieving carbon neutrality in ports by lowering net emissions through the dual effects of carbon containment in dredged soil and increased storage owing to BCE creation. Even under Scenario 1, the quantitative effect would be such that carbon neutrality would nearly be achieved by 2050. The containment effect itself is limited to the point of beneficial use (in fact, the storage effect in 2050, when the intensive dredging project ends, is smaller than that in 2030 in Scenario 2); however, because CO2 storage by BCE continues thereafter (in fact, the absorption is larger in 2050 than in 2030 in both scenarios), the earlier it is created, the more effectively it can reduce net emissions in the following years.

7.4 Future Significance of Containment of Organic Carbon in the Dredged Soil and the Creation of BCE

91

Table 7.4 Details of the two scenarios for carbon neutrality in ports Scenario 1

Scenario 2

Overview

Existing measures, such as electrification, strengthening of BCE creation based on current dredging volume

In addition to the measures under Scenario 1, promotes dredging for the development of hydrogen import bases, further strengthening BCE creation

Cargo-handling machinery

Electrification, the use of electricity derived from renewable energy sources

In addition to the measures under Scenario 1, the use of fuel cells in cargo-handling machinery

Reefer containers

Power saving through installing retractable roofs, etc., and using electricity derived from renewable energy sources

In addition to the measures under Scenario 1, installing a stand-alone hydrogen power supply

Administrative buildings

Use of electricity derived from renewable energy sources

Ships (at arrival and departure)

IMO targets: − 40% Fuel use by 2030, − 50% CO2 emissions by 2050, low-speed navigation

Ships (at berth)

Introduction of LNG ships and onshore power supply

Trucks (congestion at the gate)

Eliminated through the use of digital logistics

Eliminated through the use of digital logistics

Trucks (in-port drayage)

Modal shift to ships, vessel speed reduction, shift to EVs

In addition to the measures under Scenario 1, the introduction of FC vehicles

Trucks (hinterland transportation)

Modal shift to railroads, shift to EVs

Dredging projects

Maintaining current dredging volume

Strengthening of dredging channel projects for the development of hydrogen import bases over the next 10 years

Creation of BCE

Utilization of a certain amount of dredged soil for BCE foundation materials

Strengthening the use of dredged soil for BCE foundation materials

In addition to the measures under Scenario 1, the introduction of zero-emission ships

In Scenario 2, carbon neutrality is achieved by 2050; however, the CO2 emissions are too low to rely on the beneficial use of dredged soil and BCE. In this regard, it could be argued that emission source countermeasures are ultimately more important; however, their effectiveness as sinks for residual emissions should not be neglected. For example, the CO2 emissions in this study do not include emissions from industries in the port area, but heavy and large industries are located in ports, and residual emissions remain. To contribute to achieving carbon neutrality in Japan, the new countermeasures proposed in this chapter are significant as sinks for residual emissions. Based on the above discussion, the policy scenario of intensive implementation of dredging projects for the development of next-generation energy import bases and beneficial utilization of dredged soil generated as BCE foundation materials is fully accepted as a climate change countermeasure while achieving CNP and after realizing CNP. To achieve carbon neutrality in ports, next-generation energy sources are essential; thus, it is critical to devise an import base for them. The dredged soil generated during port development for the formation of the import base can be beneficially utilized as a BCE foundation material, further increasing the CO2 storage.

7 Carbon Containment and Creation of BCEs …

92 Table 7.5 Preconditions for emissions calculations Items

2030 (%)

2050 (%)

Source

Ratio of decarbonized power sources

59

88

Agency for Natural Resources and Energy (2021) for 2030, IEA (2021) for 2050

Electrification rate of cargo-handling machinery (Scenario 1)

50

100

Author’s setting

CO2 reduction rate by electrification of cargo-handling machinery

70

70

Ministry of the Environment (2015), assuming fossil fuel power generation

Ratio of fuel cell use in cargohandling machinery (Scenario 2)

30

99

Same as the sales volume share (heavy trucks) of Plugin hybrid electric vehicles (PHEVs), Battery electric vehicles (BEVs), and Fuel cell electric vehicles (FCEVs) in IEA (2021)

Power savings rate by installing a retractable roof

10

10

Logistics weekly (2018)

Introduction rate of stand-alone hydrogen and other power sources (Scenario 2)

30

99

Same as the sales volume share (heavy trucks) of PHEVs, BEVs, and FCEVs in IMO (2018)

Introduction rate of onshore power supply

50

100

Authors’ setting

CO2 reduction rate by the introduction of onshore power supply

40

40

MLIT (2009), assuming fossil fuel power generation

Ratio of modal shift in port-related vehicles

50

*1

Author’s setting

CO2 reduction rate from ships (Scenario 1)

40

50

IMO (2018)

Introduction rate of zero-emission ships (Scenario 2)

17.1

92.5

The International Shipping Zero Emissions Project (2022), IEA (2021)

CO2 reduction rate by introducing LNG ships (Scenario 1)

25

25

Sifakis and Tsoutsos (2021)

Electrification rate of port-related vehicles

8

59

Share of electric vehicles in stock (heavy truck) in IEA (2021)

Introduction rate of FC vehicles

30

99

Same as the sales volume share (heavy trucks) of PHEVs, BEVs, and FCEVs in IEA (2021)

*

1: Other than electrification and fuel cell use

The amount of dredging in port projects has been decreasing drastically in Japan owing to budget constraints and environmental impacts, and the promotion of dredging projects has become difficult under conventional values. Therefore, it is necessary to promote navigation dredging projects as effective climate change countermeasures, which will increase awareness about its necessity for realizing a decarbonized society, the net carbon storage effect of its beneficial utilization, and its potential for the creation of seagrass meadows, macroalgal beds, and tidal flats for BCEs. Thus, it is necessary to change the current perspective to best utilize port development projects to promote nextgeneration energy sources and climate change countermeasures for sinks. For countries such as Japan, where the enforcement approach to environmental policies has historically been problematic and enacting measures from a business perspective is relatively difficult due to port governance, as discussed in Chap. 2, the new countermeasures proposed in this chapter

7.5 Conclusion

93

Table 7.6 Preconditions for the beneficial utilization of dredged soil in each scenario Parameters

Value

Sources

Scenario 1

Scenario 2

Annual amount of dredged soil

159,500,000 m3 (per year)

259,500,000 m3 (next 10 years) 159,500,000 m3 (After 11th year)

Scenario 1: current status (Kuwae et al. 2019) Scenario 2: Additional dredging project of 10 million m3yr−1 to develop a hydrogen import base

Utilization rate for backfilling borrow pits (%)

5.9

5.9

Ministry of Finance (2018)

Utilization rate for seagrass meadows, macroalgal beds, and tidal flats (%)

25

50

Scenario 1: Kuwae et al. (2019) Scenario 2: Doubled to promote BCE creation

Utilization rate for land development (%)

50

16.1

Beneficial utilization ratio (Naito et al. 2008)—(backfilling of borrow pits + utilization for seagrass meadows, macroalgal beds, and tidal flats)

Residual rate of backfilling borrow pits (%)

95

95

Assuming innovations in construction (compared with Table 7.2)

Residual rate of seagrass meadows, macroalgal beds, and tidal flats (%)

90

90

Residual rate of land development (%)

50

50

can be effective as they require government initiatives. In Japan, large-scale port development projects are conducted directly by the national government. Accordingly, depending on the government’s determination, the feasibility of new countermeasures will be extremely high.

7.5

Conclusion

This chapter focused on the containment of organic carbon in dredged soil and carbon storage by BCE creation through the effective use of dredged soil as a foundation material for BCE. This chapter revealed the carbon storage effect of organic carbon contained in dredged soil in addition to the fact that dredged soil contributes to CO2 storage through its beneficial utilization as a BCE foundation material in port projects. However, the quantitative effects of these sinks are not sufficient to realize carbon neutrality when including emissions from ships and vehicles, which account for most CO2 released in ports. This suggests the importance of drastic measures, such as conversion to nextgeneration energy sources, for the comprehensive realization of CNPs. To this end, it is necessary to establish an import base for next-generation energy, and the effective storage of carbon contained in the dredged soil, which is expected to be generated in large quantities and utilized as a BCE foundation material, appears to have significant implications. For societies and countries aiming for carbon neutrality, sink-based measures, such as conservation and creation of BCEs, and the beneficial utilization of dredged soil are important components for offsetting residual emissions. Additional scientific support is needed to quantify the carbon storage effect in dredged soils. In addition, the carbon storage effects of BCEs can be further scrutinized. Furthermore, the combination of measures, strength of regulations, and financing methods must be addressed to realize carbon

7 Carbon Containment and Creation of BCEs … Storage by BCE

15000

15000

Carbon storage in ports ReducƟon of carbon emissions by beneficial uƟlizaƟon of dredged soil Emissions from trucks

12000

Emissions from ships Emissions from port acƟviƟes 9000

Emissions from dredging projects Emissions of carbon stored in ports

6000

Amount of emissions, storage (1,000 tCO2)

Amount of emissions, storage, and reducƟon of emissions (1,000 tCO2)

94

Storage by BCE Carbon storage by beneficial uƟlizaƟon of dredged soil

12000

Emissions from trucks Emissions from ships Emissions from port acƟviƟes

9000

Emissions from dredging projects Carbon neutrality within port acƟviƟes (including dredging projects)

6000 Carbon neutrality within port acƟviƟes (including dredging projects)

3000

3000

Total carbon neutrality 0

-3000

Total carbon neutrality 0

Current status

2030 Scenario1

2030 Scenario2

2050 Scenario1

2050 Scenario2

(a) First interpretation

-3000

Current status

2030 2030 2050 2050 Scenario1 Scenario2 Scenario1 Scenario2

(b) Second interpretation

Fig. 7.3 Calculation results of both interpretations

neutrality in ports. More specific roadmaps for individual ports, including the countermeasures provided in this chapter, can then be provided based on the specific characteristics of each country and port.

References Acciaro M, Ghiara H, Cusano MI (2014) Energy management in seaports: a new role for port authorities. Energy Policy 71:4–12 Agency for Natural Resources and Energy (2021) Basic energy plan [Draft]. Retrieved from https://www.enecho.meti. go.jp/committee/council/basic_policy_subcommittee/2021/046/046_005.pdf Bosman R, Loorbach D, Rotmans J, Van Raak R (2018) Carbon lock-out: leading the fossil port of Rotterdam into transition. Sustainability 10(7):2558 Hayashi T (2011) Estimation of carbon dioxide emissions from port development projects. National Land Technology Study Group of the Ministry of Land, Infrastructure, vol 2011. Transport and Tourism. Retrieved from https://www. mlit.go.jp/chosahokoku/h23giken/program/kadai/pdf/ippan/ippan3-02.pdf I2S2 (2013) Environmental initiatives at seaports worldwide: a snapshot of best practices. International Institute for Sustainable Seaports (I2S2). VA IEA (2021) Net zero by 2050—a road map for the global energy sector. Retrieved from https://www.iea.org/reports/netzero-by-2050 IMO (2018) Initial IMO strategy on reduction of GHG emissions from ships. Resolution MEPC.304(72)

References

95

Kubo K (2015) Carbon cycling in Tokyo Bay. Doctoral dissertation, Tokyo University of Marine Science & Technology Kusuyama K, Suto A, Nakajima S, Kawai T (2017) Improvement of marine environment by backfilling dredged depression in Hakata Bay. J Jpn Soc Civ Eng Ser B2 (Coast Eng) 73:I_1357–I_1362 Kuwae T, Yoshida G, Hori M, Watanabe K, Tanaya T, Okada T, Umezawa Y, Sasaki J (2019) Nationwide estimate of the annual uptake of atmospheric carbon dioxide by shallow coastal ecosystems in Japan. J Jpn Soc Civ Eng Ser B2 75(1):10–20 Lam JSL, Notteboom T (2014) The greening of ports: a comparison of port management tools used by leading ports in Asia and Europe. Transp Rev 34(2):169–189 Lam JSL, Li KX (2019) Green port marketing for sustainable growth and development. Transp Policy 84:73–81 Logistics weekly (2018) Hakata port terminal is becoming Japan’s most eco-friendly terminal, attracting visitors from around the world. Retrieved from https://weekly-net.co.jp/news/42166/ Lovelock CE, Duarte CM (2019) Dimensions of blue carbon and emerging perspectives. Biol Lett 15:20180781 Macreadie PI, Anton A, Raven JA, Beaumont N, Connolly RM, Friess DA, Kelleway JJ, Kennedy H, Kuwae T, Lavery PS, Lovelock CE, Smale DA, Apostolaki ET, Atwood TB, Baldock J, Bianchi TS, Chmura GL, Eyre BD, Fourqurean JW, Hall-Spencer JM, Huxham M, Hendriks IE, Krause-Jensen D, Laffoley D, Luisetti T, Marbà N, Masque P, McGlathery KJ, Megonigal JP, Murdi-yarso D, Russell BD, Santos R, Serrano O, Silliman BR, Watanabe K, Duarte CM (2019) The future of blue carbon science. Nat Commun 2019(10):3998 Ministry of Finance, Budget execution survey (FY 2018) (2018) Retrieved from https://www.mof.go.jp/policy/budget/ topics/budget_execution_audit/fy2018/sy3007/38.pdf Ministry of Land, Infrastructure, Transport and Tourism (2009) Guidelines for developing a greenhousegas emission reduction plan at ports (draft). Retrieved from https://www.mlit.go.jp/common/000043030.pdf Ministry of Land, Infrastructure, Transport and Tourism (2020) Efforts to improve the marine environment in ports. Retrieved from http://www.env.go.jp/press/108218/114447.pdf Ministry of the Environment (2015) Promotion project of CO2 reduction measures in the logistics field. Retrieved from https://www.env.go.jp/earth/ondanka/biz_local/28_a20/20gaiyo.pdf Naito R, Nakamura Y, Urase T (2008) Relationship between sediment quality and benthic macrofaunain Japanese Ports and Harbors. Technical Note of the Port and Airport Research Institute, vol 1174 Nakagawa Y (1998) A study on sedimentation of mud in a bay. REPORT of the Port and Harbor Research Institute, vol 37 Nellemann C, Corcoran E, Durate CM, Valdes L, De Young C, Fonseca L, Grimsditch G (2009) Blue carbon—a rapid response assessment, United Nations Environmental Programme. GRID-Arendal. Brikeland Trykkeri AS, Birkeland Okada T, Iseri E, Akiyama Y (2016) Sedimentation rate in coastal regions in Tokyo Bay. Technical Note of National Institute for Land and Infrastructure Management, vol 888. PIANC (2019) Carbon management for port and navigation infrastructure—EnviCom WG Report n° 188–2019. The World Association for Waterborne Transport Infrastructure PIANC Maritime Navigation Commission. Brussels, Belguim Port of Corpus Christi (2022) Port of Corpus Christi enters into agreement with TalosEnergy and Howard Energy partners for carbon capture and sequestration opportunities. Retrieved from https://natgeo.nikkeibp.co.jp/atcl/news/ 21/031900137/ Raymond CM, Berry P, Breil M, Nita MR, Kabisch N, de Bel M, Enzi V, Frantzeskaki N, Geneletti D, Cardinaletti M, Lovinger L, Basnou C, Monteiro A, Robrecht H, Sgrigna G, Munari L, Calfapietra C (2017) An impact evaluation framework to support planning and evaluation of nature-based solutions projects. Report prepared by the EKLIPSE expert working group on nature-based solutions to promote climate resilience in urban areas. Centre for Ecology & Hydrology, Wallingford, United Kingdom. Retrieved from http://www.eklipse-mechanism.eu/apps/Eklipse_data/ website/EKLIPSE_Report1-NBS_FINAL_Complete-08022017_LowRes_4Web.pdf Sifakis N, Tsoutsos T (2021) Planning zero-emissions ports through the nearly zero energy port concept. J Cleaner Prod 286:125448 Sornn-Friese H, Poulsen RT, Nowinska AU, de Langen P (2021) What drives ports around the world to adopt air emissions abatement measures? Transp Res Part D Transp Environ 90:102644 Taljaard S, Slinger JH, Arabi S, Weerts SP, Vreugdenhil H (2021) The natural environment in port development: a “green handbrake” or an equal partner? Ocean Coast Manag 199:105390 Tank News International (2020) Copenhagen Port participates in carbon capture and storage project. Retrieved from https:// portofcc.com/port-of-corpus-christi-enters-into-agreement-with-talos-energy-and-howard-energy-partners-for-carbon-captureand-sequestration-opportunities/ The International Shipping Zero Emissions Project (2022) Toward achieving carbon neutrality in international shipping by 2050. Retrieved from https://www.mlit.go.jp/maritime/content/001484435.pdf Wright C (2022) An overview of Europe’s carbon capture and storage developments. Retrieved from https://www. azocleantech.com/article.aspx?ArticleID=1452

8

Role of Ports in the Trade-Off Problem Between Circular Economy and Climate Change Action: Potential for Increased Use of Secondary Raw Materials in the Copper Industry as a Climate Change Countermeasure 8.1

Introduction

The relationship between climate change countermeasures and the logistics function of ports has received little attention. The use of next-generation energy and recyclable resources as climate change countermeasures (Poulsen et al. 2018) may improve feasibility by reducing costs through improved logistics efficiency and the role of ports as transportation hubs may be significant. For example, the spread of low-carbon technologies (electric vehicles and low-carbon power generation technologies) could significantly increase the demand for copper, which would likely increase the dependence on recyclable resources as raw materials. While transportation may have a significant impact on the cost and CO2 emissions of procuring recyclable resources with complex reverse logistics (Reddy et al. 2020), there are insufficient academic findings on the impact of the increased introduction of recyclable resources, including logistics, climate change action, and the realization of a CE, which may be conflicting goals in some cases. If ports could help resolve this conflict, then increased usage of recyclable resources, particularly for next-generation energy can be regarded as climate change countermeasures in ports. Hence, this chapter focuses on the trade-off problem between CE and climate change countermeasures. Although countries such as Japan have implemented stronger climate action plans, targeting carbon neutrality is a major challenge, especially for industries with large CO2 emissions (Zhao et al. 2021). Reducing industry emissions will entail coordinated action throughout value chains to promote all mitigation options, including energy and material efficiency, circular material flows, abatement technologies, and transformational changes in production processes (IPCC 2022). In particular, the development of a CE, which is directly related to Sustainable Development Goal 12 (responsible consumption and production), is crucial. However, the relationship between CE and climate change measures can be complex and dependent on industrial characteristics. For example, in the steel industry, the use of steel scrap as a raw material is an important measure for climate change mitigation (Suer et al. 2021). However, increasing the use of secondary raw materials (SRMs) in the copper industry can lead to increased CO2 emissions during smelting (Zhang et al. 2021). Furthermore, the collection of SRMs is complicated by reverse logistics (RL), posing a logistical challenge for reducing CO2 emissions (Reddy et al. 2020). Thus, the copper industry faces serious challenges in achieving its goals of climate change mitigation and CE development. However, decarbonization technologies, including renewable energy and electric vehicles, require a variety of metals in vast quantities for their function; in particular, copper is an essential component of decarbonization technologies (Watari et al. 2020). Therefore, the growth of low-carbon technologies © The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0_8

97

98

8

Role of Ports in the Trade-Off Problem …

could significantly increase the demand for copper by 2050 (Seck et al. 2020). As significant efforts to improve energy efficiency in the copper industry, demand growth, and declining ore grades have completely offset the improvement of energy efficiency, leading to continued increases in GHG emissions from copper production, GHG emissions associated with copper production could be important in the context of climate change (Watari et al. 2022). Despite indications that enhanced recycling is necessary based on the availability of primary copper resources, if CO2 emissions increase because of increased SRM use, contrasting effects will be observed in the overall economy as it moves toward decarbonization. This chapter examines the role of ports in the trade-off between CE and climate change countermeasures. Specifically, the potential to expand the use of SRMs as a countermeasure to climate change was examined by calculating the changes in costs and CO2 emissions resulting from the increased use of local and imported SRMs in the Japanese copper industry. In Japan, although various recycling laws are in place, domestic recycling is not guaranteed, and copper scrap is currently over-exported (Sugimura and Murakami 2016). Even if all SRMs generated locally were recycled in Japan, they would be insufficient as raw materials; thus, imports from overseas would be necessary for the increased use of SRMs. Furthermore, SRM imports are associated with RL, which exhibits unique characteristics, such as few-to-many flow and uncertainties in the quality, quantity, and timing of generation (Ilgin and Gupta 2010), because it targets recyclable resources. The scenarios considered would be more complex, especially because they deal with international RL related to maritime transportation considering Japan’s location (Sugimura and Murakami 2021). If CE and climate change countermeasures are compatible through increased use of the SRM, ports would play a very important role as import hubs, and securing port functions for this purpose could be positioned as a climate change countermeasure.

8.2

Methods

8.2.1 Scope of Analysis Currently, the Japanese copper industry imports copper concentrates from overseas and produces copper in domestic copper smelters, although some smelters predominantly employ SRMs. As of 2019, 22% SRMs based on the copper content were used. In this work, increased use of copper scrap and waste from electrical and electronic equipment (WEEE) substrates as SRMs was assumed. For copper scrap, scraps traded under HS Code 7404, which have been sorted with a high copper grade, were the focus. To increase the use of SRMs, copper scrap and WEEE must be collected from overseas. As shown in Fig. 8.1, the production of copper from copper concentrates generates CO2 and incurs costs during the mining, beneficiation, transportation, and smelting stages. However, when SRMs are utilized, CO2 emissions and costs are incurred during the recovery and transportation stages of SRMs instead of reducing CO2 emissions and costs during the mining and beneficiation stages. The current material flow of copper is described in reports by The Japan Oil, Gas, and Metals National Corporation (JOGMEC), a Japanese governmental independent administrative agency. In this chapter, data from JOGMEC (2020) are used to illustrate the assumed material flow after the increased use of SRMs (Fig. 8.2). Domestic copper production was assumed to remain unchanged, whereas the use of copper scrap converted from overseas exports for domestic use, additional copper scrap imported from overseas, WEEE substrate recovered domestically, and new WEEE substrate imported from overseas were assumed to increase, thereby reducing copper concentrate imports by an amount equivalent to the net copper content obtained from these sources.

8.2 Methods

99

Fig. 8.1 Schematic of costs incurred and CO2 emissions during the processing and use of copper ore and SRMs

The system boundary stretches from raw material procurement to copper smelting; that is, the process begins with mining when copper concentrate is used as the raw material, and with collection when SRM is used. When copper concentrates are used as raw materials, they are mined and beneficiated overseas, transported to ports, loaded and unloaded, and imported to Japan by sea. Domestic smelters that use copper concentrate as raw material are located in coastal areas, leading to negligible domestic transportation. The procurement of SRMs is complicated because of RL. In this chapter, the collection of copper scrap and WEEE in Japan and overseas was assumed to follow the hierarchy shown (Fig. 8.3), as reported by Sugimura et al. (2014). The RL of WEEE is primarily composed of quaternary RL. WEEE generated in Japan is transported to domestic smelters, whereas WEEE generated overseas is transported to ports for export. Primary RL refers to the collection of WEEE data at the municipal level. In secondary RL, WEEE is widely collected and stored in stockyards near processing facilities to secure lots for processing before transportation to processing facilities via tertiary RL. WEEE is used to remove the substrate in processing facilities, and only the substrate is transported via quaternary RL. Copper scrap is assumed to be generated in factories and, therefore, is associated with primary through tertiary RL. Copper scrap and WEEE substrates imported from overseas require additional RL for transportation from ports to smelters in Japan. RL costs include fuel and labor costs, of which fuel costs vary with the country’s area, whereas labor costs vary with economic maturity (Sugimura and Murakami 2016).

100

8

Role of Ports in the Trade-Off Problem …

Fig. 8.2 Changes in material flow resulting from increased use of SRMs Japan

Primary RL

Secondary RL

Tertiary RL

Quaternary RL

WEEE Household

Storage

Storage

Processing Facility

Secondary RL

Primary RL

Metal Smelter

Copper scrap

Tertiary RL Factory, etc.

Import partner

Storage

Storage

Primary RL

Tertiary RL

Secondary RL

Quaternary RL

WEEE

Domestic RL

Household

Storage

Storage

Processing Facility

Secondary RL

Primary RL

Metal Smelter Port

Copper scrap Factory, etc.

Storage

Storage

Tertiary RL

Fig. 8.3 Hierarchical representation of RL related to copper scrap and WEEE

8.2 Methods

101

8.2.2 Calculation Methods and Cases The case study involving the increased use of only copper scrap is referred to as Case 1 and that involving the increased use of both copper scrap and WEEE substrates is referred to as Case 2. Furthermore, we assume that in Case 2–1, all CO2 and costs are allocated to copper, and in Case 2–2, these are allocated according to the metal value. The notations of the indices and parameters used in this study are as follows. I J K P Q R REM EM o EM r RTC TC o TC r xkjp

xki xj poj s r k Ejp;p

k Ei;q Ej;r k Cjp;p

k Ci;q Cj;r k efjp;p k efi;q efj;r p1jp p1i poj

uckjp;p

Copper scrap and WEEE imported from country i Copper concentrate imported from country j SRMs (k=1: copper scrap; k=2: WEEE) Processes of SRMs generated in Japan (p=1–3: primary, secondary, and tertiary RL; 4: processing; 5: quaternary RL; 6: smelting) Processes of SRMs imported from overseas (q=1–3: primary, secondary, and tertiary RL; 4: processing; 5: quaternary RL; 6: port handling in country I; 7: maritime RL; 8: port handling in Japan; 9: RL in Japan; 10: smelting) Processes of importing copper concentrate from overseas (r=1: mining and beneficiation; 2: transportation; 3: smelting) CO2 emission reduction by shifting to SRMs (t-CO2). Current CO2 emissions (t-CO2). CO2 emissions by the increased use of SRMs (t-CO2). Cost increase because of increased use of SRMs (yen). Current cost (yen). Cost of increased use of SRMs (yen). Amount of SRM k generated in Japan (t). Amount of SRM k imported from country i (t). Amount of copper concentrate imported from country j (t). Purchase price of copper concentrate imported from country j (yen/t). Weight percentage of WEEE substrates (%). Percentage of the value of copper among metals recovered from substrates (%). CO2 emissions in process p of SRM k generated in Japan (t-CO2). CO2 emissions in process q of SRM k imported from country i (t-CO2). CO2 emissions in process r of copper concentrate imported from country j (t-CO2). Cost in process p of SRM k generated in Japan (yen). Cost in process q of SRM k imported from country i (yen). Cost of copper concentrate imported from country j in process r (yen). CO2 emission intensity in process p of SRM k generated in Japan (t-CO2/t). CO2 emission intensity in process q of SRM k imported from country i (t-CO2/t). CO2 emission intensity in process r of copper concentrate imported from country j (t-CO2/t). Purchase price of copper scrap generated in Japan (yen/t). Purchase price of copper scrap imported from country i (yen/t). Purchase price of copper concentrate imported from country j (yen/t). Cost intensity in process p of SRM k generated in Japan (yen/t).

102

ucki;q ucj;r

8

Role of Ports in the Trade-Off Problem …

Cost intensity of process q of SRM k imported from country i (yen/t). Cost intensity in process r of copper concentrate imported from country j (yen/t).

The CO2 emissions and costs generated at each stage when SRMs and copper concentrate were used as raw materials were calculated as follows: REM ¼ EM o  EM r XX EM o ¼ Ej;r

ð8:1Þ ð8:2Þ

j2J r2R

Ej;r ¼ xj  efj;r EM r ¼

X X p2P

k2K

XX

k Ejp;p þ

i2I q2Q

ð8:3Þ

! k Ei;q

ð8:4Þ

1 1 Ejp;p ¼ x1jp  efjp;p

ð8:5Þ

1 1 Ei;q ¼ x1i  efi;p

ð8:6Þ

In Case 2-1: X p2P

2 Ejp;p ¼

X q2Q

4 X

2 Ei;q ¼

p¼1

6 X

2 x2jp  efjp;p þ

4 X

2 x2i  efi;q þ

q¼1

2 s  x2jp  efjp;p

ð8:7Þ

2 s  x2i  efi;q

ð8:8Þ

p¼5 10 X q¼5

In Case 2-2: X p2P

2 Ejp;p

X q2Q

¼r

2 Ei;q

4 X p¼1

¼r

x2jp

4 X q¼1

x2i

2 efjp;p





2 efi;q

þ

6 X

! s

x2jp



s

x2i

2 efi;q

p¼5

þ

10 X q¼5

2 efjp;p

ð8:9Þ

! 

ð8:10Þ

Cost calculation: RTC ¼ TC r  TC o XX TCo ¼ Cj;r j2J r2R

ð8:11Þ ð8:12Þ

8.2 Methods

103

  Cj;r ¼ xj  poj þ ucj;r r

TC ¼

X X p2P

k2K

k Cjp;p

þ

ð8:13Þ

XX i2I q2Q

! k Ci;q

ð8:14Þ

  1 ¼ x1jp  p1jp þ uc1jp;p Cjp;p

ð8:15Þ

  1 Ci;q ¼ x1i  p1i þ uc1i;p

ð8:16Þ

In Case 2-1: X p2P

2 Cjp;p ¼

X q2Q

4 X

2 Ci;q ¼

p¼1

x2jp  uc2jp;p þ

4 X q¼1

x2i  uc2i;q þ

6 X

s  x2jp  uc2jp;p

ð8:17Þ

s  x2i  uc2i;q

ð8:18Þ

p¼5 10 X q¼5

In Case 2-2: X p2P

X q2Q

2 Cjp;p

2 Ci;q

¼r

4 X p¼1

¼r

4 X q¼1

x2jp

x2i



uc2jp;p



uc2i;q

þ

6 X

! s

p¼5

þ

10 X

x2jp



uc2jp;p

x2i

uc2i;q

ð8:19Þ !

s

q¼5



ð8:20Þ

8.2.3 Data Set As the imports of copper concentrated from importers other than the major importers fluctuated annually, imports were assumed to be only from the major importers, as shown in Table 8.1. The grade was assumed to remain constant considering the uncertainty regarding future changes. Table 8.1 Copper concentrate imports for the Japanese copper industry 2019 Actual (tons)

Percentage (%)

Corrected figures (tons)

Percentage (%)

Chile

2,191,277

45.8

2,348,528

49.1

Peru

807,396

16.9

865,337

18.1

Australia

472,112

9.9

505,992

10.6

United States of America

415,821

8.7

445,661

9.3

Canada

350,992

7.3

376,180

7.9

Papua New Guinea

229,739

4.8

246,226

5.1

Others

320,587

6.7

0

0

Total

4,787,924

100

4,787,924

100

104

8

Role of Ports in the Trade-Off Problem …

Table 8.2 Amount of copper scrap and WEEE recovered by the Japanese copper industry assuming increased use of SRMs Copper scrap (tons) Present (2019)

WEEE (tons) 2030

Present (2019)

2030

2050

66,182

166,005

203,991

304,797

13,283

47,470

993,603

2,668,015

9,534,556

39,594

117,617

500,549

1,083,576

3,218,871

16,560

33,351

92,075

107,779

217,060

599,248

5,470

6,526

9,077

27,775

33,135

46,088

2,830

8,374

30,551

139,162

411,862

1,502,553

1,982

4,739

13,798

128,209

306,579

892,635

Singapore

13,989

16,622

24,068

34,925

41,499

60,089

Thailand

27,025

40,930

99,218

192,648

291,777

707,283

Vietnam

3,076

Japan

322,989

Australia

36,046

44,294

India

4,947

Indonesia

18,290

Malaysia New Zealand Pakistan Philippines

2050

7,262

26,534

78,845

186,166

680,221

342,327

424,403

140,000

148,382

183,958

The import partners for copper scrap and WEEE assumed in this study were selected from Asian and Oceanian countries with the largest amounts of copper scrap exports and WEEE generation. China, Hong Kong, Taiwan, and South Korea were excluded, as they are unlikely to export to Japan in the future, considering the stringent metal resource policies and copper industry technologies in each country or region. The WEEE generated in Japan was assumed to be 140,000 tons, according to the target set by ASWEEE (METI, MOE 2021), and 30% of the WEEE generated overseas was assumed to be recovered, half of which can be imported into Japan. All copper scrap currently exported overseas should be retained in Japan and used as raw materials (Central Environment Council 2022). For copper scrap generated overseas, it was assumed that Japan can import 50% of the remaining amount, after excluding the amount imported by Japan from the amount exported by the relevant country. WEEE was assumed to be available for free, both in Japan and overseas. Conversely, copper scrap was assumed to be purchased at market prices both domestically and overseas. As the import volume of SRMs varies depending on the economic growth of the partner country, the economic situation of the importing partner country in 2030 and 2050 was considered in addition to the current situation as of 2019. The estimated amounts of WEEE and scrap used by the Japanese copper industry are listed in Table 8.2. • Japan is assumed to import 50% of HS Code 7404 CS exports from the concerned country, minus the proportion of Japan’s imports. Future additional imports were assumed to grow at the same rate as the gross domestic product (GDP). • Amount of WEEE for foreign countries was calculated as follows: Future per capita generation is assumed to grow at the same rate as the GDP per capita. generated per capita  population  recovery rate ðassumed as 30%Þ Future values for Japan assume the same growth rate as the following equation. GDP per capita  population growth rate

8.2 Methods

105

Table 8.3 Status and percentage of SRMs for each case study (unit: 1,000 t of pure copper equivalent) Present (2019)

Case 1 2019

Case 2 2030

2050

2019

2030

2050

Copper concentrate

1,241

886

805

496

869

767

379

Copper scrap (currently used)

353

353

353

353

353

353

353

Copper scrap (converted from exports)



253

268

332

253

268

332

Copper scrap (imported from overseas)



102

168

412

102

168

412

WEEE (recovered domestically)









1.8

2.0

2.4

WEEE (imported from overseas)









16

36

115

Total

1,594

1,594

1,594

1,594

1,594

1,594

1,594

Percentage of SRMs (%)

22

44

50

69

46

52

76

• Data Sources: Population: United Nations (UN); World Population Prospects: The 2019 Revision; GDP, GDP per capita: UN, National Accounts—Analysis of Main Aggregates (AMA); WEEE incidence: The Global E-waste Monitor 2020 Foreign CS exports; Japanese imports: UN Comtrade. The copper grade of the WEEE substrate is 0.16 t-Cu/t-substrate (0.026 t-Cu/t-WEEE) according to Sugimura et al. (2014). For copper scrap, the grade of HS Code 7404 was estimated to be 87% by the JOGMEC (2020). For a smelting yield of 0.5 and 0.9 for WEEE substrate and copper scrap, respectively (Sugimura et al. 2014), the percentage of SRMs in terms of pure copper equivalent is shown in Table 8.3. The parameter values shown in the Appendix were used for the calculation. The CO2 emissions and cost intensity for each SRM process generated in Japan (primary RL, secondary RL, tertiary RL, processing, quaternary RL, and smelting) were obtained from Sugimura et al. (2014). Among the processes of SRMs imported from overseas (primary RL, secondary RL, tertiary RL, processing, quaternary RL, port handling overseas, maritime, port handling in Japan, RL in Japan, smelting), for RL and processing overseas, CO2 emissions and cost intensities overseas by correcting the values for Japan to account for the land area (which affects RL transport distance), minimum wage (which affects labor costs for RL and processing), and CO2 emission intensity of electricity were obtained. For port handling, CO2 emissions and cost intensities overseas were obtained by correcting Japan’s actual value to account for the minimum wage and CO2 emission intensity of electricity. In Case 1, ocean transport was assumed to occur through existing container routes, whereas Case 2 involved chartering a bulk carrier, assuming that the use of SRMs increased substantially. Three bulk carriers were assumed to be imported into three regions (India, Pakistan, Indonesia, Malaysia, the Philippines, Singapore, Thailand, Vietnam, Australia, and New Zealand), halting at multiple ports along the route with the shortest transport distance. The CO2 emissions and cost intensity for maritime RL were estimated by calculating the fuel consumption with reference to Drewry Maritime Research (2011).

106

8.3

8

Role of Ports in the Trade-Off Problem …

Results

Using the method described in Sect. 2.2, the changes in costs and CO2 emissions resulting from the increased use of SRMs compared to the current situation for Cases 1, 2–1, and 2–2 were calculated. The calculation results are shown in Fig. 8.4 and Table 8.4. The carbon footprints (t-CO2/t-Cu) by stage are shown in Table 8.5. In Case 1, the CO2 emissions were reduced, but the costs increased as the percentage of SRM increased. In Case 2, the results differed depending on how much of the CO2 emissions and costs associated with the WEEE were borne by copper. In Case 2–1 wherein the allocation to other metals was not considered, CO2 emissions were larger than those in Case 1. This trend increased as the proportion of WEEE increased, such that CO2 emissions predicted for 2050 exceeded the current CO2 emissions. In Case 2–2, which considered allocation to other metals, the CO2 reduction was larger and the cost increase was smaller than that in Case 1; this trend increased as the proportion of WEEE

Fig. 8.4 Calculation results of CO2 emissions during increased use of SRMs in the Japanese copper industry

8.3 Results

107

Table 8.4 Calculation results of costs during increased use of SRMs in the Japanese copper industry (unit: billion yen) Present (2019)

Increased use of secondary raw material 2019

2030

2050

2019

2030

2050

2019

2030

2050

Copper concentrate (1,000 tCO2)

4,673

3,337

3,030

1,869

3,271

2,888

1,425

3,271

2,888

1,425

Copper scrap (currently recycled) (1,000 t-CO2)

174

174

174

174

174

174

174

174

174

174

Copper scrap (converted from exports) (1,000 t-CO2)



125

132

164

125

132

164

125

132

164

Copper scrap (imported from overseas) (1,000 t-CO2)



93

146

334

92

140

313

91

140

313

WEEE (recovered domestically) (1,000 t-CO2)



28

30

37

0.8

0.8

1.0

WEEE (imported from overseas) (1,000 t-CO2)



479

1,089

3,493

13

29

93

Total (t-CO2)

4,847

4,169

4,453

5,606

3,675

3,364

2,170

Case 1

3,729

Case 2–1

3,482

2,541

Case 2–2

t-CO2/1,000 t-Cu

3.04

2.34

2.18

1.59

2.62

2.79

3.52

2.31

2.11

1.36

Copper concentrate (billion yen)

1,165

832

755

466

815

720

355

815

720

355

Copper scrap (currently recycled) (billion yen)

338

338

338

338

338

338

338

338

338

338

Copper scrap (converted from exports) (billion yen)



242

256

318

242

256

318

242

256

318

Copper scrap (imported from overseas) (billion yen)



103

169

416

103

169

416

103

169

416

WEEE (recovered domestically) (billion yen)









70

77

103

2

2

3

(Reference; Cost of byproducts)















68

75

100

WEEE (imported from overseas) (billion yen)









461

1,073

3,488

12

29

93

(Reference; Cost of byproducts)















449

1,045

3,395

Total (billion yen)

1,502

1,513

1,518

1,537

2,029

2,634

5,019

1,512

1,514

1,524

(Reference; Total cost of byproducts)















517

1,120

3,495

(Reference; Sum of the above two)

1,502













2,029

2,634

5,019

increased. Although copper scrap is associated with a lower CO2 emission intensity per unit weight of copper than copper concentrate, the addition of land-based RL in countries with large land areas and marine RL implies that copper scrap imported from overseas had a higher CO2 emission intensity than copper scrap procured domestically. In Case 2–1, the WEEE substrate had an extremely high carbon footprint compared with the case where copper concentrates were used as the raw material. This is because land-based RL includes more steps than copper scrap; an intermediate treatment process is included, and the smelting process has a higher CO2 emission intensity than that of copper

108

8

Role of Ports in the Trade-Off Problem …

Table 8.5 Carbon footprint (t-CO2/t-Cu) calculated for each stage in the Japanese copper industry Case 1 Copper concentrate

Copper scrap (converted from exports)

Copper scrap (additional imports from overseas)

WEEE (collected domestically)

WEEE (imported from overseas)

Mining and ore dressing

2.50

Transportation

0.25

Smelting

1.02

Case 2–1

Total

3.77

Onshore RL

0.053

0.053

Smelting

0.44

0.44

Total

0.49

0.49

Onshore RL

0.39

0.39

Maritime RL

0.083

0.065

Smelting

0.44

0.44

Total

0.92

0.90

Case 2–2

Onshore RL



4.95

0.12

Intermediate processing



6.47

0.15

Smelting



4.00

0.095

Total



15.42

0.36

Onshore RL



16.99

0.45

Maritime RL



0.66

0.018

Intermediate processing



9.10

0.24

Smelting



4.00

0.11

Total



30.75

0.82

scrap. Furthermore, imported WEEE substrates had a higher CO2 emission intensity than those procured domestically. In Case 2–2, the carbon footprint was lower than when copper concentrates were used as the raw material because more CO2 was allocated to the precious metals. Thus, it is possible to reduce CO2 emissions by increasing the use of SRMs. However, as indicated by the difference between the results of Case 2–1 and Case 2–2, the increased use of WEEE substrates may conversely increase CO2 emissions. Moreover, CO2 emissions, particularly in RL, are major influencing factors.

8.4

Discussion

8.4.1 Potential Utilization of Carbon Credits Companies can achieve CO2 emissions reductions either by introducing emission reduction technologies themselves or by purchasing carbon credits (Sapkota and White 2020). If the increased use of SRMs is significantly more expensive than the use of primary copper resources, there is a likelihood that purchasing carbon credits will become economically advantageous, and the shift to increased use of SRMs will not occur. Thus, it is important to compare the cost of purchasing credit with that of reducing CO2 emissions. Japan utilizes carbon credits such as J-Credit, which can be used in legal reporting systems and international initiatives. Considering the government’s goal of promoting the use of carbon credits (METI 2021) and increasing the overall participation of the Japanese

8.4 Discussion

109

business community in international initiatives (MOE 2021), it is assumed that such credits will be implemented according to the relationship between the amount of CO2 reduction and the increased costs required to achieve them. Here, we examine whether it is more economically advantageous to increase the use of SRMs or to utilize carbon credits for Cases 1 and 2–2, wherein costs were predicted to increase. Table 8.6 summarizes the current situation in both cases, the amount of CO2 reduction and cost increase when accepting more SRMs, and the increase in the cost per unit CO2 weight. Currently, the J-Credit (energy-saving type) is approximately 1,500 yen/t CO2 (J-Credit System Office 2022). Therefore, assuming current carbon credit prices, purchasing carbon credits is more economically advantageous in Cases 1 and 2–2. Table 8.6 CO2 reduction and cost increase after accepting more SRMs (cost per unit weight of CO2) and results of sensitivity analysis for variable copper prices in Case 2–2 Case 1 2019 A: CO2 reduction (1,000 t)

1,118

Case 2–2 (Cu allocation ratio 2.67%) 2030 1,365

2050 2,306

2019

2030

2050

1,172

1,483

2,676

B: Cost increase (millions of yen)

11,424

15,667

34,678

9,269

11,011

20,326

B/A: (JPY/t-CO2)

10,216

11,477

15,036

7,906

7,425

7,594

Table 8.7 Change in CO2 emission intensity Process

2030 (%)

2050 (%)

Mining and ore dressing

38.8

0.2

Concentrate transportation

83

19

Reference Dong et al. (2020): Annual growth rate of energy consumption (0.66%) considering the decline in copper ore grade, ratio of diesel (19%) and electricity (81%) consumption in mining and ore dressing IEA (2021): CO2 emission intensity of heavy trucks (g CO2/km) (available for diesel), CO2 emission intensity (g CO2/kWh) (available for electricity) IEA (2021): Share of low-carbon fuel use in aviation and shipping For land transportation, this was assumed to be equivalent based on the share of electricity and hydrogen in total energy consumption (rail)

Onshore RL

65.1

6.0

IEA (2021): CO2 emission intensity of heavy trucks (g CO2/km)

Processing

30.1

0

IEA (2021): CO2 emission intensity (g CO2/kWh)

Port cargo handling

65.1

6.0

IEA (2021): CO2 emission intensity of heavy trucks (g CO2/km)

Maritime RL

83

19

IEA (2021): Share of low-carbon fuel use in aviation and shipping

Smelting (copper concentrate)

71.9

59.9

Inventory database IDEA*: Electricity share of CO2 emissions in smelting (40.1%) IEA (2021): CO2 emission intensity (g CO2/kWh)

Smelting (copper scrap)

57.9

39.9

Bureau of International Recycling (2016): Carbon footprint of CS in fire refining (0.21 t-CO2/t-Cu) and electrorefining (0.23 t-CO2/t-Cu) Inventory database IDEA*: Electricity share of CO2 emissions in fire refining (19.5%) and electrorefining (97.3%) IEA (2021): CO2 emission intensity (g CO2/kWh)

Smelting (WEEE substrate)

71.9

59.9

Inventory database IDEA*: Electricity share of CO2 emissions in smelting (40.1%) IEA (2021): CO2 emission intensity (g CO2/kWh)

*

LCI database IDEA Version 3.1, Research Laboratory for IDEA, AIST

110

8

Role of Ports in the Trade-Off Problem …

8.4.2 Feasibility of Carbon Neutrality in the Copper Industry Assuming Technological Innovation The calculations in Sect. 8.3 did not consider the changes in decarbonization technologies at each stage. Particularly in the case of imports from overseas, CO2 emissions related to land and sea RL had a considerable impact on the results and were particularly important in the case of WEEE, thus, highlighting the importance of changes in decarbonization technologies in transportation. In the transportation sector, the use of electric or fuel cell vehicles, zero-emission ships, and electrification or fuel cells for port loading/unloading machinery is expanding, and the CO2 emission coefficient of electricity is expected to decrease with changes in the energy mix. Therefore, technological change must be considered. Meanwhile, a decline in the grade of copper in the copper ore will increase the CO2 emissions, which must also be reflected. Table 8.7 shows the assumed changes in CO2 emission intensity for each process in 2030 and 2050, assuming that the status is equal to 100%. The emission results for Case 2–1, which had the highest CO2 emissions, are shown in Fig. 8.5. Without considering technological changes, the carbon footprint increases as the proportion of WEEE substrates increases. However, when technological changes are considered, the carbon footprint decreases, and by 2050, it is expected to be less than 1/3 of the current level. In particular, the footprint related to RL, which has a large impact on CO2 emissions, will decrease significantly, allowing the smelting portion to remain the dominant source of CO2 emissions. In the Japanese copper industry, the decarbonization process is dependent on other efforts, especially the adoption of

Fig. 8.5 Comparison of CO2 emissions assuming changes in decarbonization technology (Case 2–1)

8.5 Conclusion

111

decarbonization technologies in RL at mining sites. While measures undertaken in RL and other areas to decrease emissions can significantly affect CO2 emission reductions, carbon neutrality will ultimately depend on the decarbonization of the copper industry itself, as the smelting stage will be the main source of CO2 emissions if RL and other areas are sufficiently decarbonized. Furthermore, when the Japanese copper industry adopts the climate change countermeasure of increasing SRM use, it should recognize that the key to reducing CO2 emissions is RL and that the key to CN is its technological innovation.

8.4.3 Roles of Ports It is possible to reduce CO2 emissions by increasing the use of SRMs in the copper industry. However, as indicated by the difference between the results of Case 2–1 and Case 2–2, the increased use of WEEE substrates may conversely increase CO2 emissions. In particular, CO2 emissions from RL were the major influencing factors. Thus, although the calculated results suggest great potential to expand the use of SRMs in the copper industry, increased costs may hinder their increased use from an economic standpoint, and the purchase of carbon credits could be an alternative. Additionally, although the scenario assumes the import of SRMs from Asia and Oceania in this chapter, the cost and CO2 emissions would be higher considering long-distance RL if Japan would need to import from more distant countries such as Africa. Considering the above factors, the role that ports should play in realizing the increased use of SRMs is evident. Port facilities must be developed to accommodate imports of SRMs, and if imports are to be made by large vessels that can minimize marine transportation costs, deeper quays would be necessary. In addition, copper ore is generally imported in quays directly connected to copper smelting facilities, and imports at ports far from smelters result in CO2 emissions from domestic RL and cost disadvantages. Therefore, import hubs might differ from export hubs, which have been designated under the Recycle Port Policy (see Table 3.2). The development of such a port as an import hub for recyclable resources could be designated as a new climate change countermeasure.

8.5

Conclusion

The Japanese copper industry faces serious difficulties in achieving the two major goals of climate change mitigation and CE development. Accordingly, to address this issue, the possibility of expanding the use of SRMs, including imports, in the Japanese copper industry as a climate change countermeasure and the role of ports in resolving this issue was examined in this chapter. Assuming that the Japanese copper industry increases its use of copper scrap and WEEE substrates as raw materials, CO2 emissions can be reduced, but costs may increase. Because multiple metals can be recovered from WEEE substrates, the results vary depending on how CO2 emissions and costs are allocated, with allocations to copper contributing to an increase in its carbon footprint. As shown in this chapter, the carbon footprint of all-copper allocation will be significantly lower in the future because of the expected development of decarbonization technologies, including transportation. The results in this chapter suggest great potential to expand the use of SRMs in the copper industry; however, increased costs may hinder their increased use from an economic standpoint, and the purchase of carbon credits could be an alternative. Additionally, although the scenario assumes the import of SRMs from Asia and Oceania in this chapter, importing SRMs might become difficult depending on each country’s metal resource policy, environmental policy, or progress in smelting technology. Therefore, Japan may need to import from more distant countries such as Africa, for

112

8

Role of Ports in the Trade-Off Problem …

which the cost and CO2 emissions would be higher given the long-distance RL. Consequently, it is likely that the shift to import-based SRM use as a climate change countermeasure will not occur. Thus, there is a trade-off between the development of a CE within each country and development of a global recycling network. In contrast to the uneven distribution of natural resources, circulating resources are generated from all locations with consumers and recycling industry is not always present in that location. As it is considered optimal to recycle in countries that possess sufficient smelting technology and minimize RL, including imports, as much as possible, recycling bases in the global recycling network from the perspective of comprehensive optimization of RL, including maritime transportation, can be studied in the future. As ports can play an important role in realizing the increased use of SRMs, port facilities must be developed to accommodate SRM imports. For example, quays may need to be deepened to accomadate by large vessels that can minimize marine transportation costs. The development of such a port as an import hub for recyclable resources could be seen as a new climate change countermeasure in ports, which would solve the trade-off problem between CE and climate change action.

Appendix: Parameters Used for the Calculations See Tables 8.8, 8.9, 8.10, 8.11, 8.12, 8.13, 8.14, 8.15, 8.16, 8.17, 8.18, 8.19.

Table 8.8 CO2 emission intensity of CS collected in Japan (unit: kg-CO2/t-copper scrap) Case 1, Case 2

1 efjp;1

1 efjp;2

1 efjp;3

1 efjp;6

Common for 2019, 2030, 2050

12.50

17.90

11.00

344.52

1 1 1 1 : primary reverse logistics (RL), efjp;2 : secondary RL, efjp;3 : tertiary RL, efjp;6 : smelting efjp;1 1 1 1 , efjp;2 , and efjp;3 : Unit fuel consumption in each RL stage derived from Sugimura et al. (2014), multiplied by the efjp;1 CO2 emission factor of diesel oil (2.48 t-CO2/kL acquired from Ministry of Environment 2020) and the ratio of the bulk density of CS (1.13 t/m3) to that of WEEE (1.00 t/m3). It is assumed that the 2nd, 3rd, and 4th stages of WEEE logistics from Sugimura et al. (2014) correspond to the 1st, 2nd, and 3rd stages of CS logistics, respectively 1 efjp;6 : Carbon footprint (secondary production from scrap) 0.44 t-CO2/Cu-t acquired from the Bureau of International Recycling (2008) multiplied by 87% grade Japanese CS (from JOGMEC 2020) and 90% smelting yield (authors’ own calculations)

Table 8.9 CO2 emission intensity of CS imported from foreign countries (unit: kg-CO2/t-copper scrap) Case 1 (2019, 2030, 2050)

1 efi;1

1 efi;2

1 efi;3

1 efi;6

Australia

254.30

364.33

223.91

0.10

India

1 efi;7

49.06

1 efi;8

1 efi;9

1 efi;10

0.055

11.00

344.52

108.68

155.70

95.69

0.090

98.33

0.055

11.00

344.52

Indonesia

62.97

90.21

55.44

0.091

104.48

0.055

11.00

344.52

Malaysia

10.90

15.62

9.60

0.086

52.56

0.055

11.00

344.52

8.94

12.81

7.87

0.013

105.68

0.055

11.00

344.52

26.32

37.71

23.17

0.069

55.92

0.055

11.00

344.52

9.92

14.21

8.73

0.064

29.47

0.055

11.00

344.52

New Zealand Pakistan Philippines Singapore Thailand

0.024 16.96

0.034 24.30

0.021 14.94

0.061

91.04

0.055

11.00

0.064

43.37

0.055

11.00

344.52 344.52 (continued)

Appendix: Parameters Used for the Calculations

113

Table 8.9 (continued) Vietnam

10.95

15.69

9.64

0.060

Case 2 (2019)

1 efi;1

1 efi;2

1 efi;3

1 efi;6

Australia

254.30

364.33

223.91

0.38

India

48.11 1 efi;7

58.95

0.055

11.00

344.52

1 efi;8

1 efi;9

1 efi;10

0.38

11.00

344.52

108.68

155.70

95.69

0.38

55.76

0.38

11.00

344.52

Indonesia

62.97

90.21

55.44

0.38

44.56

0.38

11.00

344.52

Malaysia

10.90

15.62

9.60

0.38

44.56

0.38

11.00

344.52

8.94

12.81

7.87

0.38

58.95

0.38

11.00

344.52

26.32

37.71

23.17

0.38

55.76

0.38

11.00

344.52

9.92

14.21

8.73

0.38

44.56

0.38

11.00

344.52

New Zealand Pakistan Philippines Singapore

0.38

44.56

0.38

11.00

344.52

Thailand

16.96

0.024

24.30

0.034

14.94

0.021

0.38

44.56

0.38

11.00

344.52

Vietnam

10.95

15.69

9.64

0.38

44.56

0.38

11.00

344.52

1 1 1 1 1 1 efi;1 : primary RL, efi;2 : secondary RL, efi;2 : tertiary RL, efi;6 : port loading, efi;7 : maritime transport, efi;8 : port unloading, 1 1 efi;9 : RL in Japan, efi;10 : smelting

Only the CO2 emission intensity of maritime transport in Case 2 changes as follows (increased transport efficiency because of larger bulk carriers) (unit: kg-CO2/t-copper scrap): Case 2

2019

2030

2050

Australia

58.95

54.39

47.42

India

55.76

36.69

23.59

Indonesia

44.56

34.61

24.66

Malaysia

44.56

34.61

24.66

New Zealand

58.95

54.39

47.42

Pakistan

55.76

36.69

23.59

Philippines

44.56

34.61

24.66

Singapore

44.56

34.61

24.66

Thailand

44.56

34.61

24.66

Vietnam

44.56

34.61

24.66

1 efi;1 ,

1 efi;2 ,

1 efi;3 :

1 efjp;1 ,

1 efjp;2 ,

1 efjp;3

and and are adjusted by the square root of the ratio of the area of the country to that of Japan. 1 1 and efi;8 : In Case 1, the value is obtained by multiplying the electricity consumption per unit weight of 0.125 kWh/t efi;6 (calculated from Japan’s actual value of 2.5 kWh/TEU, assuming 1 TEU = 20 t) by the CO2 emission intensity for each country (Australia: 0.76, India: 0.72, Indonesia: 0.73, Malaysia: 0.69, New Zealand: 0.10, Pakistan: 0.55, Philippines: 0.51, Singapore: 0.49, Thailand: 0.51, Vietnam: 0.48, Japan: 0.44 kg-CO2/kWh), assuming a gantry crane for container transport. In Case 2, the value is obtained by multiplying the fuel consumption per unit weight 0.151 L/t (calculated from Japan’s actual value of 15.13 L/h, 100 t/h) by the emission factor for diesel oil (2.48 kg-CO2/L), assuming a crawler crane for bulk transport. 1 efi;7 : Set by calculating the fuel consumption with reference to Drewry Maritime Research (2011) and multiplying by the emission factor for fuel oil (3.00 t-CO2/kL acquired from the Ministry of Environment 2020), after setting the ship speed according to ship type (for container ships, the value is set based on 2018 results from the MDS Database, while for bulk carriers, the value is set based on interviews with shipping companies). 1 1 efi;10 : Same as efjp;6

114

8

Role of Ports in the Trade-Off Problem …

Table 8.10 CO2 emission intensity of WEEE collected in Japan (unit: kg-CO2/t-WEEE) Case 2

2 efjp;1

2 efjp;2

2 efjp;3

2 efjp;4

2 efjp;5

2 efjp;10

Common for 2019, 2030, 2050

36.61

11.06

15.84

85.07

1.63

52.61

2 2 2 2 2 2 : primary RL, efjp;2 : secondary RL, efjp;3 : tertiary RL, efjp;4 : processing, efjp;5 : quaternary RL, efjp;10 : smelting efjp;1 2 2 2 2 , efjp;2 , efjp;3 , and efjp;5 : Unit fuel consumption in each RL stage derived from Sugimura et al. (2014) multiplied by efjp;1 the CO2 emission factor of diesel oil (2.48 t-CO2/kL acquired from the Ministry of Environment 2020) 2 efjp;4 : Value of electricity consumption for mechanical demolition (1.08 kWh/kg) and separation and sorting (0.0108 kWh/kg) from Heiho et al. (2019) multiplied by the CO2 emission intensity (0.44 kg-CO2/kWh) 2 efjp;10 : Carbon footprint (secondary production from scrap) (4 t-CO2/t-Cu acquired from the Japan Mining Industry Association 2015) multiplied by the amount of copper in WEEE (0.013 t-Cu/t-WEEE from Sugimura et al. 2014)

Table 8.11 CO2 emission intensity of WEEE imported from foreign countries (unit: kg-CO2/t-WEEE) 2 efi;1

2 efi;2

Australia

36.61

225.05

322.41

138.80

33.18

0.063

9.87

0.063

1.63

52.61

India

36.61

96.18

137.79

129.34

14.18

0.063

9.34

0.063

1.63

52.61

Indonesia

36.61

55.72

79.83

131.08

8.21

0.063

7.46

0.063

1.63

52.61

Malaysia

36.61

9.65

13.83

122.43

1.42

0.063

7.46

0.063

1.63

52.61

Case 2 (2019)

2 efi;3

2 efi;4

2 efi;5

2 efi;6

2 efi;7

2 efi;8

2 efi;9

2 efi;10

New Zealand

36.61

7.91

11.34

18.52

1.17

0.063

9.87

0.063

1.63

52.61

Pakistan

36.61

23.29

33.37

99.81

3.43

0.063

9.34

0.063

1.63

52.61

Philippines

36.61

8.78

12.57

93.43

1.29

0.063

7.46

0.063

1.63

52.61

Singapore

36.61

0.021

74.52

0.0031

0.063

7.46

0.063

1.63

52.61

0.030

Thailand

36.61

15.01

21.51

80.73

2.21

0.063

7.46

0.063

1.63

52.61

Vietnam

36.61

9.69

13.88

109.88

1.43

0.063

7.46

0.063

1.63

52.61

2 2 2 2 2 2 2 efi;1 : primary RL, efi;2 : secondary RL, efi;2 : tertiary RL, efjp;4 : processing, efjp;5 : quaternary RL, efi;6 : port loading, efi;7 : 2 2 2 maritime transport, efi;8 : port unloading, efi;9 : RL in Japan, efi;10 : smelting

Only the CO2 emission intensity of maritime transport changes as follows (increased transport efficiency due to larger bulk carriers) (Unit: kg-CO2/t-WEEE): Case 2

2019

2030

2050

Australia

9.87

9.11

7.94

India

9.34

6.14

3.95

Indonesia

7.46

5.79

4.13

Malaysia

7.46

5.79

4.13

New Zealand

9.87

9.11

7.94

Pakistan

9.34

6.14

3.95

Philippines

7.46

5.79

4.13

Singapore

7.46

5.79

4.13

Thailand

7.46

5.79

4.13

Vietnam

7.46

5.79

4.13

2 2 efi;1 : Same as efjp;1 . 2 2 2 2 2 2 efi;2 , efi;3 , and efi;5 : efjp;2 , efjp;3 , and efjp;5 adjusted by the square root of the ratio of the area of the country to that of Japan. 2 2 efi;4 : Same as efjp;4 , but using the CO2 emission intensities of the country (see Table 8.9). 2 2 1 1 efi;6 and efi;8 : Same as efi;6 and efi;8 in Case 2. 2 1 : efi;7 multiplied by the substrate weight/WEEE weight (0.17 acquired from Sugimura et al. 2014) efi;7 2 2 efi;9 : Assumed to be the same as efjp;4 2 2 efi;10 : Same as efjp;10

Appendix: Parameters Used for the Calculations

115

Table 8.12 CO2 emission intensity of copper ore imported to Japan (unit: kg-CO2/t-copper ore) Common for 2019, 2030, 2050

efj;1

efj;2

efj;3

648.98

64.80

263.21

efj;1 , efj;2 , efj;3 : Futami (2015) shows the carbon footprint of the Japanese copper industry at that time (16% recycling material ratio) as 2.5 t-CO2/t-Cu for mining and mineral processing, 0.25 for transportation, and 0.92 for smelting; these values are used as are except for smelting. For smelting, considering the copper scrap smelting value of 0.44 t-CO2/t-Cu (Bureau of International Recycling 2008), the carbon footprint was assumed to be 1.02 CO2/t-Cu for copper ore

Table 8.13 Unit cost of copper scrap collected in Japan (unit: yen/kg-copper scrap) Case 1, Case 2

uc1jp;1

uc1jp;2

uc1jp;3

uc1jp;6

2019

11.46

7.41

5.65

175.73

2030

12.66

8.14

6.22

175.73

2050

17.70

11.20

8.61

175.73

uc1jp;1 : uc1jp;1 ,

uc1jp;2 :

uc1jp;3 :

uc1jp;6 :

primary RL, secondary RL, tertiary RL, smelting uc1jp;2 , uc1jp;3 : Unit cost in each RL stage derived from Sugimura et al. (2014) multiplied by the ratio of the bulk density of copper scrap (1.13 t/m3) to that of WEEE (1.00 t/m3). It is assumed that the secondary, tertiary, and quaternary RL of WEEE from Sugimura et al. (2014) correspond to the primary, secondary, and tertiary RL of copper scrap, respectively uc1jp;6 : Smelting cost per unit copper weight determined from the weighted average of TCRC + Shipment in Chile, Peru, Australia, United States of America, Canada, and Papua New Guinea (from S&P Capital IQ; https://www.capitaliq. spglobal.com. Last accessed February 23, 2022) based on Japan’s copper concentrate imports and copper production from copper concentrates

Table 8.14 Unit cost of CS imported from foreign countries (unit: yen/kg-copper scrap) Case 1 (2019)

p1i

Australia

uc1i;1

uc1i;9

uc1i;10 uc1i;10

643.66

5.65

175.73

India

610.08

5.65

175.73

Indonesia

577.06

5.65

175.73

Malaysia

615.89

5.65

175.73

New Zealand

569.00

5.65

175.73

Pakistan

588.68

5.65

175.73

Philippines

593.94

5.65

175.73

Singapore

631.33

5.65

175.73

Thailand

588.54

5.65

175.73

Vietnam

429.61

5.65

175.73

Case 1 (2030)

p1i

Australia

647.88

6.22

175.73

India

611.38

6.22

175.73

Indonesia

578.93

6.22

175.73

Malaysia

619.45

6.22

175.73

New Zealand

572.17

6.22

175.73 (continued)

uc1i;1

uc1i;2

uc1i;2

uc1i;3

uc1i;3

uc1i;6

uc1i;6

uc1i;7

uc1i;7

uc1i;8

uc1i;8

uc1i;9

uc1i;10

116

8

Role of Ports in the Trade-Off Problem …

Table 8.14 (continued) Pakistan

591.62

6.22

175.73

Philippines

594.19

6.22

175.73

Singapore

634.40

6.22

175.73

Thailand

590.44

6.22

175.73

Vietnam

433.31

Case 1 (2050)

p1i

Australia

659.81

8.61

175.73

India

616.39

8.61

175.73

Indonesia

585.33

8.61

175.73

Malaysia

631.40

8.61

175.73

New Zealand

581.14

8.61

175.73

Pakistan

600.74

8.61

175.73

Philippines

594.97

8.61

175.73

Singapore

651.95

8.61

175.73

Thailand

599.43

8.61

175.73

Vietnam

450.03

8.61

175.73

6.22 uc1i;1

uc1i;2

uc1i;3

uc1i;6

uc1i;7

uc1i;8

uc1i;9

175.73 uc1i;10

p1i : purchase price of CS, uc1i;1 : primary RL, uc1i;2 : secondary RL, uc1i;2 : tertiary RL, uc1i;6 : port loading, uc1i;7 : maritime transport, uc1i;8 : port unloading, uc1i;9 : RL in Japan, uc1i;10 : smelting

Assuming that the current cost, insurance, and freight (CIF) price are for container transportation in Case 1 whereas those in Case 2 are for bulk transportation, the portion corresponding to the CIF price in Case 2 is obtained by considering the difference in cargo handling and maritime transportation in bulk transportation, as follows (Unit: yen/kg-copper scrap).

Case 2

2019

2030

2050

Australia

648.48

652.53

664.96

India

611.63

612.09

617.26

Indonesia

577.91

579.48

586.31

Malaysia

617.71

621.05

633.66

New Zealand

572.14

575.10

584.43

Pakistan

590.96

593.14

602.60

Philippines

596.08

595.96

596.91

Singapore

633.42

636.24

654.71

Thailand

590.54

592.15

601.67

Vietnam

431.57

435.06

452.66

p1i , uc1i;1 , uc1i;2 , uc1i;2 , uc1i;6 , uc1i;7 , uc1i;8 : Calculate the import value/volume from each country to Japan for HS Code 7404 from the Uncomtrade 2019 data; the sum of these is the CIF price. Values for 2030 and 2050 are obtained by adding the difference between the sum of uc1i;1 , uc1i;2 , uc1i;2 , uc1i;6 , uc1i;7 , and uc1i;8 calculated individually (as shown in Table 8.15) and that of 2019 to the 2019 value uc1i;9 and uc1i;10 : Same as uc1jp;3 and uc1jp;6

Appendix: Parameters Used for the Calculations

117

Table 8.15 Unit cost of CS imported from foreign countries (unit: yen/kg-copper scrap) Case 1 (2019)

uc1i;1

uc1i;2

uc1i;3

uc1i;6

uc1i;7

uc1i;8

Case 2 (2019)

uc1i;6

uc1i;7

uc1i;8

Australia

28.69

25.34

17.99

0.82

0.94

0.50

Australia

2.46

3.12

1.50

India

5.01

6.82

4.25

0.019

1.89

0.50

India

0.06

2.40

1.50

Indonesia

3.56

4.31

2.77

0.041

2.01

0.50

Indonesia

0.12

1.77

1.50

Malaysia

2.60

1.84

1.42

0.10

1.15

0.50

Malaysia

0.29

1.77

1.50

13.77

7.95

6.69

0.61

2.21

0.50

New Zealand

1.84

3.12

1.50

2.05

2.12

1.43

0.043

1.20

0.50

Pakistan

0.13

2.40

1.50

Philippines

0.50

0.65

0.41

0.0038

0.65

0.50

Philippines

0.01

1.77

1.50

Singapore

13.19

7.30

6.26

0.60

1.89

0.50

Singapore

1.81

1.77

1.50

Thailand

2.49

2.01

1.47

0.081

0.93

0.50

Thailand

0.24

1.77

1.50

Vietnam

1.92

1.47

1.10

0.067

0.95

0.50

Vietnam

0.20

1.77

1.50

New Zealand Pakistan

Case 1 (2030)

uc1i;1

uc1i;2

uc1i;3

uc1i;6

uc1i;7

uc1i;8

Case 2 (2030)

uc1i;6

uc1i;7

uc1i;8

Australia

30.70

26.46

18.94

0.91

0.94

0.56

Australia

2.74

2.66

1.67

India

5.61

7.15

4.53

0.046

1.89

0.56

India

0.14

1.40

1.67

Indonesia

4.43

4.80

3.19

0.081

2.01

0.56

Indonesia

0.24

1.29

1.67

Malaysia

4.28

2.78

2.22

0.17

1.15

0.56

Malaysia

0.52

1.29

1.67

15.27

8.79

7.40

0.68

2.21

0.56

New Zealand

2.04

2.66

1.67

3.45

2.89

2.09

0.107

1.20

0.56

Pakistan

0.32

1.40

1.67

Philippines

0.60

0.70

0.45

0.0081

0.65

0.56

Philippines

0.02

1.29

1.67

Singapore

14.64

8.10

6.95

0.67

1.89

0.56

Singapore

2.01

1.29

1.67

Thailand

3.38

2.50

1.90

0.122

0.93

0.56

Thailand

0.37

1.29

1.67

Vietnam

3.68

2.44

1.93

0.147

0.95

0.56

Vietnam

0.44

1.29

1.67

New Zealand Pakistan

Case 1 (2050)

uc1i;1

uc1i;2

uc1i;3

uc1i;6

uc1i;7

uc1i;8

Case 2 (2050)

uc1i;6

uc1i;7

uc1i;8

Australia

36.34

29.58

21.62

1.17

0.94

0.79

Australia

3.51

2.18

2.36

India

7.92

8.43

5.62

0.15

1.89

0.79

India

0.45

0.89

2.36

Indonesia

7.41

6.44

4.60

0.22

2.01

0.79

Indonesia

0.65

0.98

2.36

Malaysia

9.93

5.90

4.90

0.43

1.15

0.79

Malaysia

1.30

0.98

2.36

19.49

11.12

9.40

0.87

2.21

0.79

New Zealand

2.62

2.18

2.36

7.73

5.26

4.12

0.30

1.20

0.79

Pakistan

0.91

0.89

2.36

New Zealand Pakistan Philippines

0.86

0.85

0.58

0.02

0.65

0.79

Philippines

0.06

0.98

2.36

Singapore

22.99

12.73

10.91

1.05

1.89

0.79

Singapore

3.15

0.98

2.36

Thailand

7.60

4.84

3.90

0.31

0.93

0.79

Thailand

0.94

0.98

2.36

Vietnam

11.63

6.84

5.71

0.51

0.95

0.79

Vietnam

1.53

0.98

2.36

uc2i;1 , uc2i;2 , uc2i;3 , uc2i;5 : As the cost unit from Sugimura et al. (2014) can be divided into fuel and labor costs, the cost unit for the country is uc2jp;1 , uc2jp;2 , uc2jp;3 , and uc2jp;5 , adjusting the fuel cost by the square root of the ratio of the area of Japan to the area of the country and the labor cost by the ratio of the minimum wage in Japan to that in the country uc1i;6 and uc1i;8 : uc1i;8 is set based on the actual situation in Japan (500 yen/t for containers in Case 1, 1,500 yen/t for bulk in Case 2). uc1i;6 is uc1i;8 multiplied by the ratio of the minimum wage in the country to that in Japan uc1i;7 : Fuel, capital, and operation costs are calculated based on Drewry Maritime Research (2011)

118

Role of Ports in the Trade-Off Problem …

8

Table 8.16 Unit cost of WEEE collected in Japan (unit: yen/kg-WEEE) Case 2

uc2jp;1

uc2jp;2

uc2jp;3

uc2jp;4

uc2jp;5

uc2jp;10

2019

77.10

17.30

2.10

57.34

0.94

345.93

2030

85.42

19.15

2.26

63.66

1.03

350.30

2050

120.22

26.91

2.92

90.07

1.43

318.61

uc1jp;1 : primary RL, uc1jp;2 : secondary RL, uc1jp;3 : tertiary RL, uc1jp;6 : smelting uc2jp;1 , uc2jp;2 , uc2jp;3 , and uc2jp;5 : Unit cost in each RL stage derived from Sugimura et al. (2014) uc2jp;4 : Unit cost of processing from Sugimura et al. (2014) uc2jp;10 : Value of metal sold, including by-products, minus profit (using an average profit margin of 3.6% over the past five years for the four major Japanese copper smelting companies) and costs other than smelting, divided by the WEEE weight

Table 8.17 Unit cost of WEEE imported from foreign countries (unit: yen/kg-WEEE) Case 2 (2019)

uc2i;1

uc2i;2

uc2i;3

Australia

125.42

37.13

15.99

4.38

4.70

5.89

India

uc2i;4

uc2i;5

uc2i;6

uc2i;7

uc2i;8

uc2i;9

uc2i;10

94.02

2.83

0.41

0.52

0.25

0.94

345.93

2.15

0.63

0.01

0.40

0.25

0.94

345.93

Indonesia

7.72

3.73

3.50

4.68

0.42

0.02

0.30

0.25

0.94

345.93

Malaysia

16.30

3.70

0.86

11.20

0.23

0.05

0.30

0.25

0.94

345.93

New Zealand

94.11

20.96

2.23

70.26

1.11

0.31

0.52

0.25

0.94

345.93

8.03

2.43

1.54

4.92

0.22

0.02

0.40

0.25

0.94

345.93

Pakistan Philippines

2.13

0.50

0.54

0.44

0.06

0.00

0.30

0.25

0.94

345.93

Singapore

92.68

20.30

1.73

69.17

1.05

0.30

0.30

0.25

0.94

345.93

Thailand

13.79

3.36

1.14

9.29

0.23

0.04

0.30

0.25

0.94

345.93

Vietnam

11.62

2.65

0.78

7.65

0.18

0.03

0.30

0.25

0.94

345.93

Case 2 (2030)

uc2i;1

uc2i;2

uc2i;3

uc2i;4

uc2i;5

uc2i;6

uc2i;7

uc2i;8

uc2i;9

uc2i;10

Australia

139.31

40.22

16.26

104.56

2.99

0.46

0.45

0.28

1.03

350.30

8.54

5.63

5.97

5.30

0.68

0.02

0.24

0.28

1.03

350.30

Indonesia

13.76

5.08

3.61

9.27

0.49

0.04

0.22

0.28

1.03

350.30

Malaysia

27.96

6.29

1.08

20.04

0.36

0.09

0.22

0.28

1.03

350.30

104.49

23.27

2.43

78.13

1.23

0.34

0.45

0.28

1.03

350.30

17.66

4.58

1.72

12.23

0.33

0.05

0.24

0.28

1.03

350.30

India

New Zealand Pakistan Philippines Singapore

2.77

0.64

0.56

0.93

0.07

0.00

0.22

0.28

1.03

350.30

102.71

22.54

1.92

76.79

1.16

0.34

0.22

0.28

1.03

350.30

Thailand

19.94

4.73

1.26

13.96

0.31

0.06

0.22

0.28

1.03

350.30

Vietnam

23.78

5.36

1.01

16.87

0.32

0.07

0.22

0.28

1.03

350.30

Case 2 (2050)

uc2i;1

uc2i;2

uc2i;3

uc2i;4

uc2i;5

uc2i;6

uc2i;7

uc2i;8

uc2i;9

uc2i;10

Australia

178.30

48.91

16.99

134.15

3.44

0.59

0.36

0.39

1.43

318.61

India

24.45

9.17

6.27

17.38

0.86

0.08

0.15

0.39

1.43

318.61

Indonesia

34.34

9.66

4.00

24.89

0.72

0.11

0.16

0.39

1.43

318.61

Malaysia

67.01

14.99

1.82

49.69

0.81

0.22

0.16

0.39

1.43

318.61

133.62

29.76

2.98

100.25

1.57

0.44

0.36

0.39

1.43

New Zealand

318.61 (continued)

Appendix: Parameters Used for the Calculations

119

Table 8.17 (continued) Pakistan Philippines

47.27

11.17

2.28

34.70

0.67

0.15

0.15

0.39

1.43

318.61

4.59

1.05

0.59

2.31

0.09

0.01

0.16

0.39

1.43

318.61

Singapore

160.45

35.40

3.01

120.61

1.83

0.53

0.16

0.39

1.43

318.61

Thailand

49.12

11.23

1.81

36.10

0.64

0.16

0.16

0.39

1.43

318.61

Vietnam

78.73

17.61

2.05

58.58

0.95

0.26

0.16

0.39

1.43

318.61

uc1i;1 , uc1i;2 , and uc1i;3 : As the cost unit from Sugimura et al. (2014) can be divided into fuel and labor costs, the cost unit for the country is uc1jp;1 , uc1jp;2 , and uc1jp;3 , adjusting the fuel cost by the square root of the ratio of the area of Japan to the area of the country and the labor cost by the ratio of the minimum wage in Japan to that in the country uc2i;4 : uc1jp;4 multiplied by the ratio of the minimum wage in the country to that in Japan uc2i;6 and uc2i;8 : Same as uc1i;6 and uc1i;8 uc2i;7 , uc2i;9 , and uc2jp;10 : Same as uc1i;7 , uc2jp;5 , and uc2jp;10 Table 8.18 Unit cost of copper ore imported to Japan (unit: kg-CO2/t-copper ore) pj

ucj;1

ucj;2

ucj;3 ucj;3

Chile

158.88

47.08

Peru

158.49

63.25

Australia

371.15

113.74

United States of America

146.62

35.37

Canada

177.61

63.78

Papua New Guinea

226.55

64.69

efj;1 : Mining and mineral beneficiation, efj;2 : transportation, efj;3 : smelting pj , ucj;1 , and ucj;2 : Calculate the import value/volume from each country to Japan for HS Code 2603 from the Uncomtrade 2019 data; the sum of these is the CIF price ucj;3 : TCRC + Shipment from S&P Capital IQ; https://www.capitaliq.spglobal.com/

Table 8.19 Metals in the substrates of WEEE Metal

Content (kg/t-Substrate)

Recovery rate

Metal recovery (kg)

Cu

157.10

0.5

78.55

845

Value (yen/kg)

Zn

6.92

0.6

4.15

297

Pd

0.11

0.8

0.08

8,305,502

Ag

1.31

0.9

1.18

92,815

4.40

Sb

2.04

0.6

1.23

687

0.03

6,650,605

Ratio of value (%) 2.67 0.05 26.51

Au

0.27

0.9

0.25

Pb

12.68

0.9

11.41

215

66.24 0.10

Bi

0.13

0.6

0.08

700

0.002

Content and recovery rate: Sugimura et al. (2014) Value Data of 2020/1/1 from S&P Capital IQ (https://www.capitaliq.spglobal.com/) except for Bi (from Arum Publishing Co. 2020)

120

8

Role of Ports in the Trade-Off Problem …

References Arum Publishing Co. (2020) Industrial rare metals 2020 Bureau of International Recycling (2008) Report on the environmental benefits of recycling. Retrieved from https:// www.mgg-recycling.com/wp-content/uploads/2013/06/BIR_CO2_report.pdf Central Environment Council (2022) Results of the second inspection of the progress of the fourth basic plan for establishing a sound material-cycle society. Retrieved from https://www.env.go.jp/content/000067325.pdf Dong D, van Oers L, Tukker A, van der Voet E (2020) Assessing the future environmental impacts of copper production in China: implications of the energy transition. J Cleaner Prod 274:122825 Drewry Maritime Research (2011) Ship operating costs annual review and forecast 2011/12. Drewry Maritime Research, London, UK Futami T (2015) Global contribution for climate change and CO2 emissions from copper, zine ore mining. Energy Resour 36:39–44 Heiho A, Kanematsu Y, Nagase W, Murakami S, Tokoro C, Kikuchi Y (2019) Life cycle assessment of resource recovery from waste electrical and electronic equipment: a case study of tantalum recovery by chain-using drumtyped impact mill. Kagaku Kogaku Ronbunshu 45:244–252 IEA (2021) Net zero by 2050—a roadmap for the global energy sector. Retrieved from https://www.iea.org/reports/netzero-by-2050 Ilgin MA, Gupta SM (2010) Environmentally conscious manufacturing and product recovery (ECMPRO): a review of the state of the art. J Environ Manage 91:563–591 IPCC (2022) Climate change 2022: mitigation of climate change J-Credit System Office (2022) J-credit system. Retrieved from https://japancredit.go.jp/data/pdf/credit_001.pdf JOGMEC (2020) Mineral resources material flow 2019 1. Copper (Cu). Retrieved from https://mric.jogmec.go.jp/wpcontent/uploads/2020/05/material_flow2019_Cu.pdf Ministry ofthe Environment (2020) List of calculation methods and emission factors for calculation, reporting, and publication systems. Retrieved from https://ghg-santeikohyo.env.go.jp/calc. Accessed on 5 Sep 2022 Ministry of the Environment (2021) List of companies involved in TCFD, SBT, and RE100 (as of December 31, 2021). Retrieved from http://www.env.go.jp/earth/datsutansokeiei.html Ministry of Economy, Trade and Industry (2021) Carbon pricing for growth—credit transactions, etc. Retrieved from https://www.meti.go.jp/shingikai/energy_environment/carbon_neutral_jitsugen/pdf/003_ 04_00.pdf Ministry of Economy, Trade and Industry, Ministry of the Environment (2021) Basic policy on the promotion of the recycling of small waste electrical and electronic equipment. Retrieved from https://www.env.go.jp/content/ 900536161.pdf Poulsen RT, Ponte S, Sornn-Friese H (2018) Environmental upgrading in global value chains: the potential and limitations of ports in the greening of maritime transport. Geoforum 89:83–95 Reddy KN, Kumar A, Sarkis J, Tiwari MK (2020) Effect of carbon tax on reverse logistics network design. Comput Ind Eng 139:106184 Sapkota Y, White JR (2020) Carbon offset market methodologies applicable for coastal wetland restoration and conservation in the United States: a review. Sci Total Environ 701:134497 Seck GS, Hache E, Bonnet C, Simoën M, Carcanague S (2020) Copper at the crossroads: assessment of the interactions between low-carbon energy transition and supply limitations. Resour Conserv Recycl 163:105072 Suer J, Traverso M, Ahrenhold F (2021) Carbon footprint of scenarios towards climate-neutral steel according to ISO 14067. J Cleaner Prod 318:128588 Sugimura Y, Murakami S, Ukai T (2014) Significance of cost-effectiveness analysis in establishing recycling system for small home appliances. J Jpn Soc Civ Eng Ser G Environ Res 70:II_183–II_194 Sugimura Y, Murakami S (2016) Problems in Japan’s governance system related to end-of-life electrical and electronic equipment trade. Resour Conserv Recycl 112:93–106 Sugimura Y, Murakami S (2021) Designing a resilient international reverse logistics network for material cycles: a Japanese case study. Resour Conserv Recycl 170:105603 Watari T, Nansai K, Nakajima K (2020) Review of critical metal dynamics to 2050 for 48 elements. Resour Conserv Recycl 155:104669 Watari T, Northey S, Giurco D, Hata S, Yokoi R, Nansai K, Nakajima K (2022) Global copper cycles and greenhouse gas emissions in a 1.5°C world. Resour Conserv Recycl 179:106118 Zhang W, Li Z, Dong S, Qian P, Ye S, Hu S, Xia B, Wang C (2021) Analyzing the environmental impact of copperbased mixed waste recycling-a LCA case study in China. J Cleaner Prod 284:125256 Zhao X, Ma X, Chen B, Shang Y, Song M (2022) Challenges toward carbon neutrality in China: strategies and countermeasures. Resour Conserv Recycl 176:105959

Conclusions

This book focused on climate change countermeasures at ports. As hubs for logistics and passengers, ports consume a large amount of energy; therefore, CO2 emissions are high in ports, which includes the consumption of purchased external electricity. For a society that is taking large strides toward decarbonization, climate change action in ports is receiving increasing attention, and climate change countermeasures are becoming an important component of port operations. Because CO2 emission reductions could be achieved through the global supply chain and logistics efforts, ports could play a role in optimizing the entire supply chain as nodes that connect maritime and inland transportation. Therefore, ports need to be able to apply strategies to reduce their CO2 emissions and implement countermeasures to reduce emissions from maritime and inland transportation. Many researchers have studied climate change countermeasures in ports, and owing to their efforts, climate change countermeasures in ports, including those for ships and vehicles entering and departing from ports, have been implemented to a certain extent. While some ports in the world have emerged as leaders in climate change actions, a widely adopted countermeasure in the port industry remains lacking. Moreover, the adopted countermeasures are not aligned with the long-term efforts required to achieve emission reduction targets. One of the reasons for this is that there is little research based on empirical findings in ports; this the gap between research and practical application hinders decision-making in ports. Thus, empirical studies and development of potential countermeasures other than those organized thus far have become a necessity. Based on these considerations, this book is structured as a whole, with an emphasis on empirical aspects, economic feasibility, and new countermeasures. Part 1 confirmed the sources of port-related CO2 emissions, and potential climate change countermeasures that could be adopted by ports were summarized based on practical experience and previous studies. It also discusses the relationship between port governance and climate change countermeasures and the roles and responsibilities of stakeholders considering this because the port governance of a country has a direct impact on the level of implementation of the countermeasures. Inland transportation, vessels, and port activities, in that order, were confirmed to be the largest contributors to CO2 emissions. The potential climate change countermeasures that could be adopted by ports were summarized by dividing them into those for port activities, vessels, and vehicles, and the status of application of the countermeasures in Japan was also presented. Although port reform has been promoted in Japan, port governance has several characteristics compared to global trends, such as management by local governments, a unique POC system, and the absence of GTOs. These characteristics of Japan’s port governance model make it difficult for climate change countermeasures to be actively promoted at individual ports and CTs, and it has become clear that central government initiatives are even more important than in other countries.

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 Y. Sugimura, Climate Change Countermeasures in Ports Toward Carbon Neutrality, Sustainable Development Goals Series, https://doi.org/10.1007/978-3-031-34394-0

121

122

Conclusions

Part 2 introduces the empirical CO2 emission reduction effects of the countermeasures of cargohandling machinery using Hakata Port as a case study. An economic feasibility analysis of the implementation of these cargo-handling machineries at Hakata Port was also conducted. At Hakata Port, climate change countermeasures have been implemented since 2010, and CO2 emissions have decreased while container handling volume has increased significantly. Climate change countermeasures were implemented at Hakata Port under government-sponsored pilot projects. However, without the subsidies provided by the government such countermeasures cannot be implemented on a business basis even with standard port governance models. PMBs and POCs have an important role in promoting countermeasures and balancing the implementation with port competitiveness. Part 3 focused on energy consumption reduction and decarbonization measures in RCs, and presented a simulation method to estimate the amount of energy required for cooling. The method could predict the surface temperature of container walls using CFD to calculate the benefits of installing RSs in RC areas. The proposed simulation model was used to calculate the actual energysaving effects at Hakata Port in Japan and the economic feasibility of installing RSs using the amount of energy saved was also examined. The results revealed that introducing the system from a business perspective is challenging because the installation cost cannot be recovered during the operational period based on the money saved by the reduction in power consumption alone. Because no other practical countermeasures have been sufficiently demonstrated academically, the installation of an RS is one of the few RC-related energy-saving countermeasures worth promoting, even though the energy savings are limited. In the future, it will be necessary to support technological development to maximize energy-saving effects and reduce initial costs through effective structures and materials, and ensure financial support as a policy when economic feasibility cannot be guaranteed even with such support. The major contribution of the proposed simulation model and the analysis of energy-saving effects is that it provides a powerful tool for the preliminary evaluation of roof shade installation at ports. Part 4 presents two examples of new potential climate change countermeasures in ports to address the gap between the current adoption of climate change countermeasures in ports and the ambitious emission reduction targets expressed by international organizations, eventually achieving CNPs. One is the effective use of dredged soil generated by port development projects, and the other is the role of ports as an import base for recyclable resources. Analysis revealed that it is important to accelerate transition to next-generation energy sources to achieve CNPs. To this end, necessary to establish an import base for next-generation energy, and the effective storage of carbon contained in the dredged soil, which is expected to be generated in large quantities and utilized as a BCE foundation material, appears to have significant implications. Additionally, for societies and countries aiming for carbon neutrality, these countermeasures are important for offsetting residual emissions. An important development in climate change countermeasures is the increasing use of low-carbon technologies. However, this has lead to an increase in the demand for copper and, consequently, an increase in copper recycling. Recovery of copper from copper scrap and WEEE substrates can lead to higher costs and re-covering copper from WEEE substrates can even lead to higher carbon emissions. Additionally, the use of SRMs is hindered by RL-associated costs making carbon credits a more economically viable alternative. However, ports can play a key role in making SRMs more accessible and cost-effective. Port facilities must be developed to facilitate imports of SRMs, for example, deeper quays to accommodate large vessels that can minimize marine transportation costs. The development of such a port as an import hub for recyclable resources could be seen as a new climate change countermeasure in ports, which would solve the trade-off problem between CE and climate change action.

Conclusions

123

The information presented in this book represents only a small part of the climate change countermeasures in ports. There are many other countermeasures, each of which has its challenges. It is important for authorities and researchers to play their respective roles in improving the feasibility of countermeasures and developing new ones. It is my hope that this book contributes to the promotion of climate change countermeasures in ports, and ultimately to the realization of a CN society.